content
stringlengths
1
15.9M
\section{Introduction} Close correlations between spontaneous fission half-lives and ground state masses of nuclei were noticed in 1955 by W. J. \'Swi{\c a}tecki \cite{Sw55}. He has proposed a simple formula, which has joined the observed fission lifetimes with the difference between experimental and liquid drop masses ($\delta M$). A regular dependence of $\log_{10}T^{\rm sf}_{1/2}$ obtained by him after adding an empirical correction proportional to $\delta M$ is presented in Fig.~\ref{fig1} taken from Ref.~\cite{Sw55} as function of the fissility parameter $Z^2/A$. The aim of the present paper is to check if this 58 years old \'Swi{\c a}tecki brilliant idea still works. A modern version of the liquid drop model derived in Ref.~\cite{PD03} and all up to now measured spontaneous fission half-lives \cite{nudat2} are taken in our analysis analysis. \begin{figure}[h!] \includegraphics[height=0.8\columnwidth,angle=270]{fig1.eps} \caption{Experimental (bottom part) and corrected by the ground-state shell effects (EVEN-EVEN, ODD-A and ODD-ODD) spontaneous fission half-lives of atomic nuclei as a function of fissility parameter (original figure from Ref. \cite{Sw55}.} \label{fig1} \end{figure} \section{The liquid drop model} The experimental ground-state masses of all nuclei with the proton number $Z\ge 8$ and neutron number $N\ge 8$ were calculated in Ref.~\cite{PD03} using the macroscopic-microscopic model. The macroscopic part of this Lublin-Strasbourg Drop (LSD) mass formula (in MeV units) was following: \begin{equation} \begin{array}{l} M_{\rm LSD}(Z, N, {\rm def})=\\[+0.5ex] ~~ 7.289034\cdot Z+8.071431\cdot N-0.00001433\cdot Z^{2.39}\\[+0.5ex] ~ -15.4920(1 - 1.8601 I^2) A\\[+0.5ex] ~ + 16.9707(1 - 2.2938 I^2) A^{2/3}\, B_{\rm surf}({\rm def})\\[+0.5ex] ~ + 3.8602(1 + 2.3764 I^2) A^{1/3}\,B_{\rm cur}({\rm def})\\[+0.5ex] ~ + 0.70978\,Z^2/A^{1/3}\,B_{\rm Coul}({\rm def}) - 0.9181\,Z^2/A\\[+0.5ex] ~ -10\exp({-4.2|I|})\,B_{\rm cong}({\rm def}) \,\,. \end{array} \label{LSD} \end{equation} while the ground state microscopic corrections $E_{\rm micro}$ were taken from the mass tables of M\"oller at al. \cite{MN95}. In Eq.~(\ref{LSD}) $A=Z+N$ denotes the mass number, $I=(N-Z)/A$ reduced isospin and $B_{\rm surf}$, $B_{\rm cur}$, $B_{\rm Coul}$ and $B_{\rm cong}$ are relative to the sphere: surface, curvature, Coulomb and congruence (see Ref.~\cite{MS97}) energies. The parameters in the first and the last row in Eq.~(\ref{LSD})are taken from Ref.~\cite{MN95}, while the rest 8 parameter were fitted to the data. It was shown in Refs. \cite{PD03,DP07,IP09} that the LSD model (\ref{LSD}), which parameters were fitted to the experimental ground-state masses only, is able to reproduce well the fission barrier heights of light, medium and heavy nuclei when the microscopic part of the ground-state binding energy is taken into account according to the topographical theorem of \'Swi{\c a}tecki \cite{MS96}. The liquid drop barrier height of actinides decreases almost linearly in function of $Z$ from 4.3 MeV for $Z=90$ to 0 for $Z\ge 103$. The fission barrier of finite heights appears in the super-heavy nuclei mostly due to the shell effects in the ground state. \section{Calculation details} Following Ref.~\cite{Sw55} we have subtracted from the logarithm of the spontaneous fission half-lives in years: \begin{equation} f(Z,N)=\log_{10} [T_{1/2}^{\rm sf}(Z,N)/y]+k\delta M(Z,N)\,\,. \label{fZN} \end{equation} the {\it experimental} value of ground-state microscopic energy $\delta M(Z,N)$ can be evaluated using the experimental and the liquid drop (\ref{LSD}) mass difference: \begin{equation} \delta M_{micr}^{\rm exp}(Z,N)=M_{\rm exp}(Z,N)-M_{\rm LSD}(Z,N,0) \label{dM} \end{equation} multiplied by an arbitrary factor $k$. In our analysis of function $f(Z,N)$ were evaluated for 62 even-even fissioning isotopes with $Z\le 114$. Smooth dependence of $f(Z,N)$ on $Z$, in which shell effects were almost wash out, was achieved for $k=7.5$ MeV$^{-1}$ as one can see in Fig~\ref{fig2}. \begin{figure} \includegraphics[height=\columnwidth,angle=270]{fig2.eps} \caption{ Logarithms of the observed spontaneous fission half-lives corrected with masses "shifts" as a function of proton number.} \label{fig2} \end{figure} For given element the dependence of the function $f(Z,N)$ on neutron number is very week, so we omit this dependence in further consideration. The function $f(Z)$ which represents the global trend in the fission life-times can be approximated by two crossing straight lines: \begin{equation} f(Z)=\left\{ \begin{array}{ll} -4\cdot Z +371.2 ~~~&{\rm for~~} Z < 104\,\,,\\ -42.8 &{\rm for~~} Z\ge 104\,\,. \end{array}\right. \label{fZ} \end{equation} The coefficients of these straight lines are simply given by the systematics of points representing the corrected (according to Eq. (\ref{fZN})) fission life-times. They are not free adjustable parameters. The above dependence is very similar to the dependence of the liquid-drop barrier heights in function of $Z$. It is also visible in Fig.~\ref{fig2} that the slope of similar data for odd-$A$ and odd-odd nuclei is almost the same but the corresponding curves are shifted by a constant. This shift, which can be called global hindrance factor is equal to $h=10.3$ for odd-$A$ nuclei and 17.8 for odd-odd systems. It represents the fission life-time increase given both by the odd-even effects in the ground state energy and the fission barrier enhancement due to the angular momentum and parity conservation of unpaired nucleons. Contrary to \'Swi{\c a}tecki analysis \cite{Sw55} we assume here the linear in $Z$ form of the $f(Z)$ and we have extrapolated it by a constant for the super-heavy nuclei. Using the approximation (\ref{fZ}) one can estimate the spontaneous fission half-lives by the following formula: \begin{equation}\label{Tsf} \begin{array}{ll} \log_{10}\left(\frac{T_{1/2}^{\rm sf}(Z,N)}{y}\right)& =-4 Z+371.2-7.5\delta M(Z,N)\\ &+\left\{ \begin{array}{cl} 0 &{\rm ~for~even-even \,\,,}\\ 10.3 &{\rm ~for~odd-}A\,\,,\\ 17.8&~ {\rm for~odd-odd\,\,,} \end{array}\right. \end{array} \end{equation} where the experimental microscopic correction $\delta M$ is defined in Eq.~(\ref{dM}). The estimates of $T_{1/2}^{\rm sf}$ obtained with formula (\ref{Tsf}) are compared in Fig.~\ref{fig3} with the experimental data taken from Ref.~\cite{nudat2}. Surprisingly good agreement of the model estimates with the data is achieved for all actinide and even super-heavy nuclei. The root mean square deviation of $\log T_{1/2}^{\rm sf}$ for the even-even isotopes with $Z< 104$ is 1.52 and it grows to 2.02 when odd-$A$ and odd-odd nuclei are taken into account. The rms deviation reaches 2.47 when one includes to the analysis the super-heavy isotopes (for which the liquid drop barrier vanishes). To foresee $T_{1/2}^{\rm sf}$ for unknown nuclei one can evaluated $\delta M$ using mass estimates from one of existing on the market mass tables. In order to understand this striking for the first sight result it is good to remind the topographical theorem, proposed by Myers and \'Swi{}\c{a}tecki \cite{MS96}: the mass of a nucleus in a saddle point is determined by the macroscopic part of the binding energy. The shell effects at the saddle are negligible, so the fission barrier heights is approximately determined by the difference of the macroscopic (here LSD) and the experimental masses: \begin{equation} V_{\rm B}(Z,N) = M_{\rm LSD}(Z,N,{\rm saddle}) - M_{\rm exp}^{g.s.}. \end{equation} The function $F(Z)$ (\ref{fZ}) represents the liquid-drop estimate of the fission life-time and the $k$ parameter in Eq.~(\ref{Tsf}) gives enhancement of the life-time due to increase of the fission barrier due to the ground state microscopic effects. This parameter $k$ plays a similar role as the hindrance factor $h$ which origins from the barrier augmentation due to spin and parity conservation of an odd-$A$ nucleus along the path to fission. \begin{figure}[bh!] \includegraphics[height=\columnwidth,angle=270]{fig3a.eps}\\[-4ex] \includegraphics[height=\columnwidth,angle=270]{fig3b.eps} \caption{Spontaneous fission half-lives of even-even (top) and odd (bottom) nuclei, calculated using formula (\ref{Tsf}) in comparison to the experimental values \cite{nudat2}. The open symbols correspond to the estimates evaluated on the basis of extrapolated experimental masses in Ref. \cite{nudat2}.} \label{fig3} \end{figure} \section{Summary} The following conclusions can be drawn from our investigation:\\[-5ex] \begin{itemize} \item Simple phenomenological formula for the spontaneous fission half-lives depending on proton number and microscopic energy correction in the ground state was found.\\[-5ex] \item Model estimates of the spontaneous fission half-lives of even-even nuclei are in a surprisingly good agreement with experimental data.\\[-5ex] \item Quality of evaluation for odd nuclei is much worse, what is due to fact that the effect of an odd-particle on the barrier penetrability is described here by a single constant, independent on angular momentum or parity of the odd nucleon.\\[-5ex] \item Our simple formula with the microscopic energy corrections taken e.g. from the tables \cite{MN95} can serve to rough estimates of the fission life-times of new undiscovered yet isotopes. \end{itemize} The present phenomenological formula for the spontaneous fission half-lives and similar formula for the alpha decay half-lives derived in Ref.~\cite{ZW13} can serve a useful tool to estimate stability of non discovered yet isotopes. The only input which one needs in the both formulas are the binding energy estimates which can be taken from one of the mass model existing on the market. \section*{Acknowledgements} This work was partly supported by the Polish National Science Center Grant No.~DEC-2011/01/B/ST2/03667.
\section{Evaluation Approach} \label{sec:evaluation} \vspace{-0.1in} We provide details of the three key parts given in algorithm ${\mathcal A}$: (1) Model construction and sample generation (steps A and B1) (section 3.1), (2) Score computation using metrics (step B3) (section 3.2) and (3) Ranking of algorithms with statistical significance testing (step B4 and C) (section 3.3). We make the following assumptions. (a) There is a sufficient number of examples with no missing values for learning a good true model; this is true in many practical scenarios. (b) A conditional probabilistic model powerful enough to capture the underlying distribution can be built using these examples. (c) Generating samples from the model distribution can be done easily via Gibbs sampling. \\ \noindent{\bf Notations} Let ${\mathcal D}_N$ and ${\mathcal D}_M$ denote the sets of examples with no missing values and with missing values respectively. Let ${m}_j \subset \{1,\cdots,d\}$, $j=1,\cdots,J$ denote a subset of variables for which values are missing in ${\mathcal D}_M$ with $d$ denoting the number of variables. Note that each subset $m_j$ refers to a unique missing pattern and there are $J$ missing patterns in ${\mathcal D}_M$. Also, let $o_j = \{1,\cdots,d\} \setminus m_j$ be the observed variables. Let $Y^{(i)}_{m}$ and $Y^{(i)}_{o}$ denote the sets of missing and observed variables in the $i^{th}$ example respectively. \subsection{Distribution Model Learning} \vspace{-0.1in} The first step is to build a distribution model such that for each missing pattern $m_j$, we can generate samples given the values for the corresponding observed pattern $o_j$. One possibility is to build a joint model $P(Y;\theta)$ using the set ${\mathcal D}_N$ (where $\theta$ is the model parameter). However, this may be hard and intractable; more importantly, it is not necessary. This is because we always have some observed variables that can be used to predict the missing variables. Furthermore, conditional models are relatively easier to build. Due to these reasons, for each $j$ we build a conditional distribution model $P(Y_{m_j}|Y_{o_j};\theta_j)$ for the missing pattern $m_j$ in ${\mathcal D}_M$ using the fully observed data set ${\mathcal D}_N$; here, $\theta_j$ denotes the $j^{th}$ model's parameter vector. \\ \noindent{\bf Model and Learning} While any good conditional model is sufficient for our method to work, we use the pairwise Markov random field (MRF) in this work. For ease of notation, let us remove the subscript from $m_j$. We use the model: $P(Y_m|Y_o;\theta) = \frac{1}{Z(Y_o)} \exp(s(Y_m;\theta_m) + s(Y_{m};Y_{o},\theta_{mo}))$ where $Z(Y_o) = \sum_{Y_m} \exp(s(Y_m;\theta_m) + s(Y_{m};Y_{o},\theta_{mo}))$ is the partition function and, $s(Y_m;\theta_m)$ and $s(Y_{m};Y_{o},\theta_{mo})$ denote parametrized scoring functions involving self and cross features among the missing and observed variables. In our experiments, we use a simple linear model for the scoring function: $s(Y_m;\theta_m) = \sum_{k \in m} \sum_{l: {\bar y}_{l} \in {\mathcal Y}_k} \theta_{k,l} {\mathcal I}(y_{k} = {\bar y}_{l})$ where ${\mathcal Y}_k$ is the label space for the $k^{th}$ variable and ${\mathcal I}(\cdot)$ is the indicator feature function. The scoring function $s(Y_{m};Y_{o},\theta_{mo})$ is defined using similar feature functions involving a pair of variables and associated parameters. Given a missing pattern $m$, we construct a training set comprising of input-target pairs $\{(Y^{(i)}_o,Y^{(i)}_m):i=1,...,n\}$ from ${\mathcal D}_N$ and learn the model parameter $\theta$ using the piecewise pseudo-likelihood maximization training approach suggested by Sutton and McCallum (2008). \noindent{\bf Samples from the distribution model:} Using the model parameters learned for each missing pattern $m_j$, we generate multiple imputations for each example in ${\mathcal D}_M$ (having the missing pattern $m_j$) via Gibbs sampling. For instance, sample generation for a discrete variable involves sampling a multinomial distribution with probabilities that can be easily computed from the conditional MRF distribution. \noindent{\bf Single Model} One disadvantage of the above modeling is that we need to build $J$ conditional models. However, these models can be built in parallel and is a one time effort. Alternatively, we can build a single conditional model to deal with all missing patterns. The key difference is that we learn a single model parameter vector $\theta$ (instead of learning one model parameter vector $\theta_j$ per missing pattern $m_j$). In this case, there is only one dataset. The log likelihood term for each example involves different missing patterns. Therefore, the subset of model parameters used for each example is different, depending on the missing and non-missing variables involved; clearly, parameter overlap happens across the examples. Building a single model helps in reducing the computational complexity. Though such a model is expected to be inferior in prediction quality compared to the per missing pattern model, we found it to be adequate to meet the requirement of ranking the methods correctly. \subsection{Metrics} \vspace{-0.1in} Given samples $(Y_m)$ from the model (\textit{true}) conditional distribution $(P)$ and multiple imputations $({\hat Y}_m)$ obtained from an imputation algorithm $k$ (with its distribution ${\hat P}_{k}$), we need a score/metric for measuring the the discrepancy between the samples $Y_m$ and ${\hat Y}_m)$. In this section we briefly describe the metrics (popular ones selected from the literature) used in our experiments. It is worth pointing out that the challenge here is the reliable estimation of metrics between two distributions using samples generated (independently) from them. Given its importance in applications, this problem has received much attention in the literature and several good methods have been proposed. Our work is the first one to make use of them for comparing missing value imputation algorithms. For ease of notation in the discussion below, we drop the subscript $k$ indicating the dependency on the imputation algorithm, from ${\hat P}_k$. Below we assume that $|Y_m| = |{\hat Y}_m|)$. For all metrics other than KL divergence estimation, this condition can be relaxed. \noindent{ \bf KL divergence via Convex Optimization:} Nguyen et al (2010) shows that the metric estimation problem can be posed as an empirical risk minimization problem and solved using standard convex programs. Estimating KL divergence is one instantiation of this problem; it is obtained using the solution of the following optimization problem: \begin{equation} - min_{\alpha > 0} F = - \frac{1}{L} \sum_{u=1}^L log(L \alpha_u) + \frac{1}{2\lambda_L L} g(\alpha,Y_m,{\hat Y}_m) \label{eq:KLD} \end{equation} where $\lambda_L$ is a regularization constant, $L$ is the number of imputation samples, $g(\alpha,Y_m,{\hat Y}_m) = \sum_{u,v} \alpha_u \alpha_v K(y^{(u)}_m,y^{(v)}_m) + \frac{1}{L} K({\hat y}^{(u)}_m,{\hat y}^{(v)}_m) - \sum_{u,v} \alpha_u \alpha_v K(y^{(u)}_m,{\hat y}^{(v)}_m)$ and $K(\cdot, \cdot)$ is a kernel function. We used the Gaussian kernel $K(y,{\hat y}) = \exp(-||y-{\hat y}||^2/\sigma)$ and set the kernel width parameter $\sigma$ to $1$ in our experiments. Finally, the KL divergence value is computed using the solution ${\hat \alpha}$ obtained from solving (\ref{eq:KLD}) as: $-\frac{1}{L}\sum_{u=1}^L log(L {\hat \alpha}_u)$. \noindent{\bf Symmetric KL divergence via binary classification} It is known that KL divergence is not symmetric with respect to the distributions used, and, it is often useful to consider Jensen-Shannon divergence having the symmetric property (defined as: $\frac{1}{2}(KL(P,\frac{P+{\hat P}}{2}+KL(P,\frac{P+{\hat P}}{2})$). While it is theoretically possible to adapt the estimation approach suggested by Nguyen et al (2012) to estimate the symmetric KL divergence, it does not result in a simpler convex program like (\ref{eq:KLD}). So we take a different route. Reid et al (2011) developed a unified framework to show that many machine learning problems can be viewed as binary experiments and estimation of various divergence measures can be done by solving a binary classification problem with suitably weighted loss functions for the two categories. In our context, the symmetric KL divergence can be estimated as follows: (a) assign class labels +1 and -1 for the samples from the model distribution and imputation algorithm respectively, and, (b) build a classifier using the standard logistic loss function. The intuition is that, more divergent the distributions are, easier is the classification task. Therefore, the error rate measures the discrepancy between the distributions. \noindent{\bf Maximum Mean Discrepancy} Gretton et al (2012) proposed a framework in which a statistical test can be conducted to compare samples from two distributions. The main idea is to compute a statistic known as maximum mean discrepancy (MMD) and use an asymptotic distribution of this statistic to conduct a hypothesis test ${\mathcal H}_0: P = {\hat P}$ with a suitably defined threshold. The MMD statistic is defined as the supremum over the mean difference between some function score computed from the samples of the respective distributions. Gretton et al (2012) shows that functions from reproducing kernel Hilbert space (RKHS) are rich for computing this statistic from finite samples and derives conditions under which MMD can be used to distinguish the distributions. Given the MMD function class ${\mathcal F}$ and the unit ball in a RKHS ${\mathcal H}$ with kernel $K(y,{\hat y})$ (e.g., Gaussian), the statistic is defined as: \begin{equation} {\hat MMD}^2_u({\mathcal F},Y_m,{\hat Y}_m) = g_{yy}(Y_m) + g_{{\hat y}{\hat y}}({\hat Y}_m) - g_{y{\hat y}}(Y_m,{\hat Y}_m) \label{eq:mmd} \end{equation} where $$g_{y{\hat y}}(Y_m,{\hat Y}_m) = \frac{1}{L(L-1)} \sum_{u=1}^L\sum_{v\ne u}^L K({y}^{(u)}_m,{\hat y}^{(v)}_m);$$ $g_{yy}(Y_m)$ and $g_{{\hat y}{\hat y}}({\hat Y}_m)$ are similarly defined. To reduce the computational complexity and produce low variance estimates in the finite sample issue associated with computing (\ref{eq:mmd}), Zaremba et al (2013) proposes a statistical test known as B-test. Based on this theoretical backing, we pick up two metrics: (a) the statistic MMD and (b) the B-test result, for our study on ranking imputation algorithms. The B-test gives $0$ or $1$ as the output (with $1$ indicating acceptance of ${\mathcal H}_0$); therefore, there will be ties. However, the rank aggregation algorithm that we describe in section~\ref{sec:RA} can handle this. \noindent{\bf Neighborhood Dissimilarity Score} One disadvantage associated with the metrics described above is that they are computationally expensive since we need to solve the associated problems for each example in ${\mathcal D}_M$. This becomes an issue especially when $J$ (the number of missing patterns) is large. We propose a neighborhood dissimilarity score (NDS) that is cheaper to compute and is defined as: \begin{equation} nds(h,Y_m,{\hat Y}_m) = \frac{1}{L^2}\sum_{u=1}^L \sum_{v=1}^L w_{uv} h(y^{(u)}_m,{\hat y}^{(v)}_m) \label{eq:nds} \end{equation} where $h(\cdot)$ is a distance score (e.g., Hamming loss) computed between a pair of examples on the imputed values and $w_{uv}$ is a weighting factor given by: $\lambda^{h(\cdot)} + (1-\lambda)^{(h_{max} - h(\cdot))}$; $\lambda$ is a scaling parameter; and, $h_{max}$ is the maximum Hamming distance possible. Note that $w_{uv}$ gives more weight to neighbors and $\lambda$ controls the rate at which the weight falls off in terms of distance. We used $\lambda=0.1$ in our experiments. The overall complexity is $O(JL^2)$, but $L$ is typically small (e.g., 25). NDS is closely related to (\ref{eq:mmd}), but with three key differences: (a) we consider only the cross terms, (b) there are no kernel parameters to learn in NDS, and (c) the weighting is not uniform ($w_{uv}$ compared to the uniform weighting $\frac{1}{L(L-1)}$ used in (\ref{eq:mmd})). Our experimental results show that NDS is a good metric in the sense of producing results similar to MMD and KL based metrics. \subsection{Rank Aggregation} \vspace{-0.1in} Let us take one metric ${\mathcal M}$. For each algorithm $k \in {\mathcal A}$ (a set of algorithms to compare), we compute the score $s^{(i)}_{k,{\mathcal M}}$ for the $i^{th}$ example in ${\mathcal D}_M$ using the procedures presented in the previous subsection. Using these scores, we rank the algorithms for each $i$. Collecting the ranks over all $i$ results in a rank matrix ${\mathcal R}_{\mathcal M}$ with each row representing an example and the columns representing the ranks of the algorithms in ${\mathcal A}$. To compare the algorithms, we draw analogy to the problem of statistically comparing multiple classifiers over multiple datasets (Demsar, 2006). In particular, we treat multiple imputations for each example as one dataset and we have scores for $|{\mathcal D}_M|$ datasets for a given algorithm (and the chosen metric). Our goal is to statistically compare the performance of algorithms in ${\mathcal A}$ on these datasets. Following Demsar (2006), we conduct the Friedman test with the null hypothesis that all algorithms are equivalent. This test involves computing the average ranks of each algorithm, followed by computing a statistic and checking against the ${\mathcal X}^2_F$ distribution for a given confidence level $\beta$ (we used 0.05). It turned out that the null hypothesis is always rejected for the algorithms that we considered. So, we conducted the Nemenyi test as the post-hoc test (Demsar, 2006). In this test, we compute the critical difference (CD) score, defined as: $q_{\beta} \sqrt{\frac{|{\mathcal A}|(|{\mathcal A}|+1)}{6|{\mathcal D}_M|}}$ where $q_{\beta}$ is obtained from a look-up table. The observed average rank difference between any two methods is statistically significant when it is greater than CD. We demonstrate this in our experiments later and make key observations. \label{sec:RA} \section{Evaluation Approach} In this section, we provide details of the three key parts given in algorithm ${\mathcal A}$: (1) Model construction and sample generation (steps A and B1) (section 3.1), (2) Score computation using metrics (step B3) (section 3.2) and (3) Ranking of algorithms with statistical significance testing (step B4 and C) (section 3.3). We make the following assumptions: (a) there are at least a reasonable number of examples with no missing values; this is true in many practical scenarios, (b) a conditional probabilistic model powerful enough to capture the underlying distribution can be built using these examples, and (c) generating samples from the model distribution can be done easily via Gibbs sampling. \\ \noindent{\bf Notations} Let ${\mathcal D}_N$ and ${\mathcal D}_M$ denote the sets of examples with no missing values and with missing values respectively. Let ${m}_j \subset \{1,\cdots,d\}$, $j=1,\cdots,J$ denote a subset of variables for which values are missing in ${\mathcal D}_M$ with $d$ denoting the number of variables. Note that each subset $m_j$ refers to a unique missing pattern and there are $J$ missing patterns in ${\mathcal D}_M$. Also, let $o_j = \{1,\cdots,d\} \setminus m_j$. Let $Y^{(i)}_{m}$ and $Y^{(i)}_{o}$ denote the missing and observed variables in the $i^{th}$ example respectively. \subsection{Distribution Model Learning} The first step is to build a distribution model such that for each missing pattern $m_j$, we can generate samples given the values for the corresponding observed pattern $o_j$. We can build a joint model $P(Y;\theta)$ using the set ${\mathcal D}_N$ (where $\theta$ is the model parameter). However, building such a joint model is hard and intractable; more importantly, it is not necessary. This is because we always have some observed variables that can be used to predict the missing variables. Furthermore, conditional models are relatively easier to build. Due to these reasons, we build conditional distribution model $P(Y_{m_j}|Y_{o_j};\theta_j)$ for each missing pattern $m_j$ in ${\mathcal D}_M, \forall j$ using the fully observed data set ${\mathcal D}_N$; $\theta_j$ denotes the $j^{th}$ model parameter vector. \\ \noindent{\bf Model and Learning} While any decent conditional model is sufficient for our method to work, we use pairwise Markov random field (MRF) for illustrative purpose. For ease of notation, we will remove the subscript in $m_j$. We use the model: $P(Y_m|Y_o;\theta) = \frac{1}{Z(Y_o)} \exp(s(Y_m;\theta_m) + s(Y_{m};Y_{o},\theta_{mo}))$ where $Z(Y_o) = \sum_{Y_m} \exp(s(Y_m;\theta_m) + s(Y_{m};Y_{o},\theta_{mo}))$ is the partition function and, $s(Y_m;\theta_m)$ and $s(Y_{m};Y_{o},\theta_{mo})$ denote parametrized scoring functions involving self and cross features among the missing and observed variables. In our experiments, we use a simple linear model for the scoring function: $s(Y_m;\theta_m) = \sum_{k \in m} \sum_{{\bar y}_{k,l} \in {\mathcal Y}_k} \theta_{k,l} {\mathcal I}(y_{k} = {\bar y}_{k,l})$ where ${\mathcal Y}_k$ is the label space for the $k^{th}$ variable and ${\mathcal I}(\cdot)$ is the indicator feature function. The scoring function $s(Y_{m};Y_{o},\theta_{mo})$ is defined using similar feature functions involving a pair of variables and associated parameters. Given a missing pattern $m$, we construct a training set comprising of input-target pairs $\{(Y^{(i)}_o,Y^{(i)}_m):i=1,...,n\}$ from ${\mathcal D}_N$ and learn the model parameter $\theta$ using the piecewise pseudo-likelihood maximization training approach suggested by Sutton and McCallum (2008). \noindent{\bf Multiple Imputations} Using the model parameters learned for each missing pattern $m_j$, we generate multiple imputations for each example in ${\mathcal D}_M$ (having the missing pattern $m_j$) via Gibbs sampling. For instance, sample generation for a discrete variable involves sampling a multinomial distribution with probabilities that can be easily computed from the conditional MRF distribution. \noindent{\bf Single Model} One disadvantage of the above modeling is that we need to build $J$ conditional models. However, these models can be built in parallel and is one time effort. On the other hand, we can build a single conditional model that can be used for any missing pattern. The key difference is that we learn a single model parameter vector $\theta$ (instead of learning one model parameter vector $\theta_j$ per missing pattern $m_j$). In this case, there is only one dataset and log likelihood term for each example involves different missing pattern. Therefore, the subset of linear model parameter $\theta$ described above could be different depending on the variables involved, and parameter overlapping happens across the examples. Building a single model helps in reducing the computational complexity. Though such a model is expected to be inferior in prediction quality compared to per missing pattern model, we found this to be adequate to meet the requirement of ranking the methods correctly. \subsection{Metrics} Given samples $(Y_m)$ from the model (\textit{true}) conditional distribution $(P)$ and multiple imputations $({\hat Y}_m)$ obtained from an imputation algorithm $k$ (with its distribution ${\hat P}_{k}$), we compute a score for samples $(Y_m,{\hat Y}_m)$ from each pair $(P,{\hat P}_k)$. In this section we describe several popular metrics (selected from the literature) that we use to compute the score and compare in our experiments. These metrics measure the discrepancy between samples generated from $(P,{\hat P})$. (For ease of notation, we drop the dependency on algorithm $k$ in ${\hat P}_k$ in the discussion below). \noindent{ \bf KL divergence via Convex Optimization} Estimation of divergence functionals and likelihood ratios using samples generated from two probability distributions $(P, \hat P)$ is an important problem in machine learning and information theory. Nguyen et al (2010) shows that this estimation problem can be posed as an empirical risk minimization problem and solved using standard convex programs. Computing KL divergence is an instantiation of this problem and we use their approach to compute the KL divergence between the multiple imputation samples obtained for each missing pattern. In particular, we solve the following optimization problem: \begin{equation} - min_{\alpha > 0} F = - \frac{1}{L} \sum_{u=1}^L log(L \alpha_u) + \frac{1}{2\lambda_L L} g(\alpha,Y_m,{\hat Y}_m) \label{eq:KLD} \end{equation} where $\lambda_L$ is a regularization constant, $L$ is the number of imputation samples, $g(\alpha,Y_m,{\hat Y}_m) = \sum_{u,v} \alpha_u \alpha_v K(y^{(u)}_m,y^{(v)}_m) + \frac{1}{L} K({\hat y}^{(u)}_m,{\hat y}^{(v)}_m) - \sum_{u,v} \alpha_v \alpha_j K(y^{(u)}_m,{\hat y}^{(v)}_m)$ and $K(\cdot, \cdot)$ is a kernel function. We used the Gaussian kernel $K(y,{\hat y}) = \exp(-||y-{\hat y}||^2/\sigma)$ and set the kernel width parameter $\sigma$ to $1$ in our experiments. Finally, the KL divergence value is computed using the solution ${\hat \alpha}$ obtained from solving (\ref{eq:KLD}) as: $-\frac{1}{L}\sum_{u=1}^L log(L {\hat \alpha}_u)$. \noindent{\bf Symmetric KL divergence via binary classification} It is known that the KL divergence is not symmetric with respect to the distributions used, and, it is often useful to consider Jensen-Shannon divergence having the symmetric property (defined as: $\frac{1}{2}(KL(P,\frac{P+{\hat P}}{2}+KL(P,\frac{P+{\hat P}}{2})$). While it is theoretically possible to adapt the estimation approach suggested by Nguyen et al (2012) to estimate the symmetric KL divergence, it does not result in a simpler convex program like (\ref{eq:KLD}). It is interesting to note that Reid et al (2011) developed a unified framework to show that many machine learning problems can be viewed as binary experiments and estimation of various divergence measures can be done by solving a binary classification problem with suitably weighted loss functions for the two categories. In our context, we found that the symmetric KL divergence can be estimated as follows: (a) assign class labels +1 and -1 for the samples from the model distribution and imputation algorithm respectively, and, (b) build a classifier using standard logistic loss function. The intuition is that more divergent the distributions easier the classification task. Therefore, the error rate measures the discrepancy between the distributions. \noindent{\bf Maximum Mean Discrepancy} Gretton et al (2012) proposed a framework in which a statistical test can be conducted to compare samples from two distributions. The main idea is to compute a statistic known as maximum mean discrepancy (MMD) and use an asymptotic distribution of this statistic to conduct a hypothesis test ${\mathcal H}_0: P = {\hat P}$ with a suitably defined threshold. The MMD statistic is defined as supremum over the mean difference between some function score computed from the samples of respective distributions. They show that functions from reproducing kernel Hilbert space (RKHS) are rich to compute this statistic from finite samples and derive conditions under which MMD can be used to distinguish the distributions. Given the MMD function class ${\mathcal F}$ the unit ball in a RKHS ${\mathcal H}$ with kernel $K(y,{\hat y})$ (e.g., Gaussian), the statistic is defined as: \begin{equation} {\hat MMD}^2_u({\mathcal F},Y_m,{\hat Y}_m) = g_{yy}(Y_m) + g_{{\hat y}{\hat y}}({\hat Y}_m) - g_{y{\hat y}}(Y_m,{\hat Y}_m) \label{eq:mmd} \end{equation} where $$g_{y{\hat y}}(Y_m,{\hat Y}_m) = \frac{1}{L(L-1)} \sum_{u=1}^L\sum_{v\ne u}^L K({y}^{(u)}_m,{\hat y}^{(v)}_m);$$ $g_{yy}(Y_m)$ and $g_{{\hat y}{\hat y}}({\hat Y}_m)$ are similarly defined. To reduce the computational complexity and produce low variance estimates in the finite sample issue associated with computing (\ref{eq:mmd}), Zaremba et al (2013) propose a statistical test known as B-test. With this theoretical backing, we include two metrics: (a) the statistic MMD and (b) the B-test result to rank the algorithms in our study. The B-test gives $0$ or $1$ as the output (with $1$ indicating acceptance of ${\mathcal H}_0$); therefore, there will be ties. However, the rank aggregation algorithm that we describe in section~\ref{sec:RA} can handle this. \noindent{\bf Neighborhood Dissimilarity Score} One disadvantage of the above metrics is that they are computationally expensive since we need to solve the associated problems for each example in ${\mathcal D}_M$. This becomes an issue when $J$ is large. We propose a neighborhood dissimilarity score (NDS) that is cheaper to compute and is defined as: \begin{equation} nds(h,Y_m,{\hat Y}_m) = \frac{1}{L^2}\sum_{u=1}^L \sum_{v=1}^L w_{uv} h(y^{(u)}_m,{\hat y}^{(v)}_m) \label{eq:nds} \end{equation} where $h(\cdot)$ is a distance score (e.g., Hamming loss) computed between a pair of examples on the imputed values and $w_{uv}$ is a weighting factor given by: $\lambda^{h(\cdot)} + (1-\lambda)^{(h_{max} - h(\cdot))}$; $\lambda$ and $h_{max}$ denote a scaling parameter and maximum Hamming distance possible. Note that $w_{uv}$ gives more weight to neighbors and $\lambda$ controls the rate at which the weight falls off in terms of distance. We used $\lambda=0.1$ in our experiments. Though it is quadratic in $L$, $L$ is typically small (e.g., 25) and the overall complexity is $O(JL^2)$. NDS is closely related to (\ref{eq:mmd}) with three key differences: (a) we consider only the cross terms, (b) there are no kernel parameters to learn in NDS, and (c) the weighting is not uniform ($w_{uv}$ compared to uniform weighting of $\frac{1}{L(L-1)}$ used in (\ref{eq:mmd})). Our experimental results show that NDS is a good metric in the sense of producing results similar to MMD and KL based metrics. \subsection{Rank Aggregation} Given a metric ${\mathcal M}$ and algorithm $k \in {\mathcal A}$ (a set of algorithms to compare), we compute the score $s^{(i)}_{k,{\mathcal M}}$ for $i^{th}$ example in ${\mathcal D}_M$ using procedures presented in the previous section. Using the scores, we rank the algorithms for each $i$ resulting in a rank matrix ${\mathcal R}_{\mathcal M}$ with each column representing the ranks of an algorithm in ${\mathcal A}$. To compare the algorithms, we draw analogy to the problem of statistically comparing multiple classifiers over multiple datasets (Demsar, 2006). In particular, we treat multiple imputations for each example as one dataset and we have scores for $|{\mathcal D}_M|$ datasets for a given algorithm and metric. Our goal is to statistically compare the performance of algorithms in ${\mathcal A}$ on these datasets. Following Demsar (2006), we conduct Friedman test with the null hypothesis that all algorithms are equivalent. This test involves computing the average ranks of each algorithm, and computing a statistic and checking against ${\mathcal X}^2_F$ distribution for a given confidence level $\beta$ (we used 0.05). It turned out that the null hypothesis is always rejected for the algorithms that we considered. So, we conducted Nemenyi test as the post-hoc test (Demsar, 2006). In this test, we compute critical difference (CD) score defined as: $q_{\beta} \sqrt{\frac{|{\mathcal A}|}{|{\mathcal D}_M|}}$ where $q_{\beta}$ is obtained from a look-up table. The observed average rank difference between any two methods is statistically significant when it is greater than CD. We demonstrate this in our experiments later and make key observations. \label{sec:RA} \section{Conclusion} We presented a framework to quantitatively evaluate and rank the goodness of imputation algorithms. It allows users to (a) work with any decent conditional modeling choice, (b) use metrics that are efficient and consistent (any improved quality metrics that appear in future as this field matures can be easily incorporated), (c) compare metrics using any new consistency checks, and (d) rank algorithms using well-laid out statistical significance tests. \section{Illustrative Examples} \label{sec:examples} \begin{figure}[b!] \centering \includegraphics[width=0.9\linewidth]{MoG2D_samples2}\\ \small{(a)}\\ \includegraphics[width=0.9\linewidth]{MoG2D_metrics2b}\\ \small{(b)}\\ \includegraphics[width=0.9\linewidth]{usps_samples}\\ \small{(c)}\\ \includegraphics[width=0.9\linewidth]{usps_metrics2}\\ \small{(d)} \vspace{-0.1in} \caption{\small Results of evaluating imputation algorithms on the Mixture of Gaussians (MoG) dataset and USPS. (a) Mixture components are shown as one-standard-deviation ellipses. Columns are different missing examples, rows are algorithms. The true conditional distribution of each column are: 1) $P(x,y|$\emph{label}=\emph{blue}$)$, 2) $P(x|y=4,$\emph{label}=\emph{blue}$)$, 3) $P(x,y|$\emph{label}=\emph{red}$)$, 4) $P(y|x=6,$\emph{label}=\emph{blue}$)$, 5) $P(x,y|$\emph{label}=\emph{red}$)$. The quality of the imputations can be judged by visually comparing the samples to the correct conditional distribution. (b) Average ranks computed by KL, MMD score, and NDS, for the full MoG dataset. (c) Each of three sets of images is a different USPS digit with missing values. The top row of each set shows the true image (left) and missing pixel indicators as a binary mask (right). The next four rows show five samples generated by A1 to A4 from top to bottom. (d) Average ranks computed by KL, MMD score, and NDS, for the full USPS dataset.} \label{fig:usps_mog2d} \end{figure} Before explaining the details of our framework, we first show results on two datasets for which the imputations can be visualized. This allows a visual comparison of how good the different imputation algorithms are, which can then be checked against the evaluation by our framework. \noindent\textbf{Datasets:} The first dataset is generated from a two-dimensional mixture of Gaussians with two components having equal prior probability. The Gaussians are shown in figure~\ref{fig:usps_mog2d} by their one-standard-deviation ellipses. The value of the component variable is also included in the dataset as a class label, color-coded in the figure by red and blue. We use 2000 data points sampled from the distribution. The second dataset is the USPS handwritten digits which contains $16 \times 16$ binary images flattened as 256-dimensional vectors. We select images for digits 2 and 3 to construct a dataset with 1979 rows and 256 columns. 50\% of the values are removed from both datasets completely at random. \noindent\textbf{Imputation algorithms \& Metrics:} The missing values are imputed using four different algorithms: Mean/Mode, Mixture Model, $k$ Nearest Neighbors ($k$NN), and MICE (labeled as $A1$, $A2$, $A3$, $A4$, respectively, in fig.~\ref{fig:usps_mog2d}). These algorithms are described in section~\ref{sec:evaluation}. For the 2D mixture of Gaussians, we have also included the true distribution itself as an imputation algorithm (labeled $A5$ in the plots) by using its conditionals to sample the missing variables given the observed ones. This is a good sanity check since a good metric should rank it as the best algorithm. We use KL divergence, MMD score, and NDS as the metrics (defined in section~\ref{sec:evaluation}). \noindent\textbf{Results:} Figure~\ref{fig:usps_mog2d} summarizes the results for both datasets. To give a visual feel for the quality of the imputations, we have shown as examples a few specific data vectors and the samples generated by the algorithms to fill in the vectors. To interpret these plots, figure~\ref{fig:usps_mog2d} caption lists the variables that are missing for each example from the MoG dataset and explains which pixels are missing in the USPS images. The rank plots show the average rank attained by each algorithm over all the rows with missing values in the dataset. Error bars show 95\% confidence intervals for significant differences. Visually judging the imputations, $A3$ ($k$NN) and $A4$ (MICE) produce the most plausible imputations for both datasets, while $A1$ (Mode/Mean) and $A2$ (Mixture Model) produce the worst ones. The ordering computed by all the three metrics in figure~\ref{fig:usps_mog2d} agree with this judgement. The true distribution for the MoG dataset has the best average rank for all three metrics, which is reassuring. Interestingly, MICE ties with the true distribution as the best algorithm. In the case of USPS, the ranking A1 to A4 (worst to best) is visually clear and is consistent with the average ranks and the small error bars. The results in section~\ref{sec:experiments} show a similar ordering of algorithms on datasets where the rankings cannot be verified in some other way. Doing this visual verification and corroborating its results with that of other datasets gives us more confidence that the evaluation framework is sensible. \section{Introduction} \label{sec:intro} Many commonly used machine learning algorithms assume that the input data matrix does not have any missing entries. In practice this assumption is often violated. \emph{Imputation} is one popular approach to handling missing values -- it fills in the dataset by inferring them from the observed values. Once filled, the dataset can then be passed on to an ML algorithm as if it were fully observed. \emph{Multiple imputation} \cite{lit86} takes into account the uncertainty in the missing values by sampling each missing variable from an appropriate distribution to produce multiple filled-in versions of the dataset. Multiple imputation has been an active area of research in Statistics and ML \cite{bur10,buu99,gel01,li12,rag01,su11,temp09,tem11,buu07,buu11} and has been implemented in several popular Statistics/ML packages \cite{su11,buu11}. Although there has been much work on multiple imputation, there is little work on the quantitative measurement of the relative performance of a set of imputation algorithms on a given dataset. In the special case where it is known ahead of time that the dataset is to be used only for a specific task (e.g. classification), one can use the evaluation metric for that task (e.g. 0/1 loss) to indirectly measure the performance of an imputation algorithm. But here we are interested in directly evaluating the quality of the imputation, without using a subsequent task for a proxy evaluation. Current procedures available for evaluating imputations are qualitative. Some imputation tools, such as the R packages \emph{mi} and \emph{MICE}, allow visual comparison of the histogram of observed values and imputed values for each column of the dataset. Such comparisons can be done only for one or two variables at a time and so is at best a weak sanity check. van Buuren and Groothuis-Oudshoorn mention criteria for checking whether the imputations are plausible, e.g. \emph{imputations should be values that could have been obtained had they not been missing}, \emph{imputations should be close to the data}, and they should \emph{reflect the appropriate amount of uncertainty about their `true' values}. But such criteria have not yet been translated into quantitative metrics. One possibility for quantifying imputation performance is to hold out a subset of observed values from the dataset, impute them, and measure the ``reconstruction error'' between the known and predicted values using an appropriate distance metric, e.g. Euclidean distance for real-valued variables and Hamming distance for categorical variables. As Rubin explains \cite{rub96}, this is not a good metric because there can be uncertainty in the value of the missing variables (conditioned on the values of the observed variables), and directly comparing imputed samples to a single groundtruth value cannot measure this uncertainty. The key quantity needed to evaluate an imputation is the \emph{distribution of the missing variables conditioned on the observed variables}. Consider a data vector $Y$ with observed variables $Y_{obs}$ and missing variables $Y_{mis}$, with $Y = \{Y_{obs},Y_{mis}\}$. Let $R$ be a vector of indicator variables indicating which variables in $Y$ are missing or observed. As explained in Little and Rubin \cite{lit86} (Chapter 11, page 219, equation 11.4), the joint distribution of $Y$ and $R$ can be written using Bayes rule, as \begin{equation} P(Y,R) = P(Y_{mis}|Y_{obs},R)P(Y_{obs},R). \end{equation} A dataset with missing values contains samples for $Y_{obs}$ and $R$ generated according to some ground truth joint distribution $P_{true}(Y,R)$. An imputation algorithm generates samples from an estimate of $\hat{P}(Y_{mis}|Y_{obs},R)$. Note $P(Y_{obs},R)$ does not depend on imputation. Thus, if we can compare $P_{true}(Y_{mis}|Y_{obs},R)$ to $\hat{P}(Y_{mis}|Y_{obs},R)$ that is estimated by an imputation algorithm, then it is possible to quantify how good the imputation is. Therefore the imputation evaluation problem can be turned into a problem of comparing two conditional distributions. In this work we use this connection between imputation evaluation and distribution comparison to develop a quantitative framework for evaluating imputations that can be tractably applied to real-world datasets. \noindent\textbf{Problem setup:} We have 1) a dataset with values missing according to some (possibly unknown) missingness type, and 2) the output of $K$ different imputation algorithms, with each one producing multiple filled-in versions of the dataset. The goal is to rank the $K$ algorithms according to a chosen metric. Several metric choices are discussed in section~\ref{sec:evaluation}. \noindent\textbf{Assumption:} It should be possible to estimate a model of $P_{true}(Y_{mis}|Y_{obs},R)$ from the data for each unique missingness pattern $R$. Section~\ref{sec:evaluation} shows how this can be done tractably. The samples given by an imputation algorithm from its estimated conditional $\hat{P}(Y_{mis}|Y_{obs},R)$ are then compared against this model, rather than the original data itself. \noindent\textbf{Algorithm:} The main steps are the following: \begin{itemize} \item A) Construct a model for $P_{true}(Y_{mis}|Y_{obs},R)$ for each unique $R$ in the dataset. \item B) For each row of the dataset: \begin{enumerate} \item Generate samples from the model of $P_{true}(Y_{mis}|Y_{obs},R)$. \item Generate samples from the conditional $\hat{P_k}(Y_{mis}|Y_{obs},R)$ estimated by the $k^{th}$ imputation algorithm for the current row. \item Compute a score $s_k$ for the $k^{th}$ algorithm according to a chosen evaluation metric that compares the two conditional distributions using their corresponding samples. \item Sort the scores to rank the $K$ algorithms for the current row. \end{enumerate} \item C) Use the method recommended by Demsar \cite{dem06} to combine the rankings for all the rows into a single average ranking over the $K$ algorithms. Confidence intervals are assigned to the average ranks to indicate which algorithms have significantly different imputation quality. \end{itemize} There are several choices possible for 1) the model of $P_{true}(Y_{mis}|Y_{obs},R)$, and 2) the metric for distribution comparison. Our evaluation framework is general in that it is not tied to any specific choice for either of these. Nevertheless in section~\ref{sec:evaluation} we discuss a few specific choices that are tractable and work well for high-dimensional data, e.g., problems with hundreds of features. For modeling $P_{true}(Y_{mis}|Y_{obs},R)$ we propose a Markov Random Field-based approach. We consider two choices: (a) a separate model for each missingness pattern; and (b) a single model for all missingness patterns. For distribution comparison metrics we consider 1) Kullback-Leibler (KL) divergence, 2) Symmetric KL divergence, and 3) two-sample testing for equality of two distributions. The challenge is that only samples of the true and imputed distributions are available. Recent advances in kernel-based approaches to estimating these metrics from samples have now made it possible to apply them to high-dimensional data. We estimate KL divergence using the convex risk minimization approach of \cite{ngu10}, symmetric KL using the cost-sensitive binary classification approach of \cite{rei11}, and perform two-sample testing using Maximum Mean Discrepancy (MMD) \cite{zar13,gre12,gre13}. In addition, we have proposed a new metric called Neighborhood-based Dissimilarity Score (NDS) that is faster to compute while providing similar results. We demonstrate the framework on several real datasets to compare four representative imputation algorithms from the literature. \noindent\textbf{Contributions:} To our knowledge this is the first work that proposes the following. \begin{itemize} \item A principled framework for computing quantitative metrics to evaluate imputations and rank algorithms with statistical significance tests. Its generality allows different choices for distribution modeling and distribution comparison metrics to be plugged in according to the problem. \item An efficient algorithm for jointly building distribution models for each unique missingness pattern. \item Neighborhood-based Dissimilarity Score, which is two orders of magnitude faster than KL, symmetric KL and MMD, while giving similar results. \end{itemize} We can use this framework to evaluate the quality of imputation algorithms on different datasets, or understand the sensitivity of various imputation algorithms to different missingness types. The experiments presented in this paper are a first step towards such deeper investigations. \noindent\textbf{Notes:} 1) The evaluation does not require knowledge of the missingness type, such as whether values are Missing Completely at Random (MCAR) or Not Missing at Random (NMAR) \cite{lit86}. It simply checks whether the samples generated by the imputation algorithm are distributed correctly compared to the true distribution. The imputation algorithm needs to be aware of the missingness type, but not the evaluation. 2) In our experiments we only consider up to 50\% missingness. Imputation is typically used for problems where the missingness percentage is not very high. At much higher percentages, e.g. 99\% missingness in Collaborative Filtering datasets, more specialized approaches are used and evaluation is focused more on predicting a few relevant missing entries rather than \emph{all} of them. \noindent\textbf{Outline:} Section 2 presents two motivating examples that demonstrate our evaluation framework on datasets where it is possible to visually check the quality of the imputations. Section 3 explains the details of constructing the distribution models, the metrics used for comparing distributions, and statistical significance testing for the ranking of the algorithms. Section 4 presents the experimental results on several datasets, followed by a closing discussion in section 5. \section{Introduction} \label{sec:intro} Many commonly used machine learning algorithms assume that the input data matrix does not have any of its entries missing. But in practice this is often not true. \emph{Imputation} is one popular approach to handling missing values -- it fills in the dataset by inferring them from the observed values. Once filled, the dataset can then be passed on to an ML algorithm as if it were fully observed. \emph{Multiple imputation} takes into account the uncertainty in the missing values by sampling each missing variable to produce multiple filled-in versions of the dataset. Multiple imputation has been an active area of research in Statistics and ML [refs] and has been implemented in many popular Statistics/ML packages [refs]. Although there has been much work on mulitple imputation, quantitatively measuring the relative performance of a set of imputation algorithms on a given dataset remains an open problem. In the special case where it is known ahead of time that the dataset is to be used only for a specific task (e.g. classification), one can use the evaluation metric for that task (e.g. 0/1 loss) to indirectly measure the performance of an imputation algorithm. But here we are interested in directly evaluating the quality of the imputation, without using a subsequent task for a proxy evaluation. Current procedures available for evaluating imputations are qualitative. Some imputation tools, such as the R packages \emph{mi} and \emph{MICE}, allow visual comparison of the histogram of observed values and imputed values for each column of the dataset. Such comparisons can be done only for one or two variables at a time and so is at best a weak sanity check. van Buuren and Groothuis-Oudshoorn mention criteria for checking whether the imputations are plausible, e.g. \emph{imputations should be values that could have been obtained had they not been missing}, \emph{imputations should be close to the data}, and they should \emph{reflect the appropriate amount of uncertainty about their `true' values}. But such criteria have not yet been translated into quantitative metrics. One possibility for quantifying imputation performance is to hold out a subset of observed values from the dataset, impute them, and measure the ``reconstruction error'' between the known and predicted values using an appropriate distance metric, e.g. Euclidean distance for real-valued variables and Hamming distance for categorical variables. As Rubin explains in [ref], this is not a good metric because there can be uncertainty in the value of the missing variables (conditioned on the values of the observed variables), and directly comparing imputed samples to a single groundtruth value cannot measure this uncertainty. The key quantity needed to evaluate an imputation is the \emph{distribution of the missing variables conditioned on the observed variables}. Consider a data vector $Y$ with observed variables $Y_{obs}$ and missing variables $Y_{mis}$, with $Y = \{Y_{obs},Y_{mis}\}$. Let $R$ be a vector of indicator variables indicating which variables in $Y$ are missing or not. As explained in Little and Rubin (Chapter 11, page 219, equation 11.4), the joint distribution of $Y$ and $R$ can be written as \begin{align*} P(Y,R) &= P(Y|R)P(R)\\ &= P(Y_{mis},Y_{obs}|R)P(R)\\ &= P(Y_{mis}|Y_{obs},R)P(Y_{obs}|R)P(R).\\ \end{align*} A dataset with missing values contain samples for $Y_{obs}$ and $R$ generated according to some ground truth joint distribution $P_{true}(Y,R)$. An imputation algorithm generates samples from an estimate of $\hat{P}(Y_{mis}|Y_{obs},R)$. Note that the remaining two terms in the joint distribution, $P(Y_{obs}|R)$ and $P(R)$, do not depend on imputation. If we can compare $P_{true}(Y_{mis}|Y_{obs},R)$ to $\hat{P}(Y_{mis}|Y_{obs},R)$ estimated by an imputation algorithm, then it is possible to quantify how good the imputation is. Therefore the imputation evaluation problem can be turned into a problem of comparing two distributions, for which principled quantitative evaluation is possible. From this connection to distribution comparison, we develop a framework for evaluating imputations that can be tractably applied to real-world datasets. \noindent\textbf{Problem setup:} We have 1) a dataset with values missing according to some (possibly unknown) missingness model, and 2) the output of $K$ different imputation algorithms, with each one producing multiple filled-in versions of the dataset. The goal is to rank the $K$ algorithms according to a chosen metric. (Several metric choices are discussed in section~\ref{sec:}). \noindent\textbf{Assumption:} It should be possible to estimate a model of $P_{true}(Y_{mis}|Y_{obs},R)$ from the data for each unique missingness pattern $R$. Section~\ref{sec:} shows how this can be done tractably. The samples given by an imputation algorithm from its estimated conditional $\hat{P}(Y_{mis}|Y_{obs},R)$ are then compared against this model, rather than the original data itself. \noindent\textbf{Algorithm:} The main steps are the following:\\ \begin{itemize} \item A) Construct a model for $P_{true}(Y_{mis}|Y_{obs},R)$ for each unique $R$ in the dataset. \item B) For each row of the dataset: \begin{enumerate} \item Generate samples from the model of $P_{true}(Y_{mis}|Y_{obs},R)$. \item Generate samples from the conditional $\hat{P_k}(Y_{mis}|Y_{obs},R)$ estimated by the $k^{th}$ imputation algorithm for the current row. \item Compute the score $s_k$ for the $k^{th}$ algorithm according to the chosen evaluation metric that compares the two conditional distributions using their corresponding samples. \item Sort the scores to rank the $K$ algorithms for the current row. \end{enumerate} \item C) Use the method recommended by Demsar \cite{dem06} to combine the rankings for all the rows into a single average ranking over the $K$ algorithms. Confidence intervals are assigned to the average ranks to indicate which algorithms are significantly different. \end{itemize} There are several choices possible for 1) the model of $P_{true}(Y_{mis}|Y_{obs},R)$, and 2) the metric for distribution comparison. Our evaluation framework is general in that it is not tied to any specific choice for either of these. Nevertheless in section~\ref{sec:} we discuss a few specific choices that are tractable and work well for high-dimensional data. For modeling $P_{true}(Y_{mis}|Y_{obs},R)$ we consider a Conditional Random Field-based approach, as well as propose a novel Markov Random Field-based approach to efficiently learn models for each unique $R$ in a joint manner. For distribution comparison metrics we consider 1) Kullback-Leibler (KL) divergence, 2) Symmetric KL divergence, and 3) two-sample testing for equality of two distributions. Recent advances in kernel-based approaches to estimating these metrics have now made it possible to apply them to high-dimensional data. We estimate KL divergence using the convex risk minimization approach of Nguyen et al. [ref], symmetric KL using the cost-sensitive binary classification approach of Reid and Williamson [ref], and perform two-sample testing using Maximum Mean Discrepancy (MMD) [ref]. In addition, we have proposed a new metric called Neighborhood-based Dissimilarity Score (NDS) that is faster to compute while providing similar results. We demonstrate the framework on several real datasets to compare four representative imputation algorithms from the literature. \noindent\textbf{Contributions:} To our knowledge this is the first work that defines \begin{itemize} \item A principled framework for computing quantitative metrics to evaluate imputations and rank algorithms with statistical significance tests. Its generality allows different choices for distribution modeling and distribution comparison metrics to be plugged in according to the problem. \item An efficient algorithm for jointly building distribution models for each unique missingness pattern. \item Neighborhood-based Dissimilarity Score, which is two orders of magnitude faster than KL, symmetric KL and MMD, while giving similar results. \end{itemize} Several things that were previously not possible to do in a principled way now become possible. Using our framework one can now perform evaluation studies that try to characterize which imputation algorithms perform well on what types of datasets, or try to understand the sensitivity of various imputation algorithms to different missingness models. The experiments presented in this paper are a first step towards such deeper investigations. \noindent\textbf{Notes:} 1) The evaluation does not require knowledge of the missingness model, such as whether values are Missing Completely at Random (MCAR) or Not Missing at Random (NMAR) [ref]. It simply checks whether the samples generated by the imputation algorithm are distributed correctly compared to the true distribution. The imputation algorithm needs to be aware of the missingness model, but not the evaluation. 2) In our experiments we only consider up to 50\% missingness. Imputation is typically used for problems where the missingness percentage is not very high. At much higher percentages, e.g. 99\% missingness in Collaborative Filtering datasets, more specialized approaches are used and evaluation is focused more on predicting a few relevant missing entries rather than \emph{all} of them. \noindent\textbf{Outline:} Section 2 presents two motivating examples that demonstrate our evaluation framework on datasets where it is possible to visually check the quality of the imputations. Section 3 explains the details of constructing the distribution models, the metrics used for comparing distributions, and statistical significance testing for the ranking of the algorithms. Section 4 presents the experimental results on several datasets, followed by discussion in section 5. \section{Experiments} \label{sec:experiments} We apply our framework to four UCI datasets and four representative imputation algorithms from the literature. We present the ranking results for the per-(missing)pattern model and show that the rankings are similar to those in section~\ref{sec:examples}. We then compare the results between the per-pattern model approach and the single model approach and show that the latter gives nearly identical results but is much cheaper to train. Then we consider various properties of the metrics, such as the self-consistency of the rankings for different choices of the ``true'' distribution, and their ability to discriminate among the algorithms with statistically significant rank differences. \noindent\textbf{Datasets:} See table~\ref{tab:datasets}. Note that in the case of Mushroom, we selected the first 10 columns of the original dataset to keep the running times low. None of the datasets contain missing values originally. Values are removed completely at random, i.e. the probability of a variable missing is independent of its value or the value of other observed variables. The missing percentage is varied from 10 to 50\% in steps of 10. We use the full datasets without any missing values to build the distribution models. \begin{table} \begin{center} \begin{tabular}{ | c | c | c | } \hline {\bf Dataset} & {\bf \#Rows} & {\bf \#Cols}\\ \hline Flare & 838 & 9 \\ \hline Mushroom & 2027 & 10 \\ \hline SPECT & 213 & 22 \\ \hline Yeast & 1200 & 14 \\ \hline \end{tabular} \end{center} \caption{Datasets used for experiments} \label{tab:datasets} \end{table} \begin{figure*} \centering \includegraphics[width=0.8\linewidth]{perrow_results1}\\ \caption{Average rank of the four imputation algorithms on each of the five metrics for the Yeast dataset with 30\% entries missing at random} \label{fig:per_row_results} \end{figure*} \begin{figure} \centering \includegraphics[width=0.9\linewidth]{perrow_results_missing103050}\\ \caption{Average rank of the four imputation algorithms on the KL (left column), MMD score (middle column) and NDS (right column) metrics, for the Yeast dataset with missing entries ranging from 10\% (top row), 30\% (middle row) and 50\% (bottom row).} \label{fig:per_row_results2} \end{figure} \noindent\textbf{Imputation algorithms:} The following algorithms are included in the experiments:\\ \textbf{A1: Mode/Mean:} A simple baseline that samples from the marginal distribution of each column, fit by a Gaussian for continuous variables and a multinomial distribution for categorical variables.\\ \noindent\textbf{A2: Mixture Model:} A mixture model in which each mixture component is a product of univariate distributions. Inference and learning are done using Variational Bayes inference \cite{win05}.\\ \noindent\textbf{A3: $k$-Nearest Neighbors:} We implement the $k$-NN algorithm in \cite{has99}. Multiple samples are drawn from the distribution of values for the missing variable given by the set of nearest neighbors.\\ \noindent\textbf{A4: MICE:} Multiple Imputation using Chained Equations (MICE) \cite{buu11} is a state-of-the-art imputation algorithm which builds a predictor for each variable using all other variables as inputs. The learning alternates between predicting the missing values using the current predictors and updating the predictors using the current estimates of the missing values.\\ These algorithms are selected as representatives of commonly used approaches, rather than as a comprehensive set. \noindent\textbf{Ranking results:} We have computed results using the per-pattern distribution models for 4 imputation algorithms, 5 metrics, 4 datasets, and 5 missing percentages, for a total of 400 combinations. The results can be analyzed along any of these dimensions to derive insight. For example, figure~\ref{fig:per_row_results} shows the results for the four different algorithms and five metrics for Yeast with 30\% missing values. All metrics pick MICE as the best algorithm. Or if we want to understand the effect of missing percentage on the performance of the imputation algorithms, we can compute the average ranks over all the datasets for each missing percentage and metric. Figure~\ref{fig:per_row_results2} shows the results for three missing percentages and three metrics. (We omitted some plots to save space). As the missing percentage increases, the algorithms become harder to distinguish under the KL and MMD score metrics, while NDS maintains discriminability. If we want to know the overall ranking of the imputation algorithms for each metric, we can aggregate the results across all datasets and all missing percentages. We observe that MICE is overall the best imputation algorithm, Mode/Mean is the worst, with Mixture Model and $k$-NN in between. These are some examples of the kind of analysis that become possible with our framework \noindent\textbf{Per-pattern vs. single models:} In section~\ref{sec:evaluation} we propose an alternative to learning a separate distribution model for each unique missingness pattern by learning all of them jointly with a single set of parameters. Here we verify whether learning a combined model in this manner introduces any approximation error that causes the average rankings to change. We re-run the same 400 combinations of algorithms, metrics, datasets, and missing percentages from the previous experiment, but now using the single distribution model instead of the per-pattern model. The percent difference in the average rank values between the two models for each metric is given in table~\ref{tab:per_pattern_vs_single}. The differences in the average rank values are too small to alter the rankings. (Note that the critical difference values do not depend on the choice of per-pattern vs. single model.) So both models give identical rankings over all metrics, datasets, and missing percentages. \begin{table} \begin{center} \begin{tabular}{ | c | c | c | c | c | } \hline {\bf KL} & {\bf Sym. KL} & {\bf MMD} & {\bf MMD} & {\bf NDS}\\ & & {\bf B-test} & {\bf score} & \\ \hline 2.27\% & 1.10\% & 1.40\% & 2.06\% & 0.36\%\\ \hline \end{tabular} \end{center} \caption{\% change in average rank between per-pattern model and single model, averaged over all imputation algorithms, datasets, and missing percentages.} \label{tab:per_pattern_vs_single} \end{table} \begin{figure} \vspace{-0.1in} \centering \includegraphics[width=0.7\linewidth]{consistent1}\\ \vspace{-0.1in} \caption{Rankings of imputation algorithms with respect to different `true' distributions for Flare with 10\% missing values and NDS metric. The algorithm labels are: A1 = Mode/Mean, A2 = Mixture Model, A3 = $k$-NN, A4=MICE, A5 = Single model distribution.} \label{fig:consistent1} \end{figure} \begin{figure} \vspace{-0.15in} \centering \includegraphics[width=0.7\linewidth]{consistent2}\\ \vspace{-0.2in} \caption{Inconsistency scores for various metrics on Flare with different missing percentages. The metric labels are: MMD B-test = 1, MMD score = 2, SymmKL = 3, KL = 4, NDS = 5.} \vspace{-0.2in} \label{fig:consistent2} \end{figure} \noindent\textbf{Measuring metric inconsistency:} We can use our framework to investigate whether an evaluation metric has certain desirable properties, and based on the results decide on its usefulness. One such property we examine here is how ``inconsistent'' a metric is. Suppose we have $K$ algorithms $A_1$ to $A_K$. We can rank all algorithms except $A_i$ by treating its samples as coming from the `true' distribution and comparing the rest against it. $K$ such rankings can be computed for $i = 1$ to $K$. It is possible to check how inconsistent these rankings are with each other. Consider the example in figure~\ref{fig:consistent1}. When A1 is used as the true distribution, we see that A4 has a small average rank difference with A3 and A5, and a large difference with A2. When A4 itself is used as the true distribution, the ranking of \{A2, A3, A5\} follows the same pattern. A3 and A5 are ranked as closest to A4 while A2 is ranked as distant from A4. So for the pair (A1,A4) the ranking of \{A2, A3, A5\} given by the average rank differences in the left plot matches the ranking in the right plot. When they do not match, we call them \emph{inconsistent}. We can quantify the inconsistency using the Kendall-Tau distance between the two rankings for every pair of algorithms and averaging the distances. Figure~\ref{fig:consistent2} shows the inconsistency scores computed for five metrics on Flare for different missing percentages. Note that NDS has the lowest inconsistency score among all the metrics. The difference between NDS and other metrics is particularly large at the lower missing percentages. Similar results are observed on other datasets also. This provides further evidence that NDS is a good metric for evaluating imputations. \section{Conclusion} We presented a framework to quantitatively evaluate and rank the goodness of imputation algorithms. It allows users to (a) work with any decent conditional modeling choice, (b) use metrics that are efficient and consistent (any improved quality metrics that appear in future as this field matures can be easily incorporated), (c) compare metrics using any new consistency checks, and (d) rank algorithms using well-laid out statistical significance tests.
\section{\def\@secnumfont{\mdseries}\@startsection{section}{1}% \z@{.7\linespacing\@plus\linespacing}{.5\linespacing}% {\normalfont\scshape\centering}} \def\subsection{\def\@secnumfont{\bfseries}\@startsection{subsection}{2}% {\parindent}{.5\linespacing\@plus.7\linespacing}{-.5em}% {\normalfont\bfseries}} \makeatother \def\subl#1{\subsection{}\label{#1}} \newcommand{\Hom}{\operatorname{Hom}} \newcommand{\mode}{\operatorname{mod}} \newcommand{\End}{\operatorname{End}} \newcommand{\wh}[1]{\widehat{#1}} \newcommand{\Ext}{\operatorname{Ext}} \newcommand{\ch}{\text{ch}} \newcommand{\ev}{\operatorname{ev}} \newcommand{\Ob}{\operatorname{Ob}} \newcommand{\soc}{\operatorname{soc}} \newcommand{\rad}{\operatorname{rad}} \newcommand{\head}{\operatorname{head}} \def\operatorname{Im}{\operatorname{Im}} \def\operatorname{gr}{\operatorname{gr}} \def\operatorname{mult}{\operatorname{mult}} \def\operatorname{Max}{\operatorname{Max}} \def\operatorname{Ann}{\operatorname{Ann}} \def\operatorname{sym}{\operatorname{sym}} \def\operatorname{loc}{\operatorname{loc}} \def\operatorname{\br^\lambda_A}{\operatorname{\br^\lambda_A}} \def\underline{\underline} \def$A_k(\lie{g})(\bsigma,r)${$A_k(\lie{g})(\bsigma,r)$} \newcommand{\Cal}{\cal} \newcommand{\Xp}[1]{X^+(#1)} \newcommand{\Xm}[1]{X^-(#1)} \newcommand{\on}{\operatorname} \newcommand{\Z}{{\bold Z}} \newcommand{\J}{{\cal J}} \newcommand{\C}{{\bold C}} \newcommand{\Q}{{\bold Q}} \renewcommand{\P}{{\cal P}} \newcommand{\N}{{\Bbb N}} \newcommand\boa{\bold a} \newcommand\bob{\bold b} \newcommand\boc{\bold c} \newcommand\bod{\bold d} \newcommand\boe{\bold e} \newcommand\bof{\bold f} \newcommand\bog{\bold g} \newcommand\boh{\bold h} \newcommand\boi{\bold i} \newcommand\boj{\bold j} \newcommand\bok{\bold k} \newcommand\bol{\bold l} \newcommand\bom{\bold m} \newcommand\bon{\bold n} \newcommand\boo{\bold o} \newcommand\bop{\bold p} \newcommand\boq{\bold q} \newcommand\bor{\bold r} \newcommand\bos{\bold s} \newcommand\boT{\bold t} \newcommand\boF{\bold F} \newcommand\bou{\bold u} \newcommand\bov{\bold v} \newcommand\bow{\bold w} \newcommand\boz{\bold z} \newcommand\boy{\bold y} \newcommand\ba{\bold A} \newcommand\bb{\bold B} \newcommand\bc{\bold C} \newcommand\bd{\bold D} \newcommand\be{\bold E} \newcommand\bg{\bold G} \newcommand\bh{\bold H} \newcommand\bi{\bold I} \newcommand\bj{\bold J} \newcommand\bk{\bold K} \newcommand\bl{\bold L} \newcommand\bm{\bold M} \newcommand\bn{\bold N} \newcommand\bo{\bold O} \newcommand\bp{\bold P} \newcommand\bq{\bold Q} \newcommand\br{\bold R} \newcommand\bs{\bold S} \newcommand\bt{\bold T} \newcommand\bu{\bold U} \newcommand\bv{\bold V} \newcommand\bw{\bold W} \newcommand\bz{\bold Z} \newcommand\bx{\bold x} \newcommand\KR{\bold{KR}} \newcommand\rk{\bold{rk}} \newcommand\het{\text{ht }} \newcommand\toa{\tilde a} \newcommand\tob{\tilde b} \newcommand\toc{\tilde c} \newcommand\tod{\tilde d} \newcommand\toe{\tilde e} \newcommand\tof{\tilde f} \newcommand\tog{\tilde g} \newcommand\toh{\tilde h} \newcommand\toi{\tilde i} \newcommand\toj{\tilde j} \newcommand\tok{\tilde k} \newcommand\tol{\tilde l} \newcommand\tom{\tilde m} \newcommand\ton{\tilde n} \newcommand\too{\tilde o} \newcommand\toq{\tilde q} \newcommand\tor{\tilde r} \newcommand\tos{\tilde s} \newcommand\toT{\tilde t} \newcommand\tou{\tilde u} \newcommand\tov{\tilde v} \newcommand\tow{\tilde w} \newcommand\toz{\tilde z} \newcommand\woi{w_{\omega_i}} \newcommand\chara{\operatorname{Char}} \title{Fusion Product Structure of Demazure modules} \author[R. Venkatesh]{R. Venkatesh} \address{\noindent Department of Mathematics, The Institute of Mathematical Sciences, Chennai 600113, India.} \email{<EMAIL>} \begin{abstract} Let $\lie g$ be a finite--dimensional complex simple Lie algebra. Given a non--negative integer $\ell$, we define $\mathcal{P^+_\ell}$ to be the set of dominant weights $\lambda$ of $\lie g$ such that $\ell\Lambda_0+\lambda$ is a dominant weight for the corresponding untwisted affine Kac--Moody algebra $\widehat{\lie g}$. For the current algebra $\lie g[t]$ associated to $\lie g$, we show that the fusion product of an irreducible $\lie g$--module $V(\lambda)$ such that $\lambda\in \mathcal{P^+_\ell}$ and a finite number of special family of $\lie g$--stable Demazure modules of level $\ell$ (considered in \cite{FoL2} and \cite{FoL}) again turns out to be a Demazure module. This fact is closely related with several important conjectures. We use this result to construct the $\lie g[t]$--module structure of the irreducible $\widehat{\mathfrak{g}}$--module $V(\ell \Lambda_0+\lambda)$ as a semi--infinite fusion product of finite dimensional $\lie g[t]$--modules as conjectured in \cite{FoL}. As a second application we give further evidence to the conjecture on the generalization of Schur positivity (see \cite{chfsa}). \end{abstract} \maketitle \section*{Introduction} The current algebra associated to a simple Lie algebra $\lie g$ is the special maximal parabolic subalgebra of the untwisted affine Lie algebra $\widehat{\lie g}$ associated to $\lie g$. As a vector space it is isomorphic to $\lie g\otimes \mathbb{C}[t]$, where $\mathbb{C}[t]$ is the polynomial ring in the indeterminate $t$ and the Lie bracket is given in the obvious way. Equivalently, it is just the Lie algebra of polynomial maps $\mathbb{C}\to \lie g$. The study of the category of finite dimensional graded representations of the current algebra has been the subject of many articles in the recent years. See for example \cite{eddy}, \cite{BCM}, \cite{chbdz}, \cite{deFK}, \cite{FoL2}, \cite{FoL}, \cite{Naoi}, \cite{Naoi1}. One of the original motivations for the study of this category was that it is closely related to the representation theory of the corresponding quantum affine algebra. The results of \cite{Ckir} and \cite{CMkir} showed that many interesting families of irreducible representations of the quantum affine algebra associated to $\lie g$, when specialized to $q = 1$ give indecomposable representations of the current algebra. For example, these families include the Kirillov--Reshetikhin modules. There is an important class of finite dimensional graded $\lie g[t]$--modules, called Demazure modules. Let $\lie b$ be a Borel subalgebra of $\lie g$ and $\widehat{\lie b}$ be the corresponding Borel subalgebra of $\widehat{\mathfrak{g}}$. By definition, a Demazure Module is a $\widehat{\lie b}$--submodule of an irreducible highest weight representation of $\widehat{\lie g}$ generated by an extremal weight vector. We shall only be concerned with $\lie g$--stable Demazure modules in this article. These modules naturally become finite dimensional graded modules for the associated current algebra $\lie g[t]$ by restriction. The Demazure module of level $\ell$ and weight $\lambda\in P^+$ for $\lie g[t]$ is denoted by $D(\ell, \lambda)$, where $P^+$ denotes the set of dominant weight of $\lie g$. The defining relations of these $\lie g[t]$--modules are greatly simplified in \cite{chve} using the results of \cite{mathieu} (see also \cite{FoL}, and \cite{Naoi}). We use this simplified presentation to study fusion product structure of $D(\ell, \lambda)$ for some special weights $\lambda\in P^+$. We assume that $\lie g$ is simply laced for simplicity and state our main result here. We refer the reader to Theorem \ref{mapsdem} for a more general statement. The primary goal of this paper is the following: \begin{thm*} Let $\ell, m\in \mathbb{N}$ and $z, z_1, \dots, z_m\in \mathbb{C}$ be given distinct complex numbers. Let $\mu\in P^+$, $\lambda\in \mathcal{P^+_\ell}$ and suppose that there exists $\mu_j\in P^+$, $1\le j\le m$ such that $\mu=\mu_1+\cdots+\mu_m$. Then we have an isomorphism of $\lie g[t]$--modules, $$D(\ell,\ell \mu+\lambda)\cong D^{z_1}(\ell,\ell \mu_1)*\cdots*D^{z_m}(\ell, \ell \mu_m)*\ev_{z}V(\lambda)$$ \end{thm*} This generalizes a theorem of Fourier and Littelmann (see \cite[Theorem C]{FoL}) where they only consider the case $\lambda=0$. Note in particular that the fusion product of an irreducible module and a finite number of Demazure modules of the form $D(\ell,\ell\mu)$, for sufficiently large $\ell$, is independent of the choice of the parameters that used in the fusion product construction. Since the Kirillov--Reshetikhin modules are Demazure modules (see \cite{CMkir}, \cite{FoL}), the new interpretations of our main theorem has many interesting applications related to several important conjectures. We give two important applications in this article. Let $V(\ell\Lambda_0+\lambda)$ be the irreducible highest weight module of highest weight $\ell\Lambda_0+\lambda$ for the untwisted affine Kac--Moody $\widehat{\lie g}$, where $\Lambda_0$ is the fundamental weight of $\widehat{\lie g}$ corresponding to the extra node of the extended Dykin diagram of $\lie g$. As a first application we construct the $\lie g[t]$--module structure of $V(\ell \Lambda_0+\lambda)$ as a semi--infinite fusion product of finite dimensional $\lie g[t]$--modules as conjectured in \cite{FoL}. More preciously, we prove the following: \begin{thm*} Let $\Lambda = \ell \Lambda_0 + \lambda$ be a dominant integral weight for $\widehat{\mathfrak{g}}$, then $$ V(\Lambda) \ \text{and} \ \underset{N \rightarrow \infty}{\mbox{lim }} D(\ell,\ell\Theta) \ast \cdots \ast D(\ell,\ell\Theta) \ast V(\lambda) $$ are isomorphic as $\lie g[t]$--modules. \end{thm*} The special case of this theorem (when $\Lambda$ is a multiple of $\Lambda_0$) was proved in [Theorem 9, \cite{FoL}]. The semi--infinite fusion construction was already established in \cite{FFsemi} for $\lie g= \lie sl_2$ and in \cite{kedemsemi} for $\lie g=\lie sl_n$. As a further application we prove some results on the generalization of Schur positivity, that gives some additional evidence to the conjecture that appeared in \cite{chfsa}. We now recall the conjecture on generalization of Schur positivity. The following was conjectured in \cite[Section 2.3]{chfsa} by Chari, Fourier and Sagaki: \begin{conj}\label{main--conj} Let $\lie g$ be a simple Lie algebra, and $\lambda_1, \lambda_2, \mu_1, \mu_2 \in P^+$ such that \begin{enumerit} \item[(i)] $\lambda_1+\lambda_2=\mu_1+\mu_2$, \item[(ii)]\label{schurcond} $\min\{\lambda_1(h_\alpha),\lambda_2(h_\alpha)\}\le \min\{\mu_1(h_\alpha),\mu_2(h_\alpha)\}$, for all $\alpha\in R^+$, \end{enumerit} then there exists a surjective $\lie g$--module map $V(\mu_1) \otimes V(\mu_2)\rightarrow V(\lambda_1) \otimes V(\lambda_2)\rightarrow 0.$ \end{conj} We give positive answer to this conjecture in certain important cases. Indeed we prove stronger statements in this article, we prove that the fusion product of two irreducible $\lie g$--modules in certain cases turns out be Demazure modules and use this fact to get elegant presentation for $V(\lambda_1)*V(\lambda_2)$ for some special choices of $\lambda_1, \lambda_2$. We use this elegant presentation to prove that there exists surjective $\lie g[t]$--module maps between the appropriate fusion products (see Theorem \ref{genschurpos}). The paper is organized as follows. Section \ref{preliminaries} has the basic notation and elementary results needed for the rest of the paper. In section \ref{weyl}, we give the definition of Weyl and Demazure modules and recall their presentations. In section \ref{fusion}, we prove that the fusion product of an irreducible module and the Demazure modules of the same level again turns out to be a Demazure module. The key steps of the proof are to show that there is a surjective map between appropriate modules and its dimension matches. The existence of the surjective map proved using the presentation of the Demazure module that was established in \cite{chve}. In section \ref{affine} we calculate the dimension of the Demazure modules using Demazure operators and give the limit construction of $V(\ell\Lambda_0+\lambda)$. In section \ref{schur} we give further evidence to the conjecture on the generalization of the Schur positivity. {\em Acknowledgements: The author is very grateful to Vyjayanthi Chari for many useful discussions and encouragement. This work was partially supported by a grant from the Niels Henrik Abel Board. The author thanks the Niels Henrik Abel Board and the Department of Mathematics at UCR for their financial support. The author acknowledges the hospitality and excellent working conditions at the Department of Mathematics at UCR where part of this work was done. The author likes to thank the referee for his useful comments on the previous manuscript that helped to improve the readability of this manuscript. } \section{Preliminaries}\label{preliminaries} Our base field will be the complex numbers $\mathbb{C}$ throughout. We let $\mathbb{Z}$ (resp. $\mathbb{Z}_+$, $\mathbb{N}$) be the set of integers (resp. non--negative, positive integers) and $\mathbb{C}[t]$ be the polynomial ring in a variable $t$. \subsection{} We will broadly follow the notations of \cite{chve}. Let $\lie g$ be an arbitrary finite dimensional simple Lie algebra with Cartan subalgebra $\lie h$ and Borel subalgebra $\lie b.$ We assume that the rank of $\lie g$ is $n$, i.e., $\dim\lie h=n$. Denote $I=\{1,\dots,n\}$. Let $R$ be the set of roots, $R^+$ the set of positive roots and $\{\alpha_i: i\in I\}$ the set of simple roots of $\lie g$ with respect to $\lie h$. Let $P,Q,P^+,Q^+$ be the weight lattice, the root lattice, the sets of dominant weights and non-negative integer linear combinations of simple roots respectively. The Weyl group $W$ of $\lie g$ is generated by the simple reflections $s_i=s_{\alpha_i}$ associated to the simple roots. Let $w_0$ be the longest element of $W$. Let $(\ , \ )$ be the non--degenerate, normalized $W$--invariant symmetric bilinear form on $\lie h^*$ such that the square length of a long root is two. For $\alpha\in R$, $h_\alpha\in \lie h$ denotes the corresponding co--root and set $d_\alpha=2/(\alpha,\alpha)$. Let $\{\omega_i: i\in I\}\subset\lie h^*$ be the set of fundamental weights, i.e., $(\omega_i,d_j\alpha_j)=\delta_{i,j}$, where $\delta_{i,j}$ is the Kronecker delta symbol. Let $\lie g_\alpha$ be the root space of $\lie g$ corresponding to the root $\alpha\in R$ and set $\lie n^\pm=\bigoplus_{\alpha\in R^+}\lie g_{\pm\alpha}$ and $\lie g_{\pm\alpha}=\mathbb{C}x^\pm_{\alpha}$. Denote by $\Theta\in R^+$ be the highest root in $R$ and recall that $[x^\pm_\Theta,\lie n^\pm]=0$. For $i\in I$, we simply write $x^\pm_i$, $h_i$, $d_i$ for $x^\pm_{\alpha_i}$, $h_{\alpha_i}$, $d_{\alpha_i}$. \iffalse \subsection{Affine Kac-Moody algebras} Let $\widehat{\mathfrak{g}}$ be the untwisted affine Kac--Moody algebra corresponding to the extended Dynkin diagram of $\lie g$: $$ \widehat{\mathfrak{g}}=\lie g\otimes_\mathbb{C}\mathbb{C}[t,t^{-1}]\oplus \mathbb{C} K\oplus \mathbb{C} d $$ Here $K$ is the canonical central element and $d$ denotes the derivation $d=t\frac{d}{dt}$. Naturally we consider the Lie algebra $\lie g$ as a subalgebra of $\widehat{\mathfrak{g}}$. In the same way, $\lie h$ and $\lie b$ are subalgebras of the Cartan subalgebra $\widehat{\mathfrak{h}}$ respectively the Borel subalgebra $\widehat{\mathfrak{b}}$ of $\widehat{\mathfrak{g}}$: \begin{equation}\label{hdecomp} \widehat{\mathfrak{h}}=\lie h\oplus\mathbb{C} K\oplus\mathbb{C} d,\quad \widehat{\mathfrak{b}}=\lie b\oplus\mathbb{C} K\oplus\mathbb{C} d\oplus\lie g\otimes_\mathbb{C} t\mathbb{C}[t] \end{equation} Denote by $\widehat{R}$ the root system of $\widehat{\mathfrak{g}}$ and let $\widehat{R}^+$ be the subset of positive roots. The positive non-divisible imaginary root in $\widehat{R}^+$ is denoted $\delta$. Let $\widehat{I}=\{ 0,1,\cdots, n\}$ and $\alpha_0=\delta-\Theta$. Then the set of simple roots of $\widehat{R}$ are $\{\alpha_i| \ i\in \widehat{I}\}$. Let $\{\Lambda_i| \ i\in \widehat{I}\}$ be the corresponding fundamental weights. \vskip 6pt \noindent The decomposition of $\widehat{\mathfrak{h}}$ in (\ref{hdecomp}) has its corresponding version for the dual space $\widehat{\mathfrak{h}}^*$: \begin{equation}\label{hveedecomp} \widehat{\mathfrak{h}}^* = \lie h^*\oplus\mathbb{C}\Lambda_0\oplus\mathbb{C} \delta, \end{equation} here the elements of $\lie h^*$ are extended trivially, $\langle\Lambda_0,\lie h\rangle =\langle\Lambda_0,d\rangle=0$, $\langle\Lambda_0,K\rangle=1$, $\langle\delta,\lie h\rangle=\langle\delta,K\rangle=0$ and $\langle\delta,d\rangle=1$. Let $\{h_i|\ i\in \widehat{I}\}\subset \widehat{\mathfrak{h}}$ be the corresponding basis of the coroot system, then $h_0=K-h_\Theta$. Set $\widehat{\mathfrak{h}}_\mathbb{R}^*=\mathbb{R}\delta+\sum_{i=0}^n\mathbb{R}\Lambda_i$, then we have $\lie h_\mathbb{R}^* \subseteq \widehat{\mathfrak{h}}_\mathbb{R}^*$. The affine Weyl group $\widehat{W}$ is generated by the reflections $\{s_i|\ i\in \widehat{I} \}$, where again $s_i=s_{\alpha_i}$ for a simple root. The cone $\widehat{C} = \{ \Lambda\in\widehat{\mathfrak{h}}_\mathbb{R}^* |\langle\Lambda,\alpha_i^\vee\rangle\ge 0, i\in \widehat{I}\}$ is the fundamental Weyl chamber for $\widehat{\mathfrak{g}}$. \subsection{} Let $\widehat{P}$ be the weight lattice of $\widehat{\mathfrak{g}}$, let $\widehat{P}^+$ be the subset of dominant weights and let $\mathbb{Z}[\widehat{P}]$ be the group algebra of $\widehat{P}$. We recall the following properties of $\delta$: \begin{equation}\label{Kaczitat} \langle\delta, h_i\rangle=0\,\ \text{for \ all}\, i\in \widehat{I},\quad w(\delta)=\delta \,\ \text{for \ all} \, w\in \widehat{W},\quad \text{and} \ \langle\alpha_0, h_i\rangle=-\langle\Theta, h_i\rangle\,{\rm for\ }i\ge 1. \end{equation} We have a non--degenerate symmetric bilinear form $(\cdot,\cdot)$ on $\widehat{\mathfrak{h}}$ and the corresponding isomorphism $\nu:\widehat{\mathfrak{h}}\rightarrow\widehat{\mathfrak{h}}^*$ maps $$ \nu(h_i)=d_i\alpha_i,\quad \nu(K)=\delta,\quad \nu(d)=\Lambda_0. $$ Denote by $\lie g_{\rm sc}$ the subalgebra of $\widehat{\mathfrak{g}}$ generated by $\lie g$ and $h_0=K-h_\Theta$, then $\lie h_{\rm sc}=\lie h\oplus\mathbb{C} K$ is a Cartan subalgebra of ${\lie g}_{\rm sc}$. The inclusion ${\lie h}_{\rm sc}\rightarrow\widehat{\mathfrak{h}}$ induces an epimorphism $\widehat{\mathfrak{h}}^*\rightarrow \lie h_{\rm sc}^*$ with one dimensional kernel. Now (\ref{Kaczitat}) implies that we have in fact an isomorphism $$ \widehat{\mathfrak{h}}^*/\mathbb{C}\delta\rightarrow \lie h_{\rm sc}^* $$ and we set $\lie h_{{\rm sc},\mathbb{R}}^*=\widehat{\mathfrak{h}}_\mathbb{R}^*/\mathbb{R}\delta$. Since $\mathbb{R}\delta\subset \widehat{C}$, we use the same notation $\widehat{C}$ for the image in $\lie h_{{\rm sc},\mathbb{R}}^*$. In the following we are mostly interested in characters of $\lie g$--modules respectively $\lie g_{\rm sc}$--modules obtained by restriction from $\widehat{\mathfrak{g}}$--modules, so we consider also the ring $$ \mathbb{Z}[P_{\rm sc}]:=\mathbb{Z}[\widehat{P}]/I_\delta, $$ where $I_\delta=(1-e^\delta)$ is the ideal in $\mathbb{Z}[\widehat{P}]$ generated by $(1-e^\delta)$. \subsection{The extended affine Weyl group} The group $\widehat{W}$ can be defined as the subgroup of $GL(\lie h_{{\rm sc},\mathbb{R}}^*)$ generated by the induced reflections $\{s_i| \ i\in \widehat{I} \}$, since it fixes $\delta$. Let $M\subset L\subset \lie h^*_\mathbb{R}$ be the lattices $M=\bigoplus_{i=1}^n\mathbb{Z}d_ih_i$ and $L=\bigoplus_{i=1}^n\mathbb{Z}d_i\omega_i$. An element $\Lambda\in \lie h_{{\rm sc},\mathbb{R}}^*$ can be uniquely decomposed into $\Lambda=\lambda +b\Lambda_0$ such that $\lambda\in \lie h^*_\mathbb{R}$. For an element $\mu\in L$ let $t_\mu\in GL(\lie h_{{\rm sc},\mathbb{R}}^*)$ be the map defined by \begin{equation}\label{muaction} \Lambda =\lambda +b\Lambda_0\mapsto t_\mu(\Lambda)=\lambda +b\Lambda_0 + b\mu=\Lambda+\langle \Lambda, K\rangle\mu. \end{equation} Obviously we have $t_{\mu}\circ t_{\mu'}=t_{\mu+\mu'}$, denote $t_L$ and $t_M$ the abelian subgroup of $GL(\lie h_{{\rm sc},\mathbb{R}}^*)$ consisting of the elements $t_\mu$, $\mu\in L$ and $M$ respectively. Then $\widehat{W}$ is the semidirect product $\widehat{W}= W{\,\rule[.1pt]{.4pt}{5.3pt}\hskip-1.9pt\times} t_M$ and the {\it extended affine Weyl group} $\widetilde{W}$ is the semidirect product $\widetilde{W}= W{\,\rule[.1pt]{.4pt}{5.3pt}\hskip-1.9pt\times} t_L$. Let $\Sigma$ be the subgroup of $\widetilde{W}$ stabilizing the dominant Weyl chamber $\widehat{C}$: $$ \Sigma=\{\sigma\in\widetilde{W}\mid \sigma(\widehat{C})=\widehat{C}\}. $$ Then $\Sigma$ provides a complete system of coset representatives of $\widetilde{W}/\widehat{W}$ and $\widetilde{W}=\Sigma{\,\rule[.1pt]{.4pt}{5.3pt}\hskip-1.9pt\times}\widehat{W}$. We extend the length function $\ell:\widehat{W}\rightarrow \mathbb{Z}_+$ to a length function $\ell:\widetilde{W}\rightarrow \mathbb{Z}_+$ by setting $\ell(\sigma w)=\ell(w)$ for $w\in \widehat{W}$ and $\sigma\in \Sigma$.\fi \subsection{} For a given Lie algebra $\lie a$, we let $\bu(\lie a)$ be its universal enveloping algebra. The current Lie algebra of $\lie g$ is denoted by $\lie g[t]$. As a vector space it is just $\lie g\otimes \mathbb{C}[t]$ and the Lie bracket is defined in the natural way: $[a\otimes f, b\otimes g]=[a,b]\otimes f g$, for all $a,b\in\lie g$ and $f,g\in\mathbb{C}[t]$. Denote by $\lie g[t]_+$ the ideal $\lie g\otimes t\mathbb{C}[t]$ and we freely identify $\lie g$ with the Lie subalgebra $\lie g\otimes 1$ of $\lie g[t]$. We clearly have the isomorphism of vector spaces $$\lie g[t]=\lie g[t]_+\bigoplus\lie g, \qquad \bu(\lie g[t])\cong \bu(\lie g[t]_+)\otimes\bu(\lie g).$$ The degree grading on $\mathbb{C}[t]$ defines a natural $\mathbb{Z}_+$--grading on $\lie g[t]$ and hence also on $\bu(\lie g[t])$: an element of the form $(a_1\otimes t^{r_1})\cdots (a_s\otimes t^{r_s})$ has grade $r_1+\cdots+r_s$. Denote by $\bu(\lie g[t])[r]$ the subspace of grade $r$. \subsection{} A representation $V$ of $\lie g[t]$ is {\em graded} if it admits a $\mathbb{Z}$--graded vector space decomposition which admits a compatible Lie algebra action of $\lie g[t]$, i.e., $$V=\bigoplus_{r\in\mathbb{Z}}V[r],\qquad(\lie g\otimes t^s)V[r]\subset V[r+s],\ \ r\in\mathbb{Z}, \ \ s\in\mathbb{Z}_+.$$ A morphism between two graded $\lie g[t]$--modules, by definition, is a degree zero morphism of $\lie g[t]$--modules. Let $M$ be an $\lie g$--module and $z\in\mathbb{C}$. Define a new $\lie g[t]$--module structure on $M$ with respect to $z$ by: $(a\otimes t^r)m=z^ram$, and denote this module as $\ev_z M$. It is easy to see that $\ev_zM$ is an irreducible $\lie g[t]$--module if and only if $M$ is an irreducible $\lie g$--module. Moreover the module $\ev_0M$ is a graded $\lie g[t]$--module and $\left(\ev_0M\right)[0]=M$ and in particular, $\lie g[t]_+(\ev_0M) =0$. \subsection{} Given $\mu\in P^+$, we let $V(\mu)$ be the irreducible finite--dimensional $\lie g$--module generated by an element $v_\mu$ with defining relations $$x_i^+v_\mu=0,\ \ h_iv_\mu=\mu(h_i)v_\mu,\ \ (x_i^-)^{\mu(h_i)+1}v_\mu=0,\ \ i\in I.$$ It is well--known that any finite--dimensional $\lie g$--module $V$ is isomorphic to a direct sum of irreducible modules $V(\mu)$, $\mu\in P^+$. Further, we have a weight space decomposition with respect to $\lie h$, $$V=\bigoplus_{\nu\in P} V_\nu,\ \ \ \ V_\nu=\{v\in V: hv=\nu(h)v,\ \ h\in\lie h\},$$ and we set $\operatorname{wt} V=\{\nu\in P: V_\nu\ne 0\}$. \subsection{Affine Lie algebra} Let $\widehat{\mathfrak{g}}$ be the untwisted affine Kac--Moody algebra corresponding to the extended Dynkin diagram of $\lie g$. $$ \widehat{\mathfrak{g}}=\lie g\otimes_\mathbb{C}\mathbb{C}[t,t^{-1}]\oplus \mathbb{C} K\oplus \mathbb{C} d $$ Here $K$ is the canonical central element and $d$ denotes the derivation $d=t\frac{d}{dt}$. Naturally we consider the Lie algebra $\lie g$ as a subalgebra of $\widehat{\mathfrak{g}}$. In the same way, $\lie h$ and $\lie b$ are subalgebras of the Cartan subalgebra $\widehat{\mathfrak{h}}$ respectively the Borel subalgebra $\widehat{\mathfrak{b}}$ of $\widehat{\mathfrak{g}}$: \begin{equation*} \widehat{\mathfrak{h}}=\lie h\oplus\mathbb{C} K\oplus\mathbb{C} d,\quad \widehat{\mathfrak{b}}=\lie b\oplus\mathbb{C} K\oplus\mathbb{C} d\oplus\lie g\otimes_\mathbb{C} t\mathbb{C}[t] \end{equation*} Let $\delta$ be the positive non--divisible imaginary root of $\widehat{\mathfrak{g}}$ with respect to $\widehat{\mathfrak{b}}$ and $\widehat{P}^+$ be the set of dominant weights of $\widehat{\mathfrak{g}}$. Denote $\widehat{W}$ (resp. $\widetilde{W}$) the (resp. extended) affine Weyl group of $\widehat{\mathfrak{g}}$. Then we have $\widetilde{W}=\Sigma{\,\rule[.1pt]{.4pt}{5.3pt}\hskip-1.9pt\times}\widehat{W}$, where $\Sigma$ is the group of Dynkin diagram automorphisms. We let $\Lambda_0$ be the fundamental weight corresponding to the extra node in the extended Dynkin diagram of $\lie g$. We denote $V(\Lambda)$ by the irreducible $\widehat{\mathfrak{g}}$--highest weight module of highest weight $\Lambda \in \widehat{P}^+$. For an element $\mu \in\bigoplus_{i=1}^n\mathbb{Z}d_i\omega_i$, denote by $t_\mu\in \widetilde{W}$ be the corresponding translation. In particular, we have $$t_{\mu}(b\Lambda_0+\lambda)=b\Lambda_0+\lambda+b\mu \ \ \ (\emph{mod} \ \mathbb{C}\delta)$$ for $\lambda\in P^+$ and $b\in \mathbb{C}.$ \subsection{} We conclude this section with the following simple lemma which will be needed later (see \cite[Section 1.3]{chve}), \begin{lem}\label{dalpha} Suppose that $\lambda\in P^+$ is such that $\lambda=\sum_{i\in I}d_is_i\omega_i$ for some $s_i\in \mathbb{Z}_+$. Then for $\alpha\in R^+$, there exists $s_\alpha\in \mathbb{Z}_+$ such that $\lambda(h_\alpha) =d_\alpha s_\alpha$.\end{lem} \section{Weyl and Demazure Modules}\label{weyl} In this section, we recall the definition of local Weyl modules and Demazure modules and state the required results for $\lie g$--stable Demazure modules. \subsection{Weyl Modules}\label{localweylxi} The definition of the local Weyl modules was given originally in \cite{CPweyl} and later in \cite{CFK} and \cite{FL}. Given $\lambda\in P^+,$ the local Weyl module $W_{\operatorname{loc}}(\lambda)$, is the cyclic $\lie g[t]$--module generated by an element $w_\lambda$ with the following defining relations: for $i\in I$ and $s\in\mathbb{Z}_+$, \begin{gather}\label{locweylg}(x^+_i\otimes\mathbb{C}[t])w_\lambda=0,\ \ (h_i\otimes t^s)w_\lambda=\lambda(h_i)\delta_{s,0}w_{\lambda},\\ \label{integloc} (x_i^-\otimes 1)^{\lambda(h_i)+1}w_\lambda=0.\end{gather} It is easy to see that $\operatorname{wt} W_{\operatorname{loc}}(\lambda)\subset\lambda-Q^+$ and that $\dim W_{\operatorname{loc}}(\lambda)_\lambda=1$, hence $W_{\operatorname{loc}}(\lambda)$ is an indecomposable module. Note {{that $W_{\operatorname{loc}}(0)$ is isomorphic to the trivial $\lie g[t]$--module}}. It was proved in \cite{CPweyl} (see also \cite{CFK}) that the local Weyl modules are finite--dimensional and so, in particular, we have \begin{equation}\label{intega}(x^-_\alpha\otimes 1)^{\lambda(h_\alpha)+1}w_\lambda=0,\ \ \ \alpha\in R^+.\end{equation} We declare the grade of $w_\lambda$ to be zero, so that the local Weyl module is graded by $\mathbb{Z}_+$. It follows that $W_{\operatorname{loc}}(\lambda)[0]\cong_{\lie g} V(\lambda),$ and moreover, $\ev_0 V(\lambda)$ is the unique graded irreducible quotient of $W_{\operatorname{loc}}(\lambda)$. \vskip 6pt \subsection{Demazure Modules} Let us begin with the traditional definition of Demazure modules. For $\Lambda \in \widehat{P}^+$ and $w\in\widehat{W}$, the $U(\widehat{\mathfrak{b}})$--submodule $V_w(\Lambda)=U(\widehat{\mathfrak{b}})V(\Lambda)_{w\Lambda}$ is the {\it Demazure submodule} of $V(\Lambda)$ associated to $w$. More generally, one can associate a Demazure module with any element of $\widetilde{W}$ as follows: For $\Lambda\in \widehat{P}^+$, $\sigma \in \Sigma$ and $w\in \widehat{W}$, define $$V_{\sigma w}(\Lambda)=V_{\sigma w\sigma^{-1}}(\sigma(\Lambda))\ \ \text{and} \ \ V_{w\sigma}(\Lambda)=V_w(\sigma(\Lambda)).$$ In this article, we are mainly interested in the $\lie g$--stable Demazure modules, which are in particular modules for the associated current algebra $\lie g[t]$. It is known that $V_w(\Lambda)$ is $\lie g$--stable if and only if $w\Lambda\in -P^++\ell\Lambda_0+\mathbb{Z}\delta$, where $\ell$ is the level of $\Lambda$. For $\lambda\in P^+$, $\ell\in \mathbb{N}$ and $m\in \mathbb{Z}$, there exists a unique $\Lambda\in \widehat{P}^+$ such that $w_0\lambda+\ell\Lambda_0+m\delta\in \widehat{W}\Lambda.$ For an element $w\in \widehat{W}$ such that $w\Lambda=w_0\lambda+\ell\Lambda_0+m\delta,$ we write $D(\ell, \lambda)[m]=V_w(\Lambda).$ We are indeed interested only in $\lie g[t]$--module structure of $D(\ell, \lambda)[m]$ and it is easy to see that the $\lie g[t]$--structure of $D(\ell, \lambda)[m]$ is independent of $m$. So it makes sense to denote $D(\ell, \lambda)$ by these isomorphism class of $\lie g[t]$--modules. When $\ell=0$, we simply take $\lambda=0$ and $D(\ell, \lambda)$ is the trivial module in this case. \vskip 6pt Now we give one of the equivalent definition of $D(\ell, \lambda)$ (see \cite[Proposition 3.6]{Naoi}) which is appropriate to our study. We refer the reader to \cite{FoL} and \cite{Naoi} for more details. The $\lie g[t]$--graded Demazure module $D(\ell,\lambda)$ of level $\ell\in\mathbb{Z}_+$ and weight $\lambda\in P^+$ is the graded quotient of $W_{\operatorname{loc}}(\lambda)$ generated by the elements, \begin{gather} \label{demrelo} \{(x^-_\alpha\otimes t^p)^{r+1}w_\lambda:\ \ p\in\mathbb{Z}_+,\ \ r\ge\max\{0,\lambda(h_\alpha)-d_\alpha\ell p\},\ \ \text{for} \ \alpha\in R^+\}. \end{gather} Again, we have $D(\ell,\lambda)[0]\cong_{\lie g} V(\lambda)$. The defining relations of these modules are greatly simplified in \cite{chve}. We first fix some notations to state the results of \cite{chve}. Let $\ell\in\mathbb{Z}_+$, $\lambda\in P^+$ be given. For $\alpha\in R^+$, with $\lambda(h_\alpha)>0$, let $s_\alpha,m_\alpha\in\mathbb{N}$ be the unique positive integers so that $$\lambda(h_{\alpha})=(s_{\alpha}-1)d_{\alpha}\ell+m_{\alpha},\ \ \ 0<m_{\alpha}\le d_{\alpha}\ell,$$ where we recall that $d_\alpha=2/(\alpha,\alpha)$. If $\lambda(h_\alpha)=0$ set $s_\alpha=0=m_\alpha$. With the notations above, we have the following theorem (see \cite[Theorem 2, Section 3.5]{chve}): \begin{thm}\label{genreldem1} Let $\ell\in\mathbb{Z}_+$ and $\lambda\in P^+$. Then $D(\ell,\lambda)$ is the quotient of $W_{\operatorname{loc}}(\lambda)$ by the submodule generated by the elements \begin{gather}\label{demrel00}\{ (x^-_{\alpha}\otimes t^{s_{\alpha}})w_\lambda: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t^{s_{\alpha}-1})^{m_{\alpha}+1}w_\lambda: \alpha\in R^+, \ m_\alpha<d_\alpha\ell\}\ .\end{gather}\end{thm} \vskip 6pt We record the following simple fact for future use. \begin{cor} Let $\ell\in\mathbb{Z}_+$. For $\lambda\in \mathcal{P^+_\ell}$, $D(\ell,\lambda)\cong \ev_0 V(\lambda)$ as $\lie g[t]$--modules. \end{cor} \begin{proof} The condition $\ell \Lambda_0+\lambda\in \widehat{P}^+$ implies that $\lambda(h_\Theta)\le \ell$. Now we have $\lambda(h_\alpha)\le d_\alpha\lambda(h_\Theta)\le d_\alpha \ell$, for all $\alpha\in R^+$ and hence $s_\alpha\le 1$, for all $\alpha\in R^+$. Now it is immediate from Theorem \ref{genreldem1} that, $D(\ell,\lambda)\cong {\rm ev_0} V(\lambda)$ as $\lie g[t]$--modules. \end{proof} \section{Fusion Product Structure of Demazure modules}\label{fusion} Let us begin by recalling the definition of fusion products of $\lie g[t]$--modules given in \cite{FL}. The main result of this section is Theorem \ref{mapsdem} on the fusion product of an irreducible module and the Demazure modules of the same level. \subsection{} Let $V$ be a cyclic $\lie g[t]$--module generated by an element $v\in V $. We define a filtration $\{F^rV\}_{r\in\mathbb{Z}_+}$ on $V$ as follows: $$F^rV= \sum_{0\le s\le r}\bu(\lie g[t])[s]v.$$ Clearly each $F^rV$ is a $\lie g$--module. If we set $F^{-1}V=0$, then the associated graded space $\operatorname{gr} V=\bigoplus_{r=0}^{\infty} F^rV/F^{r-1}V$ acquires a natural structure of a cyclic graded $\lie g[t]$--module with the following action: $$(x\otimes t^s)(\overline w)=\overline{(x\otimes t^s)w},\ \ \ \overline w\in F^rV/F^{r-1}V,$$ where $\overline w$ is the image of $w$ in $\operatorname{gr} V$. Moreover, $\operatorname{gr} V\cong V$ as $\lie g$--modules and $\operatorname{gr} V$ is the cyclic $\lie g[t]$--module generated by $\bar v$. The following lemma is simple and very useful. \begin{lem}\label{elemfusion} Suppose that $V$ is a cyclic $\lie g[t]$--module generated by $v\in V$. Then for all $u\in V$, $x\in\lie g$, $r\in\mathbb{N}$, $a_1,\dots, a_r\in\mathbb{C}$, we have $$(x\otimes t^r)\overline u= \overline{(x\otimes (t-a_1)\cdots (t-a_r)) u}.$$\hfill\qedsymbol \end{lem} \subsection{} Let $V$ be a $\lie g[t]$--module and $z\in\mathbb{C}$. We define a new $\lie g[t]$--module action on $V$ as follows: $$(x\otimes t^r)v=(x\otimes (t+z)^r)v,\ \ x\in \lie g,\ \ r\in\mathbb{Z}_+,\ \ v\in V.$$ We denote this new module by $V^z$. Let $V_1,\dots, V_m$ be finite--dimensional graded $\lie g[t]$--modules generated by elements $v_j$, $1\le j\le m$ and let $z_1,\dots,z_m$ be distinct complex numbers. Let $$\bv= V_1^{z_1}\otimes\cdots\otimes V_m^{z_m},$$ be the corresponding tensor product of $\lie g[t]$--modules. It is easily checked (see \cite[Proposition 1.4]{FL}) that the module $\bv$ is cyclic $\lie g[t]$--module and generated by the element $v_1\otimes \cdots\otimes v_m.$ \subsection{} The corresponding associated graded $\lie g[t]$--module $\operatorname{gr} \bv$ is defined to be the fusion product of $V_1,\dots ,V_m$ with respect to the parameters $z_1,\dots, z_m$ and denoted by $V_1^{z_1}*\cdots *V_m^{z_m}$. Clearly $V_1^{z_1}*\cdots *V_m^{z_m}$ is generated by the image of $v_1\otimes\cdots\otimes v_m$. In what follows next, for ease of notation, we suppress the dependence of the fusion product on the parameters and just write $V_1*\cdots*V_m$ instead $V_1^{z_1}*\cdots *V_m^{z_m}$. But unless explicitly stated, it should be assumed that the fusion product does depend on these parameters. Given elements $w_s\in V_s$, $1\le s\le m$, we shall denote by $w_1*\cdots *w_m\in V_1*\cdots*V_m $ the image of the element $w_1\otimes\cdots \otimes w_m\in V_1^{z_1}\otimes\cdots\otimes V_m^{z_m}$. Here, we record the following simple lemma (see \cite[Section 4.3]{chve}) for future use. \begin{lem}\label{fusionweyl} Let $\lambda_s\in P^+$ and $V_s$ be a $\lie g[t]$--module quotient of $W_{\operatorname{loc}}(\lambda_s)$ for $1\le s\le m$. Then $V_1*\cdots*V_m$ is a graded $\lie g[t]$--module quotient of $W_{\operatorname{loc}}(\lambda)$, where $\lambda=\sum_{s=1}^m\lambda_s$.\end{lem} \subsection{} We recall the definition of $\mathcal{P^+_\ell}=\{\lambda\in P^+: \ell \Lambda_0+\lambda\in \widehat{P}^+\}$. Now it is convenient to define a subset $\Gamma$ of $P^+$ as follows: $$\Gamma=\left\{\lambda\in P^+:\ \ \lambda=\sum_{i\in I}d_is_i\omega_i\right\}$$ The following is the statement of our main theorem: \begin{thm}\label{mapsdem} Let $\mu\in\Gamma$, $\lambda\in P^+$, $\ell\in \mathbb{N}$ and suppose that there exists $\mu_j\in\Gamma$, $p_j\in \mathbb{N}$, $1\le j\le m$ such that $$\ell\mu=p_1\mu_1+\cdots+p_m\mu_m, \qquad \mu(h_\alpha)\ge\sum_{j=1}^m\mu_j(h_\alpha),\ \ \alpha\in R^+.$$ There exists a non--zero surjective map of graded $\lie g[t]$--modules, $$D(\ell, \ell\mu+\lambda)\to D(p_1, p_1\mu_1)*\cdots*D(p_m,p_m\mu_m)*D(\ell,\lambda)\to 0,$$ Further more if we assume that $\lambda\in \mathcal{P^+_\ell}$ and $p_1=\cdots=p_m=\ell$ then we have an isomorphism $$D(\ell,\ell \mu+\lambda)\cong D(\ell, \ell \mu_1)*\cdots*D(\ell, \ell \mu_m)*D(\ell,\lambda)\cong D(\ell,\ell \mu_1)*\cdots*D(\ell, \ell \mu_m)*V(\lambda)$$ and, in particular, we have an isomorphism of $\lie g[t]$--modules $D(\ell, N\ell\Theta+\lambda)\cong D(\ell, \ell\Theta)^{\ast N}\ast V(\lambda)$. \end{thm} \begin{cor} The fusion product of a finite number of modules of the form $D(\ell,\ell \mu)$, $\mu\in\Gamma$ for a fixed $\ell$, and $V(\lambda)$ is independent of the choice of parameters if $\lambda\in \mathcal{P^+_\ell}$. \end{cor} \begin{rem} The special case when $\lambda=0$ of the theorem (second statement) and the corollary was proved earlier in \cite[Section 3.5]{FoL} using results of \cite{FoL2}. We note here, that these papers work with a special family of Demazure modules $D(\ell,\lambda)$. In the notation of the current paper, they only work with the modules of the form $D(\ell,\ell\mu)$, $\mu\in \Gamma$. We remark here that the condition $\ell\Lambda_0+\lambda\in \widehat{P}^+$ is equivalent to $\lambda(h_\Theta)\le \ell$. Thus we have infinite family of such examples. \end{rem} \vskip 6pt \subsection{} We need some results on the $\lie g$--structure of Demazure modules to prove the second statement of our main theorem \ref{mapsdem}. We state the required proposition here and defer its proof to section \ref{affine}. Its proof uses ideas that evolved in \cite[Section 3]{FoL2} to prove similar results. \begin{prop}\label{demprop} Let $\ell\in \mathbb{N}$, $\mu\in\Gamma$ and $\lambda\in \mathcal{P^+_\ell}$. Suppose that there exists $\mu_j\in\Gamma$, $1\le j\le m$, such that $\mu=\mu_1+\cdots+\mu_m, $ then we have \begin{enumerit} \item[(i)] $D(\ell,\ell\mu+\lambda)\cong D(\ell, \ell \mu_1)\otimes \cdots \otimes D(\ell, \ell \mu_m)\otimes D(\ell,\lambda)$ as $\lie g$--modules, \item[(ii)] in particular, $\dim D(\ell,\ell\mu+\lambda)= \dim D(\ell, \ell \mu_1)\cdots \dim D(\ell, \ell \mu_m) \dim D(\ell,\lambda).$\hfill\qedsymbol \vskip 6pt \end{enumerit} \end{prop} \noindent{\em{Proof of Theorem \ref{mapsdem}.}} \ Let $v_s\in D(p_s,p_s\mu_s)$ be the image of the generator $w_{\mu_s}$ of $W_{\operatorname{loc}}(p_s\mu_s)$ for $1\le s\le m$, and $v_{m+1}\in D(\ell,\lambda)$ the image of the generator $w_{\lambda}$ of $W_{\operatorname{loc}}(\lambda)$. Using Lemma \ref{fusionweyl}, we see that there exists a surjective map of graded $\lie g[t]$--modules, $$W_{\operatorname{loc}}(\ell \mu+\lambda)\to D(p_1,p_1\mu_1)*\cdots*D(p_m,p_m\mu_m)*D(\ell,\lambda)\to 0.$$ For each $\alpha\in R^+$, write $\lambda(h_\alpha)=(r_\alpha-1)d_\alpha \ell+m_\alpha$ with $m_\alpha\le d_\alpha \ell$ and $\mu(h_\alpha)=d_\alpha s_\alpha$ (use Lemma \ref{dalpha}). Now using Theorem \ref{genreldem1}, it suffices to show that, \begin{gather}\label{demrel} (x_\alpha^-\otimes t^{s_\alpha+r_\alpha})(v_1*\cdots *v_m*v_{m+1})=0,\ \alpha\in R^+,\\ (x^-_{\alpha}\otimes t^{s_{\alpha}+r_\alpha-1})^{m_{\alpha}+1}(v_1*\cdots *v_m*v_{m+1})=0,\ \alpha\in R^+, \ m_\alpha<d_\alpha\ell.\end{gather} Write $\mu_j(\alpha)=d_\alpha s_{\alpha,j},\ \ 1\le j\le m,$ (use Lemma \ref{dalpha}) and note that we are given that $$s_\alpha\ge\sum_{j=1}^m s_{\alpha,j},\ \ \alpha\in R^+.$$ Setting $b_\alpha=s_\alpha-\sum_j s_{\alpha,j}$ and taking $z_1,\dots,z_m, z_{m+1}$ be the parameters involved in the definition of the fusion product, we see that \begin{gather*}(x^-_\alpha\otimes t^{b_\alpha}(t-z_1)^{s_{\alpha,1}}\cdots(t-z_m)^{s_{\alpha,m}}(t-z_{m+1})^{r_\alpha})(v_1\otimes\cdots \otimes v_m\otimes v_{m+1}) \\ ={\sum_{r=1}^{m+1}v_1\otimes\cdots\otimes(x^-_\alpha\otimes (t+z_r)^{b_\alpha}(t-z_1+z_r)^{s_{\alpha,1}}\cdots t^{s_{\alpha,r}}\cdots (t+z_r-z_m)^{s_{\alpha,m}}(t+z_r-z_{m+1})^{r_\alpha}v_r)\otimes\cdots\otimes v_{m+1}}.\end{gather*} Using Theorem \ref{genreldem1}, we see that the relations $(x_\alpha^-\otimes t^b)v_r=0,\ \ b\ge s_{\alpha,r},$ hold in $D(p_r,\mu_r)$ for $1\le r\le m$, and $(x_\alpha^-\otimes t^b)v_{m+1}=0, \ \ b\ge r_\alpha$, hold in $D(\ell,\lambda)$. \vskip 6pt \noindent Hence we have $(x^-_\alpha\otimes t^{b_\alpha}(t-z_1)^{s_{\alpha,1}}\cdots(t-z_m)^{s_{\alpha,m}}(t-z_{m+1})^{r_\alpha})(v_1\otimes \cdots \otimes v_m\otimes v_{m+1})=0. $ Now it follows that $$(x^-_\alpha\otimes t^{s_\alpha+r_\alpha})(v_1\otimes \cdots \otimes v_m\otimes v_{m+1})\in F^{s_\alpha+r_\alpha-1}(\bold V),$$ where $\bold V=D^{z_1}(\ell, \ell \mu_1)\otimes \cdots \otimes D^{z_m}(\ell, \ell \mu_m)\otimes D^{z_{m+1}}(\ell,\lambda).$ This implies $$(x_\alpha^-\otimes t^{s_\alpha+r_\alpha})(v_1*\cdots*v_m*v_{m+1})=0$$ We now prove the second relation. The earlier calculation shows that, \begin{gather*} (x^-_\alpha\otimes t^{b_\alpha}(t-z_1)^{s_{\alpha,1}}\cdots(t-z_m)^{s_{\alpha,m}}(t-z_{m+1})^{r_\alpha-1})^{m_\alpha+1}(v_1\otimes \cdots \otimes v_m\otimes v_{m+1})=\\ v_1\otimes\cdots \otimes v_m\otimes (x^-_\alpha\otimes (t+z_{m+1})^{b_\alpha}(t-z_1+z_{m+1})^{s_{\alpha,1}}\cdots (t+z_{m+1}-z_m)^{s_{\alpha,m}}t^{r_\alpha-1})^{m_\alpha+1} v_{m+1}. \end{gather*} But using Theorem \ref{genreldem1} again, we have the following relations in $D(\ell, \lambda)$ $$(x^-_\alpha\otimes t^{b+r_\alpha-1})^{m_\alpha+1}v_{m+1}=0, \ \text{for} \ b\ge 0.$$ which yields $(x^-_\alpha\otimes t^{b_\alpha}(t-z_1)^{s_{\alpha,1}}\cdots(t-z_m)^{s_{\alpha,m}}(t-z_{m+1})^{r_\alpha-1})^{m_\alpha+1}(v_1\otimes \cdots \otimes v_m\otimes v_{m+1})=0.$ Again by similar argument, we have $$(x_\alpha^-\otimes t^{s_\alpha+r_\alpha-1})^{m_\alpha+1}(v_1*\cdots*v_m*v_{m+1})=0.$$ This proves the existence of the surjective map $$D(\ell, \ell \mu+\lambda)\to D(p_1, p_1\mu_1)*\cdots*D(p_m,p_m\mu_m)*D(\ell,\lambda)\to 0.$$ Now the second and third statements of the theorem are immediate from the Proposition \ref{demprop}, since the dimension of the corresponding modules matches. \hfill\qedsymbol \section{Proof of Proposition \ref{demprop} and Limit constructions }\label{affine} The first part of this section occupies the proof of Proposition \ref{demprop}. The proof that we present here is a slight generalization of Theorem 1 and Theorem 1 A in [\cite{FoL2}, Section 3.1]. We also remark that special cases of our proposition was already established in \cite{FoL2}. In the second part of this section, we reconstruct the $\lie g[t]$--module structure of the irreducible highest weight $\widehat{\mathfrak{g}}$--module $V(\ell\Lambda_0+\lambda)$ as a direct limit of fusion products of Demazure modules. This limit construction was conjectured by Fourier and Littelmann in \cite{FoL} and they proved the conjecture for the special case when $\lambda=0$. \subsection{} We begin by recalling the Demazure character formula from \cite[Chapter VIII]{skumar}. Denote $D_w$ the Demazure operator associated with an arbitrary element $w\in \widetilde{W}$. For $\Lambda\in \widehat{P}^+$, $\sigma \in \Sigma$ and $w\in \widehat{W}$ we have $$\hbox{\rm Char}\,_{\widehat{\mathfrak{h}}}\ V_{w\sigma}(\Lambda)=\hbox{\rm Char}\,_{\widehat{\mathfrak{h}}}\ V_w(\sigma(\Lambda))=D_{w}(e(\sigma\Lambda))=D_{w\sigma}(e(\Lambda))$$ We note here that we are only interested in $\lie g$--module structure of Demazure modules, and so in particular that we are interested only in $\lie h$--characters and hence it is enough to calculate the Demazure characters modulo $\delta$. We use a few results of \cite{FoL2} to calculate Demazure characters (modulo $\delta$) in what follows next, that allows us to conclude our proposition. We refer the readers to \cite{FoL2} for more details. The following is simple: \begin{lem}\label{length} For $\mu\in \Gamma$, we have $\ell(t_{w_0\mu}w_0)=\ell(t_{w_0\mu})+\ell(w_0)$, where $\ell(-)$ denotes the extended length function of $\widetilde{W}$. Hence, $D_{t_{w_0\mu}w_0}=D_{t_{w_0\mu}}D_{w_0}$. \end{lem} \subsection{Proof of Proposition \ref{demprop}} We have, by definition, $V_{t_{w_0\mu} w_0}(\ell \Lambda_0+\lambda)=D(\ell, \ell \mu+\lambda)$, since $$t_{w_0\mu} w_0(\ell \Lambda_0+\lambda)=\ell \Lambda_0+w_0(\ell \mu+\lambda) \ \ \emph{mod} \ \mathbb{Z}\delta.$$ Since $V_{t_{w_0\nu}}(\ell \Lambda_0)=D(\ell, \ell \nu)$, by definition, for any $\nu\in \Gamma$, it is sufficient to prove the following: $$D_{t_{w_0\mu}w_0}(e(\ell \Lambda_0+\lambda))=e({\ell \Lambda_0})\emph{char}_{\lie h}\overline{V_{t_{w_0\mu_1}}(\ell \Lambda_0)}\cdots \emph{char}_{\lie h}\overline{V_{t_{w_0\mu_m}}(\ell \Lambda_0)}\emph{char}_{\lie h} V(\lambda)$$ where we write $\overline{V_{t_{w_0\mu}}(\ell \Lambda_0)}$ for the Demazure module viewed as $\lie g$--module. Now using the Demazure operators and the Lemma \ref{length}, we get \begin{gather*} D_{t_{w_0\mu}w_0}(e(\ell \Lambda_0+\lambda))=D_{t_{w_0\mu}}D_{w_0}(e({\ell \Lambda_0+\lambda}))=D_{t_{w_0\mu}}(e({\ell \Lambda_0})\emph{char}_{\lie h}V(\lambda)). \end{gather*} Using the Lemma 7 in \cite[Section 3.1]{FoL2} we get $D_{t_{w_0\mu}}(e({\ell \Lambda_0})\emph{char}_{\lie h}V(\lambda))=D_{t_{w_0\mu}}(e({\ell \Lambda_0}))\emph{char}_{\lie h}V(\lambda)$ and using the Theorem 1' in \cite[Section 3.1]{FoL2} we get $$D_{t_{w_0\mu}}(e(\ell \Lambda_0))=e({\ell \Lambda_0})\emph{char}_{\lie h}\overline{V_{t_{w_0\mu_1}}(\ell \Lambda_0)}\cdots \emph{char}_{\lie h}\overline{V_{t_{w_0\mu_m}}(\ell \Lambda_0)}.$$ Putting all these together we get our desired result $$D_{t_{w_0\mu}w_0}(e(\ell \Lambda_0+\lambda))=e({\ell \Lambda_0})\emph{char}_{\lie h}\overline{V_{t_{w_0\mu_1}}(\ell \Lambda_0)}\cdots \emph{char}_{\lie h}\overline{V_{t_{w_0\mu_m}}(\ell \Lambda_0)}\emph{char}_{\lie h}V(\lambda),$$ and hence $$D(\ell, \ell\mu+\lambda)\cong D(\ell, \ell \mu_1)\otimes \cdots \otimes D(\ell, \ell \mu_m)\otimes D(\ell,\lambda),$$ as $\lie{g}$--modules. This completes the proof of the proposition. \subsection{Limit constructions}\label{seclimitconstruction} Fix a non--zero dominant weight $\lambda$ of $\lie g$ and $\ell \in \mathbb{N}$ such that $\lambda(h_\Theta)\le \ell$. In this subsection, we give a proof for the limit construction of the irreducible highest weight $U(\widehat{\mathfrak{g}})$--module $V(\ell \Lambda_0+\lambda)$ as conjectured in \cite{FoL}. Note that, in \cite{FoL2} same authors gave such a construction for $\overline{V(\ell\Lambda_0+\lambda)}$ as a semi-infinite tensor product of finite dimensional $\lie g$--module. We first recall the statement of the semi--infinite fusion product construction: \begin{thm}\label{semiinfconjecture} Let $D(\ell, N\ell\Theta+\lambda)\subset V(\Lambda)$ be the Demazure module of level $\ell$ corresponding to the element $t_{-N\Theta}w_0$ of $\widetilde{W}$. Let $w\neq0$ be a $\lie g[t]$--invariant vector of $D(\ell, \ell\Theta).$ Let $\bold{V}^\infty_{\ell, \lambda}$ be the direct limit of $$V(\lambda)\hookrightarrow D(\ell, \ell\Theta)*V(\lambda)\hookrightarrow D(\ell, \ell\Theta)*D(\ell, \ell\Theta)*V(\lambda)\hookrightarrow D(\ell, \ell\Theta)*D(\ell, \ell\Theta)*D(\ell, \ell\Theta)*V(\lambda)\hookrightarrow \cdots $$ where the inclusions are given by $v\mapsto w\otimes v.$ Then $V(\Lambda) \ \text{and} \ \bold{V}^\infty_{\ell, \lambda}$ are isomorphic as $\lie g[t]$--modules. \end{thm} \begin{proof} Here we follow the ideas of \cite{FoL}. By Theorem \ref{mapsdem} we have, for $z_1\not=z_2 \in \mathbb{C}$ and $N\in \mathbb{Z}_+$, an isomorphism of $\lie g[t]$--modules $$ D(\ell, (N+1)\ell\Theta+\lambda) \cong D(\ell, \ell \Theta)^{z_2} \ast D(\ell, N\ell\Theta+\lambda)^{z_1}.$$ Using this isomorphism of Demazure modules, the assertion can be proved in exactly the same way as \cite[Theorem 9]{FoL}. We refer the readers to \cite{FoL} for more details. \end{proof} \iffalse \vskip 6pt \noindent We first recall a few facts about inclusions of Demazure modules from \cite{FoL}. Set $\tilde{\lie b} = \lie h \oplus \lie n^+ \oplus \lie g\otimes_{\mathbb{C}} t\mathbb{C}[t] \oplus \mathbb{C} K.$ \begin{lem}\label{unique-embedding} Let $\Lambda$ be an integral dominant weight for $\widehat{\mathfrak{g}}$. Given $w\in\widehat{W} /\widehat{W} _\Lambda$, there is a unique (up to scalar multiplication) nontrivial morphism of $U(\tilde{\lie b})$-modules $$ V_{w}(\Lambda) \longrightarrow V(\Lambda). $$ In fact, this morphism is, up to scalar multiples, the canonical embedding of the Demazure module. \end{lem} \begin{cor}\label{unique-morphism-coro} Let $\Lambda\in \widehat{P}$ and $\tau < w$, then there exists (up to scalar multiples) a unique morphism of $U(\tilde{\lie b})$-modules $\iota: V_{\tau}(\Lambda) \longrightarrow V_{w}(\Lambda)$. \end{cor} \subsection{} Consider the Demazure module $D(\ell,N\ell \Theta+\lambda) = V_{t_{-N\Theta}w_0}(\ell \Lambda_0+\lambda)$. We fix a generator $w\neq0$ of the unique $U(\lie g[t])$-fixed line in $D(\ell,\ell \Theta)$. By Theorem \ref{mapsdem} we have for $c_1\not=c_2$ an isomorphism $$ D(\ell, (N+1)\ell\Theta+\lambda) \simeq D(\ell, \ell \Theta)_{c_2} \ast D(\ell, N\ell\Theta+\lambda)_{c_1}. $$ \vskip 6pt We extend this to an isomorphism of $U(\lie g[t]\oplus\mathbb{C} K)$-modules by letting $K$ operate on $D(\ell,(N+1)\ell\Theta+\lambda)$ by the level $\ell$, and letting $K$ act on the second module by $0$ on the first factor and on the second factor by the level $\ell$. Define the map $$ \tilde{\varphi}: D(\ell , N\ell\Theta+\lambda)_{c_1} \longrightarrow D(\ell,\ell \Theta )_{c_2} \otimes D(\ell, N\ell\Theta+\lambda)_{c_1} $$ by $\tilde{\varphi}(v) = w \otimes v$. This map is an $U(\lie g[t])$-module morphism because $w$ is $U(\lie g[t])$-invariant, which extends, as above, to a $U(\lie g[t]\oplus\mathbb{C} K)$-module morphism. \vskip 6pt The map respects the filtrations up to a shift: let $v_\Theta \in D(\ell ,\ell \Theta )$ be a generator and let $q$ be minimal such that $w\otimes v_\Theta \in F^q(D(\ell,\ell \Theta)_{c_2} \otimes D(\ell, N\ell\Theta+\lambda)_{c_1})$. By the $U(\lie g[t])$-equivariance it follows that $$ \tilde{\varphi}(F^j(D(\ell, N\ell\Theta+\lambda )_{c_1})\subseteq F^{j+q}(D(\ell, \ell \Theta )_{c_2} \otimes D(\ell, N\ell\Theta+\lambda)_{c_1}). $$ So we get an induced $U(\lie g[t]\oplus \mathbb{C}K)$-morphism $\varphi$ between the associated graded modules by $\varphi(\overline{v}) = \overline{ w\otimes v}$. $\varphi$ is nontrivial and so by Corollary \ref{unique-morphism-coro} it is (up to multiplication by a scalar) the embedding of Demazure modules $\iota$. We proved: \begin{lem} The map $\varphi: D(\ell, N\ell\Theta+\lambda) \longrightarrow D(\ell, \ell\Theta)_{c_2} \ast D(\ell, N\ell\Theta)_{c_1} \simeq D(\ell,(N+1)\ell\Theta)$ induced by $\varphi(v) = \overline{w \otimes v}$ is an embedding of $U( \tilde{\lie b})$-modules. \end{lem} \noindent It follows by the above: \begin{lem} Let $\lie g$ be a simple Lie algebra.\\ The following is a commutative diagram of $U(\lie g[t])$-modules% $$ \xymatrix{ D(\ell,N\ell\Theta+\lambda) \ar@{^{(}->}[rr]^\iota\ar[dd]_\wr& & D(\ell, (N+1)\ell\Theta+\lambda)\ar[dd]_\wr\\ \\ D(\ell,N\ell\Theta+\lambda)\ar[rr]^\varphi\ar[dd]_\wr & & D(\ell,\ell\Theta)\ast D(\ell, N\Theta+\lambda) \ar[dd]_\wr\\ \\ D(\ell, \ell\Theta)^{\ast N}\ast V(\lambda) \ar[rr]^\varphi&& D(\ell,\ell \Theta)^{\ast (N+1)} \ast V(\lambda)\\ } $$ where the down arrows are the isomorphism of theorem~\ref{mapsdem} \end{lem} Now one knows that $V(\ell \Lambda_0+\lambda) = \underset{N \rightarrow \infty}{\mbox{lim }} D(\ell , N\ell \Theta+\lambda)$ as $U(\lie g[t])$-modules, and also as $U(\tilde{\lie b})$-modules. \fi \section{Application to Schur positivity}\label{schur} In this section, we recall the definition of Kirillov--Reshetikhin modules and its connection with Demazure modules. It has been proved in \cite{CMkir} (see also \cite{FoL} and \cite{chve}) that Kirillov--Reshetikhin modules are indeed Demazure modules. We use this fact to establish the possible connections of our main theorem with the Schur positivity conjecture. We will freely use the notation established in the previous sections. \iffalse \subsection{} Let us begin by recalling a few facts about minuscule fundamental weights of $\lie{g}$. An element $\lambda\in P^+$ is said to be minuscule if $\lambda(h_\alpha) \in \{ 0,1 \}$ for all $\alpha \in R^+$. It can be easily seen that if $\lambda \in P^{+}$ is minuscule, then $\lambda$ must be a fundamental weight. The following is the list of minuscule weights; here, we follow the numbering of vertices of the Dynkin diagram for $\lie{g}$ in \cite{Bour}. \begin{equation*} \begin{array}{ll} A_{n} & \omega_{i},\, 1 \le i \le n \\[1mm] B_{n} & \omega_{n} \\[1mm] C_{n} & \omega_{1} \\[1mm] D_{n} & \omega_{1},\,\omega_{n-1},\,\omega_{n} \\[1mm] E_{6} & \omega_{1},\,\omega_{6} \\[1mm] E_{7} & \omega_{7} \end{array} \end{equation*}\fi \subsection{} We now recall the definition of the Kirillov--Reshetikhin modules from \cite[Section 2]{CMkir}. Thus given $i\in I$ and $m\in\bz_+$, the Kirillov--Reshetikhin module $KR(m\omega_i)$ is the quotient of the $W_{\operatorname{loc}}(m\omega_i)$ by the submodule generated by the element $(x_i^-\otimes t)w_{m\omega_i}$. We shall need the following isomorphism of $\lie g[t]$--modules (see \cite[Proposition 5.1]{chve}). \begin{prop}\label{dem=kr} For $i\in I$ and $\ell\in \mathbb{Z}_+$, we have $D(\ell, d_i\ell\omega_i)\cong KR(d_i\ell\omega_i)$ as $\lie g[t]$--modules. Further more if we assume that $d_i\omega_i(h_\alpha)\le d_\alpha, \ \alpha\in R^+$, then we have $D(\ell, d_i\ell\omega_i)\cong KR(d_i\ell\omega_i)\cong \ev_0V(d_i\ell\omega_i)$ as $\lie g[t]$--modules. \end{prop} \begin{rem}\label{minusculecoweight} We remark here that the isomorphism between the Kirillov--Reshetikhin modules and the Demazure modules was proved earlier in \cite[Section 5]{CMkir} and \cite[Section 3.2]{FoL}. The condition $d_i\omega_i(h_\alpha)\le d_\alpha, \ \alpha\in R^+$ is equivalent to saying that $d_i\omega_i(h_\Theta)\le 1.$ This forces $d_i=1$ for all such $d_i\omega_i.$ Indeed we have the following list of all such $d_i\omega_i$'s: here, we follow the numbering of vertices of the Dynkin diagram for $\lie{g}$ in \cite{Bour}. \begin{equation*} \begin{array}{ll} A_{n} & \omega_{i},\, 1 \le i \le n \\[1mm] B_{n} & \omega_{1} \\[1mm] C_{n} & \omega_{n} \\[1mm] D_{n} & \omega_{1},\,\omega_{n-1},\,\omega_{n} \\[1mm] E_{6} & \omega_{1},\,\omega_{6} \\[1mm] E_{7} & \omega_{7} \end{array} \end{equation*} \end{rem} \subsection{}New interpretations of Theorem \ref{mapsdem} using the isomorphism between Kirillov--Reshetikhin modules and Demazure modules has many interesting consequences. We fix a non--negative integer $\ell$ in what follows next in this section. The following is an immediate consequence of Theorem \ref{mapsdem} and Proposition \ref{dem=kr}. \begin{thm}\label{krdecom} Let $\lambda\in \mathcal{P^+_\ell}$ and $\mu=\sum\limits_{i\in I}d_is_i \omega_i \in \Gamma$. Then we have the following isomorphism of $\lie g[t]$--modules, $$D(\ell, \ell\mu+\lambda)\cong KR(d_1\ell\omega_1)^{*s_1}*\cdots *KR(d_n\ell\omega_n)^{*s_n}*V(\lambda).$$ In particular, the fusion product on the right hand side is independent of the choice of the parameters. \end{thm} \vskip 6pt \noindent We use Theorem \ref{krdecom} to establish an elegant presentation for the module $KR(d_i\ell\omega_i)^{\ast k}\ast V(\lambda)$, $\lambda\in \mathcal{P^+_\ell}$. \begin{prop}\label{demrelfuskr} Fix $i\in I$ and let $k\in \mathbb{N}$ and $\lambda\in \mathcal{P^+_\ell}$. Then the module $KR(d_i\ell \omega_i)^{\ast k}\ast V(\lambda)$ is the quotient of $W_{\operatorname{loc}}(kd_i\ell\omega_i+\lambda)$ by the submodule generated by the elements \begin{gather}\label{demrelschur}\{ (x^-_{\alpha}\otimes t^{ks_{\alpha}+1})w_{kd_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t^{ks_{\alpha}})^{\lambda(h_\alpha)+1}w_{kd_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ ,\end{gather} where $s_\alpha:=\frac{1}{d_\alpha}d_i\omega_i(h_\alpha)$ for all $\alpha\in R^+,$ and hence the module $KR(d_i\ell \omega_i)^{\ast k}\ast V(\lambda)$ is independent of the choice of parameters. \end{prop} \begin{proof} Using Theorem \ref{krdecom}, we get $D(\ell, kd_i\ell \omega_i+\lambda)\cong KR(d_i\ell\omega_i)^{\ast k}*V(\lambda)$ as $\lie g[t]$--modules. Since $\lambda \in \mathcal{P^+_\ell}$, we have $\lambda(h_\alpha)\le d_\alpha \ell$ for all $\alpha\in R^+$, now the proposition is immediate from Theorem \ref{genreldem1}. \end{proof} \iffalse \begin{cor}\label{demrelfusirrd} Let $(\ell, \lambda)\in \widetilde{\Gamma}$ and $\omega_i$ be minuscule. Then the module $V(d_i\ell \omega_i)^{\ast k}\ast V(\lambda)$ is the quotient of $W_{\operatorname{loc}}(kd_i\ell\omega_i+\lambda)$ by the submodule generated by the elements \begin{gather}\label{demrelschur}\{ (x^-_{\alpha}\otimes t^{ks_{\alpha}+1})w_{kd_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t^{ks_{\alpha}})^{\lambda(h_\alpha)+1}w_{kd_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ ,\end{gather} where $s_\alpha:=\frac{1}{d_\alpha}d_i\omega_i(h_\alpha)$ for all $\alpha\in R^+.$ In particular, the module $V(d_i\ell \omega_i)^{\ast k}\ast V(\lambda)$ is independent of the choice of parameters. \end{cor} \begin{proof} It is immediate from propositions \ref{dem=kr} and \ref{demrelfuskr}. \end{proof}\fi \subsection{} When we specialize Theorem \ref{krdecom} to a fundamental weight that is listed in the remark \ref{minusculecoweight}, we get simplified presentation for the fusion products of many interesting family of irreducible representations of $\lie g$. We note here that the fusion products of these special family of irreducible $\lie g$--modules are indeed Demazure modules. In particular we get presentation of 2--fold fusion product $V(\lambda)*V(\mu)$ for many special choices of $\lambda, \mu\in P^+$ using this fact, which allows us to provide further evidence to the conjecture on Schur positivity stated in \cite{chfsa}. For example, we provide presentation for $V(\ell\omega_i)*V(m\omega_j)$ when $\lie g=\lie sl_{n+1}$. We remark here that these presentations were previously known only for some special cases, but not in this generality. For example see \cite{FFq-char} for the case $\lie g=\lie sl_2$, and \cite{chve},\cite{FoL} for the case $\ell=m$ and $\lie g=\lie sl_{n+1}$ general. \begin{prop}\label{2-fold} Let $i\in I$ be such that $d_i\omega_i(h_\Theta)\le 1$ and $\lambda\in \mathcal{P^+_\ell}$. Then the module $V(d_i\ell \omega_i)\ast V(\lambda)$ is the quotient of $W_{\operatorname{loc}}(d_i\ell\omega_i+\lambda)$ by the submodule generated by the elements $$\{ (x^-_{\alpha}\otimes t^{2})w_{d_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t)^{\emph{min}\{d_i\ell \omega_i(h_\alpha),\ \lambda(h_\alpha)\}+1}w_{d_i\ell\omega_i+\lambda}: \alpha\in R^+\}\ .$$ \end{prop} \begin{proof} Let $z_1, z_2\in \mathbb{C}$ be two distinct complex numbers that used in the definition of the fusion product $V(d_i\ell\omega_i)*V(\lambda)$ and let $v_i*v_\lambda$ be the image of $v_{d_i\ell\omega_i}\otimes v_\lambda$ in $V(d_i\ell\omega_i)*V(\lambda)$. Then for any $x\in \lie g$, by the definition of the evaluation representation, we get $(x\otimes (t-z_1)(t-z_2))(v_{d_i\ell\omega_i}\otimes v_{\lambda})=0.$ Thus we get $(x^-_{\alpha}\otimes t^{2})(v_i*v_\lambda)=0$ for any $\alpha\in R^+.$ Now since $(x^-_{\alpha}\otimes t^{2})(v_i*v_\lambda)=0$ implies $(x^-_{\alpha}\otimes t^{3})(v_i*v_\lambda)=0$ for any $\alpha\in R^+,$ the result is immediate from Proposition \ref{demrelfuskr}. \end{proof} \vskip 6pt \noindent \begin{cor} Let $\lie g$ be a simple Lie algebra with a fundamental weight $\omega_i$ such that $d_i\omega_i(h_\Theta)\le 1$. Let $j \in I$ and $\ell, m\in \mathbb{N}$ such that $\ell\ge m.$ We also assume that \begin{itemize} \item if $\lie g=B_n \ \text{and} \ j\neq 1$ then $\ell\ge 2m,$ \item if $\lie g=C_n \ \text{and} \ j\neq n$ then $\ell\ge 2m,$ \item if $\lie g=D_n \ \text{and} \ j\not\in \{1,n-1,n\}$ then $\ell\ge 2m,$ \item if $\lie g=E_6$ and $j\in \{2,3,5\}$ then $\ell\ge 2m$; for $j=4$, assume that $\ell\ge 3m,$ \item if $\lie g=E_7$ and $j\in \{1,2,6\}$ then $\ell\ge 2m$; for $j=3,5$, assume that $\ell\ge 3m;$ for $j=4$, assume that $\ell\ge 4m.$ \end{itemize}Then the module $V(d_i\ell \omega_i)\ast V(d_jm\omega_j)$ is the quotient of $W_{\operatorname{loc}}(d_i\ell\omega_i+d_jm\omega_j)$ by the submodule generated by the elements $$\{ (x^-_{\alpha}\otimes t^{2})w_{d_i\ell\omega_i+d_jm\omega_j}: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t)^{\emph{min}\{d_i\ell\omega_i(h_\alpha),\ d_jm\omega_j(h_\alpha)\}+1}w_{d_i\ell\omega_i+d_jm\omega_j}: \alpha\in R^+\}\ .$$ \end{cor} \begin{proof} It is easy to see that $d_jm\omega_j(h_\Theta)\le \ell$ in all considered cases. Thus the corollary follows from the above Proposition \ref{2-fold} by setting $\lambda=d_jm\omega_j$. \end{proof} \vskip 6pt \noindent \subsection{} Proposition \ref{2-fold} and its corollary motivates us to consider the modules $\bold{V}(\lambda, \mu)$, for $\lambda, \mu\in P^+$ that defined as follows: it is the graded quotient of $W_\emph{loc}(\lambda+\mu)$ generated by the elements $$\{ (x^-_{\alpha}\otimes t^{2})w_{\lambda+\mu}: \alpha\in R^+\}\ \bigcup\ \{ (x^-_{\alpha}\otimes t)^{\emph{min}\{\lambda(h_\alpha),\mu(h_\alpha)\}+1}w_{\lambda+\mu}: \alpha\in R^+\}\ .$$ We remark that these modules studied by Fourier in \cite{Fu} for the case $\lie g=\lie sl_{n+1}$. These modules are naturally related to the conjecture on the generalization of Schur positivity of \cite{chfsa}. In fact it is easy to see that this conjecture is true for these modules. The next lemma tells us its connection with fusion products. \begin{lem}\label{2-foldlem} Let $\lambda, \mu\in P^+$ and $z_1, z_2\in \mathbb{C}$ be two distinct complex numbers. Then there exists a surjective map of $\lie g[t]$--modules $$\bold{V}(\lambda, \mu)\rightarrow \ev_{z_1}V(\lambda)*\ev_{z_2}V(\mu)\rightarrow 0$$ \end{lem} \begin{proof} The lemma easily follows from the definition of the evaluation representation and fusion products. For example, we prove that $(x^-_{\alpha}\otimes t)^{\emph{min}\{\lambda(h_\alpha),\mu(h_\alpha)\}+1}(v_{\lambda}*v_{\mu})=0$ holds in $\ev_{z_1}V(\lambda)*\ev_{z_2}V(\mu)$, where $v_{\lambda}*v_{\mu}$ is the image of $v_{\lambda}\otimes v_\mu$ in $\ev_{z_1}V(\lambda)*\ev_{z_2}V(\mu)$. Indeed, $$(x^-_{\alpha}\otimes (t-z_2))^{\lambda(h_\alpha)+1}(v_{\lambda}\otimes v_{\mu})=(z_1-z_2)^{\lambda(h_\alpha)+1}((x^-_{\alpha})^{\lambda(h_\alpha)+1}v_\lambda\otimes v_\mu)=0.$$ Hence $(x^-_{\alpha}\otimes t)^{\lambda(h_\alpha)+1}(v_{\lambda}* v_{\mu})=0$, and also similar argument shows that $(x^-_{\alpha}\otimes t)^{\mu(h_\alpha)+1}(v_{\lambda}* v_{\mu})=0$. \end{proof} \begin{prop}\label{schurpostivity} Let $i\in I$ be such that $d_i\omega_i(h_\Theta)\le 1$ and $\lambda\in \mathcal{P^+_\ell}$. Suppose that there exists $\mu_1, \mu_2\in P^+$ such that \begin{enumerit} \item[(i)] $d_i\ell\omega_i+\lambda=\mu_1+\mu_2$ and \item[(ii)]\label{schurcond} $\min\{\mu_1(h_\alpha),\mu_2(h_\alpha)\}\le \min\{d_i\ell\omega_i(h_\alpha),\lambda(h_\alpha)\}$, for all $\alpha\in R^+$, \end{enumerit} then there exists a surjective $\lie g[t]$--module map $V(d_i\ell\omega_i)*V(\lambda)\rightarrow V(\mu_1)* V(\mu_2)\rightarrow 0.$ \end{prop} \begin{proof} By Proposition \ref{2-fold}, we get a surjective map of $\lie g[t]$--modules $V(d_i\ell\omega_i)*V(\lambda)\rightarrow \bold{V}_{\mu_1,\mu_2}\rightarrow 0,$ and now using Lemma \ref{2-foldlem} we have a surjective map of $\lie g[t]$--modules $\bold{V}_{\mu_1,\mu_2}\rightarrow V(\mu_1)* V(\mu_2)\rightarrow 0.$ Putting all these together we get a surjective $\lie g[t]$--module map $$V(d_i\ell\omega_i)*V(\lambda)\rightarrow V(\mu_1)* V(\mu_2)\rightarrow 0.$$ \end{proof} \begin{cor} Let $\ell, m$ be positive integers such that $m\le \ell \le 2m-1$. Then for any dominant weight $\lambda$ of $\lie{g}$ with $\lambda(h_\Theta)\le 2m-\ell$ and a fundamental weight $\omega_i$ listed in the remark \ref{minusculecoweight}, we have a surjective map of $\lie{g}[t]$--modules \begin{gather*} V(d_im\omega_i)\ast V(\lambda+(\ell-m)d_i\omega_i)\rightarrow V(d_i\ell\omega_i)\ast V(\lambda) \rightarrow 0. \end{gather*} \end{cor} \vskip 6pt \subsection{} Here is an another important application of Proposition \ref{demrelfuskr}. \begin{thm}\label{genschurpos} Fix $k\in \mathbb{Z}_+$ and $\ell, m\in \mathbb{N}$ such that $\ell\ge m$. Suppose that there exists $\lambda\in P^+$ and $\mu\in \mathcal{P}^+_m$ such that $kd_i\ell\omega_i+\lambda=kd_im\omega_i+\mu$. \begin{itemize} \item[(i)] Then there exists a surjective map of $\lie{g}[t]-$modules \begin{gather*} KR(d_im\omega_i)^{\ast k}\ast V(\mu)\rightarrow KR(d_i\ell\omega_i)^{\ast k}\ast V(\lambda) \rightarrow 0. \end{gather*} \item[(ii)] Further more if we assume that $i\in I$ such that $d_i\omega_i(h_\Theta)\le 1$. Then there exists a surjective map of $\lie{g}[t]-$modules \begin{gather*} V(d_im\omega_i)^{\ast k}\ast V(\mu)\rightarrow V(d_i\ell\omega_i)^{\ast k}\ast V(\lambda) \rightarrow 0. \end{gather*} \end{itemize} \end{thm} \begin{proof} Observe that $kd_i\ell\omega_i(h_\Theta)+\lambda(h_\Theta)=kd_im\omega_i(h_\Theta)+\mu(h_\Theta)$ and $\mu(h_\Theta)\le m$ implies $\lambda\in \mathcal{P^+_\ell}$. Now the theorem is immediate from the Proposition \ref{demrelfuskr}. \end{proof} \begin{rem} We end this section with an important remark. We first recall the results of \cite{chfsa}. In \cite{chfsa}, the authors defined a partial order $\preceq$ on the set of $k$--tuples of dominant integral weights which add up to $\lambda$ (say $P^+(\lambda,k)$): one just requires the inequality $(ii)$ in the Proposition \ref{schurpostivity} to hold for all partial sums. Given an element of $P^{+}(\lambda,\,k)$, they considered the tensor product of the corresponding simple finite--dimensional $\lie g$--modules and showed that the dimension of this tensor product increases along this partial order $\preceq$. They also showed that in the case when $\lambda$ is a multiple of a fundamental minuscule weight ($\lie g$ and $k$ are general) or if $\lie g$ is of type $A_{2}$ and $k=2$ ($\lambda$ is general), there exists an inclusion of tensor products along with the partial order $\preceq$ on $P^{+}(\lambda,\,k)$ (see \cite[Theorem 1]{chfsa}). In particular, if $\lie g$ is of type $A_n$, this means that the difference of the characters is Schur positive. We can recover these results when $\lambda=Nd_i\omega_i$ using Proposition \ref{schurpostivity}, where $\omega_i$ is in the list of remark \ref{minusculecoweight} and $\lie g$, $k$ are general. We indeed prove that there exists a surjective map between the appropriate tensor products. The existence of such a surjective map between the tensor products can be reduced to the case when $k=2$ and this case easily follows from proposition \ref{schurpostivity}. We refer the readers to \cite{chfsa} for more details. We note here that the proof in \cite{chfsa} uses the combinatorics of LS paths and our approach avoid this. \end{rem}
\section{INTRODUCTION}{\label{sec1}} In recent years, much attention has been paid to the study of electromagnetically induced transparency (EIT), a quantum interference effect induced by a strong control field, by which the optical absorption of a probe field in resonant three-level atomic systems can be largely suppressed. In addition to the interest in fundamental research, EIT has many important applications in slow light and quantum storage, nonlinear optics at low-light level, precision laser spectroscopy, and so on~\cite{Fleischhauer2005}. The most prominent character of EIT is the opening of a transparency window in probe-field absorption spectra. However, the occurrence of transparency window is not necessarily due to EIT effect. In 1955, Autler and Townes~\cite{at} showed that the absorption spectrum of molecular transition can split into two Lorentzian lines (doublet) when one of two levels involved in the transition is coupled to a third one by a strong microwave field. Such doublet is now called Autler-Townes splitting (ATS) and has also been intensively investigated in atomic and molecular spectroscopy~\cite{coh}. Although both EIT and ATS effects can open transparency windows in probe absorption spectra, the physical mechanisms behind them are quite different. EIT is resulted from a quantum {\it destructive} interference between two competing transition pathways, whereas ATS is a dynamic Stark shift caused by a gap between two resonances. Usually, it is not easy to distinguish EIT and ATS by simply looking at the appearance of absorption spectra. Because EIT and ATS are two typical phenomena appeared widely in laser spectroscopy and have many applications, it is necessary to develop an effective technique to distinguish the difference between them. In 1997, Agarwal~\cite{Agarwal1997} proposed a spectrum decomposition method, by which the probe-field absorption spectra of cold three-level atomic systems were decomposed into two absorptive contributions plus two interference contributions. Recently, this method was used to clarify EIT and ATS in a more general way~\cite{Anisimov2008,Tony2010,Anisimov2011}. In a recent work, an experimental investigation on EIT-ATS crossover was carried out~\cite{giner}. Very recently, the spectrum decomposition method was adopted to investigate the EIT-ATS crossover in $\Lambda$- and $V$-type molecular systems with Doppler broadening~\cite{tan2013,Zhu2013}. In the study of EIT and ATS, several typical three-level systems (i.e. $\Lambda$, $V$, and ladder)~\cite{note00} are widely used. For ladder systems, there are two typical configurations, with the level diagrams and excitation schemes shown in Figs.~\ref{model}(a) and \ref{model}(b) below, called here as ladder-I (or upper-level-driven ladder system; Fig.~\ref{model}(a)\,) and ladder-II (or lower-level-driven ladder system; Fig.~\ref{model}(b)\,), respectively. The so-called upper-level-driven (lower-level-driven) means that the control field couples the two upper (lower) levels of the system. In recent years, much interest has been focused on the Rydberg excitations in cold and hot atomic gases, where ladder-type excitation schemes are widely employed~\cite{saf,pri1,sevi,moh1,moh2,wea,rai,pri}. In this article, we propose a general theoretical scheme to investigate the crossover from electromagnetically induced transparency (EIT) to Autler-Townes splitting (ATS) in open ladder-type atomic and molecular systems with Doppler broadening. We show that when the wavenumber ratio $k_c/k_p\approx -1$, EIT, ATS, and EIT-ATS crossover exist for both ladder-I and ladder-II systems, where $k_c$ ($k_p$) is the wavenumber of control (probe) field. Furthermore, when $k_c/k_p$ is far from $-1$ EIT can occur but ATS is destroyed if the upper state of the ladder-I system is a Rydberg state. In addition, ATS exists but EIT is not possible if the control field used to couple the two lower states of the ladder-II system is a microwave field. The theoretical scheme developed and the results obtained here can be applied to various ladder systems (including hot gases of Rubidium atoms, Na$_2$ molecules, and Rydberg atoms). Before proceeding, we note that many studies exist on the study of ladder systems. Except for EIT and ATS~\cite{Agarwal1997,Tony2010,saf,moh1,moh2,wea,rai,pri,pri1,sevi,Lazoudis2008,Yang1997,Yong1995,Julio1995, Lee2000,Jason2001,Qi2002,Ahmed2006,Ahmed2007,Ray2007,Moon,Kubler2010,gor,petro,ate}, other related investigations have also been carried out, including Rabi oscillations~\cite{dudin,Huber2011}, coherent population transfer~\cite{schemp}, quantum nonlinear optics at single-photon level~\cite{Peyr}, fast entanglement generation~\cite{Bari}, and microwave electrometry with Rydberg atoms~\cite{Sedl}. However, to the best of our knowledge no systematic analysis on the crossover from EIT to ATS in ladder systems has been carried out up to now; furthermore, no theory on the EIT-ATS crossover in open ladder systems with Doppler broadening has been presented. Our theoretical scheme is valid for both atoms, molecules, and other systems, and can elucidate various quantum interference characters (EIT, ATS, and EIT-ATS crossover) in a clear way. The results obtained here are not only useful for understanding the detailed feature of quantum interference in multi-level systems and guiding new experimental findings, but also may have promising applications in atomic and molecular spectroscopy, light and quantum information processing, etc. The article is arranged as follows. In the next section, we describe the theoretical model. In Sec.~\ref{sec3} and Sec.~\ref{sec4}, the quantum interference characters of the ladder-I and ladder-II systems are analyzed, respectively. Finally, in the last section we summarize the main results obtained in this work. \section{MODEL}{\label{sec2}} \begin{figure} \includegraphics[scale=0.45]{fig1}\\ \caption{\footnotesize (Color online) (a) Ladder-I system, where states $|3\rangle$ and $|2\rangle$ are coupled by the control field with center angular frequency $\omega_c$, and states $|2\rangle$ and $|1\rangle$ are coupled by the probe field with center angular frequency $\omega_p$; (b) Ladder-II system. (c) Open ladder system. The state $|2\rangle$ couples to the state $|3\rangle$ by field $a$ (with center angular frequency $\omega_a$) and the ground state $|1\rangle$ by field $b$ (with center angular frequency $\omega_b$). $\Delta_{2}$ and $\Delta_{3}$ are detunings, $\Gamma_{jl}$ are population decay rates from $|l\rangle$ to $|j\rangle$, and $\gamma$ is the transit rate. Particles occupying the state $|2\rangle$ ($|3\rangle$) may decay to other states besides $|1\rangle$ ($|2\rangle$). Levels $|4\rangle$ and $|5\rangle$ denote these other states rendering the system open.} \label{model} \end{figure} We consider a hot gas consisting of atoms or molecules, where particles have three resonant levels (i.e. ground state $|1\rangle$, intermediate state $|2\rangle$, and upper state $|3\rangle$) with a ladder configuration (Fig.~\ref{model}(c)\,)~\cite{Ahmed2007}. Especially, the upper state $|3\rangle$ may be a Rydberg state. Two laser fields with central angular frequency $\omega_a$ and $\omega_b$ couple to the transition $|2\rangle\leftrightarrow|3\rangle$ and $|1\rangle\leftrightarrow|2\rangle$, respectively. The electric field vector is $\mathbf{E}=\sum_{l=a,b}\mathbf{e}_l{\cal E}_l {\rm exp}[i({\bf k}_l\cdot {\bf r}-\omega_lt)]+$c.c., where $\mathbf{e}_l$ $(k_l)$ is the unit polarization vector (wavenumber) of the electric field component with the envelope ${\cal E}_l$ $ (l=a,b)$. We assume the system is open, i.e. particles occupying the state $|2\rangle$ ($|3\rangle$) can follow various relaxation pathways and decay into other states besides $|1\rangle$ ($|2\rangle$). For simplicity, all these other states are represented by states $|4\rangle$ and $|5\rangle$~\cite{Ahmed2007}. In the figure, $\Delta_2$ and $\Delta_3$ are detunings, $\Gamma_{jl}$ is the population decay rate from state $|l\rangle$ to state $|j\rangle$, $\gamma$ is the beam-transit rate added to account for the rate with which particles escape the interaction region (significant only for the level $|4\rangle$ since it cannot radiatively decay). Under electric-dipole and rotating-wave approximations, the interaction Hamiltonian of the system in interaction picture reads \begin{equation}\label{H_EIT} {\cal H}_{\rm int}=-\hbar(\Omega_ae^{i[\mathbf{k}_a\cdot(\mathbf{r}+\mathbf{v}t)-\omega_at]} |3\rangle\langle2|+\Omega_be^{i[\mathbf{k}_b\cdot(\mathbf{r}+\mathbf{v}t) -\omega_bt]}|2\rangle\langle1|+{\rm h.c.}), \end{equation} where $\Omega_a=\boldsymbol{\mu}_{32}\cdot{\cal E}_a/\hbar$ ($\Omega_b=\boldsymbol{\mu}_{21}\cdot{\cal E}_b/\hbar$) is the half Rabi-frequency of the field $a$ (field $b$), with $\boldsymbol{\mu}_{jl}$ being the electric-dipole matrix element associated with the transition from the state $|l\rangle$ to the state $|j\rangle$. The optical Bloch equation in the interaction picture is \begin{subequations} \label{dme1} \begin{eqnarray} &&i\frac{\partial}{\partial t}\sigma_{11}-i\Gamma_{12}\sigma_{22}-i\gamma\sigma_{44}-i\Gamma_{15}\sigma_{55} +\Omega_{b}^{*}\sigma_{21}-\Omega_{b}\sigma_{21}^{*}=0,\\ &&i\frac{\partial}{\partial t}\sigma_{22}+i\Gamma_{2}\sigma_{22}-i\Gamma_{23}\sigma_{33} +\Omega_{b}\sigma_{21}^{*}+\Omega_{a}^{*}\sigma_{32}-\Omega_{b}^{*}\sigma_{21} -\Omega_{a}\sigma_{32}^{*}=0,\\ &&i\frac{\partial}{\partial t}\sigma_{33}+i\Gamma_{3}\sigma_{33}+\Omega_{a}\sigma_{32}^{*}-\Omega_{a}^{*}\sigma_{32}=0,\\ &&i\frac{\partial}{\partial t}\sigma_{44}+i\gamma\sigma_{44}-i\Gamma_{42}\sigma_{22}-i\Gamma_{45}\sigma_{55}=0,\\ &&i\frac{\partial}{\partial t}\sigma_{55}+i\Gamma_{5}\sigma_{55}-i\Gamma_{53}\sigma_{33}=0,\\ &&(i\frac{\partial}{\partial t}+d_{21})\sigma_{21}+\Omega_{a}^{*}\sigma_{31}+\Omega_{b}(\sigma_{11}-\sigma_{22})=0,\\ &&(i\frac{\partial}{\partial t}+d_{31})\sigma_{31}-\Omega_{b}\sigma_{32}+\Omega_{a}\sigma_{21}=0,\\ &&(i\frac{\partial}{\partial t}+d_{32})\sigma_{32}-\Omega_{b}^{*}\sigma_{31}+\Omega_{a}(\sigma_{22}-\sigma_{33})=0, \end{eqnarray} \end{subequations} where $d_{21}=-\mathbf{k}_b\cdot\mathbf{v}+\Delta_2+i\gamma_{21}$, $d_{31}=-(\mathbf{k}_b+\mathbf{k}_a)\cdot\mathbf{v}+\Delta_3+i\gamma_{31}$, $d_{32}=-\mathbf{k}_a\cdot\mathbf{v}+\Delta_3-\Delta_2+i\gamma_{32}$ with $\gamma_{jl}=(\Gamma_{j}+\Gamma_{l})/2+\gamma_{jl}^{{\rm col}}\ (j,l=1,2,3)$. Here, $\bf{v}$ is the thermal velocity of the particles, $\Gamma_{j}$ denotes the total population decay rates out of levels $|j\rangle$, defined by $\Gamma_{j}=\sum_{l\neq j}\Gamma_{lj}$. The quantity $\gamma_{jl}^{{\rm col}}$ is the dephasing rate due to processes that are not associated with population transfer, such as elastic collisions. The evolution of the electric field is governed by the Maxwell equation \begin{equation}\label{ME} \nabla^2 {\bf E}-\frac{1}{c^2}\frac{\partial^2{\bf E}}{\partial t^2}=\frac{1}{\epsilon_0c^2}\frac{\partial^2 {\bf P}}{\partial t^2}. \end{equation} Due to the Doppler effect, the electric polarization intensity of the system reads \begin{equation}\label{P} {\bf P}={\cal N}\int_{-\infty}^{\infty}dv f(v)\left[ \boldsymbol{\mu}_{12}\sigma_{21} e^{i(k_b z-\omega_b t)}+\boldsymbol{\mu}_{23}\sigma_{32} e^{i(k_a z-\omega_a t)}+{\rm c.c.}\right], \end{equation} where ${\cal N}$ is particle concentration and $f(v)=e^{-(v/v_T)^2}/(\sqrt{\pi}v_T)$ is Maxwellian velocity distribution function, where $v_T=(2k_BT/M)^{1/2}$ is the most probable speed at temperature $T$ with $k_B$ the Boltzmann constant and $M$ the particle mass. For simplicity and without loss of generality, we have assumed the two laser fields propagate along the $z$ direction with a counter-propagating configuration, i.e. ${\bf k}_{a,b}=(0,0,k_{a,b})$ with $k_b=-k_a$ in order to suppress the first-order Doppler effect. Note that the model given above is valid also for a closed ladder system, which can be obtained by simply taking $\Gamma_{15}=\Gamma_{42}=\Gamma_{45}=\Gamma_{53}=\gamma=0$; furthermore, if the system is not only closed but also cold, one has $\Gamma_{15}=\Gamma_{42}=\Gamma_{45}=\Gamma_{53}=\gamma=0$ and $f(v)=\delta (v)$. \section{Quantum interference character of ladder-\uppercase\expandafter{\romannumeral1} system}\label{sec3} \subsection{Linear dispersion relation}\label{sec3a} When the laser field $a$ (field $b$) is taken as the control (probe) field, the system is the ladder-\uppercase\expandafter{\romannumeral1} system (i.e. $\omega_a=\omega_c$, $\omega_b=\omega_p$; see Fig.~\ref{model}(a)\,). In this case, under slowly varying envelope approximation (SVEA) the Maxwell Eq.~(\ref{ME}) is reduced to the form \begin{equation}\label{maxwell1} i\left(\frac{\partial}{\partial z}+\frac{1}{c}\frac{\partial}{\partial t}\right)\Omega_b +\kappa_{12}\int_{-\infty}^{\infty} dv f(v)\sigma_{21}(v)=0, \end{equation} where $\kappa_{12}={\cal N}\omega_b|\boldsymbol{\mu}_{21}|^2/(2\hbar\varepsilon_0 c)$ with $c$ is the light speed in vacuum. The base state (zero-order) solution of the system, i.e. the steady-state solution of the MB Eqs.~(\ref{dme1}) and (\ref{maxwell1}) for $\Omega_b=0$ is given by $\sigma_{11}^{(0)}=1$, $\sigma_{jl}^{(0)}=0$ ($j,l\neq 1$). When the probe field is switched on, the system will involve into time-dependent state. At the first order of $\Omega_b$, the population and the coherence between the states $|2\rangle$ and $|3\rangle$ are not changed, but \begin{subequations}\label{1order} \begin{eqnarray} &&\Omega_b^{(1)}=Fe^{i\theta}\\ &&\sigma_{21}^{(1)}=\frac{\omega+d_{31}}{|\Omega_a|^2-(\omega+d_{21})(\omega+d_{31})}F e^{i\theta},\\ &&\sigma_{31}^{(1)}=-\frac{\Omega_a}{|\Omega_a|^2-(\omega+d_{21})(\omega+d_{31})}F e^{i\theta}, \end{eqnarray} \end{subequations} here $F$ is a constant and $\theta=K(\omega)z-\omega t$. The linear dispersion relation $K(\omega)$ reads \begin{equation}\label{LDa} K(\omega)=\frac{\omega}{c}+\kappa_{12}\int_{-\infty}^{\infty} dv f(v)\frac{\omega+d_{31}}{|\Omega_a|^2-(\omega+d_{21})(\omega+d_{31})}. \end{equation} The integrand in the dispersion relation (\ref{LDa}) depends on two factors. The first is the ac Stark effect induced by the control field, reflected in the denominator, corresponding to the appearance of dressed states out of states $|2\rangle$ and $|3\rangle$, by which two Lorentzian peaks in the probe-field absorption spectrum are shifted from their original positions. The second is the Doppler effect, reflected by $d_{ij}=d_{ij}(v)$ and the velocity distribution $f(v)$, which results in an inhomogeneous broadening in Im($K$) (the imaginary part of $K$). The lineshape of Im($K$) depends strongly on the wavenumber ratio $x=k_a/k_b$. Fig.~\ref{EIT3D} shows \begin{figure} \includegraphics[scale=0.45]{fig2} \caption{\footnotesize (Color online) The probe-field absorption spectrum Im($K$) of the ladder-I system as a function of $\omega$ and the wavenumber ratio $x$.} \label{EIT3D} \end{figure} the numerical result of Im($K$) as a function of $\omega$ and $x$. The system parameters are chosen as $\Gamma_2=6$ MHz, $\Gamma_3=1$ MHz, $\gamma=0.5$ MHz, $\gamma_{ij}^{\rm col}=1$ MHz, and $\Omega_a=80$ MHz. We see that Im($K$) undergoes a transition from a deep, wide transparency window (doublet) to a single absorption peak when $x$ changes from $-1.4$ to $-0.4$. Since Fig.~\ref{EIT3D} is obtained by a numerical calculation, it is not easy to get a clear and definite conclusion on the quantum interference characters of the system. Thus we turn to an analytical approach by using the method developed in Refs.~\cite{Agarwal1997,Anisimov2008,Tony2010,Anisimov2011,tan2013,Zhu2013}. \subsection{EIT-ATS crossover in hot Rubidium atomic gases} In many experimental studies on EIT or EIT-related effects in the ladder-I system with Doppler broadening, the excitation scheme $5S_{1/2} \rightarrow 5P_{3/2} \rightarrow 5D_{5/2}$ of $^{87}$Rb atoms was adopted, such as did in Refs.~\cite{Julio1995,Moon}. In this situation, the wavenumber ratio $x=-1$, and the integration in Eq.~(\ref{LDa}) can be carried out analytically by using the residue theorem when the Maxwellian velocity distribution is replaced by the modified Lorentzian velocity distribution $f(v)=v_T/\left[\sqrt{\pi}(v^2+v_T^2)\right]$. Such technique has been widely employed by many authors~\cite{tan2013,Zhu2013,Elena2002,Lee2003,LiHuang2010}. Note that the integrand in the second term of the Eq.~(\ref{LDa}) has only one pole $k_bv=-ik_bv_T=-i\Delta\omega_D$ in the lower half complex plane of $v$. Considering the contour integration shown in Fig.~\ref{fig:LM}(a) \begin{figure} \includegraphics[scale=0.35]{fig3} \caption{(Color online) (a) The pole $(0,-ik_bv_T)$ of the integrand in Eq.~(\ref{LDa}) in the lower half complex plane of $v$. The closed curve with arrows is the contour chosen for calculating the integration in Eq.~(\ref{LDa}) by using the residue theorem. (b) Probe-field absorption spectrum Im($K$) as a function of $\omega$ for the hot ladder-I system with wavenumber ratio $x=-1$. The solid (dashed) line is for $|\Omega_a|=500$ MHz ($|\Omega_a|=0$). Definitions of Im$(K)_{\rm min}$, Im$(K)_{\rm max}$, and the width of transparency window $\Gamma_{\rm TW}$ are indicated in the figure.}\label{fig:LM} \end{figure} and using the residue theorem, we obtain the exact result \begin{equation}\label{Ka1} K(\omega)=\frac{\omega}{c}+\frac{\sqrt{\pi}\kappa_{12}(\omega +i\gamma_{31})}{|\Omega_a|^2-(\omega+i\gamma_{21}+i\Delta\omega_D)(\omega+i\gamma_{31})}, \end{equation} with$\Delta_2=\Delta_3=0$. Explicit expression of $K(\omega)$ for nonvanishing $\Delta_2$ and $\Delta_3$ can also be obtained but lengthy and thus omitted here. Fig.~\ref{fig:LM}(b) shows the profile of Im($K$) as a function of $\omega$. The dashed (solid) line is for the case of $|\Omega_a|=0$ ($|\Omega_a|=500$ MHz) for $\Gamma_{2}=6$ MHz, $\Gamma_{3}=1$ MHz, $\gamma=0.5$ MHz, $\gamma_{jl}^{{\rm col}}=1$ MHz, $\Delta\omega_D=270$ MHz and $\kappa_{12}=1\times10^{9}$ cm$^{-1}$s$^{-1}$, used in Ref.~\cite{Julio1995}. We see that the probe-field absorption spectrum for $|\Omega_a|=0$ has only a single peak. However, a transparency window is opened for $\Omega_a=500$ MHz. The minimum (Im$(K)_{\rm min}$), maximum (Im$(K)_{\rm max}$), and width of transparency window ($\Gamma_{\rm TW}$) are defined in the figure. Equation (\ref{Ka1}) can be written as the form \begin{equation}\label{Kd1} K(\omega)=\frac{\omega}{c}-\sqrt{\pi}\kappa_{12}\frac{\omega +i\gamma_{31}}{(\omega-\omega_+)(\omega-\omega_-)}, \end{equation} with \begin{equation} \omega_{\pm}=\frac{1}{2}\left\{-i(\gamma_{21}+\gamma_{31}+\Delta\omega_D) \pm 2\left[ |\Omega_a|^2-|\Omega_{\rm ref}|^2\right]^{1/2} \right\}, \end{equation} where \begin{equation} \Omega_{\rm ref}\equiv \frac{1}{2}(\gamma_{21}+\Delta\omega_{D}-\gamma_{31}). \end{equation} In order to illustrate the quantum interference effect in a simple and clear way, we decompose Im$(K)$ for different $\Omega_a$ as follows. (i). {\it Weak control field region} (i.e. $|\Omega_a|<\Omega_{\rm ref}\approx \Delta\omega_D/2$): Equation (\ref{Kd1}) can be written as \begin{equation}\label{Kd2} K(\omega)=\frac{\omega}{c}+\sqrt{\pi}\kappa_{12}\left(\frac{A_+}{\omega -\omega_+}+\frac{A_-}{\omega-\omega_-}\right), \end{equation} where $A_{\pm}=\mp(\omega_{\pm}+i\gamma_{31})/(\omega_+-\omega_-)$. Since in this region ${\rm Re}(\omega_{\pm})={\rm Im}(A_{\pm})=0$, we obtain \begin{equation}\label{form1} {\rm Im}(K)=\sqrt{\pi}\kappa_{12}\left(\frac{B_+}{\omega^2+\delta_+^2}+ \frac{B_-}{\omega^2+\delta_-^2}\right)\equiv L_2+L_1, \end{equation} with $\delta_{\pm}={\rm Im}(\omega_{\pm})$, $B_{\pm}=A_{\pm}\delta_{\pm}$, and $L_{1(2)}=\sqrt{\pi}\kappa_{12}B_{-(+)}/(\omega^2+\delta_{-(+)}^2)$. Thus the probe-field absorption profile comprises two Lorentzians centered at $\omega=0$. Shown in Fig.~\ref{figEIT1}(a) \begin{figure} \includegraphics[scale=0.35]{fig4} \caption{\footnotesize (Color online) EIT-ATS crossover for hot ladder-\uppercase\expandafter{\romannumeral1} system. (a) Probe-field absorption spectrum Im($K$) (solid line) in the region $|\Omega_a|<\Omega_{\rm ref}$ is a superposition of the positive $L_1$ (dash-dotted line) and the negative $L_2$ (dashed line). (b) Im($K$) (solid line) composed by two Lorentzians (dashed-dotted line) and destructive interference (dashed line) in the region $|\Omega_a|>\Omega_{\rm ref}$. (c) Im($K$) (solid line) composed by two Lorentzians (dashed-dotted line) and destructive interference (dashed line) in the region $|\Omega_a|>\Omega_{\rm ref}$. Panels (a), (b) and (c) correspond to EIT, EIT-ATS crossover, and ATS, respectively. (d) The ``phase diagram'' of ${\rm Im}(K)_{\omega=0}/{\rm Im}(K)_{\rm max}$ as a function of $|\Omega_a|/\Omega_{\rm ref}$ illustrating the transition from EIT to ATS for the hot ladder-I system. Three regions (EIT, EIT-ATS crossover, and ATS) are divided by two vertical dashed-dotted lines.} \label{figEIT1} \end{figure} are the results of $L_1$, which is a positive single peak (the dash-dotted line), and $L_2$, which is a negative single peak (the dashed line). When plotting the figure, we have taken $\Omega_a=100$ MHz and the other parameters are the same as those used in Fig.~\ref{fig:LM}(b). The superposition of $L_1$ and $L_2$ gives Im($K$) (the solid line), which displays a absorption doublet with a transparency window near at $\omega=0$. Because there exists a {\it destructive} interference between the positive $L_1$ and the negative $L_2$ in the probe-field absorption spectrum, the phenomenon found here belongs to EIT based on the criterion given in Refs.~\cite{Anisimov2008,Tony2010,Anisimov2011}. (ii). {\it Intermediate control field region} (i.e. $|\Omega_a|>\Omega_{\rm ref}$): In this region ${\rm Re}(\omega_{\pm})\neq0$, we obtain \begin{equation}\label{Kd3} K(\omega)=\frac{\omega}{c}-\sqrt{\pi}\kappa_{12}\left[\frac{\omega+iW}{(\omega+iW-\delta) (\omega+iW+\delta)}+\frac{i(\gamma_{31}-W)}{(\omega+iW-\delta)(\omega+iW+\delta)}\right]. \end{equation} where $W\equiv(\gamma_{21}+\gamma_{31}+\Delta\omega_D)/2$ and $\delta\equiv\sqrt{4|\Omega_a|^2-(\gamma_{21}+\Delta\omega_D-\gamma_{31})^2}/2$. The imaginary part of the Eq.~(\ref{Kd3}) is given by \begin{eqnarray}\label{form2}\nonumber {\rm Im}(K)=&&\frac{\sqrt{\pi}\kappa_{12}}{2}\left\{\frac{W}{(\omega-\delta)^2+W^2} +\frac{W}{(\omega+\delta)^2+W^2}\right.\nonumber\\ &&\left.\hspace{0.7cm}+\frac{g}{\delta}\left[\frac{\omega-\delta}{(\omega -\delta)^2+W^2}-\frac{\omega+\delta}{(\omega+\delta)^2+W^2}\right]\right\}, \end{eqnarray} with $g=W-\gamma_{31}$. The previous two terms (i.e. the two Lorentzian terms) in Eq.~(\ref{form2}) can be thought of as the net contribution coming to the absorption from two different channels corresponding to the two dressed states created by the control field $\Omega_a$~\cite{Agarwal1997}. The following terms proportional to $g$ are clearly interference terms. The interference is controlled by the parameter $g$ and it is destructive (constructive) if $g>0$ ($g<0$). Since in the ladder-\uppercase\expandafter{\romannumeral1} system with $x=-1$, $g=(\gamma_{21}+\Delta\omega_D-\gamma_{31})/2$ is always positive, thus the quantum interference induced by the control field is always destructive. Fig.~\ref{figEIT1}(b) shows the probe-field absorption spectrum Im($K$) (solid line) as a function of $\omega$ for $|\Omega_a|>\Omega_{\rm ref}$. The dashed-dotted (dashed) line denotes the contribution by the two positive Lorentzians (negative interference terms). We see that the interference is destructive. The system parameters used are the same as those in Fig.~\ref{figEIT1}(a) but with $\Omega_a=400$ MHz. A transparency window is opened due to the combined effect of EIT and ATS, which is deeper and wider than that in Fig.~\ref{figEIT1}(a). We attribute such phenomenon as EIT-ATS crossover. (iii). {\it Large control field region} (i.e., $|\Omega_a|\gg\Omega_{\rm ref}$): In this case, the quantum interference strength $g/\delta$ in Eq.~(\ref{form2}) is very weak (i.e. $g/\delta\approx 0$). Im$(K)$ reduces to \begin{equation}\label{form3} {\rm Im}(K)=\frac{\sqrt{\pi}\kappa_{12}}{2}\left[\frac{W}{(\omega-\delta)^2 +W^2}+\frac{W}{(\omega+\delta)^2+W^2}\right]. \end{equation} Fig.~\ref{figEIT1}(c) shows the result of the probe-field absorption spectrum as a function of $\omega$ for $|\Omega_a| \gg \Omega_{\rm ref}$. The dashed-dotted line represents the contribution by the sum of the two Lorentzians. For illustration, we have also plotted the contribution from the small interference terms [neglected in Eq.~(\ref{form2})\,], denoted by the dashed line. We see that the interference is still destructive but very small. The solid line is the curve of Im($K$), which has two resonances at $\omega\approx\pm\Omega_a$. Parameters used are the same as those in Fig.~\ref{figEIT1}(a) and Fig.~\ref{figEIT1}(b) but with $\Omega_a=1.2$ GHz. Obviously, the phenomenon found in this case belongs to ATS because the transparency window opened is mainly due to the contribution of the two Lorentzians. From the results given above, we see that the probe-field absorption spectrum experiences a transition from EIT to ATS as the control field is changed from small to large values. From the above result we can distinguish three different regions, i.e. the EIT ($|\Omega_a|<\Omega_{\rm ref}$), the EIT-ATS crossover ($1< |\Omega_a|/\Omega_{\rm ref}\leq 4$), and ATS $(|\Omega_a|/\Omega_{\rm ref}> 4$). Fig.~\ref{figEIT1}(d) shows a ``phase diagram'' that illustrates the transition from the EIT to ATS by plotting ${\rm Im}(K)_{\omega=0}/{\rm Im}(K)_{\rm max}$ as a function of $|\Omega_a|/\Omega_{\rm ref}$. Note that we have defined ${\rm Im}(K)_{\omega=0}/{\rm Im}(K)_{\rm max}=0.01$ as the border between EIT-ATS crossover and ATS regions. Our results on the characters of the quantum interference effect in the hot Rubidium atomic gases are consistent with those obtained in the experiments~\cite{Julio1995,Moon}. According to our analysis, the experiments carried out in Refs.~\cite{Julio1995,Moon} are mainly in the EIT region. We expect the EIT-ATS crossover and ATS may be observed experimentally if $\Omega_a$ is increased to the intermediate and the large control-field regions. \subsection{EIT-ATS crossover in hot molecular gases}\label{sec3c} In 2008, Lazoudis {\it et al.}~\cite{Lazoudis2008} made an important experimental observation on EIT and ATS in a hot Na$_2$ molecular ladder-I system for the wavenumber ratio $x=-0.896$ and $x=-1.08$~\cite{note1}. Two excitation schemes of Na$_2$ molecules were adopted in Ref.~\cite{Lazoudis2008}. The first (called the system B) is $X^1\sum_g^+(1,19)\rightarrow A^1\sum_u^+(3,18)\rightarrow 4^1\sum_g^+(0,17)$, and the second (called the system A) is $X^1\sum_g^+(0,19)\rightarrow A^1\sum_u^+(0,20)\rightarrow 2^1\Pi_g(0,19)$. Both of them correspond to the levels $|1\rangle\rightarrow |2\rangle\rightarrow |3\rangle$ in our Fig.~\ref{model}(b). We now analyze this system by using the Eq.~(\ref{LDa}). When $x$ is different from $-1$, the approach used in the last subsection is not easy to implement since the pole of the integrand in the Eq.~(\ref{LDa}) is not fixed in the lower (or upper) half complex plane of $v$. In this case, the value of the pole depends on both $x$ and $\omega$; moreover, it has an intersection with the real axis for $\omega=0$. As a result, the residue of the pole is a piecewise function, and the spectrum decomposition gives very complicated expressions not convenient for analyzing the quantum interference character of the system. Because of the above mentioned difficulty, we turn to adopt the fitting method developed from the spectrum decomposition method, proposed by Anisimov {\it et al.}~\cite{Anisimov2011}. According to the spectrum-decomposition formulas~(\ref{form1}) and (\ref{form3}), we expect: (i)if the probe-field absorption spectrum has a good fit to the function \begin{equation}\label{AEIT} A_{\rm EIT}=\frac{B_+^2}{\omega^2+\delta_+^2}-\frac{B_-^2}{\omega^2+\delta_-^2}, \end{equation} EIT dominates, where $B_+,\delta_+,B_-,\delta_-$ are fitting parameters; (ii)if the absorption spectrum has a good fit to the function \begin{equation}\label{AATS} A_{\rm ATS}=C\left[\frac{1}{(\omega-\delta)^2+W^2}+\frac{1}{(\omega+\delta)^2+W^2}\right], \end{equation} ATS dominates, with $C,\delta,W$ being fitting parameters. Based on such technique, we find that EIT, ATS, and EIT-ATS crossover exist in the open molecular ladder-I system for both $x=-1.08$ and $x=-0.896$. Fig.~\ref{sysA}(a) shows \begin{figure} \includegraphics[scale=0.35]{fig5}\\ \caption{\footnotesize (Color online) The probe absorption spectrum Im($K$) as a function of $\omega$ for (a) $x=-0.896$ and $\Omega_a=265$ MHz (corresponding to system B of Ref.~\cite{Lazoudis2008}), and (b) $x=-1.08$ and $\Omega_a=242.5$ MHz (corresponding to system A of Ref.~\cite{Lazoudis2008}). The red dashed lines are our theoretical results, and the black-solid lines are the experimental ones from Ref.~\cite{Lazoudis2008}.} \label{sysA} \end{figure} the probe-field absorption spectrum Im($K$) for $x=-0.896$ and $\Omega_a=265$ MHz (corresponding to system B in Ref.~\cite{Lazoudis2008}). The black solid line is the experimental result from Ref.~\cite{Lazoudis2008}, while the red dashed line is given by our theoretical calculation. The system parameters are given by $\Gamma_{12}=\Gamma_{42}=4.0\times10^{7}$ $\mathrm{s}^{-1}$, $\Gamma_{23}=5.6\times10^{6}$ $\mathrm{s}^{-1}$, $\Gamma_{53}=5.0\times10^{7}$ $\mathrm{s}^{-1}$, $\gamma=2.7\times10^{5}$ $\mathrm{s}^{-1}$, $\gamma_{jl}^{{\rm col}}=1\times10^{6}$ $\mathrm{s}^{-1}$, and $\Delta\omega_D=5\times10^{8}$ $\mathrm{s}^{-1}$. We see that our theoretical result agrees well with the experimental one. Note that the value of the reference Rabi frequency $\Omega_{\rm ref}$ is a function of the wavenumber ratio $x$. When $x=-0.896$, one has $\Omega_{\rm ref}\simeq 400$ MHz. Thus the system is in the weak control field region and the phenomenon found belongs to the EIT. Note in passing that here we have plotted the quantity Im$(K)$ which is proportional to the fluorescence intensity related to state $|2\rangle$ because $\sigma_{22}\simeq2|\Omega_b|^2{\rm Im}(K)/\Gamma_2$. Shown in Fig.~\ref{sysA}(b) is the absorption spectrum Im($K$) for $x=-1.08$ and $\Omega_a=242.5$ MHz (corresponding to system A in Ref.~\cite{Lazoudis2008}). The system parameters are the same as that in Fig.~\ref{sysA}(a). We see that our result also agrees well with the experimental one. Since in this case $\Omega_{\rm ref}\approx 150$ MHz, the system is in the intermediate control field region and hence the phenomenon found belongs to the EIT-ATS crossover. Note that there is a small difference for the width of the EIT transparency window between our result and that in the experiment~\cite{Lazoudis2008}. The reason is mainly due to the approximation using the modified Lorentzian velocity distribution to replace the Maxwellian velocity distribution. \subsection{EIT in hot Rydberg atomic gases}\label{sec3d} Recently, much interest has focused on the EIT in hot Rydberg atomic gases due to its promising applications for storing, manipulating quantum information and precision spectroscopy~\cite{saf,moh1,moh2,wea,rai,pri,pri1,sevi,Lazoudis2008,Yang1997,Yong1995,Julio1995, Lee2000,Jason2001,Qi2002,Ahmed2006,Ahmed2007,Ray2007,Moon,Kubler2010,gor,petro,ate}. The ladder-I system has been widely adopted in the experimental study of Rydberg EIT, in which the transition is $5S_{1/2} \rightarrow 5P_{3/2} \rightarrow nD_{5/2}$ of $^{85}$Rb atoms with $n$ being a large integer number. In this case, the upper state $|3\rangle$ in Fig.~\ref{model}(c) is a Rydberg state. If the density (e.g. lower than $10^8$ cm$^{-3}$) of a Rydberg gas is low, the interaction between Rydberg atoms can be ignored. Our theory developed in Sec.~\ref{sec2} and Sec.~\ref{sec3a} can be applied to study the probe-field propagation in such system. Shown in Fig.~\ref{fitEIT}(a) \begin{figure} \includegraphics[scale=0.3]{fig6} \caption{\footnotesize (Color online) (a) Probe-field absorption spectrum Im($K$) as a function of $\omega$. The blue solid (red dashed) line is for $|\Omega_a|=10$ MHz ($|\Omega_a|=0$). (b) Im($K$) (blue solid line), $A_{\rm EIT}$ (red dashed line) and $A_{\rm ATS}$ (black dashed-dotted line) as a function of $\omega$ for the weak control-field $\Omega_a=3$ MHz where $A_{\rm EIT}$ has a good fit. (c) The case for the intermediate control field $\Omega_a=15$ MHz where both $A_{\rm EIT}$ and $A_{\rm ATS}$ have poor fit.} \label{fitEIT} \end{figure} is the numerical result of the probe-field absorption spectrum Im($K$) as a function of $\omega$ for the hot ladder-I system with wavenumber ratio $x=-1.63$, which corresponds to the experiment carried out in 2007~\cite{moh1} by Mohapatra {\it et al}. The red dashed (blue solid) line is for the case of $|\Omega_a|=0$ ($|\Omega_a|=10$ MHz) for the system parameters $\Gamma_{2}=6$ MHz, $\Gamma_{3}=1$ kHz, $\gamma_{jl}^{{\rm col}}=1$ MHz, $\Delta\omega_D=270$ MHz, and $\kappa_{12}=1\times10^{9}$ cm$^{-1}$s$^{-1}$. We find that the line shape of Im($K$) displays enhanced absorption on both sides of the transparency window. This effect arises due to the wavelength mismatch between the control and probe fields combined with the effect of Doppler broadening. We now analyze the quantum interference character of such system. Since the spectrum decomposition method is not convenient for the analysis for the case $x\neq -1$, we employ the fitting method as done in the last subsection. Shown in Fig.~\ref{fitEIT}(b) and Fig.~\ref{fitEIT}(c) are the results of Im($K$) (blue solid line), $A_{\rm EIT}(B_+,\delta_+,B_-,\delta_-)$ (red dashed line) and $A_{\rm ATS}(C,\delta,W)$ (black dash-dotted line) as a function of $\omega$ for $\Omega_a=3$ MHz and 15 MHz, respectively. The expressions of $A_{\rm EIT}$ and $A_{\rm ATS}$ have been given by Eqs.~(\ref{AEIT}) and (\ref{AATS}). From Fig.~\ref{fitEIT}(b) we see that Im($K$) has a good fit to $A_{\rm EIT}(3.04,1.58,0.0381,0.208)$ and a poor fit to $A_{\rm ATS}(0.237,0.686,0.513)$. Thus EIT occurs in this weak control field region. However, for intermediate and large control field one can not find out the fitting parameters by which $A_{\rm EIT}$ and $A_{\rm ATS}$ can have a good fit to Im($K$) (Fig.~\ref{fitEIT}(c) shows the result for $\Omega_a=15$ MHz). Consequently, based on the criterion of Ref.~\cite{Anisimov2011}, neither EIT nor ATS dominates in the intermediate large control field regions. Note that in the system discussed here the probe-field absorption spectrum Im($K$) doesn't possess standard Lorentzian lineshape for large control field, which is due to the enhanced absorption by the Doppler effect and by the large wavenumber mismatch between the probe and control fields. Experimentally, EIT in hot Rydberg atomic gases has been observed in Ref.~\cite{moh1}. Our theoretical result given above agrees with the experimental one. We hope that the theoretical result for the intermediate and large control field region predicted here may be verified experimentally in near future. \section{Quantum interference character of ladder-II system}\label{sec4} If the probe field and the control field in the ladder-I system are exchanged, we obtain the ladder-II system (Fig.~\ref{model}(b) with $\omega_b=\omega_c$, $\omega_a=\omega_p$). In this case, the Maxwell Eq.~(\ref{ME}) under the SVEA is reduced to \begin{equation}\label{maxwell2} i\left(\frac{\partial}{\partial z}+\frac{1}{c}\frac{\partial}{\partial t}\right)\Omega_a +\kappa_{23}\int_{-\infty}^{\infty} dv f(v)\sigma_{32}(v)=0, \end{equation} with $\kappa_{23}={\cal N}\omega_a|\boldsymbol{\mu}_{32}|^2/(2\hbar\varepsilon_0 c)$. \subsection{Linear dispersion relation} The base state solution of the MB Eqs.~(\ref{dme1}) and (\ref{maxwell2}) of the ladder-II system reads \begin{subequations} \label{BS} \begin{eqnarray} &&\sigma_{11}^{(0)}=\left(\gamma\Gamma_2|d_{21}|^2 +2\gamma\gamma_{21}|\Omega_b|^2\right)\frac{1}{D_1},\\ &&\sigma_{22}^{(0)}=2\gamma\gamma_{21}|\Omega_b|^2\frac{1}{D_1},\\ &&\sigma_{44}^{(0)}=2\gamma_{21}\Gamma_{42}|\Omega_b|^2\frac{1}{D_1},\\ &&\sigma_{21}^{(0)}=-\gamma\Gamma_{2}\Omega_bd_{21}^*\frac{1}{D_1}, \end{eqnarray} \end{subequations} and $\sigma^{(0)}_{31}=\sigma^{(0)}_{32}=\sigma^{(0)}_{33}=\sigma^{(0)}_{55}=0$, with $D_1 \equiv \gamma\Gamma_2|d_{21}|^2+2\gamma_{21}(2\gamma+\Gamma_{42})|\Omega_b|^2$. By using the same method as in Sec.~\ref{sec3a}, one can obtain the solution of the MB Eqs.~(\ref{dme1}) and (\ref{maxwell2}) in linear regime, with the linear dispersion relation given by \begin{equation}\label{LDb} K(\omega)=\frac{\omega}{c}+\kappa_{23}\int_{-\infty}^{\infty} dv f(v)\frac{(\omega+d_{31})2\gamma\gamma_{21}|\Omega_b|^2-\gamma \Gamma_{2}|\Omega_b|^2d_{21}^*}{D_1\left[|\Omega_b|^2 -(\omega+d_{31})(\omega+d_{32})\right]}. \end{equation} Fig.~\ref{ATS3D} shows \begin{figure} \includegraphics[scale=0.4]{fig7} \caption{\footnotesize (Color online) The probe-field absorption spectrum Im($K$) of the ladder-II system as a function of $\omega$ and the wavenumber ratio $x=k_a/k_b$.} \label{ATS3D} \end{figure} the probe-field absorption spectrum Im($K$) as a function of $\omega$ and the wavenumber ratio $x$. We see that, similar to the ladder-I system (Fig.~\ref{EIT3D}), Im($K$) undergoes also a transition from a wide transparency window in the line center to a single absorption peak when $x$ changes from $-1.2$ to $-0.8$. The system parameters have been chosen as $\Gamma_2=6$ MHz, $\Gamma_3=1$ MHz, $\gamma=0.5$ MHz, $\gamma_{ij}^{\rm col}=1$ MHz, and $\Omega_b=100$ MHz. \subsection{EIT-ATS crossover in hot Sodium atomic gases} In 1978, Gray and Stroud~\cite{Gray1978} made an experimental observation on ATS in a ladder-II type hot sodium atomic system with $|1\rangle=|3S_{1/2},F=2,M_F=2\rangle$, $|2\rangle=|3P_{3/2},F=3,M_F=3\rangle$, $|3\rangle=4D_{5/2},F=4,M_F=4\rangle$, and the wavenumber ratio $x\approx-1$. Such system can be described by the MB Eqs.~(\ref{dme1}) and (\ref{maxwell2}), and hence the linear dispersion relation (\ref{LDb}) can be used to describe the probe-field propagation. To get an analytical insight, we replace the Maxwellian velocity distribution by the modified Lorentzian velocity distribution and calculate the integration (\ref{LDb}) using the residue theorem. We find two poles of the integrand in the lower half complex plane of $v$, which are $k_av=-ik_av_T=-i\Delta\omega_D$ and $k_av=-iC=-i[\gamma_{21}^2+2\gamma_{21}(2\gamma +\Gamma_{42})|\Omega_b|^2/\gamma\Gamma_{2}]^{1/2}$. By taking the contour consisting of the lower half complex plane of $v$ and its real axis, we can calculate the integration exactly, with the result given by \begin{equation}\label{Kb1} K(\omega)=\omega/c+{\cal K}_1+{\cal K}_2, \end{equation} with \begin{subequations} \label{LS} \begin{eqnarray} &&{\cal K}_1=\frac{2\sqrt{\pi}\kappa_{23}\gamma_{21}|\Omega_b|^2\left\{\omega +i[\gamma_{31}+\Gamma_{2}(\Delta\omega_D+\gamma_{21})/(2\gamma_{21})] \right\}}{\Gamma_{2}(C^2-\Delta\omega_D^2)\left[|\Omega_b|^2-(\omega +i\gamma_{31})(\omega+i\gamma_{32}+i\Delta\omega_D)\right]},\\ &&{\cal K}_2=\frac{2\sqrt{\pi}\kappa_{23}\gamma_{21}\Delta\omega_D|\Omega_b|^2 \left\{\omega+i[\gamma_{31}+\Gamma_{2}(C+\gamma_{21})/(2\gamma_{21})] \right\}}{\Gamma_{2}C(\Delta\omega_D^2-C^2)\left[|\Omega_b|^2 -(\omega+i\gamma_{31})(\omega+i\gamma_{32}+iC)\right]}. \end{eqnarray} \end{subequations} We can also carry out a spectrum decomposition for ${\cal K}_j$ ($j=1,2$), like that done in Ref.~\ref{sec3a}. The explicit expressions of the decomposition have been given in Appendix~\ref{AppD}. Similarly, three different control field regions (i.e. the weak control field region $|\Omega_b|<\Omega_{\rm ref}$, the intermediate control field region $|\Omega_b|>\Omega_{\rm ref}$, and the strong control field region $|\Omega_b|\gg \Omega_{\rm ref}$; $\Omega_{\rm ref}\equiv \Delta\omega_D/2$) can also be obtained. Fig.~\ref{figAT1}(a) \begin{figure} \includegraphics[scale=0.3]{fig8} \caption{\footnotesize (Color online) EIT-ATS crossover for the hot atoms in the ladder-II system for the wavenumber ratio $x=-1$. (a) Probe-field absorption spectrum Im($K$) in the weak control field region ($|\Omega_a|<\Omega_{\rm ref}$). The dashed-dotted line is the contribution by positive $L_1$, the dashed line is by negative $L_2$. The sum of $L_1$ and $L_2$ gives Im($K$) (solid line). (b) Probe-field absorption spectrum Im($K$) (solid line) composed by two Lorentzians (dashed-dotted line) and the destructive interference (dashed line) in the intermediate control field region $|\Omega_a|>\Omega_{\rm ref}$. (c) Probe-field absorption spectrum Im($K$) (solid line) composed by two Lorentzians (dashed-dotted line) and the destructive interference (dashed line) in the strong control field region $|\Omega_a|\gg \Omega_{\rm ref}$. Panels (a), (b) and (c) correspond to EIT, EIT-ATS crossover, and ATS, respectively.} \label{figAT1} \end{figure} shows the absorption spectrum Im($K$) in the weak control field region ($|\Omega_b|=100$ MHz, which is smaller than $\Omega_{\rm ref}=150$ MHz). The dashed-dotted line is the contribution by positive $L_1$, and the dashed line is by negative $L_2$. The superposition (sum) of $L_1$ and $L_2$ gives Im($K$) (solid line). The expressions of $L_1$ and $L_2$ have been presented in Appendix~\ref{AppD}. System parameters are chosen as $\Gamma_2=10$ MHz, $\Gamma_3=3.15$ MHz, $\Delta\omega_D=300$ MHz~\cite{Stroud96}, with other parameters the same as those in the last section. We see that in the curve of Im($K$) a deep transparency window is opened, resulting from the {\it destructive} quantum interference (because $L_1$ is positive and $L_2$ is negative). Hence in this region EIT exists. Fig.~\ref{figAT1}(b) shows Im($K$) (solid line) in the intermediate control field region ($|\Omega_b|=200$ MHz), which is the sum of the two Lorentzians (dashed-dotted line) and the destructive interference (dashed line). In this region, a large dip appears in Im($K$) due to the contribution of the destructive interference. This region belongs to an EIT-ATS crossover. Fig.~\ref{figAT1}(c) illustrates Im($K$) (solid line), the two Lorentzians (dashed-dotted line), and the destructive interference (dashed line) in the large control field region ($|\Omega_b|=800$ MHz). We see that in this region the contribution of the quantum interference is too small to be neglected. Obviously, the phenomenon found in this situation belongs to ATS because the transparency window opened is mainly due to the contribution by the two Lorentzians. From the above analysis, we see that EIT, EIT-ATS crossover, and ATS exist in the ladder-II system with the Doppler broadening for the wavenumber ratio $x=-1$. This is different from cold ladder-II systems where no EIT and thus EIT-ATS crossover exist~\cite{Tony2010}. Although the experiment on ATS in a hot atomic system with the ladder-II configuration for $x=-1$ has been realized~\cite{Gray1978,Stroud96}, it seems that up to now no experimental study has been carried out on EIT, and EIT-ATS crossover in the ladder-II system with Doppler broadening. We hope new experiments can be designed to verify our predictions given here. \subsection{Microwave induced transparency} We now discuss the case when the control field in the ladder-\uppercase\expandafter{\romannumeral2} system is a microwave field, i.e. $x \rightarrow 0$. The relevant experimental result, named by Zhao {\it et al.}~\cite{Yang1997} as microwave induced transparency, was first reported in 1997. In this case, the level diagram and excitation scheme is given by Fig.~\ref{micro}, \begin{figure} \includegraphics[scale=0.4]{fig9} \caption{\footnotesize (Color online) Microwave field driven ladder-\uppercase\expandafter{\romannumeral2} configuration. All notations are given in the text.} \label{micro} \end{figure} in which the optical transition between the two lower states $|1\rangle$ and $|2\rangle$ is forbidden, but the optical transitions between the highest state $|3\rangle$ and the two lower states $|1\rangle$, $|2\rangle$ are allowed, so $\Gamma_{12}=\Gamma_{42}=0$. All spontaneous emission decay rates $\Gamma_{31}$, $\Gamma_{32}$, and $\Gamma_{34}$ (corresponding to the decay pathways $|3\rangle\rightarrow |1\rangle$, $|3\rangle\rightarrow |2\rangle$, and $|3\rangle\rightarrow |4\rangle$, respectively), and the transit rate $\gamma$ from $|4\rangle\rightarrow |3\rangle$ have been indicated in the figure. The base state solution of the MB equations for the present case reads $\sigma_{11}^{(0)}=\sigma_{22}^{(0)}=1/2$ and other $\sigma_{jl}^{(0)}=0$. The linear dispersion relation of the system is given by \begin{equation}\label{Kmw} K(\omega)=\frac{\omega}{c}+\frac{\kappa_{23}}{2}\int_{-\infty}^{\infty} dv f(v)\frac{\omega+d_{31}}{|\Omega_b|^2-(\omega+d_{31})(\omega+d_{32})}, \end{equation} with $d_{31}=-k_av+\Delta_3+i\gamma_{31}$. Because $\gamma_{31}=\gamma_{32}$, the integrand in Eq.(\ref{Kmw}) has only one pole in the lower half complex plane of $v$, given by $k_av=-ik_av_T=-i\Delta\omega_D$. When replacing the Maxwellian distribution by the modified Lorentzian distribution, the integration can be calculated exactly by using the residue theorem. One obtains \begin{equation}\label{Kmw2} K(\omega)=\frac{\omega}{c}+\frac{\sqrt{\pi}\kappa_{23}}{2} \frac{\omega+i\gamma_{31}+i\Delta\omega_D}{|\Omega_b|^2-(\omega +i\gamma_{31}+i\Delta\omega_D)^2}. \end{equation} It is easy to get the probe-field absorption spectrum Im($K$) from Eq.~(\ref{Kmw2}), which reads \begin{equation}\label{Kmw3} {\rm Im}(K)=\frac{\sqrt{\pi}\kappa_{23}}{2}\left[\frac{W}{(\omega-\delta)^2 +W^2}+\frac{W}{(\omega+\delta)^2+W^2}\right], \end{equation} with $W=\gamma_{31}+\Delta\omega_D$ and $\delta=|\Omega_b|$. Equation (\ref{Kmw3}) consists of two pure Lorentzians, which means that there is no quantum interference occurring in the system and the phenomenon found is an ATS one. Consequently, we conclude that there is no EIT and EIT-ATS crossover in the ladder-\uppercase\expandafter{\romannumeral2} system when the control field used is a microwave one. \section{Summary}\label{sec5} In Sec.~\ref{sec3} and Sec.~\ref{sec4}, we have analyzed the quantum interference characters in the hot ladder-I and ladder-II systems with Doppler broadening for many different cases. For clearness and for comparison, in Table~\ref{table:SUM} \begin{table} \caption{Quantum interference characters for various ladder systems with different wavenumber ratio $x$. ``Hot'' (``Cold'') means hot (cold) atoms or molecules. ``Any'' means any value of $x$. The last column gives some references in which related experiments have been carried out. \label{table:SUM}} \begin{ruledtabular} \begin{tabular}{ccccc} System & Wavenumber ratio $x$ & EIT & ATS & Reference \\ \hline \multirow{4}*{Ladder-\uppercase\expandafter{\romannumeral1} (Hot)} & $-0.896$ & Yes & Yes & \cite{Lazoudis2008} \\ & $-1$ & Yes & Yes & \cite{Julio1995,Moon} \\ & $-1.08$ & Yes & Yes & \cite{Lazoudis2008} \\ & $-1.63$ & Yes & No & \cite{moh1}\\\hline \multirow{2}*{Ladder-\uppercase\expandafter{\romannumeral2} (Hot)} & $-1$ & Yes & Yes & \cite{Gray1978} \\ & 0 & No & Yes & \cite{Yang1997}\\\hline Ladder-\uppercase\expandafter{\romannumeral1} (Cold)& Any & Yes & Yes & \cite{Jason2001,Weatherill2008} \\\hline Ladder-\uppercase\expandafter{\romannumeral2} (Cold)& Any & No & Yes & \cite{Teo2003,Hao2013} \\ \end{tabular} \end{ruledtabular} \end{table} we have summarized the main results obtained for different ladder configurations with different wavenumver ratio $x$. The first four lines are for the hot ladder-I system; the next two lines are for the hot ladder-II system. The seventh and eighth lines are for cold ladder-I system and cold ladder-II system, for which relevant theoretical analysis has been given in Refs.~\cite{Agarwal1997,Tony2010} and related experiments were made in Refs.~\cite{Jason2001,Weatherill2008,Teo2003,Hao2013}. If in the table there is ``Yes'' in the same line for both EIT and ATS, an EIT-ATS crossover also exists in the system. The last column of the table gives some references in which related experimental results were reported. In summary, in this work we have proposed a general theoretical scheme for studying the crossover from EIT to ATS in the open systems of ladder-type level configuration with Doppler broadening. We have elucidated various mechanisms of the EIT, ATS, and their crossover in such systems in a clear and unified way. We have obtained the following conclusions. First, when the wavenumber ratio $x\approx -1$, EIT, ATS, and EIT-ATS crossover exist for both ladder-I and ladder-II systems. Second, when $x$ is far from $-1$, EIT can occur but ATS is destroyed if the upper state of the ladder-I system is a Rydberg state. Third, ATS exists but EIT is not possible if the control field that couples the two lower states of the ladder-II system is a microwave field. Our theoretical analysis have applied to various ladder systems (including hot gases of Rubidium atoms, molecules, and Rydberg atoms, and so on), and the results obtained on the quantum interference characters agree well with experimental ones reported up to now. The results obtained here may have practical applications in optical information processing and transmission. \begin{acknowledgments} This work was supported by NSF-China under Grant Nos. 10874043 and 11174080. \end{acknowledgments}
\section{#1}} \newcommand\tr{\mathop{\mathrm tr}} \newcommand\Tr{\mathop{\mathrm Tr}} \newcommand\partder[2]{\frac{{\partial {#1}}}{{\partial {#2}}}} \newcommand\partderd[2]{{{\partial^2 {#1}}\over{{\partial {#2}}^2}}} \newcommand\partderh[3]{{{\partial^{#3} {#1}}\over{{\partial {#2}}^{#3}}}} \newcommand\partderm[3]{{{\partial^2 {#1}}\over{\partial {#2} \partial{#3} }}} \newcommand\partderM[6]{{{\partial^{#2} {#1}}\over{{\partial {#3}}^{#4}{\partial {#5}}^{#6} }}} \newcommand\funcder[2]{{{\delta {#1}}\over{\delta {#2}}}} \newcommand\Bil[2]{\Bigl\langle {#1} \Bigg\vert {#2} \Bigr\rangle} \newcommand\bil[2]{\left\langle {#1} \bigg\vert {#2} \right\rangle} \newcommand\me[2]{\left\langle {#1}\right|\left. {#2} \right\rangle} \newcommand\sbr[2]{\left\lbrack\,{#1}\, ,\,{#2}\,\right\rbrack} \newcommand\Sbr[2]{\Bigl\lbrack\,{#1}\, ,\,{#2}\,\Bigr\rbrack} \newcommand\Gbr[2]{\Bigl\lbrack\,{#1}\, ,\,{#2}\,\Bigr\} } \newcommand\pbr[2]{\{\,{#1}\, ,\,{#2}\,\}} \newcommand\Pbr[2]{\Bigl\{ \,{#1}\, ,\,{#2}\,\Bigr\}} \newcommand\pbbr[2]{\lcurl\,{#1}\, ,\,{#2}\,\rcurl} \renewcommand\a{\alpha} \renewcommand\b{\beta} \renewcommand\c{\chi} \renewcommand\d{\delta} \newcommand\D{\Delta} \newcommand\eps{\epsilon} \newcommand\vareps{\varepsilon} \newcommand\g{\gamma} \newcommand\G{\Gamma} \newcommand\grad{\nabla} \newcommand\h{\frac{1}{2}} \renewcommand\k{\kappa} \renewcommand\l{\lambda} \renewcommand\L{\Lambda} \newcommand\m{\mu} \newcommand\n{\nu} \renewcommand\o{\over} \newcommand\om{\omega} \renewcommand\O{\Omega} \newcommand\p{\phi} \newcommand\vp{\varphi} \renewcommand\P{\Phi} \newcommand\pa{\partial} \newcommand\tpa{{\tilde \partial}} \newcommand\bpa{{\bar \partial}} \newcommand\pr{\prime} \newcommand\ra{\rightarrow} \newcommand\lra{\longrightarrow} \renewcommand\r{\rho} \newcommand\s{\sigma} \renewcommand\S{\Sigma} \renewcommand\t{\tau} \renewcommand\th{\theta} \newcommand\bth{{\bar \theta}} \newcommand\Th{\Theta} \newcommand\z{\zeta} \newcommand\ti{\tilde} \newcommand\wti{\widetilde} \newcommand\twomat[4]{\left(\begin{array}{cc} {#1} & {#2} \\ {#3} & {#4} \end{array} \right)} \newcommand\threemat[9]{\left(\begin{array}{ccc} {#1} & {#2} & {#3}\\ {#4} & {#5} & {#6}\\ {#7} & {#8} & {#9} \end{array} \right)} \newcommand\cA{{\mathcal A}} \newcommand\cB{{\mathcal B}} \newcommand\cC{{\mathcal C}} \newcommand\cD{{\mathcal D}} \newcommand\cE{{\mathcal E}} \newcommand\cF{{\mathcal F}} \newcommand\cG{{\mathcal G}} \newcommand\cH{{\mathcal H}} \newcommand\cI{{\mathcal I}} \newcommand\cJ{{\mathcal J}} \newcommand\cK{{\mathcal K}} \newcommand\cL{{\mathcal L}} \newcommand\cM{{\mathcal M}} \newcommand\cN{{\mathcal N}} \newcommand\cO{{\mathcal O}} \newcommand\cP{{\mathcal P}} \newcommand\cQ{{\mathcal Q}} \newcommand\cR{{\mathcal R}} \newcommand\cS{{\mathcal S}} \newcommand\cT{{\mathcal T}} \newcommand\cU{{\mathcal U}} \newcommand\cV{{\mathcal V}} \newcommand\cX{{\mathcal X}} \newcommand\cW{{\mathcal W}} \newcommand\cY{{\mathcal Y}} \newcommand\cZ{{\mathcal Z}} \newcommand{\noindent}{\noindent} \newcommand{\ct}[1]{\cite{#1}} \newcommand{\bib}[1]{\bibitem{#1}} \newcommand\PRL[3]{\textsl{Phys. Rev. Lett.} \textbf{#1}, #3 (#2)} \newcommand\NPB[3]{\textsl{Nucl. Phys.} \textbf{B#1}, #3 (#2)} \newcommand\NPBFS[4]{\textsl{Nucl. Phys.} \textbf{B#2} [FS#1], #4 (#3)} \newcommand\CMP[3]{\textsl{Commun. Math. Phys.} \textbf{#1}, #3 (#2)} \newcommand\PRD[3]{\textsl{Phys. Rev.} \textbf{D#1}, #3 (#2)} \newcommand\PLA[3]{\textsl{Phys. Lett.} \textbf{#1A}, #3 (#2)} \newcommand\PLB[3]{\textsl{Phys. Lett.} \textbf{#1B}, #3 (#2)} \newcommand\CQG[3]{\textsl{Class. Quantum Grav.} \textbf{#1}, #3 (#2)} \newcommand\JMP[3]{\textsl{J. Math. Phys.} \textbf{#1}, #3 (#2)} \newcommand\PTP[3]{\textsl{Prog. Theor. Phys.} \textbf{#1}, #3 (#2)} \newcommand\SPTP[3]{\textsl{Suppl. Prog. Theor. Phys.} \textbf{#1}, #3 (#2)} \newcommand\AoP[3]{\textsl{Ann. of Phys.} \textbf{#1}, #3 (#2)} \newcommand\RMP[3]{\textsl{Rev. Mod. Phys.} \textbf{#1}, #3 (#2)} \newcommand\PR[3]{\textsl{Phys. Reports} \textbf{#1}, #3 (#2)} \newcommand\FAP[3]{\textsl{Funkt. Anal. Prilozheniya} \textbf{#1}, #3 (#2)} \newcommand\FAaIA[3]{\textsl{Funct. Anal. Appl.} \textbf{#1}, #3 (#2)} \newcommand\TAMS[3]{\textsl{Trans. Am. Math. Soc.} \textbf{#1}, #3 (#2)} \newcommand\InvM[3]{\textsl{Invent. Math.} \textbf{#1}, #3 (#2)} \newcommand\AdM[3]{\textsl{Advances in Math.} \textbf{#1}, #3 (#2)} \newcommand\PNAS[3]{\textsl{Proc. Natl. Acad. Sci. USA} \textbf{#1}, #3 (#2)} \newcommand\LMP[3]{\textsl{Letters in Math. Phys.} \textbf{#1}, #3 (#2)} \newcommand\IJMPA[3]{\textsl{Int. J. Mod. Phys.} \textbf{A#1}, #3 (#2)} \newcommand\IJMPD[3]{\textsl{Int. J. Mod. Phys.} \textbf{D#1}, #3 (#2)} \newcommand\TMP[3]{\textsl{Theor. Math. Phys.} \textbf{#1}, #3 (#2)} \newcommand\JPA[3]{\textsl{J. Physics} \textbf{A#1}, #3 (#2)} \newcommand\JSM[3]{\textsl{J. Soviet Math.} \textbf{#1}, #3 (#2)} \newcommand\MPLA[3]{\textsl{Mod. Phys. Lett.} \textbf{A#1}, #3 (#2)} \newcommand\JETP[3]{\textsl{Sov. Phys. JETP} \textbf{#1}, #3 (#2)} \newcommand\JETPL[3]{\textsl{ Sov. Phys. JETP Lett.} \textbf{#1}, #3 (#2)} \newcommand\PHSA[3]{\textsl{Physica} \textbf{A#1}, #3 (#2)} \newcommand\PHSD[3]{\textsl{Physica} \textbf{D#1}, #3 (#2)} \newcommand\JPSJ[3]{\textsl{J. Phys. Soc. Jpn.} \textbf{#1}, #3 (#2)} \newcommand\JGP[3]{\textsl{J. Geom. Phys.} \textbf{#1}, #3 (#2)} \newcommand\hepth[1]{\textsl{hep-th/#1}} \newcommand\nlin[1]{\textsl{nlin.SI/#1}} \newcommand\solvint[1]{\textsl{solv-int/#1}} \newcommand\Xdot{\stackrel{.}{X}} \newcommand\ydot{\stackrel{.}{y}} \newcommand\yddot{\stackrel{..}{y}} \begin{document} \title{Confining Boundary conditions from dynamical Coupling Constants} \author {E. I. Guendelman \footnote{e-mail: <EMAIL>}} \address{Physics Department, Ben Gurion University of the Negev, Beer Sheva 84105, Israel} \author {R. Steiner \footnote{e-mail: <EMAIL>}} \address{Physics Department, Ben Gurion University of the Negev, Beer Sheva 84105, Israel} \begin{abstract} It is shown that it is possible to consistently and gauge invariantly formulate models where the coupling constant is a non trivial function of a scalar field . In the $U(1)$ case the coupling to the gauge field contains a term of the form $g(\phi)j_\mu (A^{\mu} +\partial^{\mu}B)$ where $B$ is an auxiliary field and $j_\mu$ is the Dirac current. The scalar field $\phi$ determines the local value of the coupling of the gauge field to the Dirac particle. The consistency of the equations determine the condition $\partial^{\mu}\phi j_\mu = 0$ which implies that the Dirac current cannot have a component in the direction of the gradient of the scalar field. As a consequence, if $\phi$ has a soliton behaviour, like defining a bubble that connects two vacuua, we obtain that the Dirac current cannot have a flux through the wall of the bubble, defining a confinement mechanism where the fermions are kept inside those bags. Consistent models with time dependent fine structure constant can be also constructed \end{abstract} \pacs{14.70.Bh, 12.20.-m, 11.40.Dw \maketitle \date{\today \maketitle \section{Introduction} In this paper it will be shown that an alternative coupling of gauge fields to charged particles is possible in such a way that the the coupling constants can be dynamical. The gauge coupling has a term of the form $g(\phi)j_\mu (A^{\mu} +\partial^{\mu}B)$ where $B$ is an auxiliary field and the current $j_\mu$ is the Dirac current. Before studying the issue of dynamical gauging, we review how the $ B $ field can be used in a gauge theory playing the role of a scalar gauge field \cite{Scalar gauge field}.That can be used to define a new type of convariante derivative. Starting with a complex scalar field we now gauge the phase symmetry of $\phi$ by introducing a real, scalar $B( x _\mu)$ and two types of covariant derivatives as \begin{equation} \label{cov-ab} D ^A _\mu = \partial _\mu + i e A_\mu ~~~;~~~ D ^B _\mu = \partial _\mu + i e \partial _\mu B ~. \end{equation} The gauge transformation of the complex scalar, vector gauge field and scalar gauge field have the following gauge transformation \begin{equation} \label{gauge-trans} \phi \rightarrow e^{i e \Lambda} \phi ~~~;~~~ A_\mu \rightarrow A_\mu + \partial _\mu \Lambda ~~~;~~~ B \rightarrow B - \Lambda ~. \end{equation} It is easy to see that terms like $D ^A _\mu \phi$ and $D ^B _\mu \phi$, will be covariant under \ref{gauge-trans} that is they transform the same way as the scalar field $ \phi $ and their complex conjugates will transfor as $\phi^*$ does. Thus one can generate kinetic energy type terms like $(D ^A _\mu \phi) (D ^{A \mu} \phi)^*$, $(D ^B _\mu \phi) (D ^{B \mu} \phi)^*$, $(D ^A _\mu \phi) (D ^{B \mu} \phi)^*$, and $(D ^B _\mu \phi) (D ^{A \mu} \phi)^*$. Unlike $A_\mu$ where one can add a gauge invariant kinetic term involving only $A_\mu$ (i.e. $F_{\mu \nu} F^{\mu \nu}$) this is apparently not possible to do for the scalar gauge field $B$. However note that the term $A_\mu + \partial _\mu B$ is invariant under the gauge field transformation alone (i.e. $A_\mu \rightarrow A_\mu + \partial _\mu \Lambda$ and $B \rightarrow B - \Lambda$). Thus one can add a term like $(A_\mu + \partial _\mu B)(A^\mu + \partial ^\mu B)$ to the Lagrangian which is invariant with respect to the gauge field part only of the gauge transformation in \ref{gauge-trans}. This gauge invariant term will lead to both mass-like terms for the vector gauge field and kinetic energy-like terms for the scalar gauge field. In total a general Lagrangian which respects the new gauge transformation and is a generalization of the usual gauge Lagrangian, which has the form \begin{eqnarray} \label{u2} & {\cal L} = c_1 D^A _\mu \phi (D ^{A \mu} \phi ) ^* + c_2 D^B _\mu \phi (D ^{B \mu} \phi )^* + c_3 D^A _\mu \phi (D ^{B \mu} \phi )^* \nonumber \\ & + c_4 D^B _\mu \phi (D ^{A \mu} \phi )^* - V(\phi) \nonumber \\ & - \frac{1}{4} F_{\mu \nu} F^{\mu \nu} + c_5 (A_\mu + \partial _\mu B)(A^\mu + \partial ^\mu B)~, \end{eqnarray} where $c_i$'s are constants that should be fixed to get a physically acceptable Lagrangian where $ c_{3}=c^{*}_{4} $ and $ c_{1}\, , c_{2}\, , c_{5} $ are real.\\ At first glance one might conclude that $B(x)$ is not a physical field, it appears that one could "gauge" it away by taking $\Lambda = B(x)$ in \ref{gauge-trans}. However in the case of symmetry breaking when one introduces a complex charged scalar field that get expectation value which is not zero, one must be careful since this would imply that the gauge transformation of the field $\phi$ would be of the form $\phi \rightarrow e^{i e B} \phi$ i.e. the phase factor would be fixed by the gauge transformation of $B(x)$. In this situation one would no longer to able to use the usual unitary gauge transformation to eliminate the Goldstone boson in the case when one has spontaneous symmetry breaking. \\ Indeed in the case when there is spontaneous symmetry breaking, the physical gauge (the generalization of the unitary gauge) is not the gauge $B=0$, as discussed in \cite{Scalar gauge field}, it is a gauge where the scalar gauge field $B$ has to be taken proportional to the phase of the scalar field, with a proportionality constant that depends on the expectation value of the Higgs field. Also, in general there are the three degrees of freedom of a massive vector field and the Higgs field, and therefore all together five degrees of freedom. If there is no spontaneous symmetry breaking, fixing the gauge $B=0$ does not coincide with the gauge that allows us to display that the photon has two polarizations, this gauge being Coulomb gauge. This is true even if we do not add a gauge invariant mass term (possible given the existence of the $B$ field). By fixing the Coulomb gauge, which will make the the photon manifestly having only two polarizations, we will have already exhausted the gauge freedom and cannot in general in addition require the gauge $B=0$. So, in Coulomb gauge where the photon will have two polarizations, the $B$ field and in addition the two other scalars, the real and imaginary part of $\phi$ all represent true degrees of freedom, so altogether we have five degrees of freedom, the same as the case displaying spontaneous symmetry breaking. If we add a gauge invariant mass term, even when there is no spontaneous symmetry breaking (the $c_5$ term), in the gauge $B=0$ we have three pollarizations of the massive vector field and still the real and imaginary parts of the complex scalar field $\phi$, still five degrees of freedom altogether. Also, one can use this kind of field to define a coupling of electrodynamics to charged scalar field which enjoys only global $ U(1) $ symmetry \cite{GLOBALQED} In "Global scalar QED" , we work with the following Lagrangian density \begin{equation} \label{GlobalQED} \mathcal{L}= g^{\mu\nu}\frac{\partial\psi^{*}}{\partial x^{\mu}}\frac{\partial \psi}{\partial x^{\nu}}-U(\psi^{*}\psi) -\frac{1}{4} F^{\mu \nu}F_{\mu \nu} + j_\mu (A^{\mu} +\partial^{\mu}B) \end{equation} where \begin{equation} j_\mu = ie(\psi^{*}\frac{\partial\psi}{\partial x^\mu} -\psi\frac{\partial\psi^{*}}{\partial x^\mu}) \end{equation} and where we have also allowed an arbitrary potential $U(\psi^{*}\psi)$ to allow for the possibility of spontaneous breaking of symmetry. The model is separately invariant under local gauge transformations \begin{equation} \label{GTGlobalQED} A^\mu \rightarrow A^\mu + \partial^\mu \Lambda \textrm{; } \ \ \ B \rightarrow B - \Lambda \end{equation} and the independent global phase transformations \begin{equation} \psi \rightarrow exp (i\chi) \psi \end{equation} The use of a gauge invariant combination $(A^{\mu} +\partial^{\mu}B)$ can be utilized for the construction of mass terms\cite{Stueckelberg} or both mass terms and couplings to a current defined from the gradient of a scalar in the form $(A^{\mu} +\partial^{\mu}B)\partial_{\mu}A$ \cite{Guendelman}. In the non abelian case mass terms constructed along these lines have been considered by Cornwall \cite{Cornwall}. Since the subject of this paper is electromagnetic couplings of photons and there is absolutely no evidence for a photon mass, we will disregard such type of mass terms and concentrate on the implications of the $(A^{\mu} +\partial^{\mu}B)j_{\mu}$ couplings. It is also interesting to point out the use of scalars instead of vectors fields has been studied in \cite{Singleton} in their general study of gauge procedure with gauge fields of various ranks. \section{Confining Boundary conditions from dynamical Coupling Constants} In this chapter and the following one we will show that dynamical Coupling Constants can lead to confinement. The dynamical Coupling Constants is dynamical mostly at the boundary of the confinement and out side the boundary. Lets proceed with the same consideration as in the chapter before, but with Dirac field $ \psi $ and real scalar field $ \phi $, with the action: \begin{eqnarray}\label{Dirac:boundary} & S=\int{\bar{\psi}(i\gamma^{\mu}\partial_{\mu}-m+e\gamma^{\mu}A_{\mu})\psi \,d^{4}x} - \frac{1}{4} \int{F^{\mu\nu}F_{\mu\nu}\,d^{4}x} \nonumber \\ & +\int d^{4}x [ g(\phi)\bar{\psi}\gamma^{\mu}\psi(A_{\mu}+\partial_{\mu}B)\nonumber \\ & +\frac{1}{2}\partial_{\mu}\phi\partial^{\mu}\phi - V(\phi)] \end{eqnarray} The model is invariant under local gauge transformations \begin{equation} \label{GTGlobalQED2} A^\mu \rightarrow A^\mu + \partial^\mu \Lambda \textrm{; } \ \ \ B \rightarrow B - \Lambda \end{equation} \begin{equation} \psi \rightarrow exp (ie\Lambda) \psi \end{equation} The Noether current conservation law for global symmetry $ \psi \rightarrow e^{i\theta}\psi $, $\theta= constant$ is, \begin{equation} \partial_{\mu} j^{\mu}_{N}=(\partial_{\mu})(\frac{\partial \mathcal{L}}{\partial \psi_{,\mu}}\delta\psi) = \partial_{\mu}(\bar{\psi}\gamma^{\mu}\psi)=0 \end{equation} The gauge field equation, containing in the right hand side the current which is the source of the gauge field is: \begin{eqnarray} \partial_{\mu}F^{\mu\nu}=(e+g(\phi))\bar{\psi}\gamma^{\nu}\psi = j^{\nu}_{Source} \end{eqnarray} By considering the divergence of the above equation, we obtain the additional conservation law: \begin{eqnarray}\label{bag conservation law} & \partial_{\mu}j^{\mu}_{Source}=\partial_{\mu}(g(\phi))\bar{\psi}\gamma^{\mu}\psi + g(\phi) \partial_{\mu}(\bar{\psi}\gamma^{\mu}\psi)\nonumber \\ & = \partial_{\mu}(g(\phi))\bar{\psi}\gamma^{\mu}\psi = 0 \end{eqnarray} If we have scalar potential $ V(\phi) $ with domain wall between two false vacuum state (see figure 2) , than because of the transition of the scalar field on the domain wall $ \partial_{\mu}(g(\phi))=\frac{\partial g(\phi)}{\partial \phi}\partial_{\mu}\phi=\frac{\partial g(\phi)}{\partial \phi}n_{\mu}f \neq 0 $ (see figure 3) We must conclude that $ n_{\mu}(\bar{\psi}\gamma^{\mu}\psi)\mid_{x=domain\,wall}=0 $. This means that on the domain wall there is no communication between the two sector of the domain, which give a confinement (see figure 1). \begin{figure}[h!] \centering \includegraphics[scale=.4]{paint1.jpg} \caption[Caption for LOF]{{\tiny on the domain wall there is no communication between the two sector of the domain, which give a confinement}} \end{figure} \begin{figure}[h!] \centering \includegraphics[scale=.5]{Vphi.jpg} \caption[Caption for LOF]{{\tiny Scalar potential $ V(\phi) $ with domain wall between two false vacuum state }} \includegraphics[scale=.5]{tanh.png} \caption[Caption for LOF]{{\tiny transition of the scalar field on the domain wall $ \partial_{\mu}(g(\phi))=\frac{\partial g(\phi)}{\partial \phi}\partial_{\mu}\phi=\frac{\partial g(\phi)}{\partial \phi}n_{\mu}f \neq 0 $}} \end{figure} \section{Confining Boundary conditions holding at a specific region in a domain wall} The constraint that we have gotten in the last chapter is too strong and non trivial and holds everywhere even if the gradients are small , but we wants to have constraint only in the region were the domain wall is located, so lets proceed with the same consideration as above but with coupling of the gauge field and the scalar field so the constraint will follow from an equation of motion and only at a specific location. We will see that if in the action we add an additional term (the $ 1/l_0 $ term): \begin{eqnarray}\label{Dirac:boundary2} & S=\int{\bar{\psi}(i\gamma^{\mu}\partial_{\mu}-m+e\gamma^{\mu}A_{\mu})\psi \,d^{4}x} - \frac{1}{4} \int{F^{\mu\nu}F_{\mu\nu}\,d^{4}x} \nonumber \\ & +\int d^{4}x [ g(\phi)\bar{\psi}\gamma^{\mu}\psi(A_{\mu}+\partial_{\mu}B)+\frac{1}{2}\partial_{\mu}\phi\partial^{\mu}\phi - V(\phi) \nonumber \\ & +\frac{1}{l_{0}}\partial_{\mu} \phi (A^{\mu} +\partial^{\mu}B)] \end{eqnarray} The model is invariant under local gauge transformations as in equation \ref{GTGlobalQED2}. The Noether current conservation law for global symmetry $ \psi \rightarrow e^{i\theta}\psi $, $\theta= constant$ is, \begin{equation}\label{Noether conservation law 2} \partial_{\mu} j^{\mu}_{N}=(\partial_{\mu})(\frac{\partial \mathcal{L}}{\partial \psi_{,\mu}}\delta\psi) = \partial_{\mu}(\bar{\psi}\gamma^{\mu}\psi)=0 \end{equation} The gauge field equation, containing in the right hand side the current which is the source of the gauge field is: \begin{eqnarray} \partial_{\mu}F^{\mu\nu}=(e+g(\phi))\bar{\psi}\gamma^{\nu}\psi +\frac{1}{l_{0}}\partial^{\nu}\phi = j^{\nu}_{Source} \end{eqnarray} We can see that we have additional long range term to the constraint \ref{bag conservation law}. By considering the divergence of the above equation, we obtain the additional conservation law: \begin{eqnarray}\label{bag conservation law 2} & \partial_{\mu}j^{\mu}_{Source}=\partial_{\mu}(g(\phi))\bar{\psi}\gamma^{\mu}\psi + g(\phi) \partial_{\mu}(\bar{\psi}\gamma^{\mu}\psi) +\frac{1}{l_{0}}\partial^{\mu}\partial_{\mu}\phi \nonumber \\ & = \partial_{\mu}(g(\phi))\bar{\psi}\gamma^{\mu}\psi \nonumber \\ & +\frac{1}{l_{0}}\partial^{\mu}\partial_{\mu}\phi = 0 \end{eqnarray} The variation on the action by $ \phi $ gives: \begin{eqnarray}\label{bag EOM phi} & \partial^{\mu}\partial_{\mu}\phi + \frac{\partial V}{\partial \phi} - \frac{1}{l_{0}} \partial_{\mu}(A^{\mu}+\partial^{\mu}B) \nonumber \\ & + \frac{\partial g(\phi)}{\partial \phi}\bar{\psi}\gamma^{\mu}\psi(A_{\mu}+\partial_{\mu}B)=0 \end{eqnarray} Lets consider a scalar potential $ V(\phi) $ with domain wall between two false vacuum state $ V(\nu_{1}) $ and $ V(\nu_{2}) $ (see figure 2) , and statically solution. Than for finite energy solution we need to demand that $ \partial_{i}\phi(\pm \infty)=0 $ and $ \phi(\infty)=\nu_{1} $ and $ \phi(-\infty)=\nu_{2} $. From Rolle's mathematical theorem, we must conclude that at some point of the transition of the scalar field on the domain wall we have that $ \partial_{i}\partial_{i}\phi=0 $ , so equation \ref{bag conservation law 2} on some point on the transition reads: \begin{eqnarray} \partial_{\mu}(g(\phi))\bar{\psi}\gamma^{\mu}\psi=0 \end{eqnarray} because on the point of the transition were $ \partial_{i}\partial_{i}\phi=0 $, $ \partial_{i}\phi \neq 0 $ than $ \partial_{\mu}(g(\phi))=\frac{\partial g(\phi)}{\partial \phi}\partial_{\mu}\phi=\frac{\partial g(\phi)}{\partial \phi}n_{\mu}f \neq 0 $ (see figure 3) \\So we must conclude that $ n_{\mu}(\bar{\psi}\gamma^{\mu}\psi)\mid_{x=domain\,wall}=0 $. This means that on the domain wall there is no communication between the two sector of the domain, which give a confinement (see figure 1). Also we can see that the coupling constant far from the domain wall is constant. \section{Consistent models with time dependent fine structure constant } The formalism developed here provides the possibility of formulating a consistent formalism where the effective electric charge can change with space and time such possibility have been considered in cosmological contexts. Many papers have been published on the subject of the variation of the fine structure constant. There are some clues that show that the structure constant has been slightly variable, although this is not generally agreed. Bekenstein \cite{Bekenstein} has shown a different approach to formulate consistently a theory with a variable coupling constant. The Oklo natural geological fission reactor has lead to a measurement that some claim it implies the structure constant has changed by a small amount of the order of $ \frac{\dot{\alpha}}{\alpha} \approx 1\times 10^{-7} $ \cite{Uzan}. \section{Discussion and Conclusions} It is shown that it is possible to consistently and gauge invariantly formulate models where the coupling constant is a non trivial function of a scalar field . In the $U(1)$ case the coupling to the gauge field contains a term of the form $g(\phi)j_\mu (A^{\mu} +\partial^{\mu}B)$ where $B$ is an auxiliary field and $j_\mu$ is the Dirac current. The scalar field $\phi$ determines the local value of the coupling of the gauge field to the Dirac particle. The consistency of the equations determine the condition $\partial^{\mu}\phi j_\mu = 0$ which implies that the Dirac current cannot have a component in the direction of the gradient of the scalar field. As a consequence, if $\phi$ has a soliton behaviour, like defining a bubble that connects two vacuua, we obtain that the Dirac current cannot have a flux through the wall of the bubble, defining a confinement mechanism where the fermions are kept inside those bags. This gives rise to a condition that was considered also for example in M.I.T bag model \cite{mit bag 2} (for a review see \cite{mit bag}), but the way to obtain it is quite different. It will interesting to study new effects that may appear when the gauge symmetry is broken. For example a modification of the M.I.T bag model when there is symmetry breaking as been study in ref \cite{Unconfined quarks and gluons}, as we pointed before the physical gauge in this case is not $ B=0 $ but it is a gauge where the $ B $ field and the phase of the Higgs field are proportional with the proportionality constant that depends on the expectation value of the Higgs field \cite{Scalar gauge field}. The formalism developed here provides the possibility of formulating a consistent formalism where the effective electric charge can change with space and time such possibility have been considered in cosmological contexts. \section{Acknowledgements} We are very grateful to Douglas Singleton, Holger Nielsen, Jon Chkareuli and Rahul Kumar for very useful conversations.
\section{Introduction} The geometry and topology of $3$-manifolds with non-negative curvature has been an active field of research during the last century. One of the properties of these $3$-manifolds is the fact that they generally admit compact (without boundary) embedded minimal surfaces. Many authors have contributed to the study of compact minimal surfaces in order to understand the geometry of the $3$-manifold (see, for instance,~\cite{AR,GR,MSY, SY}). In 1970, Lawson~\cite{Lawson} proved that any compact orientable surface can be minimally embedded in the constant sectional curvature $3$-sphere $\mathbb{S}^3$. Lawson's result has also been extended to different $3$-manifolds such as the Berger spheres (see~\cite{Torralbo}). The aim of this paper is to determine which compact surfaces admit minimal embeddings in the homogeneous product $3$-manifolds $\mathbb{S}^2\times\mathbb{S}^1(r)$, $r>0$, where $\mathbb{S}^2$ denotes the constant curvature one sphere, and $\mathbb{S}^1(r)$ is the radius $r$ circle, i.e., $\mathbb{S}^1(r)$ has length $2\pi r$ and will be identified with $\mathbb{R}/2\pi r$. These $3$-manifolds have a distinguished unit vector field $\xi_{(p,t)} = (0, 1)$, which is parallel and generates the \emph{fibers} of the natural Riemannian submersion $\pi\colon \mathbb{S}^2 \times \mathbb{S}^1(r) \rightarrow \mathbb{S}^2$. Given a vector field $X$ we will say that $X$ is \emph{horizontal} if it is orthogonal to $\xi$ and \emph{vertical} if it is parallel to $\xi$. Also, given a differentiable curve $\alpha$ in $\mathbb{S}^2\times \mathbb{S}^1(r)$ we will say that it is \emph{horizontal} (resp.\ \emph{vertical}) if its tangent vector is pointwise horizontal (resp.\ vertical). We can express the curvature tensor in terms of the metric and $\xi$ as \begin{equation}\label{eq:curvature} \begin{split} R(X, Y)Z = &\langle Y, Z \rangle X - \langle X, Z\rangle Y + \langle X, \xi\rangle \langle Z, \xi \rangle Y - \langle Y, \xi\rangle \langle Z, \xi \rangle X + \\ &(\langle X, Z \rangle \langle Y, \xi\rangle - \langle Y, Z \rangle \langle X, \xi \rangle )\xi, \end{split} \end{equation} and hence $\mathrm{Ric}(X) = \|X\|^2 - \langle X, \xi\rangle^2\geq 0$ for each vector field $X$. We will begin by briefly discussing known examples with low genus. Recall that a non-orientable compact surface $\Sigma$ has (non-orientable) genus $g$ if it is topologically equivalent to the connected sum of $g$ projective planes (equivalently, $\chi(\Sigma) = 2 - g$, where $\chi(\Sigma)$ stands for the Euler characteristic of the surface $\Sigma$). First of all, the only spheres minimally immersed in $\mathbb{S}^2\times\mathbb{S}^1(r)$ are the horizontal slices $\mathbb{S}^2\times\{t_0\}$ (it follows from lifting such a sphere to a minimal sphere in $\mathbb{S} ^2\times\mathbb{R}$ and applying the maximum principle with respect to a foliation of $\mathbb{S} ^2\times\mathbb{R}$ by slices), which are always embedded. This implies that the projective plane cannot be minimally immersed in $\mathbb{S}^2\times\mathbb{S}^1(r)$ for any $r>0$. Moreover, they are the only compact minimal stable surfaces in $\mathbb{S}^2 \times \mathbb{S}^1(r)$ (cf.~\cite[Corollary~1]{TU13}). If the Euler characteristic is zero the situation is quite different in both orientable and non-orientable cases: \begin{enumerate}[$\bullet$] \item On the one hand, \emph{vertical helicoids} in $\mathbb{S}^2\times\mathbb{R}$ form a $1$-parameter family of minimal surfaces ruled by ho\-ri\-zon\-tal geo\-de\-sics invariant under a $1$-parameter group of ambient screw-motions, see~\cite{Rosenberg}. Each vertical helicoid is determined by its \emph{pitch} $\ell \in [0, +\infty[$ (i.e., twice the distance between successive sheets of the helicoids). Hence, taking quotients by the vertical translation $(p,t)\mapsto(p,t+2\pi r)$ in $\mathbb{S}^2\times\mathbb{R}$, vertical helicoids with appropriate pitches give rise to embedded minimal tori ($\ell = 2\pi r$) and Klein bottles ($\ell = 4\pi r$) in $\mathbb{S}^2\times\mathbb{S}^1(r)$ for any $r > 0$. If $\ell \to +\infty$, the helicoid converges to the \emph{vertical cylinder} $\Gamma \times \mathbb{R}$, $\Gamma$ being a geodesic of $\mathbb{S}^2$. \item On the other hand, one can consider examples foliated by circles: if the centers of the circles lie in the same vertical axis $\{p_0\} \times \mathbb{R}$, then they are the rotationally invariant examples studied by Pedrosa and Ritoré~\cite{PR99}. They are called \emph{unduloids} by analogy to the Delaunay surfaces and form a $1$-parameter family of minimal surfaces in $\mathbb{S}^2\times\mathbb{R}$. Moreover, they are singly periodic in the $\mathbb{R}$ factor with period $T \in \,]0, 2\pi[$ (if $T \to 0$ the unduloid converges to a double cover of a slice, whereas if $T \to 2\pi$ the unduloid converges to the vertical cylinder). If the centers of the circle do not lie in the same vertical axis, we get the Riemann-type examples studied by Hauswirth~\cite[Theorem 4.1]{Haus06}. They are also singly periodic and produce embedded minimal tori in $\mathbb{S}^2\times \mathbb{S}^1(r)$ for all $r > 0$ (see also~\cite[\S 7]{Ros03} for a beautiful description). Note that, thanks to \cite[Theorem 4.8]{Urb12}, a compact minimal surface $\Sigma$ immersed in $\mathbb{S}^2 \times \mathbb{S}^1(r)$, $r \geq 1$, foliated by circles has index greater than or equal to one (the index is one only for $r=1$ and $\Sigma=\Gamma\times\mathbb{S}^1(1)$, $\Gamma\subset\mathbb{S}^2$ being a great circle). A similar estimation of the index for $r < 1$ remains an open question (see~\cite[Remark 4.9]{Urb12}). \end{enumerate} In the higher genus case, Rosenberg~\cite{Rosenberg} constructed compact minimal surfaces in $\mathbb{S}^2\times\mathbb{S}^1(r)$, for all $r>0$, by a technique similar to Lawson's. More specifically, he defined a polygon in $\mathbb{S}^2\times\mathbb{R}$ made out of vertical and horizontal geodesics depending on two parameters $\gamma=\frac{\pi}{d}$, $d\in\mathbb{N}$, and $\tilde{h}>0$ (see Figure~\ref{f:boundary1}). Then he solved the Plateau problem with respect to this contour and obtained a minimal disk that can be reflected across its boundary in virtue of Schwarz's reflection principle. This procedure gives a complete surface that projects to the quotient $\mathbb{S}^2\times\mathbb{S}^1(r)$, $r=\frac{\tilde{h}}{\pi}$, as a compact embedded minimal surface. Unfortunately, it fails to be orientable as stated in~\cite{Rosenberg}, being a non-orientable surface of Euler characteristic $\chi=2(1-d)$ so its (non-orientable) genus $k=2-\chi=2d$ is always even. One can solve this issue by considering $r=2\frac{\tilde{h}}{\pi}$ (i.e., doubling the vertical length, which represents a $2$-fold cover of the non-orientable examples), so the resulting minimal surface has Euler characteristic $\chi=4(1-d)$ and its (orientable) genus is $g=1-\frac{\chi}{2}=2d-1$, which turns out to be odd. When $d = 2m$ is even, Rosenberg's examples induce one-sided compact non-orientable embedded minimal surfaces in the quotient $\R\mathrm{P}^2 \times \mathbb{S}^1(r)$ of odd genus $1+2m$, $m \geq 1$. Very recently, Hoffman, Traizet and White obtained a class of properly embedded minimal surfaces in $\mathbb{S}^2\times\mathbb{R}$ called \emph{periodic} genus $g$ helicoids. More explicitly, given a genus $g\geq 1$, a radius $r>0$, and a helicoid $H\subset\mathbb{S}^2\times\mathbb{R}$ containing the horizontal geodesics $\Gamma\times\{0\}$ and $\Gamma\times\{\pm\pi r\}$ for some great circle $\Gamma\subset\mathbb{S}^2$, \cite[Theorem 1]{HTW} yields the existence of two compact orientable embedded minimal surfaces $M^+$ and $M^-$ with genus $g$ in the quotient of $\mathbb{S}^2\times\mathbb{R}$ by the translation $(p,t)\mapsto(p,t+2\pi r)$. The existence of infinitely-many non-congruent helicoids $H$ satisfying the conditions above implies the existence of infinitely-many non-congruent compact orientable embedded minimal surfaces in $\mathbb{S}^2\times\mathbb{S}^1(r)$ with genus $g$. The even genera examples of Hoffman-Traizet-White produce compact minimal surfaces in the quotient $\R\mathrm{P}^2 \times \mathbb{S}^1(r)$, for all $r>0$. Finally, it is worth mentioning the work of Coutant~\cite[Theorem 1.0.2]{Coutant12}, where Riemann-Wei type minimal surfaces in $\mathbb{S}^2\times \mathbb{R}$ are constructed. Some of them give rise to compact embedded orientable minimal surfaces of arbitrary genus $g \geq 4$ in the quotient $\mathbb{S}^2 \times \mathbb{S}^1(r)$ for $r$ small enough. In Section~\ref{sec:non-existence} we will state the main result of the paper proving that there are no compact embedded minimal surfaces in $\mathbb{S}^2\times\mathbb{S}^1(r)$ of the remaining topological types (i.e., non-orientable odd-genus surfaces) for any $r>0$: \begin{quote} \emph{Every compact surface but the odd Euler characteristic ones can be minimally embedded in $\mathbb{S}^2\times \mathbb{S}^1(r)$ for any $r > 0$.} \end{quote} We will also discuss some interesting topological properties of compact minimal surfaces in $\mathbb{S}^2\times\mathbb{R}$ that will lead to the non-existence result. Although we have shown that there exist compact embedded minimal examples of all allowed topological types in $\mathbb{S}^2\times\mathbb{S}^1(r)$, Section~\ref{sec:examples} will be devoted to obtain Schwarz P-type examples in $\mathbb{S}^2\times\mathbb{S}^1(r)$. More precisely: \begin{enumerate}[$\bullet$] \item \emph{Triply-periodic odd-genus orientable examples invariant by the isometry group of tilings of the sphere and a vertical translation (see Figures~\ref{f:cube-tessellation} and~\ref{f:balloon-tessellation})}. \item \emph{Arbitrary genus orientable surfaces in $\mathbb{S}^2\times \mathbb{S}^1(r)$ for $r$ small enough (cf.\ Proposition~\ref{prop:orientable-arbitrary-genus-examples}).} \end{enumerate} All the examples will be constructed via Lawson's technique~\cite{Lawson}, and complementing it by considering the conjugate minimal surface (see~\cite{HST}) instead of the original solution to the Plateau problem for a geodesic polygon. In this case, it is well known that the boundary lines of the conjugate surface are curves of planar symmetry since the initial surface is bounded by geodesic curves. Again, reflecting the conjugate surface across its edges will produce a complete example. This technique, known as \emph{conjugate Plateau construction} has been applied in different situations (e.g., see~\cite{MT,Plehnert,Plehnert2}). It is worth mentioning that a similar construction leads to classical P-Schwarz minimal surfaces in $\mathbb{R}^3$, but our construction can be also implemented in the product space $\mathbb{H}^2\times\mathbb{R}$ to produce embedded triply-periodic minimal surfaces with the symmetries of a tessellation of $\mathbb{H}^2$ by regular polygons together with a $1$-parameter group of vertical translations. We would also like to mention the minimal surfaces of $\mathbb{H}^2\times \mathbb{R}$ constructed in~\cite{MRR} also by similar techniques, specially the ones showed in~\cite[Section~6]{MRR} that are invariant by the isometry group of a regular tiling of $\mathbb{H}^2$. We would like to thank Harold Rosenberg for his valuable comments that helped us improve this paper, as well as the referee, whose thorough report has encouraged us to clarify some aspects of the original manuscript. \section{Topological non-existence results}\label{sec:non-existence} We will begin by describing the intersection of two compact minimal surfaces in $\mathbb{S}^2\times\mathbb{S}^1(r)$. The ideas in the proof are adapted from those of Frankel~\cite{Frankel} (see also~\cite{AR,GR}). In~\cite[Theorem 4.3]{Rosenberg}, a similar result is proved for properly embedded minimal surfaces in $\mathbb{S}^2\times\mathbb{R}$ by using a different approach. \begin{proposition}\label{prop:intersection} Let $\Sigma_1$ and $\Sigma_2$ be two compact minimal surfaces immersed in $\mathbb{S}^2\times\mathbb{S}^1(r)$. If $\Sigma_1\cap\Sigma_2=\emptyset$, then $\Sigma_1$ and $\Sigma_2$ are two horizontal slices. \end{proposition} \begin{proof} Let $\gamma\colon [a,b]\to\mathbb{S}^2\times\mathbb{S}^1(r)$ be a unit-speed curve satisfying $\gamma(a)\in\Sigma_1$, $\gamma(b)\in\Sigma_2$ and minimizing the distance from $\Sigma_1$ to $\Sigma_2$. This guarantees that $\gamma$ is orthogonal to $\Sigma_1$ at $\gamma(a)$ and to $\Sigma_2$ at $\gamma(b)$. For any parallel vector field $X$ along $\gamma$ and orthogonal to $\gamma'$, we can produce a variation $\gamma_t$ of $\gamma=\gamma_0$ with variational field $X$. Note that $\gamma_t$ can be chosen so that $\gamma_t(a)\in\Sigma_1$ and $\gamma_t(b)\in\Sigma_2$ for all $t$, since $X$ is orthogonal to $\gamma'$. Let $\{e_1, e_2, \gamma'(a)\}$ an orthonormal reference at $T_{\gamma(a)}\mathbb{S}^2\times \mathbb{S}^1(r)$ and $X_j$ the parallel transported vector field along $\gamma$ with $X_j(\gamma(a)) = e_j$, $j = 1, 2$. The function $\ell_X(t)=\mathrm{Length}(\gamma_t)$ satisfies $\ell_X'(0)=0$ and \begin{equation}\label{eqn:second-variation} \ell''_{X_1}(0) + \ell''_{X_2}(0) = H_1 - H_2 -\int_\gamma\mathrm{Ric}(\gamma') = -\int_\gamma \mathrm{Ric}(\gamma'), \end{equation} where $H_j$ is the mean curvature of $\Sigma_j$, $j = 1, 2$ (see~\cite[pages~69-70]{Frankel}). The minimization property tells us that $\ell''_{X_1}(0) + \ell''_{X_2}(0)\geq 0$. Since $\mathrm{Ric}(Y)\geq 0$ for any vector field $Y$ in $\mathbb{S}^2\times\mathbb{S}^1(r)$ and $\mathrm{Ric}(Y)=0$ if and only if $Y$ is vertical (see~\eqref{eq:curvature}), we deduce from equation~\eqref{eqn:second-variation} that $\gamma'$ is vertical. This means that the distance $d$ from $\Sigma_1$ to $\Sigma_2$ is realized by a vertical geodesic. Hence, translating $\Sigma_1$ vertically by distance $d$, we produce a contact point at which the translated surface is locally at one side of $\Sigma_2$. The maximum principle for minimal surfaces tells us that $\Sigma_2$ coincides with the vertical translated copy of $\Sigma_1$ at distance $d$. Finally, we will suppose that $\Sigma_1$ is not horizontal at some point $p$ and reach a contradiction. In that case, note that the previous argument guarantees the existence of a vertical segment of distance $d$ joining $p$ with a point in $\Sigma_2$, but, since this segment is not orthogonal to any of the two surfaces, we could shorten it and provide a curve from $\Sigma_1$ to $\Sigma_2$ whose length is strictly less than $d$, which is a contradiction. \end{proof} As a consequence of this result, any compact embedded minimal surface in $\Sigma\subset\mathbb{S}^2\times\mathbb{S}^1(r)$ is either connected or a finite union of horizontal slices. Moreover, the lift of a connected compact embedded minimal surface of $\mathbb{S}^2\times \mathbb{S}^1(r)$ different from an horizontal slice to $\mathbb{S}^2\times\mathbb{R}$ is connected and periodic. Let us now study the orientability of a compact minimal surface in $\mathbb{S}^2\times\mathbb{S}^1(r)$. \begin{lemma}\label{lema:separation-S2xR} Let $\Sigma\subset\mathbb{S}^2\times\mathbb{R}$ be a connected properly embedded minimal surface. Then $(\mathbb{S}^2\times\mathbb{R})\smallsetminus\Sigma$ has two connected components and $\Sigma$ is orientable. \end{lemma} \begin{proof} If $\Sigma$ is a horizontal slice, the result is trivial. Otherwise, given $t\in\mathbb{R}$, let us consider $S_t=\mathbb{S}^2\times\{t\}$. Then the set $\Sigma\cap S_t$ is the (non-empty) intersection of two minimal surfaces, hence it consists of a equiangular system of curves in the $2$-sphere $S_t$. We will denote by $C_t$ the set of intersection points of these curves (possibly empty), and it is well-known that an even number of such curves meet at each point of $C_t$. The fact that $\Sigma$ is properly embedded guarantees that $\cup_{t\in\mathbb{R}}C_t$ consists of isolated points. We can decompose $S_t\smallsetminus\Sigma=A^1_t\cup A^2_t$ in such a way that $A^1_t,A^2_t\subset S_t$ are open and, given $i\in\{1,2\}$, the intersection of the closures of two connected components of $A^i_t$, for $i = 1, 2$, is contained in $C_t$. In other words, we are painting the components of $S_t\smallsetminus\Sigma$ in two colors so that adjacent components have different color. Observe that the sets $\Sigma\cap S_t$ depend continuously on $t\in\mathbb{R}$ so it is clear that $A^1_t$ and $A^2_t$ can be chosen in such a way that $W_i=\cup_{t\in\mathbb{R}}A^i_t$ is open for $i\in\{1,2\}$. As $W_1\cap W_2=\emptyset$ and $W_1\cup W_2=(\mathbb{S}^2\times\mathbb{R})\smallsetminus\Sigma$, we get that $W_1$ and $W_2$ are the connected components of $(\mathbb{S}^2\times\mathbb{R})\smallsetminus\Sigma$. In particular, $\Sigma$ is orientable. \end{proof} If $\Sigma$ is a connected surface embedded in a orientable $3$-manifold $M$ and $M\smallsetminus\Sigma$ has two connected components, then $\Sigma$ is well-known to be orientable (the surface $\Sigma$ is said to \emph{separate} $M$). The converse is false in $\mathbb{S}^2\times\mathbb{S}^1(r)$ as horizontal slices show, but they turn out to be the only minimal counterexamples. \begin{proposition}\label{lema:separation-S2xS1} Let $\Sigma\subset\mathbb{S}^2\times\mathbb{S}^1(r)$ be a compact embedded minimal surface, different from a finite union of horizontal slices. Then $\Sigma$ separates $\mathbb{S}^2\times\mathbb{S}^1(r)$ if and only if $\Sigma$ is orientable. \end{proposition} \begin{proof} We will suppose that $\Sigma$ is orientable and prove that it separates $\mathbb{S}^2\times\mathbb{S}^1(r)$. In order to achieve this, we consider the projection $\pi\colon \mathbb{S}^2\times\mathbb{R}\to\mathbb{S}^2\times\mathbb{S}^1(r)$ and the lifted surface $\widetilde\Sigma\subset\mathbb{S}^2\times\mathbb{R}$ such that $\pi(\widetilde\Sigma)=\Sigma$. Then $\widetilde\Sigma$ is properly embedded, and connected as a consequence of Proposition~\ref{prop:intersection}. Lemma~\ref{lema:separation-S2xR} states that we can decompose it in connected components $(\mathbb{S}^2\times\mathbb{R})\smallsetminus\widetilde\Sigma=W_1\cup W_2$. Let $\phi_r\colon \mathbb{S}^2\times\mathbb{R}\to\mathbb{S}^2\times\mathbb{R}$ be the vertical translation $\phi_r(p,t)=(p,t+2\pi r)$. As $\phi_r(\widetilde\Sigma)=\widetilde\Sigma$ there are two possible cases, $\phi_r$ either preserves $W_1$ and $W_2$ or swaps them: \begin{enumerate}[(1)] \item If $\phi_r(W_1)=W_1$ and $\phi_r(W_2)=W_2$, then $\pi(W_1)\cap\pi(W_2)=\emptyset$. In this case $\Sigma$ separates $\mathbb{S}^2\times\mathbb{S}^1(r)$. \item If $\phi_r(W_1)=W_2$ and $\phi_r(W_2)=W_1$, then $\pi(W_1)=\pi(W_2)=(\mathbb{S}^2\times\mathbb{S}^1(r))\smallsetminus\Sigma$. In particular, $(\mathbb{S}^2\times\mathbb{S}^1(r))\smallsetminus\Sigma$ is connected and $\Sigma$ does not separate $\mathbb{S}^2\times\mathbb{S}^1(r)$. Since $\phi_r$ swaps $W_1$ and $W_2$ the unit normal vector field $\tilde{N}$ of $\tilde{\Sigma}$ does not induce a unit normal vector field on $\Sigma$. This contradicts the fact that $\Sigma$ is orientable because $\tilde{N}$ must be the lift of the unit normal vector field $N$ of $\Sigma$.\qedhere \end{enumerate} \end{proof} We now state our main result, which is inspired by~\cite[Proposition 2]{Ros}. \begin{theorem} Let $\Sigma$ be a compact embedded non-orientable minimal surface in $\mathbb{S}^2\times\mathbb{S}^1(r)$. Then $\Sigma$ has an even (non-orientable) genus. \end{theorem} \begin{proof} We can lift $\Sigma$ in a natural way to a compact minimal surface $\Sigma_2\subset\mathbb{S}^2\times\mathbb{S}^1(2r)$ so $\Sigma_2$ is a $2$-fold cover of $\Sigma$. We also lift $\Sigma$ to a minimal surface $\widetilde\Sigma\subset\mathbb{S}^2\times\mathbb{R}$ and decompose $(\mathbb{S}^2\times\mathbb{R})\smallsetminus\widetilde\Sigma=W_1\cup W_2$ as in the proof of Proposition~\ref{lema:separation-S2xS1}. The map $\phi_r\colon \mathbb{S}^2\times\mathbb{R}\to\mathbb{S}^2\times\mathbb{R}$ given by $\phi_r(p,t)=(p,t+2\pi r)$ swaps $W_1$ and $W_2$ so $\phi_{2r}\defeq\phi_r\circ\phi_r$ preserves $W_1$ and $W_2$, which means that $\Sigma_2$ is orientable. We will prove that $\Sigma_2$ has odd orientable genus, from where it follows that $\Sigma$ has even non-orientable genus. Let us consider the isometric involution $F\colon \mathbb{S}^2\times\mathbb{S}^1(2r)\to\mathbb{S}^2\times\mathbb{S}^1(2r)$ given by $F(p,t)=(p,t+2\pi r)$, and the projection $\pi_2\colon \mathbb{S}^2\times\mathbb{R}\to\mathbb{S}^2\times\mathbb{S}^1(2r)$ such that $\pi_2(\widetilde\Sigma)=\Sigma_2$. It satisfies $F(\Sigma_2)=\Sigma_2$ and swaps $\pi_2(W_1)$ and $\pi_2(W_2)$. Given a horizontal slice $S\subset\mathbb{S}^2\times\mathbb{S}^1(2r)$ intersecting $\Sigma_2$ transversally, the surface $\Sigma_2\smallsetminus(S\cup F(S))$ can be split in two isometric surfaces $\Sigma_2'$ and $F(\Sigma_2')$. Decomposing $\Sigma_2'$ in connected components $P_1,\ldots,P_k$, we can compute $\chi(P_i)=2-2g_i-r_i$, where $g_i$ is the genus and $r_i$ the number of boundary components of $P_i$. Thus \begin{equation}\label{eqn:genus} \chi(\Sigma_2)=2\chi(\Sigma_2')=2\sum_{i=1}^k\chi(P_i)=4\left(k-\sum_{i=1}^kg_i\right)-2\sum_{i=1}^kr_i. \end{equation} The last term in equation~\eqref{eqn:genus} is a multiple of $4$ since the total number of boundary components $\sum_{i=1}^kr_i$ is even (note that each of such components appears twice, once in $S$ and once in $F(S)$), say $4m=\chi(\Sigma_2)=2(1-g)$. It follows that $g$, the genus of $\Sigma_2$, is odd and we are done. \end{proof} \section{Construction of examples}\label{sec:examples} Given an isometric minimal immersion $\tilde{\phi}: \Sigma \rightarrow \mathbb{S}^2\times \mathbb{R}$ of a simply connected surface $\Sigma$, Hauswirth, Sa Earp and Toubiana~\cite{HST} constructed an associated isometric minimal immersion $\phi: \Sigma \to \mathbb{S}^2\times \mathbb{R}$ called the \emph{conjugate immersion}. The following properties are well-known and will be used repeatedly in the sequel (see~\cite{Daniel07} and~\cite[Lemma~1]{MT}): \begin{enumerate}[(i)] \item \label{lm:properties-conjugate:item:angle-function} $\phi$ and $\tilde{\phi}$ have the same angle function, i.e., $\nu = \prodesc{N}{\xi} = \prodesc{\tilde{N}}{\xi}$, where $N$ and $\tilde{N}$ are the unit normal vector fields to $\phi$ and $\tilde{\phi}$ respectively. \item \label{lm:properties-conjugate:item:tangent-projection} The tangential projections of $\xi$ are rotated: $\,\mathrm{d} \phi^{-1}(T)=J\,\mathrm{d}\tilde{\phi}^{-1}(\tilde{T})$, where $T=\xi-\nu N$, $\tilde{T}=\xi-\nu \tilde{N}$, and $J$ is the $\tfrac{\pi}{2}$-rotation in $T\Sigma$. \item \label{lm:properties-conjugate:item:shape-operator} The shape operators $S$ and $\tilde{S}$ of $\phi$ and $\tilde{\phi}$ respectively are related by $S = J\tilde{S}$, where $J$ is the $\tfrac{\pi}{2}$-rotation in $T\Sigma$. \item \label{lm:properties-conjugate:item:symmetry-curves} Any geodesic curvature line in the initial surface becomes a planar line of symmetry in the conjugate one. More precisely, given a curve $\alpha$ in $\Sigma$, if $\tilde{\phi}\circ\alpha$ is a horizontal (resp.\ vertical) geodesic, then $\phi\circ\alpha$ is contained in a vertical plane (resp.\ slice), which the immersions meets orthogonally. \end{enumerate} \begin{lemma}\label{lm:properties-conjugate} Let $\alpha$ be a curve in $\Sigma$ such that $\tilde{\gamma} = \tilde{\phi}\circ\alpha$ is a vertical unit-speed geodesic in $\mathbb{S}^2\times \mathbb{R}$, and express $\tilde{N}_{\tilde{\gamma}(s)} = \cos (\theta(s)) P_{\tilde{\gamma}(s)} + \sin (\theta(s)) Q_{\tilde{\gamma}(s)}$, where $\{P,Q,\tilde{\gamma}'\}$ is an orthonormal frame of parallel vector fields along $\tilde{\gamma}$. Then the conjugate curve $\gamma = \phi\circ\alpha$ lies in a slice $\mathbb{S}^2\times \{t_0\}$ and its geodesic curvature $\kappa$ as a curve of $\mathbb{S}^2\times \{t_0\}$ satisfies $\kappa(s) = \theta'(s)$. \end{lemma} \begin{proof} The Levi-Civita connection of $\mathbb{S}^2\times \mathbb{R}$ and the slice $\mathbb{S}^2\times \{t_0\}$ where $\gamma$ lies will be denoted by $\overline{\nabla}$ and $\nabla^{\mathbb{S}^2\times \{t_0\}}$, respectively. Since $\tilde{\gamma}$ is a vertical geodesic, the normal vector field $\tilde{N}$ along $\tilde{\gamma}$ is always horizontal. Hence, we can write $\tilde{N}_{\tilde{\gamma}(s)} = \cos (\theta(s)) P_{\tilde{\gamma}(s)} + \sin (\theta(s)) Q_{\tilde{\gamma}(s)}$ for some smooth function $\theta$. Hence \[ \overline{\nabla}_{\tilde{\gamma}'} \tilde{N} = \theta' \bigl( -\sin(\theta) P + \cos(\theta) Q\bigr) = \theta' \tilde\phi_*(J\alpha'), \] because $P$ and $Q$ are parallel vector fields and $-\sin (\theta) P + \cos(\theta) Q$ is tangent to $\tilde{\phi}(\Sigma)$ (orthogonal to $N$) and orthogonal to $\tilde{\gamma}'$ (both $P$ and $Q$ are horizontal while $\tilde{\gamma}'$ is vertical by assumption) so it is colinear with $\tilde\phi_*(J\alpha')$ and, up to a change of orientation in $\Sigma$ if necessary, we can suppose that they are equal. Finally \[ \begin{split} \theta' &= \prodesc{\overline{\nabla}_{\tilde{\gamma}'} \tilde{N}}{\tilde\phi_*(J\alpha')} = -\prodesc{\tilde{S}\alpha'}{J\alpha'} = \prodesc{J\tilde{S}\alpha'}{\alpha'} = \prodesc{S \alpha'}{\alpha'} =\\ &= \prodesc{\overline{\nabla}_{\gamma'} \gamma'}{N} = \prodesc{\nabla^{\mathbb{S}^2\times\{t_0\}}_{\gamma'}\gamma'}{N} = \kappa, \end{split} \] where we have used that $J$ is skew-adjoint and the properties of the conjugation stated above. \end{proof} In this section we are going to apply the \emph{conjugate Plateau technique} in order to construct compact embedded minimal surfaces in $\mathbb{S}^2 \times \mathbb{S}^1(r)$. This technique consists in solving the Plateau problem over a geodesic polygon producing a minimal surface in $\mathbb{S}^2\times \mathbb{R}$, and considering its \emph{conjugate surface}. We must ensure that, after successive reflections over its boundary, we obtain a periodic (in the $\mathbb{R}$ factor) minimal surface, giving rise to a compact one in the quotient $\mathbb{S}^2 \times \mathbb{S}^1(r)$. \subsection{Initial minimal piece and conjugate surface} \label{subsec:contours} Consider a geodesic triangle with angles $\tilde{\alpha}$, $\tilde{\beta}$ and $\gamma$ in $\mathbb{S}^2\times\{0\}$, lift the hinge given by the angle $\gamma$ in vertical direction and add two vertical geodesics (each of length $\tilde{h} > 0$) to produce a closed curve $\tilde{\Gamma}$ in $\mathbb{S}^2 \times \mathbb{R}$ (see Figure~\ref{f:boundary1}). For $\tilde{h}\geq0$, $0<\gamma<\pi$ and $\tilde{\alpha},\tilde{\beta}\leq\frac{\pi}{2}$ (so $\tilde{a}, \tilde{b}\leq \tfrac{\pi}{2}$) the polygonal Jordan curve $\tilde{\Gamma}$ bounds a minimal graph $\tilde{M}$ over $\mathbb{S}^2 \times \{0\}$ (the existence follows from Radó's theorem, see~\cite{Rosenberg} for a particular example). Hence the angle function of $\tilde{M}$ does not vanish and we can suppose, up to a change of the normal to $\tilde{M}$, that $\tilde{\nu} > 0$. Its conjugate surface $M$ is a minimal surface also in $\mathbb{S}^2\times\mathbb{R}$, bounded by a closed Jordan curve $\Gamma$ consisting of three symmetry curves in vertical planes and two horizontal symmetry curves. Two successive curves in vertical planes enclose an angle $\gamma$ (this angle is intrinsic to the surface), the remaining angles in the vertexes of $\Gamma$ are equal to $\frac{\pi}{2}$. We will denote by $\tilde{\imath}$ (resp.\ $i$), $i \in \{1, 2, 3, 4, 5\}$, each of the vertexes of $\tilde{\Gamma}$ (resp.\ $\Gamma$). Moreover, we will denote $\widetilde{\imath\jmath}$ (resp.\ $\overline{\imath\jmath}$) the segment of $\tilde{\Gamma}$ (resp.\ $\Gamma$) that joints the points $\tilde{\imath}$ and $\tilde{\jmath}$ (resp.\ $i$ and $j$). Furthermore, $\Pi_{ij}$ will stand for the vertical plane or horizontal slice where $\overline{\imath\jmath}$ lies in the conjugate piece. Up to a translation in $\mathbb{S}^2\times \mathbb{R}$, we can suppose that the slice $\Pi_{34}$ is $\mathbb{S}^2\times \{0\}$. \begin{lemma}\label{lm:basic-properties-conjugate-polygon} In the previous setting, let $\tilde{h} \leq \tfrac{\pi}{2}$. Then: \begin{enumerate}[(i)] \item The segments of $\Gamma$ are embedded curves. Moreover, each of the segment of $\Gamma$ projects injectively to $\mathbb{S}^2\times\{0\} = \Pi_{34}$. \item The vertical planes $\Pi_{23}$, $\Pi_{45}$ and $\Pi_{51}$ and the slices $\Pi_{12}$ and $\Pi_{34}$ determine a prism $\Omega$ defined by the angles $\alpha$ (between $\Pi_{51}$ and $\Pi_{23}$), $\beta$ (between $\Pi_{23}$ and $\Pi_{45}$), $\gamma$, and the height $h$ (see Figure~\ref{f:boundary2}). Moreover, $\alpha > \tilde{\alpha}$ and $\beta > \tilde{\beta}$. \item The polygon $\Gamma$ is contained in $\Omega$ and so $\Gamma$ projects injectively to $\mathbb{S}^2\times \{0\}$. \item The minimal surface $M$ is embedded and lies in the interior of $\Omega$. \end{enumerate} \end{lemma} Thanks to Lemma~\ref{lm:basic-properties-conjugate-polygon} we are able to draw more precisely what the polygon $\Gamma$ looks like (see Figure~\ref{f:boundary2}) provided $\tilde{h}$ is short enough, i.e., $\tilde{h}\leq \tfrac{\pi}{2}$. \begin{figure}[htbp] \begin{minipage}[b]{.5\textwidth} \begin{center} \includegraphics{Gamma-tilde} \end{center} \caption{\small{The boundary curve $\tilde{\Gamma}$. If $\tilde{a} = \tilde{b}$, then $\tilde{\Gamma}$ and $\tilde{M}$ are symmetric with respect to a vertical plane containing $\tilde{\delta}$.}} \label{f:boundary1} \end{minipage} \begin{minipage}[b]{.49\textwidth} \begin{center} \includegraphics{Gamma-Va-Vb} \end{center} \caption{\small{The boundary curve $\Gamma$ of the conjugate minimal surface $M$ inside the prism $\Omega$ with data $(\alpha, \beta, \gamma, h)$.}}\label{f:boundary2} \end{minipage} \end{figure} \begin{proof} (i) The angle function $\tilde{\nu}$ does not vanish in the interior of the horizontal segments of $\tilde{\Gamma}$ because this would contradict the boundary maximum principle for minimal surfaces ($\tilde\Sigma$ would be tangent to a vertical plane containing this horizontal segment). Hence the angle function $\nu$ of the conjugate piece (which is the same as $\tilde{\nu}$) does not vanish in the interior points of the curves contained in vertical planes. Let us focus on the segment $\overline{23}$ parametrized by a curve $\sigma:[a, b] \rightarrow \mathbb{S}^2\times \mathbb{R}$ with unit speed. The length of the projection $\pi(\overline{23})$ between the point $2 = \sigma(a)$ and $\sigma(t)$ is given by $\ell(t) = \int_a^t \nu(\sigma(s)) \,\mathrm{d} s$. Hence $\ell(t)$ is strictly increasing so the curve $\overline{23}$ projects injectively to $\mathbb{S}^2\times \{0\}$. In particular, it is embedded. The same argument works for $\overline{45}$ and $\overline{51}$. The curves $\overline{12}$ and $\overline{34}$ are also embedded. To prove it, we realize that Lemma~\ref{lm:properties-conjugate} implies that the geodesic curvature of both curves does not change sign because the rotation of the normal vector field $\tilde{N}$ along $\widetilde{12}$ and $\widetilde{34}$ is monotonic (otherwise a contradiction with the boundary maximum principle would be found). Let us suppose that the curve $\overline{12}$ is not embedded. Then the curve contains a loop enclosing a domain $D$. By the Gau\ss-Bonnet theorem, \[ \mathrm{area}(D)\geq \pi - \int_{\text{loop}} \kappa \geq \pi - \abs{\int_{\text{loop}} \kappa }\geq \pi - \tilde{\alpha} \geq \frac{\pi}{2}, \quad \text{since } \tilde{\alpha} \in\ ]0,\tfrac{\pi}{2}], \] where $\kappa$ is the geodesic curvature of $\overline{12}$ in $\Pi_{12}$ and we have taken into account that total curvature of $\overline{12}$ is exactly $\tilde{\alpha}$ (see Lemma~\ref{lm:properties-conjugate}). Hence, by the isoperimetric inequality, $\tilde{h} = \ell(\overline{12}) \geq \ell(\mathrm{loop}) \geq (\tfrac{7}{4}\pi^2)^{1/2} > \tfrac{\pi}{2}$, contradicting the assumption $\tilde{h}<\frac{\pi}{2}$. A similar reasoning implies that the curve $\overline{34}$ is embedded for $\tilde{h} < \tfrac{\pi}{2}$. (ii) First, the slices $\Pi_{12}$ and $\Pi_{34}$ do not coincide. Otherwise, we get a contradiction to the maximum principle by considering a slice $\mathbb{S}^2\times\{t\}$ for $|t|$ large enough so $(\mathbb{S}^2\times \{t\})\cap M = \emptyset$. Move such slice downwards or upwards until a first contact point appears. That point cannot be at the interior so it has to be either in $\overline{45}$, $\overline{51}$ or $\overline{23}$, and it would yield a contradiction to the boundary maximum principle). Let $h$ be the distance between $\Pi_{12}$ and $\Pi_{34}$ so $\Pi_{12} = \mathbb{S}^2\times \{h\}$. Note also that $\Pi_{45}$ and $\Pi_{51}$ are different since they form an angle of $\gamma \not\in\{0,\pi\}$. Finally, we will prove that $\Pi_{23}$ is different from $\Pi_{45}$ (resp.\ $\Pi_{51}$). Were it not the case, the endpoints of the horizontal curve $\sigma = \overline{34}$ (resp.\ $\sigma = \overline{12}$) would lie in a geodesic of $\mathbb{S}^2\times\{0\}$ (resp.\ $\mathbb{S}^2\times\{h\}$) meeting it orthogonally. Hence, the curve $\sigma$ and the segment of geodesic that joints the points $3$ and $4$ (resp.\ $1$ and $2$) would define a domain $D$ in $\mathbb{S}^2\times\{0\}$ (resp.\ $\mathbb{S}^2\times \{h\}$). Gau\ss-Bonnet formula in $D$ yields \begin{equation}\label{eqn:geod} \int_{\sigma} \kappa = \pi - \mathrm{area}(D), \end{equation} where $\kappa$ is the geodesic curvature of $\sigma$ in the slice. The total rotation $\Theta$ of the normal vector field to $\tilde{M}$ along $\tilde{\sigma} = \widetilde{34}$ (resp. $\tilde{\sigma} = \widetilde{12}$) is such that $|\Theta|= \tilde{\beta}$ (resp.\ $|\Theta| = \tilde{\alpha}$), so $|\Theta|\leq\frac{\pi}{2}$. By Lemma~\ref{lm:properties-conjugate} and equation~\eqref{eqn:geod}, it follows that $\tfrac{\pi}{2}\leq\mathrm{area}(D) <\frac{3\pi}{2}$. Again the isoperimetric inequality in the sphere would imply that $4\tilde{h}^2 \geq \mathrm{area}(D)(4\pi - \mathrm{area}(D)) > \tfrac{7}{4}\pi^2$, contradicting that $\tilde h<\frac{\pi}{2}$. The last assertion of (ii) will be showed at the end of the proof. (iii) First, the height of the point $5$ is strictly between $0$ and $h$. In other case, for instance if the height of the point $5$ is bigger than or equal to $h$, we consider a slice $S = \mathbb{S}^2 \times \{t_0\}$ with $t_0$ sufficiently large so $M \cap S = \emptyset$. Then, we move $S$ downwards until it first touches $M$ at a point $p$. Then $p$ should be either an interior point of $M$, or at the interior of $\overline{45}$, $\overline{51}$ or $\overline{23}$, or $p = 5$. Anyway we will find a contradiction to the maximum principle at the interior or at the boundary. Likewise the curves $\overline{45}$, $\overline{51}$ and $\overline{23}$ are also contained in $\mathbb{S}^2\times[0,h]$. Since the vertical planes $\Pi_{23}$, $\Pi_{45}$ and $\Pi_{51}$ are pairwise different (see (ii)), they divide the slice $\mathbb{S}^2\times \{0\}$ in eight triangles. Hence the vertical planes $\Pi_{23}$, $\Pi_{45}$ and $\Pi_{51}$ and the slices $\Pi_{12}$ and $\Pi_{34}$ define eight different prism. Since we have shown in (i) that the curves contained in vertical planes project one-to-one, (iii) will follow from the fact that $\overline{12}$ and $\overline{34}$ lie in the same prism, that we will call $\Omega$. \noindent\textbf{Claim 1}: \emph{The points $1$ and $2$ (resp.\ $3$ and $4$) lie in the same triangle.} Otherwise the geodesic curvature of $\overline{12}$ (resp.\ $\overline{34}$) in $\mathbb{S}^2\times \{h\}$ (resp.\ $\mathbb{S}^2\times \{0\}$) will change sign but that is impossible (see the proof of (i)). \noindent\textbf{Claim 2}: \emph{The curve $\overline{12}$ (resp.\ $\overline{34}$) is completely contained in a triangle.} This claim follows from the fact that the geodesic curvature of $\overline{12}$ (resp.\ $\overline{34}$) does not change sign, as well as the fact the angle function of $\Sigma$ does not vanish. This means that the projection of $\Sigma$ to the $\mathbb{S}^2$-factor is an immersed domain $G\subset\mathbb{S}^2$, so we can define a unit normal to $\Sigma$ along $\overline{12}$ (resp.\ $\overline{34}$) that project to a normal to $G$ pointing towards the interior of $G$. If $\overline{12}$ or $\overline{34}$ are not contained in a triangle, it is easy to see that such an immersed domain $G$ cannot exist, which is a contradiction. \noindent\textbf{Claim 3}: \emph{The curves $\overline{12}$ and $\overline{34}$ lie in the same prism.} Let us assume that $\overline{12}$ and $\overline{34}$ do not lie in the same prism and suppose, without loss of generality, that $\overline{12}$ does not lie in the same prism as the point $5$. In that case the surface $M$ will contain two different points in the same vertical geodesic $\Pi_{51}\cap \Pi_{23}$. Hence there is one point $p$ in the interior of $M$ such that $\nu(p) = 0$ which is impossible (recall that $\nu = \tilde{\nu} > 0$, see the beginning of Section~\ref{subsec:contours}, and that $M$ is orthogonal to $\Pi_{51}$ and $\Pi_{23}$). Once we know that $\Gamma$ is contained in the boundary of a well-defined prism $\Omega$ (see (ii)), (i) ensures that the whole polygon $\Gamma$ projects injectively to $\mathbb{S}^2\times\{0\}$. (iv) This assertion follows from (iii) by a classical application of the maximum principle. It remains to prove that $\alpha > \tilde{\alpha}$ and $\beta > \tilde{\beta}$. First, let us apply the Gau\ss{}-Bonnet-theorem to the domain $V_\alpha$ bounded by $\overline{12}$ and the edges of $\Omega$ as in Figure~\ref{f:boundary2} (it is well defined thanks to (i) and (iii)). We get $\alpha=\operatorname{area}(V_\alpha)+\tilde{\alpha}>\tilde{\alpha}$. Likewise, $\beta>\tilde\beta$ by using the corresponding domain $V_\beta$. \end{proof} Our aim is to define $\tilde{\Gamma}$ such that the conjugate surface has the desired properties, i.e., it can be smoothly extended without branch points by Schwarz reflection about its boundary producing a compact embedded surface in $\mathbb{S}^2\times\mathbb{R}$ after a finite number of reflections. Depending on the surface we want to construct we have to ensure different data $(\alpha,\beta,\gamma, h)$ for the prism $\Omega$ (see Lemma~\ref{lm:basic-properties-conjugate-polygon}). We want to remark that if $m$ copies of $M$ are needed to produce a compact orientable minimal surface $\Sigma$ by Schwarz reflection, then the genus $g$ of $\Sigma$ is $g = 1 + \frac{m}{4\pi}(\pi - \gamma)$. This can be worked out by the Gau\ss-Bonnet theorem: The total curvature of $M$ is $\gamma - \pi$ and in order to close the surface we need by assumption $m$ copies of $M$. Hence the total curvature of $\Sigma$ is $m(\gamma - \pi)$ and so the genus is $g = 1 + \frac{m}{4\pi}(\pi - \gamma)$. \begin{remark} One could consider the same boundary curve $\tilde{\Gamma}$ consisting of horizontal and vertical geodesics in $\mathbb{R}^3$. This leads to the well-known Schwarz P-surface, whose name is motivated by the fact that it is invariant under a primitive cubic lattice $\Lambda$. Moreover, there exists another lattice under which the surface is preserved but its orientation is not; $\Lambda$ is a subset of it. The quotient $P$ of the surface under the lattice $\Lambda$ has genus $3$ and consists of 16 copies of $M$ with $\alpha=\beta=\frac{\pi}{4}$ and $\gamma=\frac{\pi}{2}$. If we set $a=b=h$ (cp.\ Figure \ref{f:boundary2}), the quotient $P$ is contained in a cube of edge length $2a$, and these cubes tessellate $\mathbb{R}^3$. \end{remark} \subsection{Odd genus Schwarz P-type examples.}\label{subsec:odd-genus} This section is devoted to investigate the most symmetric case in the construction above, which follows from choosing $\tilde a=\tilde b$ in the initial contour. Since the solution of the Plateau problem with boundary $\tilde\Gamma$ is unique and in this particular case $\tilde\Gamma$ is symmetric about a vertical plane, we get that the initial minimal surface $\tilde M$ is also symmetric about the same vertical plane, intersecting $\tilde M$ orthogonally along a curve $\tilde\delta$ (see Figure~\ref{f:boundary1}). From~\cite[\S2]{MT}, this implies that the conjugate surface $M$ is symmetric about a horizontal geodesic $\delta$ lying in the interior of $M$ (see Figure \ref{f:boundary2}). In particular, $a=b$ and $\alpha=\beta$. \begin{figure}[h] \begin{minipage}[b]{.5\textwidth} \centering \includegraphics{cube-tessellation-tags} \caption{\small{Regular tiling of $\mathbb{S}^2$ by quadrilaterals}}\label{f:cube-tessellation} \end{minipage} \begin{minipage}[b]{.49\textwidth} \centering \includegraphics{beach-balloon-tessellation-tags} \caption{\small{Tessellation of the sphere by isosceles triangles}}\label{f:balloon-tessellation} \end{minipage} \end{figure} We are going to produce new compact orientable embedded minimal examples in $\mathbb{S}^2 \times \mathbb{S}^1(r)$, for sufficiently small radius $r > 0$, coming from two different tessellations of the sphere by isosceles triangles (see the shaded triangles in Figures~\ref{f:cube-tessellation} and \ref{f:balloon-tessellation}). The first one comes from the regular quadrilateral tiling of the sphere once we decompose each square in four isosceles triangles with angles $\alpha = \beta = \frac{\pi}{3}$ and $\gamma = \frac{\pi}{2}$. The second one, as depicted in Figure~\ref{f:balloon-tessellation}, follows from decomposing $\mathbb{S}^2$ in $4k$ isosceles triangles with angles $\alpha = \beta = \frac{\pi}{2}$ and $\gamma = \frac{\pi}{k}$, for any $k\geq 1$. Let us assume that we can produce, via conjugate Plateau technique, an embedded minimal surface $M_j$ that fits inside a prism $\Omega_j$ with data $(\frac{\pi}{3}, \frac{\pi}{3},\frac{\pi}{2}, h)$ if $j = 1$ and $(\frac{\pi}{2}, \frac{\pi}{2}, \frac{\pi}{k}, h)$ if $j = 2$ for arbitrary small $h$ (see Figure~\ref{f:boundary2}). Then: \begin{enumerate}[$\bullet$] \item It is clear that 24 congruent copies of $\Omega_1$ tessellate $\mathbb{S}^2\times [0,h]$. Since the tessellation comes from reflections about vertical planes, it implies a smooth continuation of $M_1$, which we call $M_1'$. The surface $M_1'$ is connected, topologically it is $\mathbb{S}^2\setminus\{p_1,\dots,p_8\}$, where each $p_i$ corresponds to one vertex of the quadrilateral tiling. After reflecting $M_1'$ about one of the horizontal planes $\mathbb{S}^2\times \{0\}$ or $\mathbb{S}^2\times \{h\}$ we get a surface consisting of 48 copies of $M_1$. The vertical translation $T$ about $2h$ yields a simply periodic embedded minimal surface $S$. It follows that the quotient $S/T$ is a compact minimal surface with genus $7$ in $\mathbb{S}^2 \times \mathbb{S}^1(\frac{h}{\pi})$. \item We can reflect the surface $M_2$ about the vertical planes containing the edges $a$ and $b$. Repeating this reflection successively the surface closes up in such a way that $2k$ copies of $M_2$ build a smooth surface in the product of an hemisphere of $\mathbb{S}^2$ and an interval $I$ of length $h$. The surface meets the vertical plane above the bounding great circle orthogonally and a reflection about this plane yields a minimal surface $M_2'$ which is topologically $\mathbb{S}^2\setminus\{p_1,\dots,p_{2k}\}$, where each $p_j$ corresponds to one vertex of the tiling. As before reflecting $M_2'$ about one of the horizontal planes $\mathbb{S}^2\times \partial I$ we get a surface $S$ which is invariant under the vertical translation $T$ of length $2h$. Moreover it is invariant under a rotation about the fiber over the north or south pole by an angle of $\frac{\pi}{k}$. The quotient $S/T$ is compact and consists of $8k$ copies of $M$, so its genus is given by $g=2k-1$, i.e., an arbitrary odd positive integer. \end{enumerate} Now, we are going to show the existence of the embedded minimal surfaces $M_1$ (resp.\ $M_2$) associated to $\Omega_1$ (resp.\ $\Omega_2$). Let $\gamma = \frac{\pi}{2}$ (resp.\ $\gamma = \frac{\pi}{k}$). Recall that we have fixed $\tilde{a} = \tilde{b}$. As mentioned at the beginning of the section, the conjugate piece contains a horizontal geodesic $\delta$ (so $a = b$ and $\alpha = \beta$, see Figures~\ref{f:boundary1} and~\ref{f:boundary2}). Therefore, $\delta$ and its projection onto the base of $\Omega$ have the same length $\ell(\pi(\delta)) = \ell(\delta) = \ell(\tilde{\delta})$, and we can decompose the base of $\Omega$ in two right triangles as in Figure~\ref{fig:base-triangle-Omega}. By spherical trigonometry, we get that \begin{equation}\label{eq:relation-alpha-delta} \cos(\alpha) = \sin(\tfrac{\gamma}{2}) \cos(\ell(\delta)). \end{equation} \begin{figure}[htbp] \centering \includegraphics{triangulo} \caption{Base triangle of the prism $\Omega$ in the special case $\tilde{a} = \tilde{b}$.} \label{fig:base-triangle-Omega} \end{figure} We claim that \emph{for each $\tilde{a} \in\ ]0, \tfrac{\pi}{2}]$ there exists $\tilde{h}(\tilde{a}) \leq \tfrac{\pi}{2}$ such that $\alpha = \tfrac{\pi}{3}$ (resp.\ $\alpha = \tfrac{\pi}{2}$)}. Let us check the claim for $M_1$, i.e., in the case $\gamma = \tfrac{\pi}{2}$ and $\alpha = \tfrac{\pi}{3}$. Fixing $\tilde{a}\in\ ]0, \tfrac{\pi}{2}]$, the length of $\tilde{\delta}$ varies continuously in $\tilde{h}$ from zero (when $\tilde{h} \rightarrow 0$) to $+\infty$ (when $\tilde{h} \rightarrow +\infty$). Moreover, $\tilde{h} < \ell(\tilde{\delta})$ by construction. Since $\gamma = \tfrac{\pi}{2}$ and $\alpha = \tfrac{\pi}{3}$ then, in view of equation~\eqref{eq:relation-alpha-delta}, $\ell(\tilde\delta) = \ell(\delta) = \arccos(\sqrt{2}\cos(\tfrac{\pi}{3}))< \tfrac{\pi}{2}$ (in the case $\gamma=\tfrac{\pi}{k}$ and $\alpha = \tfrac{\pi}{2}$, we have that $\ell(\delta) = \tfrac{\pi}{2}$). Hence the height in the initial polygon satisfies $\tilde{h} \leq \ell(\tilde{\delta}) \leq \tfrac{\pi}{2}$ and the assumption of Lemma~\ref{lm:basic-properties-conjugate-polygon} is fulfilled. Finally, varying $\tilde{h} \in\ ]0, \ell(\tilde{\delta})[$ we can find $\tilde{h}(\tilde{a})$ such that $\alpha = \tfrac{\pi}{3}$ and $\tilde{h}(\tilde{a}) \leq \tfrac{\pi}{2}$. Lemma~\ref{lm:basic-properties-conjugate-polygon} applies and the claim is proved. Since $h \leq \ell(\overline{23}) = \ell(\widetilde{23}) = 2 \arctan\bigl(\sin(\tilde{a})( {\cot^2(\tfrac{\gamma}{2}) + \cos^2(\tilde{a})})^{-1/2}\bigr)$, we can make $h$ arbitrarily small by choosing $\tilde{a}$ small enough. To sum up, we can always choose a polygon $\tilde{\Gamma}$ with data $(\tilde{\alpha}, \tilde{\alpha}, \frac{\pi}{2}, \tilde{h})$ (resp.\ $(\tilde{\alpha}, \tilde{\alpha}, \frac{\pi}{k}, \tilde{h})$) satisfying the hypothesis of Lemma~\ref{lm:basic-properties-conjugate-polygon}. Thus the conjugate of the Plateau solution of $\tilde{\Gamma}$ is embedded and its boundary is contained in the prism $\Omega_1$ (resp.\ $\Omega_2$), for sufficiently small height $h$. \subsection{Arbitrary genus Schwarz P-type examples.}\label{subsec:arbitrary-genus} To construct a compact orient\-able minimal surface $M_k$ with arbitrary genus, let us guarantee the existence of a minimal surface $M$ that matches a prism $\Omega$ with data $(\frac\pi k,\frac\pi 2,\frac\pi 2,h)$ for some $h>0$. This leads to less symmetric examples since $\alpha\ne\beta$ and we have to use a degree argument as in \cite{KPS88} to guarantee existence. If $M\subset\Omega$ exists, then Schwarz reflection about the vertical mirror planes continues the surface smoothly and $4k$ copies of $M$ build a minimal surface $M'$ in the product $\mathbb{S}^2\times ]0,h[$ (see Figure~\ref{f:arbitrary-genus}). Topologically $M'$ is $\mathbb{S}^2\setminus\{p_1,\dots,p_{k+2}\}$, where $p_j$ are the vertexes of the tiling. As in the symmetric cases above, reflecting $M'$ about the bounding horizontal mirror planes gives a complete surface $M''$, which is invariant under vertical translation about $2h$. The quotient is the desired surface $M_k$. Moreover, $M''$ is also invariant under rotation about those vertical fibers which are intersection of vertical mirror planes by angles $\frac{2\pi}{k}$ (resp.\ $\pi$). As in the previous cases one computes the genus and gets $1+k$, since $8k$ copies of $M$ build the compact surface $M_k$. \begin{figure}[htbp] \centering \includegraphics[height=4.5cm]{arbitrary-genus} \caption{\small{The polygon $\Gamma$ for the configuration $(\frac\pi k,\frac\pi 2,\frac\pi 2,h)$. The resulting compact surface $M_k$ has genus $k+1$. }} \label{f:arbitrary-genus} \end{figure} \begin{proposition}\label{prop:orientable-arbitrary-genus-examples} There exists a compact orientable minimal surface $M_k$ with genus $k+1$, embedded in $\mathbb{S}^2\times\mathbb{S}^1(r)$, for any $k\geq3$ and $r$ small enough. \end{proposition} \begin{proof} We only need to prove the existence of the minimal surface $M$ with data $(\frac\pi k,\frac\pi 2,\frac\pi 2,h)$, the existence of $M_k$ follows directly by Schwarz reflection. We start with a polygonal Jordan curve $\tilde{\Gamma}$ as in Figure \ref{f:boundary1} with $\gamma=\frac{\pi}{2}$ and some $\tilde{h}>0$. We have seen that the minimal surface $M$ is then uniquely defined by $(\tilde{a},\tilde{b})$, so this defines a map $f\colon \mathbb{R}_{+}\times\,]0,\frac{\pi}{2}]^2\to \mathbb{R}^3,\,(\tilde{h},\tilde{a},\tilde{b})\mapsto(h,\alpha,\beta)$. The map $f$ is continuous, since a sequence $\{\tilde{M}_n\}$ of solutions to the Plateau problem with data $(\tilde{h}_n,\tilde{a}_n,\tilde{b}_n)$, it has a converging subsequence with limit given by the solution to the Plateau problem with data $(\tilde{h}_\infty, \tilde{a}_\infty, \tilde{b}_\infty)=\lim_n(\tilde{h}_n,\tilde{a}_n,\tilde{b}_n)$. In order to justify this, observe that each $\tilde{M}_n$ can be extended by Schwarz reflection to a complete minimal surface $\hat{M}_n$. Since the ambient geometry is bounded and the surfaces have uniformly bounded second fundamental forms, general convergence arguments show that there exists a subsequence of $\{\hat{M}_n\}$ converging to a complete minimal surface $\hat{M}_\infty$ in the $\mathcal{C}^k$-topology on compact subsets for every $k\geq 0$. Since every surface $\hat{M}_n$ contains the polygon $\tilde{\Gamma}_n$ associated to $(\tilde{h}_n,\tilde{a}_n,\tilde{b}_n)$, the limit surface $\hat{M}_\infty$ contains the polygon $\tilde{\Gamma}_\infty$ associated to $(\tilde{h}_\infty, \tilde{a}_\infty, \tilde{b}_\infty)$. Moreover, $\hat{M}_\infty$ is also a graph in the interior of the polygon, so it coincides with the solution to the Plateau problem with respect to $\tilde{\Gamma}_\infty$ (note that such a solution is unique). In order to prove the existence of a minimal surface $M$, i.e., the existence of a triple $(\tilde{h}_0,\tilde{a}_0,\tilde{b}_0)$ such that $f(\tilde{h}_0,\tilde{a}_0,\tilde{b}_0)=(h_0,\frac\pi k,\frac\pi 2)$ for some $h_0>0$, we use a degree argument (recall that $\tilde{\alpha}, \tilde{\beta}\in\ ]0, \tfrac{\pi}{2}]$ so $\tilde{a}, \tilde{b} \in\ ]0, \tfrac{\pi}{2}]$). We consider the map \[ f_{\tilde{h}}\colon]0,\tfrac\pi2]\times\ ]0,\tfrac\pi2]\to \mathbb{R}^2,\,(\tilde{a},\tilde{b})\mapsto(\alpha,\beta) \] and show there exists a closed Jordan curve $c\colon \mathbb{R}\to ]0,\frac\pi2]^2$ such that the image $f_{\tilde{h}}\circ c$ is a closed curve around $(\frac\pi k,\frac \pi 2)$. We compose $c$ of four straight lines $c_i$, $i\in\{1,\ldots, 4\}$, namely, \begin{multicols}{2} \begin{itemize} \item $c_1(t)=\!(\frac\pi2,t)$ with $t\in[\frac{1}{2k},\frac{\pi}{k}]$, \item $c_2(t)=(t,\frac{\pi}{k})$ with $t\in[\frac12,\frac\pi2]$, \item $c_3(t)=(\frac12,t)$ with $t\in[\frac{1}{2k},\frac\pi k]$, \item $c_4(t)=(t,\frac1{2k})$ with $t\in[\frac12,\frac\pi2]$, \end{itemize} \end{multicols} \noindent and we claim there exists $\tilde{h}>0$ such that $f_{\tilde{h}}\circ c$ has the desired property. Since $\tilde{\gamma}=\tilde{a}=\frac\pi2$ along $c_1$, we have $\beta>\tilde{\beta}=\frac\pi2$ (by the cosine rule $\cos(\tilde{\beta}) = 0$ since $\gamma = \tilde{a} = \tfrac{\pi}{2}$) along $f_{\tilde{h}}\circ c_1$ (see Lemma~\ref{lm:basic-properties-conjugate-polygon}). Likewise, $\alpha>\tilde\alpha>\frac\pi k$ along $f_{\tilde{h}}\circ c_2$ (by cosine rule $\cos(\tilde{\alpha})= \sin(\tilde{\beta})\cos(\tilde{b}) < \cos(\tilde{b}) = \cos(\tfrac{\pi}{k})$). Now, $\tilde{\beta} < \tfrac{\pi}{2}$ along $c_3$ since, by the cosine rule, $\cos(\tilde{\beta}) = \cos(\tilde{a}) \sin(\tilde{\alpha}) = \cos(\tfrac{1}{2})\sin(\tilde{\alpha})> 0$. Moreover, we will also show that $\tilde{\alpha} < \tfrac{\pi}{k}$ along $c_4$. First, by the sine rule, \[ \sin(\tilde{\beta}) = \frac{\sin(\tilde{\alpha})}{\sin(\tilde{b})}\sin{(\tilde{a})} > \frac{\sin(\tilde{\alpha})}{\sin(\tfrac{1}{2k})} \sin(\tfrac{1}{2}), \quad \text{since } \tilde{a} \in\ ]\tfrac{1}{2},\tfrac{\pi}{2}[ \text{ along $c_4$}. \] By the cosine rule, $\cos(\tilde{\alpha}) = \sin(\tilde{\beta})\cos(\tilde{b}) = \sin(\tilde{\beta})\cos(\tfrac{1}{2k})$. Hence \[ \cos(\tilde{\alpha})> \sin(\tfrac{1}{2})\cot(\tfrac{1}{2k}) \sin(\tilde{\alpha}), \] so $\tilde{\alpha} < \arccot\bigl(\sin(\tfrac{1}{2})\cot(\tfrac{1}{2k})\bigr) < \tfrac{\pi}{k}$. Therefore, we have showed that $\tilde\beta<\frac\pi2$ (resp. $\tilde\alpha<\frac{\pi}{k}$) along $c_3$ (resp.\ $c_4$). Since $\alpha\to\tilde{\alpha}$ and $\beta\to\tilde{\beta}$ for $\tilde{h}\to 0$, we deduce that there exists $0 < \tilde{h}_0\leq \tfrac{\pi}{2}$ such that $\beta<\frac\pi 2$ along $f_{\tilde{h}}\circ c_3$ and $\alpha<\frac\pi k$ along $f_{\tilde{h}}\circ c_4$ for all $\tilde{h}\leq\tilde{h}_0$. Note that we can choose $\tilde{h}_0$ such that Lemma~\ref{lm:basic-properties-conjugate-polygon} applies and we are done. \end{proof} \begin{remark} Note that the Schwarz P-type example of genus $2k-1$, $k\geq 2$, (see Section~\ref{subsec:odd-genus}) induces a compact embedded minimal surface in $\R\mathrm{P}^2\times \mathbb{S}^1(r)$ if and only if $k = 2m$. In this case, we get a two-sided non-orientable embedded minimal surface of genus $4m$, $m \geq 1$. The orientable arbitrary genus embedded minimal surface $M_k$, $k \geq 3$, constructed in Proposition~\ref{prop:orientable-arbitrary-genus-examples} induces a compact embedded minimal surface in $\R\mathrm{P}^2\times \mathbb{S}^1(r)$ if and only if $k = 2m$. In this case, the induce surface is two-sided non-orientable embedded and minimal with genus $2(1+m)$, $m \geq 2$. \end{remark} \subsection{Final remarks} In this section we are going to point out the reason why it is difficult to obtain a better result in $\mathbb{S}^2 \times \mathbb{S}^1(r)$ using the Plateau construction technique for all $r$. First we state the following general properties: \begin{proposition}\label{prop:symmetries} \begin{enumerate}[(i)] \item Let $\Sigma$ be a properly embedded minimal surface of $\mathbb{S}^2 \times \mathbb{R}$. If $\Sigma$ contains a vertical geodesic, then it also contains the antipodal geodesic. More precisely if $\{p\} \times \mathbb{R} \subset \Sigma$, then $\{-p\}\times \mathbb{R} \subset \Sigma$. \item Let $\Sigma$ be an oriented, embedded minimal surface of $\mathbb{S}^2 \times \mathbb{S}^1(r)$ different from a finite union of horizontal slices. If $\Sigma$ contains a horizontal geodesic, then it also contains another horizontal geodesic at vertical distance $\pi r$. More precisely, if $\Gamma\times \{0\} \subset \Sigma$, $\Gamma$ a great circle of $\mathbb{S}^2$, then $\Gamma\times\{\pi r\} \subset \Sigma$. \end{enumerate} \end{proposition} \begin{proof} Both assertions follow from two facts: (1) under the hypothesis the surface $\Sigma$ separates the ambient space in two different connected components; (2) there exists an isometry $\rho$ of the ambient space which preserves the surface by the Schwarz reflection principle but interchanges the connected components of the complement of $\Sigma$. Hence, the fixed points of $\rho$ must be contained in $\Sigma$. In the first case, $\rho$ is the reflection around $\{p\} \times \mathbb{R}$ whereas in the second case $\rho$ is the reflection around $\Gamma \times \{0\}$. The first assertion appears in~\cite[\S 2]{HW}. \end{proof} One important consequence of this proposition is that, generically, the oriented compact embedded minimal surfaces constructed solving and reflecting a Plateau problem have odd genus. More precisely, let $M$ a minimal disk (different from an open subset of a slice) spanning a polygon $\Gamma$ made of horizontal and vertical geodesics. Suppose that $M$ produces a compact surface $\Sigma$ by Schwarz reflection, and the angle function $\nu = \prodesc{N}{\xi}$, where $N$ is the unit normal to $M$, satisfies $\nu^2\lvert_{M} < 1$ and $\nu^2\lvert_\Gamma = 1$ only at the points of $\Gamma$ where two horizontal geodesics meet. Then $\Sigma$ has odd genus if it is orientable or it has even genus if it is not. In order to prove this, we consider the vector field $T = \xi - \nu N$ which is the tangent part of $\xi$, so $\abs{T}^2 = 1-\nu^2$, i.e., $T$ only vanishes in the points where $\nu^2 = 1$ that turns out to be the points where the Hopf differential associated to $\Sigma$ vanishes. The zeros of $T$ appear in multiples of $4$ since if $(p,0) \in \Sigma$ is such a point then it is an intersection point of two horizontal geodesics $\gamma_1 \times \{0\}$ and $\gamma_2\times \{0\}$ contained in the surface $\Sigma$ and so the point $(-p,0) \in (\gamma_1\cap \gamma_2) \times \{0\}$ is also in $\Sigma$ and $T_{(-p,0)} = 0$. Moreover, by Proposition~\ref{prop:symmetries}, the horizontal geodesics $\gamma_j \times\{\pi r\}$, $j = 1, 2$, are contained in $\Sigma$ so $(p, \pi r)$ and $(-p, \pi r)$ are also zeros of $T$. Applying the Poincaré-Hopf theorem to the vector field $T$, the Euler characteristic of $\Sigma$ must be $-4k$, that is, the genus of $\Sigma$ is $g = 2k + 1$ if $\Sigma$ is orientable or $g = 2(1+2k)$ if it is not. Note that this trick is not useful in general for the conjugate surface of $M$ (we assume that the conjugate piece can be reflected to obtain a compact surface): Although the angle function is preserved by conjugation and so are the zeros of the vector field $T$, we can not guarantee in general that the zeros of $T$ occur in the intersection of two horizontal geodesics in the conjugate piece.
\section{Introduction} We are concerned with finite strings over a finite alphabet $\Sigma_k$ having $k \geq 2$ letters. A {\it palindrome} is a string $x$ equal to its reversal $x^R$, like the English word {\tt radar}. If $T, U$ are sets of strings over $\Sigma_k$ then (as usual) $TU = \lbrace tu \ : \ t \in T, u \in U \rbrace$. Also $T^i = \overbrace{TT \cdots T}^i$ and $T^* = \bigcup_{i \geq 0} T^i$ and $T^+ = \bigcup_{i \geq 1} T^i$. We define $$P = \lbrace x\, x^R \ : \ x \in \Sigma_k^+ \rbrace,$$ the language of nonempty even-length palindromes. Following Knuth, Morris, and Pratt \cite{Knuth&Morris&Pratt:1977}, we call a string $x$ a {\it palstar} if it belongs to $P^*$, that is, if it can be written as the concatenation of elements of $P$. Clearly every palstar is of even length. We call $x$ a {\it prime palstar} if it is a nonempty palstar, but not the concatenation of two or more palstars; alternatively, if $x \in P^+ - P^2 P^* $ where $-$ is set difference. Thus, for example, the the English word {\tt noon} is a prime palstar, but the English word {\tt appall} and the French word {\tt assailli} are palstars that are not prime. Knuth, Morris, and Pratt \cite{Knuth&Morris&Pratt:1977} proved that no prime palstar is a proper prefix of another prime palstar, and, consequently, every palstar has a unique factorization as a concatenation of prime palstars. A nonempty string $x$ is a {\it border} of a string $y$ if $x$ is both a prefix and a suffix of $y$ and $x \not= y$. We say a string $y$ is {\it bordered} if it has a border. Thus, for example, the English word {\tt ionization} is bordered with border {\tt ion}. Otherwise a word is {\it unbordered}. Rampersad et al.~\cite{Rampersad&Shallit&Wang:2011} recently gave a bijection between the unbordered strings of length $n$ and the prime palstars of length $2n$. As a consequence they obtained a formula for the number of prime palstars. Despite some interest in the palstars themselves \cite{Manacher:1975,Galil&Seiferas:1978}, it seems no one has enumerated them. Here we observe that bijection mentioned previously, together with the unique factorization of palstars, provides an asymptotic enumeration for the number of palstars. \section{Generating function for the palstars} Again, let $k \geq 2$ denote the size of the alphabet. Let $p_k(n)$ denote the number of palstars of length $2n$, and let $u_k (n)$ denote the number of unbordered strings of length $n$. \begin{lemma} For $n \geq 1$ and $k \geq 2$ we have $$ p_k (n) = \sum_{1 \leq i \leq n} u_k(i) p_k(n-i) .$$ \label{one} \end{lemma} \begin{proof} Consider a palstar of length $2n > 0$. Either it is a prime palstar, and by \cite{Rampersad&Shallit&Wang:2011} there are $u_k(n) = u_k (n) p_k (0)$ of them, or it is the concatenation of two or more prime palstars. In the latter case, consider the length of this first factor; it can potentially be $2i$ for $1 \leq i \leq n$. Removing this first factor, what is left is also a palstar. This gives $u_k (i) p_k (n-i)$ distinct palstars for each $i$. Since factorization into prime palstars is unique, the result follows. \end{proof} Now we define generating functions as follows: \begin{eqnarray*} P_k (X) &=& \sum_{n \geq 0} p_k (n) X^n \\ U_k (X) &=& \sum_{n \geq 0} u_k (n) X^n . \end{eqnarray*} The first few terms are as follows: \begin{eqnarray*} P_k (X) &=& 1 + kX + (2k^2-k)X^2 + (4k^3-3k^2)X^3 + (8k^4-8k^3+k)X^4 + \cdots \\ U_k (X) &=& 1 + kX + (k^2-k)X^2 + (k^3-k^2)X^3 + (k^4-k^3-k^2+k)X^4 + \cdots . \end{eqnarray*} \begin{theorem} $$P_k (X) = {1 \over {2-U_k (X)}}.$$ \label{two} \end{theorem} \begin{proof} From Lemma~\ref{one}, we have \begin{eqnarray*} U_k (X) P_k (X) &=& \left(\sum_{n \geq 0} u_k (n) X^n \right) \left(\sum_{n \geq 0} p_k (n) X^n \right) \\ &=& 1 + \sum_{n \geq 1} \left(\sum_{0 \leq i \leq n} u_k(i) p_k (n-i) \right) X^n \\ &=& 1+ \left( \sum_{n \geq 1} \sum_{1 \leq i \leq n} u_k(i) p_k (n-i) X^n \right) + \sum_{n \geq 1} p_k(n) X^n \\ &=& 1 + \left( \sum_{n \geq 1} p_k (n) X^n \right) + \sum_{n \geq 1} p_k (n) X^n \\ &=& 2 P_k (X) - 1, \end{eqnarray*} from which the result follows immediately. \end{proof} \section{The main result} \begin{theorem} For all $k \geq 2$ there is a constant $\alpha_k$ with $2k -1 < \alpha_k < 2k-{1 \over 2}$ such that the number of palstars of length $2n$ is $\Theta(\alpha_k^n)$. \end{theorem} \begin{proof} From Theorem~\ref{two} and the ``First Principle of Coefficient Asymptotics'' \cite[p.~260]{Flajolet&Sedgewick:2009}, it follows that the asymptotic behavior of $[X^n] P_k (X)$, the coefficient of $X^n$ in $P_k (X)$, is controlled by the behavior of the roots of $U_k (X) = 2$. Since $u_{k}(0) = 1$ and $U_{k}(X) \rightarrow \infty$ as $X \rightarrow \infty$, the equation $U_k (X) = 2$ has a single positive real root, which is $\rho = \rho_k = \alpha_k^{-1}$. We first show that $2k -1 < \alpha_k < 2k-{1 \over 2}$. Recalling that $u_k(n)$ is the number of unbordered strings of length $n$ over a $k$-letter alphabet, we see that $u_k(n) \leq k^n - k^{n-1}$ for $n \geq 2$, since $k^n$ counts the total number of strings of length $n$, and $k^{n-1}$ counts the number of strings with a border of length $1$. Similarly $$u_k(n) \geq \begin{cases} k^n - k^{n-1} - \cdots - k^{n/2}, & \text{if $n \geq 2$ is even}; \\ k^n - k^{n-1} - \cdots - k^{(n+1)/2}, & \text{if $n \geq 2$ is odd}, \end{cases} $$ since this quantity represents removing strings with borders of lengths $1, 2, \ldots, n/2$ (resp., $1,2, \ldots, (n+1)/2$) if $n$ is even (resp., odd) from the total number. Here we use the classical fact that if a word of length $n$ has a border, it has one of length $\leq n/2$. It follows that for real $X > 0$ we have \begin{align*} U_k (X) &= \sum_{n \geq 0} u_k(n) X^n \\ &= 1 + kX + \sum_{n \geq 2} u_k(n) X^n \\ &\leq 1 + kX + \sum_{n \geq 2} (k^n - k^{n-1}) X^n \\ & = {{kX^2 - 1} \over {kX - 1 }} . \end{align*} Similarly for real $X > 0$ we have \begin{align*} U_k(X) & = \sum_{n \geq 0} u_k(n) X^n \\ &= 1 +kX + \sum_{l \geq 1} u_k (2l) X^{2l} + \sum_{m \geq 1} u_k (2m+1) X^{2m+1} \\ &\geq 1 + kX + \sum_{l \geq 1} (k^{2l} - k^{2l-1} - \cdots - k^l) X^{2l} + \sum_{m \geq 1} (k^{2m+1} - \cdots - k^{m+1}) X^{2m+1} \\ &= {{1-2kX^2} \over {(kX-1)(kX^2-1)}} . \end{align*} This gives, for $k \geq 2$, that $$ 2 < {{(2k-1) (4k^2 - 6k+1) } \over {(k-1)^2 (4k-1)}} \leq U_k \left({1 \over {2k-1}} \right) $$ and $$ U_k\left({ 1 \over {2k-{1\over 2}}}\right) \leq {{16k^2-12k+1} \over {(4k-1)(2k-1)}} < 2 .$$ It follows that ${1 \over {2k -{1 \over 2}}} \leq \rho_k \leq {1 \over {2k-1}} $ and hence $2k -1 < \alpha_k < 2k-{1 \over 2}$. To understand the asymptotic behavior of $[X^n] P_k (X)$, we need to rule out other (complex) roots with the same absolute value as $\rho$. Suppose there is another solution $X = \rho e^{i\psi}$ with $-\pi < \psi < 0$ or $0 < \psi \le \pi $. Then $$2 = \sum_{n \ge 0}u_{k}(n)\rho^{n}e^{in\psi} = \sum_{n \ge 0}u_{k}(n)\rho^{n}\cos n\psi + i\sum_{n \ge 0}u_{k}(n)\rho^{n}\sin n\psi.$$ We must have $\sum_{n \ge 0}u_{k}(n)\rho^{n}\sin n\psi = 0$. Hence $$2 = \sum_{n \ge 0}u_{k}(n)\rho^{n}\cos n\psi.$$ Now if $| \cos n\psi | < 1$ for some $n$ then $$2 = \left|\sum_{n \ge 0}u_{k}(n)\rho^{n}\cos n\psi \right| \le \sum_{n \ge 0}u_{k}(n)\rho^{n}|\cos n\psi| < \sum_{n \ge 0}u_{k}(n)\rho^{n} = 2 .$$ This is a contradiction, so $|\cos n\psi| = 1$ for all $n$. Hence $\cos n\psi = \pm 1$ for all $n$. Thus $n\psi = \pm \pi + 2\pi l_{n}$ for all $n$ and $l_{n}$ is an integer for all $n$. Since $\cos x = \cos (-x)$, we may suppose that $0 < \psi \le \pi$. If $\psi = \pi$ then $X = -\rho$. But then, using the fact that $u_{k}(1) = k$, we get $$2 = \sum_{n = 0}^{\infty}u_{k}(n)\rho^{n}(-1)^{n} < \sum_{n = 0}u_{k}(n)\rho^{n} = 2.$$ This contradiction shows that we may suppose $0 < \psi < \pi$. Suppose $\cos n\psi = \pm 1$ for all $n$. Then for all $n$ $$n\psi = \pm \pi + 2\pi l_{n}$$ where $l_{n}$ is an integer. Thus $n\psi/\pi = \pm 1 + 2l_{n}$ for all $n$. From Dirichlet's diophantine approximation theorem (e.g., \cite[Thm.~185]{Hardy&Wright:1971}), given $\psi/\pi$ and an integer $q \ge 1$, there are infinitely many $n$ and integers $x$ such that $$\left|n\frac{\psi}{\pi} - x\right| < \frac{1}{q}.$$ Thus $x - 1/q < n\psi/\pi < x + 1/q$ and $$\left|\pm 1 + 2l_{n} - x\right| < \frac{1}{q}.$$ Choosing $q > 1$ we see that $\pm 1 + 2l_{n} - x$ is an integer $<1$ in absolute value. Thus $\pm 1 + 2l_{n} -x = 0$, and so $n\psi/\pi$ is an integer for infinitely many $n$. Thus $\psi/\pi = p/m$ or $\psi = (p/m)\pi$ and we may suppose $p$ and $m$ are coprime. We have seen that if $X = \rho^{n} e^{nip/m}$ and if $|\cos np/m| < 1$ for any $n$ we have a contradiction. Therefore for all $n$ we have $$\cos\left(\frac{n p \pi}{m}\right) = \pm 1.$$ Thus $n p \pi/m = l\pi$ for all $n$. Thus $m$ divides $n$ for all $n$. This is a contradiction if $n = m + 1$. Thus $$\frac{1}{2 - U_{k}(x)}$$ has only one singularity $x = \rho > 0$ with $|x| = \rho$. It remains to determine the order of the zero $\rho$. From above $U_{k}(X) = 2$ has a solution $\alpha_{k}^{-1}$ which satisfies $2k - 1 < \alpha_{k} < 2k - {1 \over 2}$. Nielsen \cite{Nielsen:1973} showed that $u_k (n) \sim c_{k}k^{n}$ for a constant $c_k$. Thus $U_{k}(X)$ has radius of convergence $1/k$. Thus $1/\alpha_{k}$ is in the region where $U_{k}$ is analytic. Thus $2 - U_{k}(X)$ is analytic at $1/\alpha_{k}$ and has a zero at $1/\alpha_{k}$ of multiplicity $m$. If $m \ge 2$ then the derivative of $2 - U_{k}(X)$ equals $0$ at $X = 1/\alpha_{k}$. However $U_{k}^{'}(X) > 0$ since $u_{k}(n) > 0$ for some $n$. Thus $2 - U_{k}(X)$ has a simple zero at $X = 1/\alpha_{k}$, and so $P_{k}(X)$ has a simple pole at $X = 1/\alpha_{k}$. Near $\alpha_{k}^{-1}$ the generating function $ U_{k}(X)$ has the expansion $2 + C_k(X - \alpha_{k}^{-1}) + C'(X - \alpha_{k}^{-1})^{2} + \cdots$ with $C_k > 0$. Furthermore $$P_{k}(X) = \frac{1}{2 - U_{k}(X)} = \frac{1}{-C_k(X - 1/\alpha_{k}) - C'(X - 1/\alpha_{k})^{2} + \cdots}.$$ Now $$P_{k}^{'}(X) = \frac{U_{k}^{'}(X)}{(2 - U_{k}(X))^{2}},$$ so there is a positive $\delta$ such that $$[X^{n}]P_{k}(X) = [X^{n}]\frac{1}{C_k(1/\alpha_{k} -X)} +\cdots = [X^{n}]\frac{\alpha_{k}}{C_k}\left(\frac{1}{1 - \alpha_{k}X}\right) + \cdots = \frac{\alpha_{k}^{n + 1}}{C_k} + O\left(\left(\alpha_{k} - \delta\right)^{n}\right),$$ since $$P_{k}(X) - \frac{\alpha_{k}}{C_k}\left(\frac{1}{1 - \alpha_{k}X}\right)$$ has no singularity on $|X| = 1/\alpha_{k}$ so has radius of convergence $> 1/\alpha_{k}$. Here $$C_k = U_{k}^{'}\left(\frac{1}{\alpha_{k}}\right).$$ It now follows from standard results (e.g., \cite[Thm.~IV.7, p.~244]{Flajolet&Sedgewick:2009}) that $$ [X^n] P_k (X) = {{\alpha_k^{n+1}} \over {C_k}} + O((\alpha_k - \delta)^n) = \Theta(\alpha_k^n) .$$ \end{proof} \section{Numerical results} Here is a table giving the first few values of $P_k (n)$. \begin{figure}[H] \begin{center} \begin{tabular}{cccccccccccc} \hline $n = $ & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 \\ \hline $k = 2$& 1 & 2 & 6 & 20 & 66 & 220 & 732 & 2440 & 8134 & 27124 & 90452 \\ $k = 3$& 1 & 3 & 15 & 81 & 435 & 2349 & 12681 & 68499 & 370023 & 1998945 & 10798821 \\ $k = 4$& 1 & 4 & 28 & 208 & 1540 & 11440 & 84976 & 631360 & 4690972 & 34854352 & 258971536 \\ \hline \end{tabular} \end{center} \end{figure} By truncating the power series $U_k(X)$ and solving the equation $U_k (X) = 2$ we get better and better approximations to $\alpha_k^{-1}$. For example, for $k = 2$ we have \begin{eqnarray*} \alpha_2^{-1} &\doteq & 0.29983821359352690506155111814579603919303182364781730366339199333065202 \\ \alpha_2 & \doteq & 3.3351319300335793676678962610376244842363270634405611577104447308511860 \\ C_2 & \doteq & 6.278652437421018217684895562492005276088368718322063642652328654828673 \end{eqnarray*} To determine an asymptotic expansion for $\alpha_k$ as $k \rightarrow \infty$, we compute the Taylor series expansion for $P_k(n)/P_k(n+1)$, treating $k$ as an indeterminate, for $n$ large enough to cover the error term desired. For example, for $O(k^{-10})$ it suffices to take $k = 16$, which gives $$ \alpha_k^{-1} = {1 \over {2k}} + {1 \over {8k^2}} + {3 \over {32k^3}} + {1 \over {16k^4}} + {{27} \over {512k^5}} + {{93} \over {2048k^6}} + {{83} \over {2048k^7}} + {{155} \over {4096k^8}} + {{4735} \over {131072k^9}} + O(k^{-10}) $$ and hence $$ \alpha_k = 2k - {1 \over 2} - {1 \over {4k}} - {3 \over {32k^2}} - {5 \over {64k^3}} - {{31} \over {512k^4}} - {{25} \over {512k^5}} - {{23} \over {512 k^6}} - {{683} \over {16384k^7}} + O(k^{-8}).$$
\section{Introduction} In this work we study the evolution of the interface between two different incompressible fluids with the same viscosity coefficient in a porous medium with two different permeabilities. This problem is of practical importance because it is used as a model for a geothermal reservoir (see \cite{CF} and references therein). The velocity of a fluid flowing in a porous medium satisfies Darcy's law (see \cite{bear,Muskat,bn}) \begin{equation} \frac{\mu}{\kappa(\vec{x})}v=-\nabla p-g\rho(\vec{x}) (0,1), \label{IIeq1} \end{equation} where $\mu$ is the dynamic viscosity, $\kappa(\vec{x})$ is the permeability of the medium, $g$ is the acceleration due to gravity, $\rho(\vec{x})$ is the density of the fluid, $p(\vec{x})$ is the pressure of the fluid and $v(\vec{x})$ is the incompressible velocity field. In our favourite units, we can assume $g=\mu=1.$ The spatial domains considered in this work are $S=\mathbb R^2,\mathbb T\times\mathbb R$ (infinite depth) and $\mathbb R\times(-\pi/2,\pi/2)$ (finite depth). We have two immiscible and incompressible fluids with the same viscosity and different densities; $\rho^1$ fill in the upper domain $S^1(t)$ and $\rho^2$ fill in the lower domain $S^2(t)$. The curve $$ z(\alpha,t)=\{(z_1(\alpha,t),z_2(\alpha,t)): \: \alpha\in\mathbb R\} $$ is the interface between the fluids. In particular we are making the ansatz that $S^1$ and $S^2$ are a partition of $S$ and they are separated by a curve $z$. The system is in the stable regime if the denser fluid is below the lighter one, \emph{i.e.} $\rho^2>\rho^1$. This is known in the literature as the Rayleigh-Taylor condition. The function that measures this condition is defined as $$ RT(\alpha,t)=-(\nabla p^2(z(\alpha,t))-\nabla p^1(z(\alpha,t)))\cdot\partial_\alpha^\bot z(\alpha,t)>0. $$ In the case with $\kappa(\vec{x})\equiv\text{costant}>0$, the motion of a fluid in a two-dimensional porous medium is analogous to the Hele-Shaw cell problem (see \cite{cheng2012global, constantin1999global, escher1997classical, H-S} and the references therein) and if the fluids fill the whole plane (in the case with the same viscosity but different densities) the contour equation satisfies (see \cite{c-g07}) \begin{equation}\label{IIfull} \partial_t f=\frac{\rho^2-\rho^1}{2\pi}\text{P.V.}\int_\mathbb R \frac{(\partial_x f(x)-\partial_x f(x-\eta))\eta}{\eta^2+(f(x)-f(x-\eta))^2}d\eta. \end{equation} They show the existence of classical solution locally in time (see \cite{c-g07} and also \cite{ambrose2004well, e-m10, escher2011generalized, KK}) in the Rayleigh-Taylor stable regime which means that $\rho^2>\rho^1$, and maximum principles for $\|f(t)\|_{L^\infty}$ and $\|\partial_x f(t)\|_{L^\infty}$ (see \cite{c-g09}). Moreover, in \cite{ccfgl} the authors show that there exists initial data in $H^4$ such that $\|\partial_x f\|_{L^\infty}$ blows up in finite time. Furthermore, in \cite{castro2012breakdown} the authors prove that there exist analytic initial data in the stable regime for the Muskat problem such that the solution turns to the unstable regime and later no longer belongs to $C^4$. In \cite{ccgs-10} the authors show an energy balance for $L^2$ and that if initially $\|\partial_x f_0\|_{L^\infty}<1$, then there is global lipschitz solution and if the initial datum has $\|f_0\|_{H^3}<1/5$ then there is global classical solution. In \cite{c-c-g10, SCH} the authors study the case with different viscosities. In \cite{knupfer2010darcy} the authors study the case where the interface reach the boundary in a moving point with a constant (non-zero) angle. \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{scheme.eps} \end{center} \caption{Physical situation} \label{IIscheme} \end{figure} The case where the fluid domain is the strip $\mathbb R\times(-l,l)$, with $0<l$, has been studied in \cite{CGO, e-m10, escher2011generalized}. In this regime the equation for the interface is \begin{multline} \partial_t f(x,t) = \frac{\rho^2-\rho^1}{8l}\text{P.V.}\int_\mathbb R\bigg{[}\frac{\left (\partial_xf\left (x\right )-\partial_xf\left (x-\eta\right )\right)\sinh\left(\frac{\pi}{2l}\eta\right)}{\cosh \left(\frac{\pi}{2l}\eta\right)-\cos(\frac{\pi}{2l}(f(x)-f(x-\eta)))}\\ + \frac{(\partial_xf\left (x\right )+\partial_xf\left (x-\eta\right )\sinh\left(\frac{\pi}{2l}\eta\right)}{\cosh \left(\frac{\pi}{2l}\eta\right)+\cos(\frac{\pi}{2l}(f(x)+f(x-\eta)))}\bigg{]}d\eta. \label{IIeq0.1} \end{multline} For equation \eqref{IIeq0.1} the authors in \cite{CGO} obtain the existence of classical solution locally in time in the stable regime case where the initial interface does not reach the boundaries, and the existence of finite time singularities. These singularities mean that the curve is initially a graph in the stable regime, and in finite time, the curve can not be parametrized as a graph and the interface turns to the unstable regime. Also the authors study the effect of the boundaries on the evolution of the interface, obtaining the maximum principle and a decay estimate for $\|f\|_{L^\infty}$ and the maximum principle for $\|\partial_x f\|_{L^\infty}$ for initial datum satisfying smallness conditions on $\|\partial_x f_0\|_{L^\infty}$ and on $\|f_0\|_{L^\infty}$. So, not only the slope must be small, also amplitude of the curve plays a role. Both result differs from the results corresponding to the infinite depth case \eqref{IIfull}. We note that the case with boundaries can also be understood as a problem with different permeabilities where the permeability outside vanishes. In the forthcoming work \cite{GG} the authors compare the different models \eqref{IIfull}, \eqref{IIeq0.1} and \eqref{IIeq13} from the point of view of the existence of turning waves. In this work we study the case where permeability $\kappa(\vec{x})$ is a step function, more precisely, we have a curve $$ h(\alpha)=\{(h_1(\alpha),h_2(\alpha)): \: \alpha\in\mathbb R\} $$ separating two regions with different values for the permeability (see Figure \ref{IIscheme}). We study the regime with infinite depth, for periodic and for "flat at infinity" initial datum, but also the case where the depth is finite and equal to $\frac{\pi}{2}$. In the region above the curve $h(\alpha)$ the permeability is $\kappa(\vec{x})\equiv\kappa^1$, while in the region below the curve $h(\alpha)$ the permeability is $\kappa(\vec{x})\equiv\kappa^2\neq\kappa^1$. Note that the curve $h(\alpha)$ is known and fixed. Then it follows from Darcy's law that the vorticity is $$ \omega(\vec{x})=\varpi_1(\alpha,t)\delta(\vec{x}-z(\alpha,t))+\varpi_2(\alpha,t)\delta(\vec{x}-h(\alpha)), $$ where $\varpi_1$ corresponds to the difference of the densities, $\varpi_2$ corresponding to the difference of permeabilities and $\delta$ is the usual Dirac's distribution. In fact both amplitudes for the vorticity are quite different, while $\varpi_1$ is a derivative, the amplitude $\varpi_2$ has a nonlocal character (see \eqref{IIeq10A}, \eqref{IIeq14A} and Section \ref{IIsec2}). The equation for the interface, when $h(x)=(x,-h_2)$ and the fluid fill the whole plane, is \begin{multline} \label{IIeq9} \partial_t f(x)=\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\frac{(\partial_x f(x)-\partial_x f(\beta))(x-\beta)}{(x-\beta)^2+(f(x)-f(\beta))^2}d\beta\\ +\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(\beta)(x-\beta+\partial_x f(x)(f(x)+h_2))}{(x-\beta)^2+(f(x)+h_2)^2}d\beta, \end{multline} with \begin{eqnarray} \varpi_2(x)&=&\frac{\kappa^1-\kappa^2}{\kappa^2+\kappa^1}\frac{\kappa^1(\rho^2-\rho^1)}{\pi}\text{P.V.}\int_\mathbb R\frac{\partial_x f(\beta)(h_2+f(\beta))}{(x-\beta)^2+(-h_2-f(\beta))^2}d\beta\label{IIeq10A} \end{eqnarray} If the fluids fill the whole space but the initial curve is periodic the equation reduces to \begin{multline}\label{IIeq13} \partial_t f(x)=\frac{\kappa^1(\rho^2-\rho^1)}{4\pi} \text{P.V.}\int_\mathbb T\frac{\sin(x-\beta)(\partial_x f(x)-\partial_x f(\beta))d\beta}{\cosh(f(x)-f(\beta))-\cos(x-\beta)}\\ +\frac{1}{4\pi} \text{P.V.}\int_\mathbb T\frac{(\partial_x f(x)\sinh(f(x)+h_2)+\sin(x-\beta))\varpi_2(\beta)d\beta}{\cosh(f(x)+h_2)-\cos(x-\beta)}, \end{multline} where the second vorticity amplitude can be written as \begin{eqnarray} \varpi_2(x)&=&\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\frac{\kappa^1-\kappa^2}{\kappa^1+\kappa^2} \text{P.V.}\int_\mathbb T\frac{\sinh(h_2+f(\beta))\partial_x f(\beta)d\beta}{\cosh(h_2+f(\beta))-\cos(x-\beta)}.\label{IIeq14A} \end{eqnarray} If we consider the regime where the amplitude of the wave and the depth of the medium are of the same order then the equation for the interface, when the depth is chosen to be $\pi/2$, is \begin{eqnarray} \partial_t f(x)&=&\frac{\kappa^1(\rho^2-\rho^1)}{4\pi}\text{P.V.}\int_\mathbb R\frac{(\partial_x f(x)-\partial_x f(\beta))\sinh(x-\beta)}{\cosh(x-\beta)-\cos(f(x)-f(\beta))}d\beta\nonumber\\ &&+\frac{\kappa^1(\rho^2-\rho^1)}{4\pi}\text{P.V.}\int_\mathbb R\frac{(\partial_x f(x)+\partial_x f(\beta))\sinh(x-\beta)}{\cosh(x-\beta)+\cos(f(x)+f(\beta))}d\beta\nonumber\\ &&+\frac{1}{4\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(\beta)(\sinh(x-\beta)+\partial_x f(x)\sin(f(x)+h_2))}{\cosh(x-\beta)-\cos(f(x)+h_2)}d\beta\nonumber\\ &&+\frac{1}{4\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(\beta)(-\sinh(x-\beta)+\partial_x f(x)\sin(f(x)-h_2))}{\cosh(x-\beta)+\cos(f(x)-h_2)}d\beta\label{IIeqv2}, \end{eqnarray} where \begin{eqnarray} \varpi_2(x)&=&\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\partial_x f(\beta)\frac{\sin(h_2+f(\beta))}{\cosh(x-\beta)-\cos(h_2+f(\beta))}d\beta\nonumber\\ &&-\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\partial_x f(\beta)\frac{\sin(-h_2+f(\beta))}{\cosh(x-\beta)+\cos(-h_2+f(\beta))}d\beta\nonumber\\ &&+\frac{\mathcal{K}^2}{\sqrt{2\pi}}\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}G_{h_2,\mathcal{K}}*\text{P.V.}\int_\mathbb R\frac{\partial_x f(\beta)\sin(h_2+f(\beta))}{\cosh(x-\beta)-\cos(h_2+f(\beta))}d\beta\nonumber\\ &&-\frac{\mathcal{K}^2}{\sqrt{2\pi}}\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}G_{h_2,\mathcal{K}}*\text{P.V.}\int_\mathbb R\frac{\partial_x f(\beta)\sin(-h_2+f(\beta))}{\cosh(x-\beta)+\cos(-h_2+f(\beta))}d\beta, \label{IIw2defc} \end{eqnarray} with $$ G_{h_2,\mathcal{K}}(x)=\mathcal{F}^{-1}\left(\frac{\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}{1+\frac{\mathcal{K}}{\sqrt{2\pi}}\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}\right) $$ a Schwartz function. \begin{coment} For notational simplicity, we denote $\mathcal{K}=\frac{\kappa^1-\kappa^2}{\kappa^1+\kappa^2}$ and we drop the $t$ dependence. \end{coment} The plan of the paper is as follows: in Section \ref{IIsec2} we derive the contour equations \eqref{IIeq9},\eqref{IIeq13} and \eqref{IIeqv2}. In Section \ref{IIsec3} we show the local in time solvability and an energy balance for the $L^2$ norm. In Section \ref{IIsec5} we perform numerics and in Section \ref{IIsec4} we obtain finite time singularities for equations \eqref{IIeq9} \eqref{IIeq13} and \eqref{IIeqv2} when the physical parameters are in some region and numerical evidence showing that, in fact, every value is valid for the physical parameters. \section{The contour equation}\label{IIsec2} In this section we derive the contour equations \eqref{IIeq9}, \eqref{IIeq13} and \eqref{IIeqv2}, \emph{i.e.} the equations for the interface. First we obtain the equation in the infinite depth case, both, flat at infinity and periodic. Given $\omega$ a scalar, $\gamma,z,$ curves, and a spatial domain $\Omega=\mathbb T$ or $\Omega=\mathbb R$, we denote the Birkhoff-Rott integral as \begin{equation} \label{IIeq2} BR(\omega,z)\gamma=\text{P.V.}\int_\Omega \omega(\beta) BS(\gamma_1(\alpha),\gamma_2(\alpha),z_1(\beta),z_2(\beta))d\beta, \end{equation} where $BS$ denotes the kernel of $\nabla^\perp\Delta^{-1}$ (which depends on the domain). If the domain is $\mathbb R^2$ we have \begin{equation}\label{IIBSplane} BS(x,y,\mu,\nu)=\frac{1}{2\pi}\left(-\frac{y-\nu}{(y-\nu)^2+(x-\mu)^2}, \frac{x-\mu}{(y-\nu)^2+(x-\mu)^2}\right), \end{equation} for $\mathbb T\times \mathbb R$ we have \begin{equation}\label{IIBSperiodic} BS(x,y,\mu,\nu)=\frac{1}{4\pi}\left( \frac{-\sinh(y-\nu)}{\cosh(y-\nu)-\cos(x-\mu)}, \frac{\sin(x-\mu)}{\cosh(y-\nu)-\cos(x-\mu)}\right), \end{equation} and for $\mathbb R\times(-\pi/2,\pi/2)$ the kernel is (see \cite{CGO}) \begin{multline}\label{IIBSconf} BS(x,y,\mu,\nu)=\frac{1}{4\pi}\left(-\frac{\sin(y-\nu)}{\cosh(x-\mu)-\cos(y-\nu)}-\frac{\sin(y+\nu)}{\cosh(x-\mu)+\cos(y+\nu)},\right.\\ \left.\frac{\sinh(x-\mu)}{\cosh(x-\mu)-\cos(y-\nu)}-\frac{\sinh(x-\mu)}{\cosh(x-\mu)+\cos(y+\nu)}\right). \end{multline} \subsection{Infinite depth} \subsubsection{Assuming $S=\mathbb R^2$:} Using the kernel \eqref{IIBSplane}, we obtain \begin{equation}\label{IIBS} v(\vec{x})=\frac{1}{2\pi}\text{P.V.}\int_\mathbb R \varpi_1(\beta)\frac{(\vec{x}-z(\beta))^\perp}{|\vec{x}-z(\beta)|^2}d\beta+\frac{1}{2\pi}\text{P.V.}\int_\mathbb R \varpi_2(\beta)\frac{(\vec{x}-h(\beta))^\perp}{|\vec{x}-h(\beta)|^2}d\beta, \end{equation} where $(a,b)^\perp=(-b,a).$ We have \begin{equation} \label{IIeq3} v^\pm(z(\alpha))=\lim_{\epsilon\rightarrow0}v(z(\alpha)\pm\epsilon\partial_\alpha^\perp z(\alpha))=BR(\varpi_1,z)z+BR(\varpi_2,h)z\mp\frac{1}{2}\frac{\varpi_1(\alpha)}{|\partial_\alpha z(\alpha)|^2}\partial_\alpha z(\alpha), \end{equation} and \begin{equation} \label{IIeq4} v^\pm(h(\alpha))=\lim_{\epsilon\rightarrow0}v(h(\alpha)\pm\epsilon\partial_\alpha^\perp h(\alpha))=BR(\varpi_1,z)h+BR(\varpi_2,h)h\mp\frac{1}{2}\frac{\varpi_2(\alpha)}{|\partial_\alpha h(\alpha)|^2}\partial_\alpha h(\alpha). \end{equation} We observe that $v^+(z(\alpha))$ is the limit inside $S^1$ (the upper subdomain) and $v^-(z(\alpha))$ is the limit inside $S^2$ (the lower subdomain). The curve $z(\alpha)$ doesn't touch the curve $h(\alpha)$, so, the limit for the curve $h$ are in the same domain $S^i$. Using Darcy's Law and assuming that the initial interface $z(\alpha,0)$ is in the region with permeability $\kappa^1$, we obtain \begin{eqnarray*} (v^-(z(\alpha))-v^+(z(\alpha)))\cdot\partial_\alpha z(\alpha)&=&\kappa^1\left(-\partial_\alpha(p^-(z(\alpha))-p^+(z(\alpha)))\right)-\kappa^1(\rho^2-\rho^1)\partial_\alpha z_1(\alpha)\\ &=&0-\kappa^1(\rho^2-\rho^1)\partial_\alpha z_2(\alpha), \end{eqnarray*} where in the last equality we have used the continuity of the pressure along the interface (see \cite{c-c-g10}). Using \eqref{IIeq3} we conclude \begin{equation} \label{IIeq5} \varpi_1(\alpha)=-\kappa^1(\rho^2-\rho^1)\partial_\alpha z_2(\alpha). \end{equation} We need to determine $\varpi_2$. We consider \begin{eqnarray*} \left[\frac{v}{\kappa}\right]&=&\left(\frac{v^-(h(\alpha))}{\kappa^2}-\frac{v^+(h(\alpha))}{\kappa^1}\right)\cdot\partial_\alpha h(\alpha)\\ &=&-\partial_\alpha (p^-(h(\alpha))-p^+(h(\alpha)))\\ &=&0, \end{eqnarray*} where the first equality is due to Darcy's Law. Using the expression \eqref{IIeq4} we have \begin{equation}\label{w2eq} \left[\frac{v}{\kappa}\right]=\left(\frac{1}{\kappa^2}-\frac{1}{\kappa^1}\right)\left(BR(\varpi_1,z)h+BR(\varpi_2,h)h\right)\cdot\partial_\alpha h(\alpha)+\left(\frac{1}{2\kappa^2}+\frac{1}{2\kappa^1}\right)\varpi_2. \end{equation} We take $h(\alpha)=(\alpha,-h_2)$, with $h_2>0$ a fixed constant. Then $$ BR(\varpi_2,h)h\cdot\partial_\alpha h=\left(0,\frac{1}{2}H(\varpi_2)\right)\cdot(1,0)=0, $$ where $H$ denotes the Hilbert transform. Finally, we have \begin{equation} \label{IIeq6} \varpi_2(\alpha)=-2\mathcal{K}BR(\varpi_1,z)h\cdot(1,0)=\mathcal{K}\frac{1}{\pi}\text{P.V.}\int_\mathbb R\varpi_1(\beta)\frac{-h_2-z_2(\beta)}{|h(\alpha)-z(\beta)|^2}d\beta, \end{equation} (see Remark 1 for the definition of $\mathcal{K}$). The identity $$ \int_{\mathbb R}\partial_\beta\log((A-z_1(\beta))^2+(B-z_2(\beta))^2)=0, $$ gives us $$ \frac{1}{2\pi}\text{P.V.}\int_\mathbb R (-\partial_\alpha z_2(\beta)) \frac{z_2(\alpha)-z_2(\beta)}{|z(\alpha)-z(\beta)|^2}d\beta=\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\partial_\alpha z_1(\beta) \frac{z_1(\alpha)-z_1(\beta)}{|z(\alpha)-z(\beta)|^2}d\beta, $$ and $$ \frac{1}{2\pi}\text{P.V.}\int_\mathbb R \partial_\alpha z_2(\beta) \frac{h_2+z_2(\beta)}{|h(\alpha)-z(\beta)|^2}d\beta=\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\partial_\alpha z_1(\beta) \frac{h_1(\alpha)-z_1(\beta)}{|h(\alpha)-z(\beta)|^2}d\beta. $$ Thus, \begin{multline} \label{IIeq7} \varpi_2(\alpha)=\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{\pi}\text{P.V.}\int_\mathbb R\partial_\alpha z_2(\beta)\frac{h_2+z_2(\beta)}{|h(\alpha)-z(\beta)|^2}d\beta\\ =\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{\pi}\text{P.V.}\int_\mathbb R\partial_\alpha z_1(\beta) \frac{h_1(\alpha)-z_1(\beta)}{|h(\alpha)-z(\beta)|^2}d\beta, \end{multline} and $$ BR(\varpi_1,z)z=\frac{-\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\frac{z_1(\alpha)-z_1(\beta)}{|z(\alpha)-z(\beta)|^2}\partial_\alpha z(\beta)d\beta $$ Due to the conservation of mass the curve $z$ is advected by the flow, but we can add any tangential term in the equation for the evolution of the interface without changing the shape of the resulting curve (see \cite{c-c-g10}), \emph{i.e.} we consider that the equation for the curve is $$ \partial_t z(\alpha)=v(\alpha)+c(\alpha,t)\partial_\alpha z(\alpha). $$ Taking $c(\alpha)=-v_1(\alpha)$, we conclude \begin{multline} \label{IIeq8} \partial_t z=\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\frac{z_1(\alpha)-z_1(\beta)}{|z(\alpha)-z(\beta)|^2}(\partial_\alpha z(\alpha)-\partial_\alpha z(\beta))d\beta\\ +\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\varpi_2(\beta)\frac{(z(\alpha)-h(\beta))^\perp}{|z(\alpha)-h(\beta)|^2}d\beta\\ +\partial_\alpha z(\alpha)\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\varpi_2(\beta)\frac{z_2(\alpha)+h_2}{|z(\alpha)-h(\beta)|^2}d\beta. \end{multline} By choosing this tangential term, if our initial datum can be parametrized as a graph, we have $\partial_t z_1=0.$ Therefore the parametrization as a graph propagates. Finally we conclude \eqref{IIeq9} as the evolution equation for the interface (which initially is a graph above the line $y\equiv-h_2$). We remark that the second vorticity \eqref{IIeq10A} can be written in equivalent ways \begin{eqnarray} \varpi_2(x)&=&\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{\pi}\text{P.V.}\int_\mathbb R\partial_x f(\beta)\frac{h_2+f(\beta)}{(x-\beta)^2+(-h_2-f(\beta))^2}d\beta\label{IIeq10}\\ &=&\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{\pi}\text{P.V.}\int_\mathbb R \frac{x-\beta}{(x-\beta)^2+(-h_2-f(\beta))^2}d\beta\label{IIeq10.b}\\ &=&\mathcal{K}\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\text{P.V.}\int_\mathbb R\partial_x\log ((x-\beta)^2+(-h_2-f(\beta))^2)d\beta.\nonumber \end{eqnarray} \begin{coment} Notice that in the case with different viscosities the expression for the amplitude of the vorticity located at the interface $z(\alpha)$ (see equation \eqref{IIeq5}) is no longer valid. Instead, we have $$ -\kappa^1(\rho^2-\rho^1)\partial_\alpha z_2(\alpha)=\left(\mu^2-\mu^1\right)\left(BR(\varpi_1,z)z+BR(\varpi_2,h)z\right)\cdot\partial_\alpha z(\alpha)+\left(\frac{\mu^2+\mu^1}{2}\right)\varpi_1. $$ To this integral equation, we add the equation \eqref{w2eq} or \eqref{IIeq7}. Thus, one needs to invert an operator. This is a rather delicate issue that is beyond the scope of this paper (see \cite{c-c-g10} for further details in the case $\kappa^1=\kappa^2$). \end{coment} \subsubsection{Assuming $S=\mathbb T\times\mathbb R$:} We have that \eqref{IIBS} is still valid, but now $\varpi_i$ are periodic functions and $z(\alpha+2k\pi)=z(\alpha)+(2k\pi,0)$. Using complex variables notation we have \begin{multline*} \bar{v}(\vec{x})=\frac{1}{2\pi i}\text{P.V.}\int_\mathbb R\frac{\varpi_1(\beta)}{\vec{x}-z(\beta)}d\beta+\frac{1}{2\pi i}\text{P.V.}\int_\mathbb R\frac{\varpi_2(\beta)}{\vec{x}-h(\beta)}d\beta\\ =\frac{1}{2\pi i}\left(\text{P.V.}\int_{-\pi}^{\pi}+\sum_{k\geq1}\left(\int_{(2k-1)\pi}^{(2k+1)\pi}+\int_{-(2k+1)\pi}^{-(2k-1)\pi}\right)\right)\frac{\varpi_1(\beta)}{\vec{x}-z(\beta)}+\frac{\varpi_2(\beta)}{\vec{x}-h(\beta)}d\beta. \end{multline*} Changing variables and using the identity $$ \frac{1}{z}+\sum_{k\geq1}\frac{2z}{z^2-(2k\pi)^2}=\frac{1}{2\tan(z/2)},\;\;\forall z\in\mathbb C, $$ we obtain $$ \bar{v}(\vec{x})=\frac{1}{4\pi i}\left(\text{P.V.}\int_\mathbb T\frac{\varpi_1(\beta)}{\tan((\vec{x}-z(\beta))/2)}d\beta + \text{P.V.}\int_\mathbb T\frac{\varpi_2(\beta)}{\tan((\vec{x}-h(\beta))/2)}d\beta\right). $$ Equivalently, \begin{multline*} v(\vec{x})=\frac{1}{4\pi} \left(\text{P.V.}\int_\mathbb T\frac{-\sinh(y-z_2(\beta))\varpi_1(\beta)d\beta}{\cosh(y-z_2(\beta))-\cos(x-z_1(\beta))}\right.\\ \left.+\text{P.V.}\int_\mathbb T\frac{-\sinh(y-h_2(\beta))\varpi_2(\beta)d\beta}{\cosh(y-h_2(\beta))-\cos(x-h_1(\beta))}\right)\\ +\frac{i}{4\pi}\left( \text{P.V.}\int_\mathbb T\frac{\sin(x-z_1(\beta))\varpi_1(\beta)d\beta}{\cosh(y-z_2(\beta))-\cos(x-z_1(\beta))}\right.\\ \left.+\text{P.V.}\int_\mathbb T\frac{\sin(x-h_1(\beta))\varpi_2(\beta)d\beta}{\cosh(y-h_2(\beta))-\cos(x-h_1(\beta))}\right). \end{multline*} Recall that \eqref{IIeq5} and \eqref{IIeq7} are still valid if $h(\alpha)=(\alpha,-h_2)$ for $0<h_2$ a fixed constant. We have $$ \int_{\mathbb T}\partial_\beta\log(\cosh(B-z_2(\beta))-\cos(A-z_1(\beta)))d\beta=0, $$ thus, the velocity in the curve when the correct tangential terms are added is \begin{multline}\label{IIeq11} \partial_t z(\alpha)=\frac{1}{4\pi} \left(\kappa^1(\rho^2-\rho^1)\text{P.V.}\int_\mathbb T\frac{\sin(z_1(\alpha)-z_1(\beta))(\partial_\alpha z(\alpha)-\partial_\alpha z(\beta))d\beta}{\cosh(z_2(\alpha)-z_2(\beta))-\cos(z_1(\alpha)-z_1(\beta))}\right.\\ \left.+(\partial_\alpha z_1(\alpha)-1)\text{P.V.}\int_\mathbb T\frac{\sinh(z_2(\alpha)+h_2)\varpi_2(\beta)d\beta}{\cosh(z_2(\alpha)+h_2)-\cos(z_1(\alpha)-h_1(\beta))}\right)\\ +\frac{i}{4\pi} \text{P.V.}\int_\mathbb T\frac{(\partial_\alpha z_2(\alpha)\sinh(z_2(\alpha)+h_2)+\sin(z_1(\alpha)-h_1(\beta)))\varpi_2(\beta)d\beta}{\cosh(z_2(\alpha)+h_2)-\cos(z_1(\alpha)-h_1(\beta))}. \end{multline} We can do the same in order to write $\varpi_2$ as an integral on the torus. \begin{multline}\label{IIeq12} \varpi_2(\alpha)=-2\mathcal{K}BR(\varpi_1,z)h\cdot(1,0)\\ =\frac{1}{2\pi}\mathcal{K} \text{P.V.}\int_\mathbb T\frac{\sinh(-h_2-z_2(\beta))\varpi_1(\beta)d\beta}{\cosh(-h_2-z_2(\beta))-\cos(h_1(\alpha)-z_1(\beta))}\\ =\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\mathcal{K} \text{P.V.}\int_\mathbb T\frac{\sinh(h_2+z_2(\beta))\partial_\alpha z_2(\beta)d\beta}{\cosh(-h_2-z_2(\beta))-\cos(h_1(\alpha)-z_1(\beta))}. \end{multline} If the initial datum can be parametrized as a graph the equation for the interface reduces to \eqref{IIeq13}, where the second vorticity amplitude \eqref{IIeq14A} can be written as \begin{eqnarray} \varpi_2(x)&=&\frac{1}{2\pi}\mathcal{K} \text{P.V.}\int_\mathbb T\frac{\sinh(-h_2-f(\beta))\varpi_1(\beta)d\beta}{\cosh(-h_2-f(\beta))-\cos(x-\beta)}\nonumber\\ &=&\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\mathcal{K} \text{P.V.}\int_\mathbb T\frac{\sinh(h_2+f(\beta))\partial_x f(\beta)d\beta}{\cosh(h_2+f(\beta))-\cos(x-\beta)}\label{IIeq14}\\ &=&\frac{\kappa^1(\rho^2-\rho^1)}{2\pi}\mathcal{K} \text{P.V.}\int_\mathbb T\frac{\sin(x-\beta)d\beta}{\cosh(h_2+f(\beta))-\cos(x-\beta)}.\label{IIeq14.1} \end{eqnarray} \subsection{Finite depth} Now we consider the bounded porous medium $\mathbb R\times (-\pi/2,\pi/2)$ (see Figure \ref{IIscheme}). This regime is equivalent to the case with more than two $\kappa^i$ because the boundaries can be understood as regions with $\kappa=0$. As before, $$ v(x,y)=\text{P.V.}\int_{\mathbb R}\varpi_1(\beta)BS(x,y,z_1(\beta),z_2(\beta))d\beta+\text{P.V.}\int_{\mathbb R}\varpi_2(\beta)BS(x,y,h_1(\beta),h_2(\beta))d\beta. $$ We assume that $h(\alpha)=(\alpha,-h_2)$ with $0<h_2<\pi/2$. We have that $\varpi_1$ is given by \eqref{IIeq5}. The main difference between the finite depth and the infinite depth is at the level of $\varpi_2$. As in the infinite depth case we have $$ 0=\left(\frac{1}{\kappa^2}-\frac{1}{\kappa^1}\right)\left(BR(\varpi_1,z)h+BR(\varpi_2,h)h\right)\cdot\partial_\alpha h(\alpha)+\left(\frac{1}{2\kappa^2}+\frac{1}{2\kappa^1}\right)\varpi_2, $$ where now $BR$ has the usual definition \eqref{IIeq2} in terms of $BS$ in expression \eqref{IIBSconf}. In the unbounded case we have an explicit expression for $\varpi_2$ \eqref{IIeq7} in terms of $z$ and $h$, but now we have a Fredholm integral equation of second kind: \begin{equation}\label{IIw2def} \varpi_2(\alpha)+\frac{\mathcal{K}}{2\pi}\;\text{P.V.}\int_\mathbb R\frac{\varpi_2(\beta)\sin(2h_2)}{\cosh(\alpha-\beta)+\cos(2h_2)}d\beta=-2\mathcal{K}BR(\varpi_1,z)h\cdot(1,0). \end{equation} After taking the Fourier transform, denoted by $\mathcal{F}(\cdot)(\zeta)$, and using some of its basic properties, we have \begin{equation*} \mathcal{F}(\varpi_2)(\zeta)\left(1+\frac{\mathcal{K}}{\sqrt{2\pi}}\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)\right)=-2\mathcal{K}\mathcal{F}(BR(\varpi_1,z)h\cdot(1,0))(\zeta). \end{equation*} We can solve the equation for $\varpi_2$ for any $|\mathcal{K}|<\delta(h_2)$ with \begin{equation}\label{IIdelta} \delta(h_2)=\min\left\{1,\frac{\sqrt{2\pi}}{\max_{\zeta}\left|\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)\right|}\right\}. \end{equation} We obtain \begin{multline}\label{IIw2defb} \varpi_2(\alpha)=-2\mathcal{K}BR(\varpi_1,z)h\cdot(1,0)\\ +\frac{2\mathcal{K}^2}{\sqrt{2\pi}}BR(\varpi_1,z)h\cdot(1,0)*\mathcal{F}^{-1}\left(\frac{\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}{1+\frac{\mathcal{K}}{\sqrt{2\pi}}\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}\right). \end{multline} Now we observe that if $s(\zeta)$ is a function in the Schwartz class, $\mathcal{S}$, such that $1+s(\zeta)>0$ we have that $$ \frac{s(\zeta)}{1+s(\zeta)}\in\mathcal{S}, $$ and we obtain $$ G_{h_2,\mathcal{K}}(x)=\mathcal{F}^{-1}\left(\frac{\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}{1+\frac{\mathcal{K}}{\sqrt{2\pi}}\mathcal{F}\left(\frac{\sin(2h_2)}{\cosh(x)+\cos(2h_2)}\right)(\zeta)}\right)\in\mathcal{S}. $$ Recall here that in order to obtain $\varpi_2$ we invert an integral operator. In general this is a delicate issue (compare with \cite{c-c-g10}), but with our choice of $h$ this point can be addressed in a simpler way. Using $$ \int_\mathbb R\partial_\beta\log\left(\cosh(x-z_1(\beta))\pm\cos(y\pm z_2(\beta))\right)d\beta=0, $$ and adding the correct tangential term, we obtain \begin{eqnarray} \partial_t z(\alpha)&=&\frac{\kappa^1(\rho^2-\rho^1)}{4\pi}\text{P.V.}\int_\mathbb R\frac{(\partial_\alpha z(\alpha)-\partial_\alpha z(\beta))\sinh(z_1(\alpha)-z_1(\beta))}{\cosh(z_1(\alpha)-z_1(\beta))-\cos(z_2(\alpha)-z_2(\beta))}d\beta\nonumber\\ &&+\frac{\kappa^1(\rho^2-\rho^1)}{4\pi}\text{P.V.}\int_\mathbb R\frac{(\partial_\alpha z_1(\alpha)-\partial_\alpha z_1(\beta),\partial_\alpha z_2(\alpha)+\partial_\alpha z_2(\beta))\sinh(z_1(\alpha)-z_1(\beta))}{\cosh(z_1(\alpha)-z_1(\beta))+\cos(z_2(\alpha)+z_2(\beta))}d\beta\nonumber\\ &&+\frac{1}{4\pi}\text{P.V.}\int_\mathbb R\varpi_2(\beta)BS(z_1(\alpha),z_2(\alpha),\beta,-h_2)d\beta\nonumber\\ &&+\frac{\partial_\alpha z(\alpha)}{4\pi}\text{P.V.}\int_\mathbb R\varpi_2(\beta)\frac{\sin(z_2(\alpha)+h_2)}{\cosh(z_1(\alpha)-\beta)-\cos(z_2(\alpha)+h_2)}d\beta\nonumber\\ &&+\frac{\partial_\alpha z(\alpha)}{4\pi}\text{P.V.}\int_\mathbb R\varpi_2(\beta)\frac{\sin(z_2(\alpha)-h_2)}{\cosh(z_1(\alpha)-\beta)+\cos(z_2(\alpha)-h_2)}d\beta\label{IIeqv}. \end{eqnarray} If the initial curve can be parametrized as a graph the equation reduces to \eqref{IIeqv2} where $\varpi_2$ is defined in \eqref{IIw2defc}. \begin{coment} If $h_2=\pi/4$ by an explicit computation we obtain $\delta(\pi/4)=1$, thus, any $\mathcal{K}$ is valid. Moreover, we have tested numerically that the same remains valid for any $0<h_2<\pi/2$, so \eqref{IIw2defc} would be correct for any $\mathcal{K}$. \end{coment} \section{Well-posedness in Sobolev spaces}\label{IIsec3} \subsection{Energy balance for the $L^2$ norm}\label{IIsec3.0} Here we obtain an energy balance inequality for the $L^2$ norm of the solution of equation \eqref{IIeqv2}. We define $\Omega^1=\{(x,y),f(x,t)<y<\pi/2\}$, $\Omega^2=\{(x,y),-h_2<y<f(x,t))\}$ and $\Omega^3=\{(x,y),-\pi/2<y<-h_2\}$. \begin{lem}For every $0<\kappa^ 1,\kappa^ 2$ the smooth solutions of \eqref{IIeqv2} in the stable regime, \emph{i.e.} $\rho^2>\rho^1$, case verifies \begin{equation}\label{IIEB2} \|f(t)\|^2_{L^2(\mathbb R)}+\int_0^t\frac{\|v\|_{L^2(\mathbb R\times(-h_2,\pi/2))}^2}{\kappa^1(\rho^2-\rho^1)}+\frac{\|v\|^2_{L^2(\mathbb R\times(-\pi/2,-h_2))}}{\kappa^2(\rho^2-\rho^1)}ds=\|f_0\|^2_{L^2(\mathbb R)}. \end{equation} \end{lem} \begin{proof} We define the potentials $$ \phi^1(x,y,t)=\kappa^1(p(x,y,t)+\rho^1y),\;\;\text{if }(x,y)\in\Omega^1, $$ $$ \phi^2(x,y,t)=\kappa^1(p(x,y,t)+\rho^2y),\;\;\text{if }(x,y)\in\Omega^2, $$ $$ \phi^3(x,y,t)=\kappa^2(p(x,y,t)+\rho^2y),\;\;\text{if }(x,y)\in\Omega^3. $$ We have $v^i=-\nabla\phi^i$ in each subdomain $S^i$. Since the velocity is incompressible we have $$ 0=\int_{\Omega^i}\Delta\phi^i\phi^i dxdy=-\int_{\Omega^i}|v^i|^2dxdy+\int_{\partial \Omega^i}\phi^i\partial_n\phi^i ds. $$ Moreover, the normal component of the velocity is continuous through the interface $(x,f(x))$ and the line where permeability changes $(x,-h_2)$. Using the impermeable boundary conditions, we only need to integrate over the curve $(x,f(x,t))$ and $(x,-h_2)$. Indeed, we have \begin{equation}\label{IIMP1} 0=-\int_{\Omega^1}|v^1|^2dxdy+\kappa^1\int_\mathbb R (p(x,f(x,t),t)+\rho^1f(x,t))(-v(x,f(x,t),t)\cdot(\partial_x f(x,t),-1))dx, \end{equation} \begin{multline}\label{IIMP2} 0=-\int_{\Omega^2}|v^2|^2dxdy+\kappa^1\int_\mathbb R (p(x,f(x,t),t)+\rho^2f(x,t))(-v(x,f(x,t),t)\cdot(-\partial_x f(x,t),1))dx\\ +\kappa^1\int_\mathbb R (p(x,-h_2,t)-\rho^2h_2)(-v(x,-h_2,t)\cdot(0,-1))dx, \end{multline} \begin{equation}\label{IIMP3} 0=-\int_{\Omega^3}|v^3|^2dxdy+\kappa^2\int_\mathbb R (p(x,-h_2,t)-\rho^2h_2)(-v(x,-h_2,t)\cdot(0,1))dx. \end{equation} Inserting \eqref{IIMP3} in \eqref{IIMP2} we get \begin{multline}\label{IIMP2b} 0=-\int_{\Omega^2}|v^2|^2dxdy-\frac{\kappa^1}{\kappa^ 2}\int_{\Omega^3}|v^3|^2dxdy\\ +\kappa^1\int_\mathbb R (p(x,f(x,t),t)+\rho^2f(x,t))(-v(x,f(x,t),t)\cdot(-\partial_x f(x,t),1))dx, \end{multline} Thus, summing \eqref{IIMP2b} and \eqref{IIMP1} together and using the continuity of the pressure and the velocity in the normal direction, we obtain \begin{equation}\label{IIEB} \int_{\Omega^ 1\cup \Omega^2}|v|^2dxdy+\frac{\kappa^1}{\kappa^ 2}\int_{\Omega^3}|v|^2dxdy=\kappa^1\int_\mathbb R (\rho^2-\rho^1)f(x,t)(-\partial_t f(x,t))dx. \end{equation} Integrating in time we get the desired result \eqref{IIEB2}. \end{proof} \subsection{Well-posedness for the infinite depth case} Let $\Omega$ be the spatial domain considered, \emph{i.e.} $\Omega=\mathbb R$ or $\Omega=\mathbb T$. In this section we prove the short time existence of classical solution for both spatial domains. We have the following result: \begin{teo} Consider $0<h_2$ a fixed constant and the initial datum $f_0(x)=f(x,0)\in H^k(\Omega)$, $k\geq3$, such that $-h_2<\min_x f_0(x)$. Then, if the Rayleigh-Taylor condition is satisfied, \emph{i.e.} $\rho^2-\rho^1>0$, there exists an unique classical solution of \eqref{IIeq9} $f\in C([0,T],H^k(\Omega))$ where $T=T(f_0)$. Moreover, we have $f\in C^1([0,T],C(\Omega))\cap C([0,T],C^2(\Omega)).$ \label{IIteo1} \end{teo} \begin{proof} We prove the result in the case $\Omega=\mathbb R$, being the case $\Omega=\mathbb T$ similar. Let us consider the usual Sobolev space $H^3(\mathbb R)$ endowed with the norm $$ \|f\|_{H^3}=\|f\|_{L^2}+\|\Lambda^3 f\|_{L^2}, $$ where $\Lambda=\sqrt{-\Delta}$. Define the energy \begin{equation}\label{IIeq15} E[f]:=\|f\|_{H^3}+\|d^h[f]\|_{L^\infty}, \end{equation} with \begin{equation}\label{IIeq16} d^h[f](x,\beta)=\frac{1}{(x-\beta)^2+(f(x)+h_2)^2}. \end{equation} To use the classical energy method we need \emph{a priori} estimates. To simplify notation we drop the physical parameters present in the problem by considering $\kappa^1(\rho^2-\rho^1)=2\pi$ and $\mathcal{K}=\frac{1}{2}$. The sign of the difference between the permeabilities will not be important to obtain local existence. We denote $c$ a constant that can changes from one line to another. \textbf{Estimates on $\|\varpi_2\|_{H^3}$:} Given $f(x)$ such that $E[f]<\infty$ we consider $\varpi_2$ as defined in \eqref{IIeq10}. Then we have that $\|\varpi_2\|_{H^3}\leq c(E[f]+1)^k$ for some constants $c,k$. We proceed now to prove this claim. We start with the $L^2$ norm. Changing variables in \eqref{IIeq10} we have \begin{multline*} \|\varpi_2\|^2_{L^2}\leq c\left\|\text{P.V.}\int_{B(0,1)}\frac{\partial_x f(x-\beta)(h_2+f(x-\beta))}{\beta^2+(h_2+f(x-\beta))^2}d\beta\right\|_{L^2}^2\\ +c\left\|\text{P.V.}\int_{B^c(0,1)}\frac{\partial_x f(x-\beta)(h_2+f(x-\beta))}{\beta^2+(h_2+f(x-\beta))^2}d\beta\right\|_{L^2}^2\\ =A_1+A_2.\qquad\qquad\qquad\qquad \end{multline*} The inner term, $A_1,$ can be bounded as follows \begin{multline*} A_1=\int_\mathbb R\text{P.V.}\int_{B(0,1)}\frac{\partial_x f(x-\beta)(h_2+f(x-\beta))}{\beta^2+(h_2+f(x-\beta))^2}d\beta dx\\ \times\text{P.V.}\int_{B(0,1)}\frac{\partial_x f(x-\xi)(h_2+f(x-\xi))}{\xi^2+(h_2+f(x-\xi))^2}d\xi dx\\ \leq c\|d^h[f]\|_{L^\infty}^2(1+\|f\|_{L^\infty})^2\|\partial_x f\|_{L^2}^2. \end{multline*} In the last inequality we have used Cauchy-Schwartz inequality and Tonelli's Theorem. For the outer part we have \begin{multline*} A_2=\int_\mathbb R\text{P.V.}\int_{B^c(0,1)}\frac{\partial_x f(x-\beta)(h_2+f(x-\beta))}{\beta^2+(h_2+f(x-\beta))^2}d\beta dx\\ \times\text{P.V.}\int_{B^c(0,1)}\frac{\partial_x f(x-\xi)(h_2+f(x-\xi))}{\xi^2+(h_2+f(x-\xi))^2}d\xi dx\\ \leq c(1+\|f\|_{L^\infty})^2\|\partial_x f\|_{L^2}^2, \end{multline*} where we have used that $\int_{1}^\infty\frac{d\beta}{\beta^2}<\infty$ and Cauchy-Schwartz inequality. We change variables in \eqref{IIeq10.b} to obtain $$ \varpi_2(x)=\text{P.V.}\int_\mathbb R \frac{\beta}{\beta^2+(h_2+f(x-\beta))^2}d\beta. $$ Now it is clear that $\varpi_2$ is at the level of $f$ in terms of regularity and the inequality follows using the same techniques. Using Sobolev embedding we conclude this step. \textbf{Estimates on $\|d^h[f]\|_{L^\infty}$:} The first integral in \eqref{IIeq9} can be bounded as follows $$ I_1\leq\left\|\text{P.V.}\int_\mathbb R\frac{(x-\beta)(\partial_x f(x)-\partial_x f(\beta))}{(x-\beta)^2+(f(x)-f(\beta))^2}d\beta\right\|_{L^\infty}\leq c(E[f]+1)^k, $$ for some positive and finite $k$. The new term is the second integral in \eqref{IIeq9}. \begin{multline*} I_2\leq\left\|\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(x-\beta)(\beta+\partial_x f(x)(f(x)+h_2))}{\beta^2+(f(x)+h_2)^2}d\beta\right\|_{L^\infty}\\ \leq \left\|\frac{1}{2\pi}\text{P.V.}\int_{B(0,1)}d\beta\right\|_{L^\infty} +\left\|\frac{1}{2\pi}\text{P.V.}\int_{B^c(0,1)}d\beta\right\|_{L^\infty}= A_1+A_2. \end{multline*} Easily we have $$ A_1\leq c\|\varpi_2\|_{L^\infty}\|d^h[f]\|_{L^\infty}(1+\|\partial_x f\|_{L^\infty}(\|f\|_{L^\infty}+1)). $$ We split $A_2=B_1+B_2$ \begin{multline*} B_1=\frac{1}{2\pi}\text{P.V.}\int_{B^c(0,1)}\frac{\varpi_2(x-\beta)\beta}{\beta^2+(f(x)+h_2)^2}\pm\frac{\varpi_2(x-\beta)\beta}{\beta^2}d\beta\\ \leq c\|\varpi_2\|_{L^\infty}(\|f\|_{L^\infty}+1)^2+c\|H\varpi_2\|_{L^\infty}+c\|\partial_x \varpi_2\|_{L^\infty}, \end{multline*} where $H$ denotes the Hilbert transform. Now we conclude the desired bound using the previous estimate on $\|\varpi_2\|_{H^3}$ and Sobolev embedding. The second term can be bounded as $$ B_2=\frac{1}{2\pi}\text{P.V.}\int_{B^c(0,1)}\frac{\varpi_2(x-\beta)\partial_x f(x)(f(x)+h_2)}{\beta^2+(f(x)+h_2)^2}d\beta\leq c\|\varpi_2\|_{L^\infty}(\|f\|_{L^\infty}+1)\|\partial_x f\|_{L^\infty}. $$ We obtain the following useful estimate \begin{equation}\label{IIdtf} \|\partial_t f\|_{L^\infty}\leq c(E[f]+1)^k. \end{equation} We have $$ \frac{d}{dt}d^h[f]=\frac{-\partial_t f(x)2(f(x)+h_2)}{(\beta^2+(f(x)+h_2)^2)^2}\leq cd^h[f]\|d^h[f]\|_{L^\infty}(\|f\|_{L^\infty}+1)\|\partial_t f\|_{L^\infty}. $$ Thus, integrating in time and using \eqref{IIdtf}, $$ \|d^h[f](t+h)\|_{L^\infty}\leq\|d^h[f](t)\|_{L^\infty}e^{c\int_t^{t+h}(E[f]+1)^k}, $$ and we conclude this step $$ \frac{d}{dt}\|d^h[f]\|_{L^\infty}=\lim_{h\rightarrow0}\frac{\|d^h[f](t+h)\|_{L^\infty}-\|d^h[f](t)\|_{L^\infty}}{h}\leq c(E[f]+1)^k. $$ \textbf{Estimates on $\|\partial_x^3 f\|_{L^2}$:} As before, the bound for the term coming from the first integral in \eqref{IIeq9} can be obtained as in \cite{c-g07}, so it only remains the term coming from the second integral. We have $$ I_2=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\partial_x^3\left(\frac{\varpi_2(x-\beta)(\beta+\partial_x f(x)(f(x)+h_2))}{\beta^2+(f(x)+h_2)^2}\right)d\beta dx. $$ For the sake of brevity we only bound the terms with higher order, being the remaining terms analogous. We have $$ I_2=J_3+J_4+J_5+J_6+J_7+\text{l.o.t.}, $$ with $$ J_3=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\frac{\partial_x^3\varpi_2(x-\beta)\beta}{\beta^2+(f(x)+h_2)^2}d\beta dx, $$ $$ J_4=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\frac{\partial_x^3\varpi_2(x-\beta)\partial_x f(x)(f(x)+h_2)}{\beta^2+(f(x)+h_2)^2}d\beta dx, $$ $$ J_5=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\frac{2\varpi_2(x-\beta)(\beta+\partial_x f(x)(f(x)+h_2))(-f(x)-h_2)\partial_x^3 f(x)}{(\beta^2+(f(x)+h_2)^2)^2}d\beta dx, $$ $$ J_6=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\frac{\varpi_2(x-\beta)(f(x)+h_2)\partial_x^4 f(x)}{\beta^2+(f(x)+h_2)^2}d\beta dx, $$ and $$ J_7=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\frac{4\varpi_2(x-\beta)\partial_x f(x)\partial_x^3 f(x)}{\beta^2+(f(x)+h_2)^2}d\beta dx. $$ In order to bound $J_3$ we use the symmetries in the formulae ($\partial_x=-\partial_\beta$) and we integrate by parts: \begin{multline*} J_3=\frac{1}{2\pi}\int_\mathbb R \partial_x^3 f(x)\text{P.V.}\int_{\mathbb R}\partial_x^2\varpi_2(x-\beta)\partial_\beta\left(\frac{\beta}{\beta^2+(f(x)+h_2)^2}\right)d\beta dx\\ \leq c\|\partial_x^3 f\|_{L^2}\|\partial_x^2\varpi_2\|_{L^2}(\|d^h[f]\|_{L^\infty}^2+\|d^h[f]\|_{L^\infty}+1).\hspace{4cm} \end{multline*} In $J_4$ we use Cauchy-Schwartz inequality to obtain $$ J_4\leq c(\|d^h[f]\|_{L^\infty}+1)\|\partial_x^3 f\|_{L^2}\|\partial_x^3 \varpi_2\|_{L^2}\|\partial_x f\|_{L^\infty}(\|f\|_{L^\infty}+h_2) $$ The bounds for $J_5$ and $J_7$ are similar: $$ J_5\leq c(\|d^h[f]\|_{L^\infty}^2+1)\|\partial_x^3 f\|^2_{L^2}\|\varpi_2\|_{L^\infty}(1+\|\partial_x f\|_{L^\infty}(\|f\|_{L^\infty}+h_2))(\|f\|_{L^\infty}+h_2), $$ $$ J_7\leq c(\|d^h[f]\|_{L^\infty}+1)\|\partial_x^3 f\|^2_{L^2}\|\varpi_2\|_{L^\infty}\|\partial_x f\|_{L^\infty}. $$ Finally, we integrate by parts in $J_6$ and we get \begin{multline*} J_6\leq c\|\partial_x^3 f\|^2_{L^2}(\|d^h[f]\|_{L^\infty}+1)\left(\|\partial_x\varpi_2\|_{L^\infty}(\|f\|_{L^\infty}+1)+\|\varpi_2\|_{L^\infty}\|\partial_x f\|_{L^\infty}\right)\\ +c\|\partial_x^3 f\|^2_{L^2}(\|d^h[f]\|_{L^\infty}^2+1)\|\varpi_2\|_{L^\infty}\|\partial_x f\|_{L^\infty}(\|f\|_{L^\infty}+1)^2. \end{multline*} As a conclusion, we obtain $$ \frac{d}{dt}\|\partial_x^3 f\|_{L^2}\leq c(E[f]+1)^k. $$ Putting all the estimates together we get the desired bound for the energy: \begin{equation}\label{IIenergybound} \frac{d}{dt}E[f]\leq c(E[f]+1)^k. \end{equation} \textbf{Regularization:} This step is classical, so, we only sketch this part (see \cite{bertozzi-Majda} for the details). We regularize the problem and we show that the regularized problems have a solution using Picard's Theorem on a ball in $H^3$. Using the previous energy estimates and the fact that the initial energy is finite, these solutions have the same time of existence ($T$ depending only on the initial datum) and we can show that they are a Cauchy sequence in $C([0,T],L^2)$. From here we obtain $f\in C([0,T],H^s(\Omega))\cap L^\infty([0,T],H^3(\Omega))$ where $T=T(f_0)$ and $0<s<3$, a solution to \eqref{IIeq9} as the limit of these regularized solutions. The continuity of the strongest norm $H^3$ for positive times follows from the parabolic character of the equation. The continuity of $\|f(t)\|_{H^3}$ at $t=0$ follows from the fact that $f(t)\rightharpoonup f_0$ in $H^3$ and from the energy estimates. \textbf{Uniqueness:} Only remains to show that the solution is unique. Let us suppose that for the same initial datum $f_0$ there are two smooth solutions $f^1$ and $f^2$ with finite energy as defined in \eqref{IIeq15} and consider $f=f^1-f^2$. Following the same ideas as in the energy estimates we obtain $$ \frac{d}{dt}\|f\|_{L^2}\leq c(f_0,E[f^1],E[f^2])\|f\|_{L^2}. $$ Now we conclude using Gronwall inequality. \end{proof} \subsection{Well-posedness for the finite depth case} In this section we prove the short time existence of classical solution in the case where the depth is finite. We have the following result: \begin{teo} Consider $0<h_2<\pi/2$ a constant and $f_0(x)=f(x,0)\in H^k(\mathbb R)$, $k\geq3$, an initial datum such that $\|f_0\|_{L^\infty}<\pi/2$ and $-h_2<\min_x f_0(x)$. Then, if the Rayleigh-Taylor condition is satisfied, \emph{i.e.} $\rho^2-\rho^1>0$, there exists an unique classical solution of \eqref{IIeqv2} $f\in C([0,T],H^k(\mathbb R))$ where $T=T(f_0)$. Moreover, we have $f\in C^1([0,T],C(\mathbb R))\cap C([0,T],C^2(\mathbb R)).$ \label{IIteo3} \end{teo} \begin{proof} Let us consider the usual Sobolev space $H^3(\mathbb R)$, being the other cases analogous, and define the energy \begin{equation}\label{IIeq23} E[f]=\|f\|_{H^3}+\|d^h[f]\|_{L^\infty}+\|d[f]\|_{L^\infty}, \end{equation} with \begin{equation}\label{IIeq24} d^h[f](x,\beta)=\frac{1}{\cosh(x-\beta)-\cos(f(x)+h_2)}, \end{equation} and \begin{equation}\label{IIeq25} d[f](x,\beta)=\frac{1}{\cosh(x-\beta)+\cos(f(x)+f(\beta))}. \end{equation} We note that $d^h[f]$ represents the distance between $f$ and $h$ and $d[f]$ the distance between $f$ and the boundaries. To simplify notation we drop the physical parameters present in the problem by considering $\kappa^1(\rho^2-\rho^1)=4\pi$ and $\mathcal{K}=\frac{1}{2}$. Again, the sign of the difference between the permeabilities will not be important to obtain local existence. We write \eqref{IIeqv2} as $\partial_t f= I_1+I_2+I_3+I_4$, being $I_1,I_2$ the integrals corresponding $\varpi_1$ and $I_3,I_4$ the integrals involving $\varpi_2$. We denote $c$ a constant that can changes from one line to another. \textbf{Estimate on $\|\varpi_2\|_{H^3}$:} Given $f(x)$ such that $E[f]<\infty$ and consider $\varpi_2$ as defined in \eqref{IIw2defc}. Then we have that $\|\varpi_2\|_{H^3}\leq c(E[f]+1)^k$. We need to bound $\|J_1\|_{H^3}$ and $\|J_2\|_{H^3}$ with $$ J_1=\text{P.V.}\int_\mathbb R\partial_x f(x-\beta)\frac{\sin(h_2+f(x-\beta))}{\cosh(\beta)-\cos(h_2+f(x-\beta))}d\beta $$ $$ J_2=-\text{P.V.}\int_\mathbb R\partial_x f(x-\beta)\frac{\sin(-h_2+f(x-\beta))}{\cosh(\beta)+\cos(-h_2+f(x-\beta))}d\beta. $$ We have \begin{multline*} \|J_1\|_{L^2}\leq\left\|\text{P.V.}\int_{B(0,1)}\frac{\partial_x f(x-\beta)\sin(h_2+f(x-\beta))}{\cosh(\beta)-\cos(h_2+f(x-\beta))}d\beta\right\|_{L^2}\\ +\left\|\text{P.V.}\int_{B^c(0,1)}\frac{\partial_x f(x-\beta)\sin(h_2+f(x-\beta))}{\cosh(\beta)-\cos(h_2+f(x-\beta))}d\beta\right\|_{L^2}\\ \leq c\|\partial_x f\|_{L^2}\|d^h[f]\|_{L^\infty}+c\|\partial_x f\|_{L^2},\hspace{3cm} \end{multline*} where we have used Tonelli's Theorem and Cauchy-Schwartz inequality. Recall that $f-h_2\in \left(-2h_2,\frac{\pi}{2}-h_2\right)$, thus $$ \frac{1}{\cosh(x-\beta)+\cos(f(x)-h_2)}<\frac{1}{\cosh(x-\beta)-c(h_2)}, $$ and the kernel corresponding to $\varpi_2$ can not be singular and we also obtain $$ \|J_2\|_{L^2}\leq c\|\partial_x f\|_{L^2}. $$ Now, as $G_{h_2,\mathcal{K}}\in\mathcal{S}$, we can use the Young's inequality for the convolution terms obtaining bounds with an universal constant depending on $h_2$ and $\mathcal{K}$. Indeed, we have $$ \|G_{h_2,\mathcal{K}}*J_i\|_{L^2}\leq c\|J_i\|_{L^2}, $$ and we obtain $$ \|\varpi_2\|_{L^2}\leq c(E[f]+1)^k. $$ Now we observe that $$ J_1=\text{P.V.}\int_\mathbb R\frac{\sinh(\beta)}{\cosh(\beta)-\cos(h_2+f(x-\beta))}d\beta,\;\;J_2=\text{P.V.}\int_\mathbb R\frac{\sinh(\beta)}{\cosh(\beta)+\cos(-h_2+f(x-\beta))}d\beta, $$ and we obtain $\|\partial_x^3 J_i\|_{L^2}\leq c(E[f]+1)^k$. Using Young inequality we conclude $$ \|\varpi_2\|_{H^3}\leq c(E[f]+1)^k. $$ \textbf{Estimates on $\|d^h[f]\|_{L^\infty}$ and $\|d[f]\|_{L^\infty}$:} The integrals corresponding to $\varpi_1$ in \eqref{IIeqv2} can be bounded (see \cite{CGO}) as $$ |I_1+I_2|\leq c(E[f]+1)^k. $$ The new terms are the integrals $I_3$ and $I_4$, those involving $\varpi_2$ in \eqref{IIeqv2}. We have, when splitted accordingly to the decay at infinity, $$ I_3+I_4=J_3+J_4, $$ where \begin{multline*} |J_3|\leq\left\|\frac{1}{4\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(x-\beta)\sinh(\beta)}{\cosh(\beta)-\cos(f(x)+h_2)}-\frac{\varpi_2(x-\beta)\sinh(\beta)}{\cosh(\beta)+\cos(f(x)-h_2)}d\beta\right\|_{L^\infty}\\ \leq c\|\varpi_2\|_{L^\infty}\left(\|d^h[f]\|_{L^\infty}+1\right), \end{multline*} and \begin{multline*} |J_4|\leq\left\|\frac{1}{4\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(x-\beta)\partial_x f(x)\sin(f(x)+h_2)}{\cosh(\beta)-\cos(f(x)+h_2)}+\frac{\varpi_2(x-\beta)\partial_x f(x)\sin(f(x)-h_2)}{\cosh(\beta)+\cos(f(x)-h_2)}d\beta\right\|_{L^\infty}\\ \leq c\|\varpi_2\|_{L^\infty}\|\partial_x f\|_{L^\infty}\left(\|d^h[f]\|_{L^\infty}+1\right). \end{multline*} We conclude the following useful estimate \begin{equation}\label{IIdtfFD} \|\partial_t f\|_{L^\infty}\leq c(E[f]+1)^k. \end{equation} We have $$ \frac{d}{dt}d^h[f]=-\frac{\sin(f(x)+h_2)\partial_t f(x)}{(\cosh(x-\beta)-\cos(f(x)+h_2))^2}\leq d^h[f]\|d^h[f]\|_{L^\infty}\|\partial_t f\|_{L^\infty}. $$ Thus, using \eqref{IIdtfFD} and integrating in time, we obtain the desired bound for $d^h[f]$: $$ \frac{d}{dt}\|d^h[f]\|_{L^\infty}=\lim_{h\rightarrow0}\frac{\|d^h[f](t+h)\|_{L^\infty}-\|d^h[f](t)\|_{L^\infty}}{h}\leq c(E[f]+1)^k. $$ To obtain the corresponding bound for $d[f]$ we proceed in the same way and we use \eqref{IIdtfFD} (see \cite{CGO} for the details) \textbf{Estimates on $\|\partial_x^3 f\|_{L^2}$:} As before, see \cite{CGO} for the details concerning the terms coming from $\varpi_1$ in \eqref{IIeqv2}. It only remains the terms coming from $\varpi_2$: \begin{multline*} I=\int_\mathbb R \text{P.V.}\int_\mathbb R \partial_x^3f(x)\partial_x^3\left(\frac{\varpi_2(\beta)(\sinh(x-\beta)+\partial_x f(x)\sin(f(x)+h_2))}{\cosh(x-\beta)-\cos(f(x)+h_2)}\right.\\ \left.+\frac{\varpi_2(\beta)(-\sinh(x-\beta)+\partial_x f(x)\sin(f(x)-h_2))}{\cosh(x-\beta)+\cos(f(x)-h_2)}\right)d\beta dx. \end{multline*} We split $$ I=J_7+J_8+J_9+\text{l.o.t.}. $$ The lower order terms (l.o.t.) can be obtained in a similar way, so we only study the terms $J_i$. We have \begin{multline*} J_7\leq \int_\mathbb R \text{P.V.}\int_\mathbb R \frac{\partial_x^3 f(x)\partial_x^3\varpi_2(x-\beta)\sinh(\beta)}{\cosh(\beta)-\cos(f(x)+h_2)}-\frac{\partial_x^3f(x)\partial_x^3\varpi_2(x-\beta)\sinh(\beta)}{\cosh(\beta)+\cos(f(x)-h_2)}d\beta dx\\ \leq c\|\partial_x^3 f\|_{L^2}\|\partial_x^3 \varpi_2\|_{L^2}(\|d^h[f]+1\|), \end{multline*} \begin{multline*} J_8\leq \int_\mathbb R \text{P.V.}\int_\mathbb R \frac{\partial_x^3 f(x)\partial_x^3\varpi_2(x-\beta)\partial_x f(x)\sin(f(x)+h_2)}{\cosh(\beta)-\cos(f(x)+h_2)}\\ -\frac{\partial_x^3f(x)\partial_x^3\varpi_2(x-\beta)\partial_x f(x)\sin(f(x)+h_2)}{\cosh(\beta)+\cos(f(x)-h_2)}d\beta dx\\ \leq c\|\partial_x^3 f\|_{L^2}\|\partial_x^3 \varpi_2\|_{L^2}\|\partial_x f\|_{L^\infty}(\|d^h[f]+1\|). \end{multline*} The term $J_9$ is given by \begin{multline*} J_9=\frac{1}{2}\int_\mathbb R \text{P.V.}\int_\mathbb R \partial_x(\partial_x^3f(x))^2\left(\frac{\varpi_2(\beta)\sin(f(x)+h_2)}{\cosh(x-\beta)-\cos(f(x)+h_2)}\right.\\ \left.+\frac{\varpi_2(\beta)\sin(f(x)-h_2)}{\cosh(x-\beta)+\cos(f(x)-h_2)}\right)d\beta dx. \end{multline*} Integrating by parts \begin{multline*} |J_9| \leq c\|\partial_x^3 f\|_{L^2}(\|d^h[f]\|_{L^\infty}+1)(\|\partial_x \varpi_2\|_{L^\infty}+\|\varpi_2\|_{L^\infty}\|\partial_x f\|_{L^\infty})\\ +c\|\partial_x^3 f\|_{L^2}(\|d^h[f]\|^2_{L^\infty}+1)\|\varpi_2\|_{L^\infty}(1+\|\partial_x f\|_{L^\infty}) \end{multline*} \textbf{Regularization and uniqueness:} These steps follow the same lines as in Theorem \ref{IIteo1}. This concludes the result. \end{proof} \section{Numerical simulations}\label{IIsec5} In this section we perform numerical simulations to better understand the role of $\varpi_2$. We consider equation \eqref{IIeq13} where $\kappa^1=1$, $\rho^2-\rho^1=4\pi$ and $h_2=\pi/2$. For each initial datum we approximate the solution of \eqref{IIeq13} corresponding to different $\mathcal{K}$. Indeed, we take different $\kappa^2$ to get $\mathcal{K}=\frac{-999}{1001},\frac{-1}{3},0,\frac{1}{3}$ and $\frac{999}{1001}$. To perform the simulations we follow the ideas in \cite{c-g-o08}. The interface is approximated using cubic splines with $N$ spatial nodes. The spatial operator is approximated with Lobatto quadrature (using the function \emph{quadl} in Matlab). Then, three different integrals appear for a fixed node $x_i$. The integral between $x_{i-1}$ and $x_i$, the integral between $x_i$ and $x_{i+1}$ and the nonsingular ones. In the two first integrals we use Taylor theorem to remove the zeros present in the integrand. In the nonsingular integrals the integrand is made explicit using the splines. We use a classical explicit Runge-Kutta method of order 4 to integrate in time. In the simulations we take $N=120$ and $dt=10^{-3}$. The case 1 (see Figure \ref{IIcase1} and \ref{IIcase1b}) approximates the solution corresponding to the initial datum $$ f_0(x)=-\left(\frac{\pi}{2}-0.000001\right)e^{-x^{12}}. $$ \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{minf1.eps} \end{center} \caption{Evolution of $-\|f\|_{L^\infty}$ for different $\mathcal{K}$ in case 1.} \label{IIcase1} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{df1.eps} \end{center} \caption{Evolution of $\|\partial_x f\|_{L^\infty}$ for different $\mathcal{K}$ in case 1.} \label{IIcase1b} \end{figure} The case 2 (see Figure \ref{IIcase2} and \ref{IIcase2b}) approximates the solution corresponding to the initial datum $$ f_0(x)=-\left(\frac{\pi}{2}-0.000001\right)\cos(x^2). $$ \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{minf2.eps} \end{center} \caption{Evolution of $-\|f\|_{L^\infty}$ for different $\mathcal{K}$ in case 2.} \label{IIcase2} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{df2.eps} \end{center} \caption{Evolution of $\|\partial_x f\|_{L^\infty}$ for different $\mathcal{K}$ in case 2.} \label{IIcase2b} \end{figure} The case 3 (see Figure \ref{IIcase3} and \ref{IIcase3b}) approximates the solution corresponding to the initial datum $$ f_0(x)=-\left(\frac{\pi}{2}-0.000001\right)e^{-(x-2)^{12}}-\left(\frac{\pi}{2}-0.000001\right)e^{-(x+2)^{12}}+e^{-x^2}\cos^2(x). $$ \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{minf3.eps} \end{center} \caption{Evolution of $-\|f\|_{L^\infty}$ for different $\mathcal{K}$ in case 3.} \label{IIcase3} \end{figure} \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{df3.eps} \end{center} \caption{Evolution of $\|\partial_x f\|_{L^\infty}$ for different $\mathcal{K}$ in case 3.} \label{IIcase3b} \end{figure} In these simulations we observe that $\|f\|_{C^1}$ decays but rather differently depending on $\mathcal{K}$. If $\mathcal{K}<0$ the decay of $\|f\|_{L^\infty}$ is faster when compared with the case $\mathcal{K}=0$. In the case where $\mathcal{K}>0$ the term corresponding to $\varpi_2$ slows down the decay of $\|f\|_{L^\infty}$ but we observe still a decay. Particularly, we observe that if $\mathcal{K}\approx 1$ ($\kappa^2\approx 0$) the decay is initially almost zero and then slowly increases. When the evolution of $\|\partial_x f\|_{L^\infty}$ is considered the situation is reversed. Now the simulations corresponding to $\mathcal{K}>0$ have the faster decay. With these result we can not define a \emph{stable} regime for $\mathcal{K}$ in which the evolution would be \emph{smoother}. Recall that we know that there is not any hypothesis on the sign or size of $\mathcal{K}$ to ensure the existence (see Theorem \ref{IIteo1} and \ref{IIteo3}). \section{Turning waves}\label{IIsec4} In this section we prove finite time singularities for equations \eqref{IIeq9}, \eqref{IIeq13} and \eqref{IIeqv2}. These singularities mean that the curve turns over or, equivalently, in finite time they can not be parametrized as graphs. The proof of turning waves follows the steps and ideas in \cite{ccfgl} for the homogeneus infinitely deep case where here we have to deal with the difficulties coming from the boundaries and the delta coming from the jump in the permeabilities. \subsection{Infinite depth}\label{IIsec4.1} Let $\Omega$ be the spatial domain considered, \emph{i.e.} $\Omega=\mathbb R$ or $\Omega=\mathbb T$. We have \begin{teo} Let us suppose that the Rayleigh-Taylor condition is satisfied, \emph{i.e.} $\rho^2-\rho^1>0$. Then there exists $f_0(x)=f(x,0)\in H^3(\Omega)$, an admissible (see Theorem \ref{IIteo1}) initial datum, such that, for any possible choice of $\kappa^1,\kappa^2>0$ and $h_2>>1$, there exists a solution of \eqref{IIeq9} and \eqref{IIeq13} for which $\lim_{t\rightarrow T^*}\|\partial_x f(t)\|_{L^\infty}=\infty$ in finite time $0<T^*<\infty$. For short time $t>T^*$ the solution can be continued but it is not a graph. \label{IIteo2} \end{teo} \begin{proof} To simplify notation we drop the physical parameters present in the problem by considering $\kappa^1(\rho^2-\rho^1)=2\pi$. The proof has three steps. First we consider solutions which are arbitrary curves (not necesary graphs) and we \emph{translate} the singularity formation to the fact $\partial_\alpha v_1(0)=\partial_t \partial_\alpha z_1(0)<0$. The second step is to construct a family of curves such that this expression is negative. Thus, we have that \emph{if there exists, forward and backward in time, a solution in the Rayleigh-Taylor stable case corresponding to initial data which are arbitrary curves} then, we have proved that there is a singularity in finite time. The last step is to prove, using a Cauchy-Kovalevsky theorem, that there exists local in time solutions in this unstable case. \textbf{Obtaining the correct expression:} Consider the case $\Omega=\mathbb R$. Due to \eqref{IIeq8} we have $$ \partial_\alpha \partial_t z_1(\alpha)=I_1+I_2+I_3, $$ where $$ I_1(\alpha)=\partial_\alpha\text{P.V.}\int_\mathbb R\frac{z_1(\alpha)-z_1(\alpha-\beta)}{|z(\alpha)-z(\alpha-\beta)|^2}(\partial_\alpha z_1(\alpha)-\partial_\alpha z_1(\alpha-\beta))d\beta, $$ $$ I_2(\alpha)=\partial_\alpha \frac{1}{2\pi}\text{P.V.}\int_\mathbb R\varpi_2(\alpha-\beta)\frac{-z_2(\alpha)-h_2}{|z(\alpha)-h(\alpha-\beta)|^2}d\beta, $$ $$ I_3(\alpha)=\partial_\alpha\left(\partial_\alpha z_1(\alpha)\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\varpi_2(\alpha-\beta)\frac{z_2(\alpha)+h_2}{|z(\alpha)-h(\alpha-\beta)|^2}d\beta\right). $$ Assume now that the following conditions for $z(\alpha)$ holds: \begin{itemize} \item $z_i(\alpha)$ are odd functions, \item $\partial_\alpha z_1(0)=0,\partial_\alpha z_1(\alpha)>0$ $\forall \alpha\neq0$, and $\partial_\alpha z_2(0)>0$, \item $z(\alpha)\neq h(\alpha)$ $\forall \alpha$. \end{itemize} The previous hypotheses mean that $z$ is a curve satisfying the arc-chord condition and $\partial_\alpha z(0)$ only has vertical component. Due to these conditions on $z$ we have $\partial_\alpha z_1(0)=0$ and $\partial_\alpha^2 z_1$ is odd (and then the second derivative at zero is zero) and we get that $I_3(0)=0$. For $I_1$ we get \begin{multline*} I_1(0)=\text{P.V.}\int_\mathbb R\frac{\partial_\alpha^2z_1(\beta)z_1(\beta)+(\partial_\alpha z_1(\beta))^2}{(z_1(\beta))^2+(z_2(\beta))^2}d\beta-2\text{P.V.}\int_\mathbb R\frac{(\partial_\alpha z_1(\beta)z_1(\beta))^2}{((z_1(\beta))^2+(z_2(\beta))^2)^2}d\beta\\ +2\text{P.V.}\int_\mathbb R\frac{\partial_\alpha z_1(\beta)z_1(\beta)z_2(\beta)\left(\partial_\alpha z_2(0)-\partial_\alpha z_2(\beta)\right)}{((z_1(\beta))^2+(z_2(\beta))^2)^2}d\beta. \end{multline*} We integrate by parts and we obtain, after some lengthy computations, \begin{equation}\label{IIsing1R} I_1(0)=4\partial_\alpha z_2(0)\text{P.V.}\int_0^\infty\frac{\partial_\alpha z_1(\beta)z_1(\beta)z_2(\beta)}{((z_1(\beta))^2+(z_2(\beta))^2)^2}d\beta. \end{equation} For the term with the second vorticity we have \begin{multline*} I_2(0)=\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{\partial_\beta\varpi_2(-\beta)h_2}{\beta^2+h_2^2}d\beta+\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{-\varpi_2(-\beta)\partial_\alpha z_2(0)}{\beta^2+h_2^2}d\beta\\ -\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{2\varpi_2(-\beta)\beta h_2}{(\beta^2+h_2^2)^2}d\beta-\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{\partial_\alpha z_2(0)\varpi_2(-\beta)(-h_2^2)}{(\beta^2+h_2^2)^2}d\beta, \end{multline*} and, after an integration by parts we obtain \begin{equation}\label{IIsing2R} I_2(0)=-\frac{\partial_\alpha z_2(0)}{2\pi}\text{P.V.}\int_0^\infty\frac{(\varpi_2(\beta)+\varpi_2(-\beta))\beta^2}{(\beta^2+h_2^2)^2}d\beta. \end{equation} Putting all together we obtain that in the flat at infinity case the important quantity for the singularity is \begin{multline}\label{IIsing3R} \partial_\alpha v_1(0)=\partial_\alpha z_2(0)\left(4\text{P.V.}\int_0^\infty\frac{\partial_\alpha z_1(\beta)z_1(\beta)z_2(\beta)}{((z_1(\beta))^2+(z_2(\beta))^2)^2}d\beta\right.\\ \left.-\frac{1}{2\pi}\text{P.V.}\int_0^\infty\frac{(\varpi_2(\beta)+\varpi_2(-\beta))\beta^2}{(\beta^2+h_2^2)^2}d\beta\right), \end{multline} where, due to \eqref{IIeq7}, $\varpi_2$ is defined as \begin{equation}\label{IIw2f} \varpi_2(\beta)=2\mathcal{K}\text{P.V.}\int_\mathbb R\frac{(h_2+z_2(\gamma))\partial_\alpha z_2(\gamma)}{(h_2+z_2(\gamma))^2+(\beta-z_1(\gamma))^2}d\gamma. \end{equation} We apply the same procedure to equation \eqref{IIeq11} and we get the importat quantity in the periodic setting (recall the superscript $p$ in the notation denoting that we are in the periodic setting): \begin{multline}\label{IIsing3T} \partial_\alpha v^p_1(0)=\partial_\alpha z_2(0)\left(\int_0^\pi\frac{\partial_\alpha z_1(\beta)\sin(z_1(\beta))\sinh(z_2(\beta))}{(\cosh(z_2(\beta))-\cos(z_1(\beta)))^2}d\beta\right.\\ \left.+\frac{1}{4\pi}\int_0^\pi\frac{(\varpi^p_2(\beta)+\varpi^p_2(-\beta))(-1+\cosh(h_2)\cos(\beta))}{(\cosh(h_2)-\cos(\beta))^2}d\beta\right), \end{multline} and, due to \eqref{IIeq12}, \begin{equation}\label{IIw2p} \varpi^p_2(\beta)=\mathcal{K}\int_{\mathbb T}\frac{\sin(\beta-z_1(\gamma))\partial_\alpha z_1(\gamma)}{\cosh(h_2+z_2(\gamma))-\cos(\beta-z_1(\gamma))}d\gamma. \end{equation} \textbf{Taking the appropriate curve:} To clarify the proof, let us consider first the periodic setting. Given $1<h_2$, we consider $a,b,$ constants such that $2<b\leq a$ and let us define $$ z_1(\alpha)=\alpha-\sin(\alpha), $$ and \begin{equation}\label{IIzT} z_2(\alpha)=\left\{ \begin{array}{lllll}\displaystyle \frac{\sin(a\alpha)}{a} & \hbox{ if } \displaystyle 0\leq \alpha\leq \frac{\pi}{a},\\ \displaystyle \frac{\sin\left(\pi\frac{\alpha-(\pi/a)}{(\pi/a)-(\pi/b)}\right)}{b} & \text { if } \displaystyle\frac {\pi}{a}<\alpha<\frac{\pi}{b},\\ \displaystyle \left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\alpha -\frac{\pi}{b}\right) & \text{ if }\displaystyle \frac{\pi}{b}\leq\alpha<\frac{\pi} {2},\\ \displaystyle -\left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\alpha -\pi+\frac{\pi}{b}\right) & \text{ if } \displaystyle\frac{\pi}{2}\leq\alpha<\pi (1-\frac{1}{b}),\\ \displaystyle 0 & \text{ if } \displaystyle \pi(1-\frac{1}{b})\leq\alpha.\\ \end{array}\right. \end{equation} \begin{figure}[t] \begin{center} \includegraphics[scale=0.4]{curva.eps} \end{center} \caption{$z_2$ in \eqref{IIzT} for $a=5,b=3,h_2=\pi/2$} \label{IIturning} \end{figure} Due to the definition of $z_2$, we have $$ \frac{h_2}{2}\leq h_2 + z_2(\alpha)\leq \frac{3h_2}{2}, $$ and using \eqref{IIw2p}, we get $$ |\varpi^p_2(\beta)|\leq \frac{4\pi}{\cosh(h_2/2)-1}. $$ Inserting this curve in \eqref{IIsing3T} we obtain $$ \partial_\alpha v_1^p(0)\leq I_a+I^{h_2}_b+I^{h_2}_2, $$ with $$ I_a=\int_0^{\pi/a}\frac{(1-\cos(\beta))\sin(\beta-\sin(\beta))\sinh\left(\frac{\sin(a\beta)}{a}\right)}{\left(\cosh\left(\frac{\sin(a\beta)}{a}\right)-\cos(\beta-\sin(\beta))\right)^2}d\beta, $$ \begin{multline*} I_b^{h_2}=\int_{(\pi/b+\pi)/3}^{\pi/2}\frac{(1-\cos(\beta))\sin(\beta-\sin(\beta)) \sinh\left(\left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\beta -\frac{\pi}{b}\right)\right)}{\left(\cosh\left(\left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\beta -\frac{\pi}{b}\right)\right)-\cos(\beta-\sin(\beta))\right)^2}d\beta\\ +\int_{\pi/2}^{(2\pi-\pi/b)/3)}\frac{(1-\cos(\beta))\sin(\beta-\sin(\beta)) \sinh\left(-\left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\alpha -\pi+\frac{\pi}{b}\right)\right)}{\left(\cosh\left(-\left(\frac{-h_2/2}{\frac{\pi}{2}-\frac{\pi}{b}}\left(\alpha -\pi+\frac{\pi}{b}\right)\right)\right)-\cos(\beta-\sin(\beta))\right)^2}d\beta, \end{multline*} and $I^{h_2}_2$ is the integral involving the second vorticity $\varpi^p_2$. We remark that $I_a$ does not depend on $h_2$. The sign of $I^{h_2}_b$ is the same as the sign of $z_2$, thus we get $I_b^{h_2}<0$ and this is independent of the choice of $a$ and $h_2$. Now we fix $b$ and we take $h_2$ sufficiently large such that $$ I^{h_2}_b+I^{h_2}_2\leq I^{h_2}_b+\frac{2\pi}{\cosh(h_2/2)-1}\frac{1+\cosh(h_2)}{(\cosh(h_2)-1)^2}<0. $$ We can do that because $$ \frac{c_b\sinh(h_2/3)}{(\cosh(h_2/2)+1)^2} \leq |I^{h_2}_b| $$ or, equivalently, $$ I^{h_2}_b+I^{h_2}_2=-|I^{h_2}_b|+I^{h_2}_2\leq -\frac{c_b\sinh(h_2/3)}{(\cosh(h_2/2)-1)^2}+\frac{2\pi}{\cosh(h_2/2)-1}\frac{1+\cosh(h_2)}{(\cosh(h_2)-1)^2}<0, $$ if $h_2$ is large enough. The integral $I_a$ is well defined and positive, but goes to zero as $a$ grows. Then, fixed $b$ and $h_2$ in such a way $I^{h_2}_b+I^{h_2}_2<0$, we take $a$ sufficiently large such that $I_a+I^{h_2}_b+I^{h_2}_2<0$. We are done with the periodic case. We proceed with the flat at infinity case. We take $2<b\leq a$ as before and $0<\delta<1$ and define \begin{equation}\label{IIzRI} z_1(\alpha)=\alpha-\sin(\alpha)\exp(-\alpha^2), \end{equation} and \begin{equation}\label{IIzR} z_2(\alpha)=\left\{ \begin{array}{lllll}\displaystyle \frac{\sin(a\alpha)}{a} & \hbox{ if } \displaystyle 0\leq \alpha\leq \frac{\pi}{a},\\ \displaystyle \frac{\sin\left(\pi\frac{\alpha-(\pi/a)}{(\pi/a)-(\pi/b)}\right)}{b} & \text { if } \displaystyle\frac {\pi}{a}<\alpha<\frac{\pi}{b},\\ \displaystyle \left(\frac{-h_2^\delta}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\alpha -\frac{\pi}{b}\right) & \text{ if }\displaystyle \frac{\pi}{b}\leq\alpha<\frac{\pi} {2},\\ \displaystyle -\left(\frac{-h_2^\delta}{\frac{\pi}{2}-\frac{\pi}{b}}\right)\left(\alpha -\pi+\frac{\pi}{b}\right) & \text{ if } \displaystyle\frac{\pi}{2}\leq\alpha<\pi (1-\frac{1}{b}),\\ \displaystyle 0 & \text{ if } \displaystyle \pi(1-\frac{1}{b})\leq\alpha.\\ \end{array}\right. \end{equation} We have $$ h_2-h_2^\delta<h_2+z_2(\beta)<h_2+h_2^\delta, $$ and we assume $1<h_2-h_2^\delta$. Inserting the curve \eqref{IIzRI} and \eqref{IIzR} in \eqref{IIw2f} and changing variables, we obtain \begin{eqnarray*} |\varpi_2(\beta)|\leq 2\text{P.V.}\int_\mathbb R\frac{(h_2+h_2^\delta)h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}}{(h_2-h_2^\delta)^2+(\gamma-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})^2}d\gamma. \end{eqnarray*} We split the integral in two parts: $$J_1=2\text{P.V.}\int_{B(0,2(h_2-h_2^\delta))}\frac{(h_2+h_2^\delta)h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}}{(h_2-h_2^\delta)^2+(\gamma-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})^2}d\gamma\leq 8h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}, $$ and $$ J_2=2\text{P.V.}\int_{B^c(0,2(h_2-h_2^\delta))}\frac{(h_2+h_2^\delta)h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}}{(h_2-h_2^\delta)^2+(\gamma-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})^2}d\gamma. $$ We have \begin{multline*} K_1=\text{P.V.}\int_{2(h_2-h_2^\delta)}^\infty\frac{1}{(h_2-h_2^\delta)^2+(\gamma-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})^2}d\gamma\\ \leq\text{P.V.}\int_{2(h_2-h_2^\delta)}^\infty\frac{1}{(h_2-h_2^\delta)^2+\gamma^2-2\gamma\sin(\beta-\gamma)e^{-(\beta-\gamma)^2}}d\gamma\\ \leq \text{P.V.}\int_{2(h_2-h_2^\delta)}^\infty\frac{1}{(h_2-h_2^\delta-\gamma)^2+2\gamma(h_2-h_2^\delta-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})}d\gamma, \end{multline*} and using that $h_2$ is such that $1<h_2-h_2^\delta$, we get $$ K_1\leq\text{P.V.}\int_{2(h_2-h_2^\delta)}^\infty\frac{1}{(h_2-h_2^\delta-\gamma)^2}d\gamma=\frac{1}{h_2-h_2^\delta}. $$ The remaining integral is \begin{multline*} K_2=\text{P.V.}\int^{-2(h_2-h_2^\delta)}_{-\infty}\frac{1}{(h_2-h_2^\delta)^2+(\gamma-\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})^2}d\gamma\\ \leq\text{P.V.}\int^{-2(h_2-h_2^\delta)}_{-\infty}\frac{1}{(h_2-h_2^\delta)^2+\gamma^2-2\gamma\sin(\beta-\gamma)e^{-(\beta-\gamma)^2}}d\gamma\\ \leq \text{P.V.}\int^{-2(h_2-h_2^\delta)}_{-\infty}\frac{1}{(h_2-h_2^\delta+\gamma)^2-2\gamma(h_2-h_2^\delta+\sin(\beta-\gamma)e^{-(\beta-\gamma)^2})}d\gamma, \end{multline*} and using that $h_2$ is such that $1<h_2-h_2^\delta$, we get $$ K_2\leq\text{P.V.}\int^{-2(h_2-h_2^\delta)}_{-\infty}\frac{1}{(h_2-h_2^\delta+\gamma)^2}d\gamma=\frac{1}{h_2-h_2^\delta}. $$ Putting all together we get $$ J_2\leq 4h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}, $$ and $$ |\varpi_2(\beta)|\leq 12h_2^\delta\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}. $$ Using this bound in \eqref{IIsing3R} we get $$ |I_2^{h_2}|\leq 3h_2^{\delta-1}\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}. $$ Then, as before, $$ \partial_\alpha v_1(0)\leq I_a+I_b^{h_2}+I_2^{h_2}, $$ where $I_a,I_b^{h_2}$ are the integrals $I_1(0)$ on the intervals $(0,\pi/a)$ and $((\pi/b+\pi)/3,(2\pi-\pi/b)/3)$, respectively. We have $$ c_b\frac{2h_2^\delta}{3(h_2^\delta)^4}\leq|I_b^{h_2}| $$ thus, $$ I^{h_2}_b+I^{h_2}_2=-|I^{h_2}_b|+I^{h_2}_2\leq -c_b\frac{2h_2^{-3\delta}}{3}+3h_2^{\delta-1}\left(\frac{\pi}{2}-\frac{\pi}{b}\right)^{-1}. $$ To ensure that the decay of $I_2^{h_2}$ is faster than the decay of $I_b^{h_2}$ we take $\delta<1/4$. Now, fixing $b$, we can obtain $1<h_2$ and $0<\delta<1/4$ such that $1<h_2-h_2^\delta$ and $I^{h_2}_b+I^{h_2}_2<0$. Taking $a>>b$ we obtain a curve such that $\partial_\alpha v_1(0)<0$. In order to conclude the argument it is enough to approximate these curves \eqref{IIzT} and \eqref{IIzR} by analytic functions. We are done with this step of the proof. \textbf{Showing the forward and backward solvability:} At this point, we need to prove that there is a solution forward and backward in time corresponding to these curves \eqref{IIzT} and \eqref{IIzR}. Indeed, if this solution exists then, due to the previous step, we obtain that, for a short time $t<0$, the solution is a graph with finite $H^3(\Omega)$ energy (in fact, it is analytic). This graph at time $t=0$ has a blow up for $\|\partial_x f\|_{L^\infty}$ and, for a short time $t>0$, the solution can not be parametrized as a graph. We show the result corresponding to the flat at infinity case, being the periodic one analogous. We consider curves $z$ satisfying the arc-chord condition and such that $$ \lim_{|\alpha|\rightarrow\infty}|z(\alpha)-(\alpha,0)|=0. $$ We define the complex strip $\mathbb B_r=\{\zeta+i\xi, \zeta\in\mathbb R, |\xi|<r\},$ and the spaces \begin{equation}\label{IIXdef} X_r=\{z=(z_1,z_2) \text{ analytic curves satisfying the arc-chord condition on } \mathbb B_r\}, \end{equation} with norm $$ \|z\|^2_r=\|z(\gamma)-(\gamma,0)\|^2_{H^3(\mathbb B_r)}, $$ where $H^3(\mathbb B_r)$ denotes the Hardy-Sobolev space on the strip with the norm \begin{equation}\label{IIXnorm} \|f\|^2_r=\sum_{\pm}\int_\mathbb R|f(\zeta\pm r i)|^2d\zeta+\int_\mathbb R|\partial_\alpha^3 f(\zeta\pm r i)|^2d\zeta, \end{equation} (see \cite{bakan2007hardy}). These spaces form a Banach scale. For notational convenience we write $\gamma=\alpha\pm ir$, $\gamma'=\alpha\pm ir'$. Recall that, for $0<r'<r$, \begin{equation}\label{IIcauchy2} \|\partial_\alpha \cdot\|_{L^2(\mathbb B_{r'})}\leq\frac{C}{r-r'}\|\cdot\|_{L^2(\mathbb B_r)}. \end{equation} We consider the complex extension of \eqref{IIeq7} and \eqref{IIeq8}, which is given by \begin{multline} \label{IIeq18} \partial_t z(\gamma)=\text{P.V.}\int_\mathbb R\frac{(z_1(\gamma)-z_1(\gamma-\beta))(\partial_\alpha z(\gamma)-\partial_\alpha z(\gamma-\beta))}{(z_1(\gamma)-z_1(\gamma-\beta))^2+(z_2(\gamma)-z_2(\gamma-\beta))^2}d\beta\\ +\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{\varpi_2(\gamma-\beta)(z(\gamma)-h(\gamma-\beta))^\perp}{(z_1(\gamma)-(\gamma-\beta))^2+(z_2(\gamma)+h_2)^2}d\beta\\ +\partial_\alpha z(\gamma)\frac{1}{2\pi}\text{P.V.}\int_\mathbb R\frac{(z_2(\gamma)+h_2)\varpi_2(\gamma-\beta)}{(z_1(\gamma)-(\gamma-\beta))^2+(z_2(\gamma)+h_2)^2}d\beta, \end{multline} with \begin{equation} \label{IIeq17} \varpi_2(\gamma)=2\mathcal{K}\,\text{P.V.}\int_\mathbb R\frac{(h_2+z_2(\gamma-\beta))\partial_\alpha z_2(\gamma-\beta)}{(\gamma-z_1(\gamma-\beta))^2+(h_2+z_2(\gamma-\beta))^2}d\beta. \end{equation} Recall the fact that in the case of a real variable graph $\varpi_2$ has the same regularity as $f$, but in the case of an arbitrary curve $\varpi_2$ is, roughly speaking, at the level of the first derivative of the interface. This fact will be used below. We define \begin{equation} \label{IIeq19} d^-[z](\gamma,\beta)=\frac{\beta^2}{(z_1(\gamma)-z_1(\gamma-\beta))^2+(z_2(\gamma)-z_2(\gamma-\beta))^2}, \end{equation} \begin{equation} \label{IIeq19b} d^h[z](\gamma,\beta)=\frac{1+\beta^2}{(z_1(\gamma)-(\gamma-\beta))^2+(z_2(\gamma)+h_2)^2}. \end{equation} The function $d^-$ is the complex extension of the \emph{arc chord condition} and we need it to bound the terms with $\varpi_1$. The function $d^h$ comes from the different permeabilities and we use it to bound the terms with $\varpi_2$. We observe that both are bounded functions for the considered curves. Consider $0< r'<r$ and the set $$ O_R=\{z\in X_r\text{ such that }\|z\|_r<R, \|d^-[z]\|_{L^\infty(\mathbb B_r)}<R, \|d^h[z]\|_{L^\infty(\mathbb B_r)}<R\}, $$ where $d^-[z]$ and $d^h[z]$ are defined in \eqref{IIeq19} and \eqref{IIeq19b}. Then we claim that, for $z,w\in O_R$, the righthand side of \eqref{IIeq18}, $F:O_R\rightarrow X_{r'}$ is continuous and the following inequalities holds: \begin{eqnarray}\label{IIeq20}&&\|F[z]\|_{H^3(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z\|_r,\\ \label{IIeq21}&&\|F[z]-F[w]\|_{H^3(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z-w\|_{H^3(\mathbb B_r)},\\ \label{IIeq22}&&\sup_{\gamma\in \mathbb B_r,\beta\in\mathbb R}|F[z](\gamma)-F[z](\gamma-\beta)|\leq C_R|\beta|. \end{eqnarray} The claim for the spatial operator corresponding to $\varpi_1$ has been studied in \cite{ccfgl}, thus, we only deal with the new terms containing $\varpi_2$. For the sake of brevity we only bound some terms, being the other analogous. Using Tonelli's theorem and Cauchy-Schwartz inequality we have that $$ \|\varpi_2\|_{L^2(\mathbb B_{r'})}\leq c\|d^h[z]\|_{L^\infty}(1+\|z_2\|_{L^\infty(\mathbb B_{r'})})\|\partial_\alpha z_2\|_{L^2(\mathbb B_{r'})}. $$ Moreover, we get \begin{equation}\label{IIeqw2} \|\varpi_2\|_{H^2(\mathbb B_{r})}\leq C_R\|z\|_r. \end{equation} For $\partial_\alpha^3 \varpi_2$ the procedure is similar but we lose one derivative. Using \eqref{IIcauchy2} and Sobolev embedding we conclude \begin{equation}\label{IIeqw2b} \|\varpi_2\|_{H^3(\mathbb B_{r'})}\leq \frac{C_R}{r-r'}\|z\|_r. \end{equation} From here inequality \eqref{IIeq20} follows. Inequality \eqref{IIeq21}, for the terms involving $\varpi_1$, can be obtained using the properties of the Hilbert transform as in \cite{ccfgl}. Let's change slightly the notation and write $\varpi_2[z](\gamma)$ for the integral in \eqref{IIeq17}. We split \begin{eqnarray*} A_1&=&\text{P.V.}\int_\mathbb R\frac{(\varpi_2[z](\gamma'-\beta)-\varpi_2[w](\gamma'-\beta))(z(\gamma')-h(\gamma'-\beta))^\perp}{(z_1(\gamma')-(\gamma'-\beta))^2+(z_2(\gamma')+h_2)^2}d\beta\\ &&+\text{P.V.}\int_\mathbb R\frac{\varpi_2[w](\gamma'-\beta)\left((z(\gamma')-h(\gamma'-\beta))^\perp-(w(\gamma')-h(\gamma'-\beta))^\perp\right)}{(z_1(\gamma')-(\gamma'-\beta))^2+(z_2(\gamma')+h_2)^2}d\beta\\ &&+\text{P.V.}\int_\mathbb R\varpi_2[w](\gamma'-\beta)(w(\gamma')-h(\gamma'-\beta))^\perp\frac{d^h[z](\gamma',\beta)-d^h[w](\gamma',\beta)}{1+\beta^2}d\beta\\ &&=B_1+B_2+B_3. \end{eqnarray*} In $B_3$ we need some extra decay at infinity to ensure the finiteness of the integral. We compute $$ |d^h[z]-d^h[w]|\leq C_R\frac{|d^h[z]d^h[w]|}{1+\beta^2}\left|(1+\beta)(z_1-w_1)+z_2-w_2\right|<C_R|z-w|\frac{|1+\beta|}{1+\beta^2}, $$ and, due to Sobolev embedding, we get $$ \|B_3\|_{L^2(\mathbb B_{r'})}\leq C_R \|\varpi_2[w]\|_{L^2(\mathbb B_{r'})}\|z-w\|_{L^\infty(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z-w\|_{H^3(\mathbb B_r)}. $$ For the second term we obtain the same bound $$ \|B_2\|_{L^2(\mathbb B_{r'})}\leq C_R \|\varpi_2[w]\|_{L^2(\mathbb B_{r'})}\|z-w\|_{L^\infty(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z-w\|_{H^3(\mathbb B_r)}. $$ We split $B_1$ componentwise. In the first coordinate we have \begin{multline*} \|C_1\|_{L^2(\mathbb B_{r'})}=\left\|\text{P.V.}\int_\mathbb R\frac{(\varpi_2[z](\gamma'-\beta)-\varpi_2[w](\gamma'-\beta))(-z_2(\gamma')-h_2)}{(z_1(\gamma')-(\gamma'-\beta))^2+(z_2(\gamma')+h_2)^2}d\beta\right\|_{L^2(\mathbb B_{r'})}\\ \leq C_R\|\varpi_2[z]-\varpi_2[w]\|_{L^2(\mathbb B_{r'})}. \end{multline*} In the second coordinate we need to ensure the integrability at infinity. We get \begin{eqnarray*} C_2&=&\text{P.V.}\int_\mathbb R\frac{(\varpi_2[z](\gamma'-\beta)-\varpi_2[w](\gamma'-\beta))(z_1(\gamma')-\gamma')}{(z_1(\gamma')-(\gamma'-\beta))^2+(z_2(\gamma')+h_2)^2}d\beta\\ &&+\text{P.V.}\int_\mathbb R(\varpi_2[z](\gamma'-\beta)-\varpi_2[w](\gamma'-\beta))\left(\frac{\beta d^h[z]}{1+\beta^2}-\frac{1}{\beta}\right)d\beta\\ &&+H\varpi_2[z](\gamma')-H\varpi_2[w](\gamma'), \end{eqnarray*} and, with this splitting and the properties of the Hilbert transform, we obtain $$ \|C_2\|_{L^2(\mathbb B_{r'})}\leq C_R\|\varpi_2[z]-\varpi_2[w]\|_{L^2(\mathbb B_{r'})}. $$ We get $$ \varpi_2[z]-\varpi_2[w]=C_3+C_4+C_5, $$ where $$ C_3=2\mathcal{K}\;\text{P.V.}\int_\mathbb R\frac{(z_2(\gamma-\beta)-w_2(\gamma-\beta))\partial_\alpha z_2(\gamma-\beta)}{(\gamma-z_1(\gamma-\beta))^2+(h_2+z_2(\gamma-\beta))^2}d\beta, $$ $$ C_4=2\mathcal{K}\;\text{P.V.}\int_\mathbb R\frac{(h_2+w_2(\gamma-\beta))(\partial_\alpha z_2(\gamma-\beta)-\partial_\alpha w_2(\gamma-\beta))}{(\gamma-z_1(\gamma-\beta))^2+(h_2+z_2(\gamma-\beta))^2}d\beta, $$ and $$ C_5=2\mathcal{K}\;\text{P.V.}\int_\mathbb R(h_2+w_2(\gamma-\beta))\partial_\alpha w_2(\gamma-\beta)\frac{d^h[z](\gamma-\beta,-\beta)-d^h[w](\gamma-\beta,-\beta)}{1+\beta^2}d\beta. $$ From these expressions we obtain $$ \|C_3\|_{L^2(\mathbb B_{r'})}\leq C_R\|z-w\|_{L^\infty}\|\partial_\alpha z_2\|_{L^2(\mathbb B_{r'})}, $$ $$ \|C_4\|_{L^2(\mathbb B_{r'})}\leq C_R\|\partial_\alpha(z-w)\|_{L^2(\mathbb B_{r'})}, $$ and $$ \|C_5\|_{L^2(\mathbb B_{r'})}\leq C_R\|z-w\|_{L^\infty}\|\partial_\alpha z_2\|_{L^2(\mathbb B_{r'})}. $$ Collecting all these estimates, and due to Sobolev embedding and \eqref{IIcauchy2} we obtain $$ \|B_1\|_{L^2(\mathbb B_{r'})}\frac{C_R}{r-r'}\|z-w\|_{H^3(\mathbb B_r)}. $$ We are done with \eqref{IIeq21}. Inequality \eqref{IIeq22} is equivalent to the bound $|\partial_t\partial_\alpha z|<C_R$. Such a bound for the terms involving $\varpi_2$ can be obtained from \eqref{IIeq19b} and \eqref{IIeqw2}. For instance \begin{multline*} A_2=\text{P.V.}\int_\mathbb R\frac{\partial_\alpha\varpi_2(\gamma-\beta)(z(\gamma)-h(\gamma-\beta))^\perp}{(z_1(\gamma)-(\gamma-\beta))^2+(z_2(\gamma)+h_2)^2}d\beta=\\ -\text{P.V.}\int_\mathbb R\varpi_2(\gamma-\beta)\partial_\beta\left(\frac{(z(\gamma)-h(\gamma-\beta))^\perp}{(z_1(\gamma)-(\gamma-\beta))^2+(z_2(\gamma)+h_2)^2}\right)d\beta\\ \leq C_R\|\varpi_2\|_{H^2(\mathbb B_r)}\|d^h[z]\|_{L^\infty}. \end{multline*} The remaining terms can be handled in a similar way. Now we can finish with the forward and backward solvability step. Take $z(0)$ the analytic extension of $z$ in \eqref{IIzR} (\eqref{IIzT} for the periodic case). We have $z(0)\in X_{r_0}$ for some $r_0>0$, it satisfies the arc-chord condition and does not reach the curve $h$, thus, there exists $R_0$ such that $z(0)\in O_{R_0}$. We take $r<r_0$ and $R>R_0$ in order to define $O_R$ and we consider the iterates $$ z_{n+1}=z(0)+\int_0^tF[z_n]ds,\;z_0=z(0), $$ and assume by induction that $z_k\in O_R$ for $k\leq n$. Then, following the proofs in \cite{ccfgl,CGO,nirenberg1972abstract,nishida1977note}, we obtain a time $T_{CK}>0$ of existence. It remains to show that $$ \|d^-[z_{n+1}]\|_{L^\infty(\mathbb B_r)},\|d^h[z_{n+1}]\|_{L^\infty(\mathbb B_r)}<R, $$ for some times $T_A, T_B>0$ respectively. Then we choose $T=\min\{T_{CK},T_A,T_B\}$ and we finish the proof. As $d^-$ has been studied in \cite{ccfgl} we only deal with $d^h$. Due to \eqref{IIeq20} and the definition of $z(0)$, we have $$ (d^{+}[z_{n+1}])^{-1}>\frac{1}{R_0}-C_R(t^2+t), $$ and, if we take a sufficiently small $T_B$ we can ensure that for $t<T_B$ we have $d^h[z_{n+1}]<R$. We conclude the proof of the Theorem. \end{proof} We observe that in the periodic case the curve $z$ is of the same order as $h_2$, so, even if $h_2>>1$, this result is not some kind of \emph{linearization}. The same result is valid if $\mathcal{K}<<1$ for any $h_2$ (see Theorem \ref{IIteo4}). Moreover, we have numerical evidence showing that for every $|\mathcal{K}|<1$ and $h_2=\pi/2$ (and not $h_2>>1$) there are curves showing turning effect. \begin{NE} There are curves such that for every $|\mathcal{K}|<1$ and $h_2=\pi/2$ turn over. \end{NE} Let us consider first the periodic setting. Recall the fact that $h_2=\pi/2$ and let us define \begin{equation}\label{IIzTNE} z_1(\alpha)=\alpha-\sin(\alpha), \;\;z_2(\alpha)=\frac{\sin(3\alpha)}{3}-\sin(\alpha)\left(e^{-(\alpha+2)^2}+e^{-(\alpha-2)^2}\right)\text{ for $\alpha\in\mathbb T$}. \end{equation} Inserting this curve in \eqref{IIsing3T} we obtain that for any possible $-1<\mathcal{K}<1$, $$ I_1(0)+|I_2(0)|<0. $$ In particular $$ \partial_\alpha v_1^p(0)=I_1(0)+I_2(0)<I_1(0)+|I_2(0)|<0. $$ Let us introduce the algorithm we use. We need to compute $$ \partial_\alpha v^p_1(0)=\int_0^\pi\mathcal{I}_1+\int_0^\pi\mathcal{I}_2, $$ where $\mathcal{I}_i$ means the $i-$integral in \eqref{IIsing3T}. Recall that $\mathcal{I}_i$ is two times differentiable, so, we can use the sharp error bound for the trapezoidal rule. We denote $dx$ the mesh size when we compute the first integral. We approximate the integral of $\mathcal{I}_1$ using the trapezoidal rule between $(0.1,\pi)$. We neglect the integral in the interval $(0,0.1)$, paying with an error denoted by $|E^1_{PV}|=O(10^{-3})$. The trapezoidal rule gives us an error $|E^1_I|\leq\frac{dx^2(\pi-0.1)}{12}\|\partial_\alpha^2 \mathcal{I}_1\|_{L^\infty}$. As we know the curve $z$, we can bound $\partial_\alpha^2 \mathcal{I}_1$. We obtain $$ |E^1_I|\leq dx^2\frac{(\pi-0.1)}{6}10^{5}. $$ We take $dx=10^{-7}$. Putting all together we obtain \begin{equation*} |E^1|\leq|E^1_{PV}|+|E^1_I|+\leq 3O(10^{-3})= O(10^{-2}). \end{equation*} Then, we can ensure that \begin{equation}\label{III1err} \partial_\alpha z_2(0)\int_0^\pi\frac{\partial_\alpha z_1(\beta)\sin(z_1(\beta))\sinh(z_2(\beta))}{(\cosh(z_2(\beta))-\cos(z_1(\beta)))^2}d\beta\leq -0.7+|E^1|<-0.6. \end{equation} We need to control analytically the error in the integral involving $\varpi^p_2$. This second integral has the error coming form the numerical integration, $E^2_I$ and a new error coming from the fact that $\varpi^p_2$ is known with some error. We denote this new error as $E^2_\varpi$. Let us write $\tilde{dx}$ the mesh size for the second integral. Then, using the smoothness of $\mathcal{I}_2$, we have $$ |E^2_I|\leq\frac{\tilde{dx}^2}{16}\|\varpi^p_2\|_{C^2}\leq\frac{\tilde{dx}^2}{4}\cdot50. $$ We take $\tilde{dx}=10^{-4}.$ It remains the error coming from $\varpi_2^p$. The second vorticity, $\varpi_2^p$, is given by the integral \eqref{IIw2p}. We compute the integral \eqref{IIw2p} using the same mesh size as for $I_2$, $\tilde{dx}$. Thus, the errors are $$ |E_\varpi^2|\leq O(10^{-3}), $$ Putting all together we have $$ |E^2|\leq |E^2_I|+|E^2_\varpi|\leq O(10^{-2}), $$ and we conclude \begin{equation}\label{III2err} \left|\frac{\partial_\alpha z_2(0)}{4\pi}\int_0^\pi\frac{(\varpi^p_2(\beta)+\varpi^p_2(-\beta))(-1+\cosh(h_2)\cos(\beta))}{(\cosh(h_2)-\cos(\beta))^2}d\beta\right|\leq 0.1+|E^2|<0.2. \end{equation} Now, using \eqref{III1err} and \eqref{III2err}, we obtain $\partial_\alpha v_1^p(0)<0$, and we are done with the periodic case. We proceed with the flat at infinity case. We have to deal with the unboundedness of the domain so we define \begin{equation}\label{IIzRNE} z_1(\alpha)=\alpha-\sin(\alpha)\exp(-\alpha^2/100), \;\;z_2(\alpha)=\frac{\sin(3\alpha)}{3}-\sin(\alpha)\left(e^{-(\alpha+2)^2}+e^{-(\alpha-2)^2}\right)\textbf{1}_{\{|\alpha|<\pi\}}. \end{equation} Inserting this curve in \eqref{IIsing3R} we obtain that for any possible $-1<\mathcal{K}<1$, $$ I_1(0)+|I_2(0)|<0. $$ Then, as before, $$ \partial_\alpha v_1(0)=I_1(0)+I_2(0)<I_1(0)+|I_2(0)|<0. $$ The function $z_2$ is Lipschitz, so the same for $\mathcal{I}_1$, where now $\mathcal{I}_i$ are the expressions in \eqref{IIsing3R} and the second integral $I_2$ is over an unbounded interval. To avoid these problems we compute the numerical aproximation of $$ \int_{0.1}^{\pi-dx}\mathcal{I}_1+\int_0^{L_2}\mathcal{I}_2. $$ Recall that $\varpi_2$ is given by \eqref{IIw2f} and then, due to the definition of $z_2$, we can approximate it by an integral over $(0,\pi-\tilde{dx})$. The lack of analyticity of $z_2$ and the truncation of $I_2(0)$ introduces two new sources of error. We denote them by $E^1_{z_2}$ and $E^2_\mathbb R$. We take $dx=10^{-7},\tilde{dx}=10^{-4}$ and $L_2=2\pi$. Using the bounds $z_1\leq\pi$, $\partial_\alpha z_1\leq 2$ and $z_2\leq h_2$ we obtain $$ |E^1_{z_2}|\leq\left|\int_{\pi-dx}^\pi\mathcal{I}_1\right|\leq dx\cdot0.2\cdot4\pi^2\leq 8\cdot10^{-7}. $$ We have $$ |\varpi_2(\beta)|\leq 4\pi\frac{(h_2+\max_\gamma|z_2(\gamma)|)\max_\gamma|\partial_\alpha z_2(\gamma)|}{\min_\gamma(h_2+z_2(\gamma))^2+(\beta-z_1(\gamma))^2}d\gamma\leq \frac{4\pi\cdot3\cdot2}{\min_\gamma(h_2+z_2(\gamma))^2+(\beta-z_1(\gamma))^2}=C(\beta), $$ and we get $C(\beta)<C(L_2)$ for $\beta>L_2$. Using this inequality we get the desired bound for the second error as follows: $$ |E^2_{\mathbb R}|\leq \frac{|C(L_2)|}{\pi}\int_{L_2}^\infty\frac{\beta^2}{\left(\beta^2+\left(\frac{\pi}{2}\right)^2\right)^2}\leq \frac{4\pi\cdot3\cdot2}{10}\cdot 0.05<4\cdot 10^{-1}. $$ The other errors can be bounded as before, obtaining, $$ |E^1|\leq |E^1_{PV}|+|E^1_I|+|E^1_{z_2}|=O(10^{-2}), $$ $$ |E^2|\leq |E^2_{\varpi}|+|E^2_I|+|E^2_{\mathbb R}|=0.42. $$ We conclude \begin{equation}\label{III1errf} \partial_\alpha z_2(0)\cdot4\text{P.V.}\int_0^\infty\frac{\partial_\alpha z_1(\beta)z_1(\beta)z_2(\beta)}{((z_1(\beta))^2+(z_2(\beta))^2)^2}d\beta\leq -0.7+|E^1|<-0.6, \end{equation} and \begin{equation}\label{III2errf} \left|-\frac{1}{2\pi}\text{P.V.}\int_0^\infty\frac{(\varpi_2(\beta)+\varpi_2(-\beta))\beta^2}{(\beta^2+h_2^2)^2}d\beta\right|<0.02+|E^2|<0.5. \end{equation} Putting together \eqref{III1errf} and \eqref{III2errf} we conclude $\partial_\alpha v_1(0)<0$. In order to complete a rigorous enclosure of the integral, we are left with the bounding of the errors coming from the floating point representation and the computer operations and their propagation. In a forthcoming paper (see \cite{GG}) we will deal with this matter. By using interval arithmetics we will give a computer assisted proof of this result. \subsection{Finite depth} In this section we show the existence of finite time singularities for some curves and physical parameters in an explicit range (see \eqref{IIeqK1}). This result is a consequence of Theorem 4 in \cite{CGO} by means of a continuous dependence on the physical parameters. As a consequence the range of physical parameters plays a role. Indeed, we have \begin{teo} Let us suppose that the Rayleigh-Taylor condition is satisfied, \emph{i.e.} $\rho^2-\rho^1>0$, and take $0<h_2<\frac{\pi}{2}$. There are $f_0(x)=f(x,0)\in H^3(\mathbb R)$, an admissible (see Theorem \ref{IIteo3}) initial datum, such that, for any $|\mathcal{K}|$ small enough, there exists solutions of \eqref{IIeqv2} such that $\lim_{t\rightarrow T^*}\|\partial_x f(t)\|_{L^\infty}=\infty$ for $0<T^*<\infty$. For short time $t>T^*$ the solution can be continued but it is not a graph. \label{IIteo4} \end{teo} \begin{proof} The proof is similar to the proof in Theorem \ref{IIteo2}. First, using the result in \cite{CGO} we obtain a curve, $z(0)$, such that the integrals in $\partial_\alpha v_1(0)$ coming from $\varpi_1$ have a negative contribution. The second step is to take $\mathcal{K}$ small enough, when compared with some quantities depending on the curve $z(0)$, such that the contribution of the terms involving $\varpi_2$ is small enough to ensure the singularity. Now, the third step is to prove, using a Cauchy-Kovalevsky theorem, that there exists local in time solutions corresponding to the initial datum $z(0)$. To simplify notation we take $\kappa^1(\rho^2-\rho^1)=4\pi$. Then the parameters present in the problem are $h_2$ and $\mathcal{K}$. \textbf{Obtaining the correct expression:} As in \cite{CGO} and Theorem \ref{IIteo2} we obtain $$ \partial_\alpha v_1(0)=\partial_t\partial_\alpha z_1(0)=I_1+I_2, $$ where $$ I_1=2\partial_\alpha z_2(0)\int_0^\infty \frac{\partial_\alpha z_1(\beta)\sinh(z_1(\beta))\sin(z_2(\beta))}{\left(\cosh(z_1(\beta))-\cos(z_2(\beta))\right)^2}+\frac{\partial_\alpha z_1(\beta)\sinh(z_1(\beta))\sin(z_2(\beta))}{\left(\cosh(z_1(\beta))+\cos(z_2(\beta))\right)^2}d\beta, $$ and \begin{multline*} I_2=\frac{\partial_\alpha z_2(0)}{4\pi}\int_\mathbb R\frac{\varpi_2(-\beta)(-\cosh(\beta)\cos(h_2)+1)}{(\cosh(\beta)-\cos(h_2))^2}d\beta\\ +\frac{\partial_\alpha z_2(0)}{4\pi}\int_\mathbb R\frac{\varpi_2(-\beta)(-\cosh(\beta)\cos(h_2)-\cos^2(h_2)+\sin^2(h_2))}{(\cosh(\beta)+\cos(h_2))^2}d\beta. \end{multline*} \textbf{Taking the appropriate curve and $\mathcal{K}$:} From Theorem 4 in \cite{CGO} we know that there are initial curves $w_0$ such that $I_1=-a^2,$ $a=a(w_0)>0$. We take one of this curves and we denote this smooth, fixed curve as $z(0)$. We need to obtain $\mathcal{K}=\mathcal{K}(z(0),h_2)$ such that $\partial_\alpha v_1(0)=-a^2+I_2<0$. As in \eqref{IIeq19b} we define \begin{equation}\label{IIeq19d} d_1^h[z](\gamma,\beta)=\frac{\cosh^2(\beta/2)}{\cosh(z_1(\gamma)-(\gamma-\beta))-\cos(z_2(\gamma)+h_2)}, \end{equation} and \begin{equation}\label{IIeq19f} d_2^h[z](\gamma,\beta)=\frac{\cosh^2(\beta/2)}{\cosh(z_1(\gamma)-(\gamma-\beta))+\cos(z_2(\gamma)-h_2)}. \end{equation} From the definition of $I_2$ it is easy to obtain $$ |I_2|\leq C(h_2)\partial_\alpha z_2(0)\|\varpi_2\|_{L^\infty}, $$ where $$ C(h_2)=\frac{1}{4\pi}\int_\mathbb R\frac{\cosh(\beta)\cos(h_2)+1}{(\cosh(\beta)-\cos(h_2))^2}d\beta+\frac{1}{4\pi}\int_\mathbb R\frac{\cosh(\beta)\cos(h_2)+\cos(2h_2)}{(\cosh(\beta)+\cos(h_2))^2}d\beta. $$ From the definition of $\varpi_2$ for curves (which follows from \eqref{IIw2defc} in a straightforward way) we obtain $$ \|\varpi_2\|_{L^\infty}\leq8\mathcal{K}\|\partial_\alpha z_2\|_{L^\infty}\left(\|d_1^h[z]\|_{L^\infty}+\|d_2^h[z]\|_{L^\infty}\right)\left(1+\frac{\mathcal{K}}{\sqrt{2\pi}}\|G_{h_2,\mathcal{K}}\|_{L^1}\right). $$ Fixing $0<h_2<\pi/2$ and collecting all the estimates we obtain \begin{equation*} |I_2|\leq C(h_2)8\partial_\alpha z_2(0)\mathcal{K}\|\partial_\alpha z_2\|_{L^\infty}\left(\|d_1^h[z]\|_{L^\infty}+\|d_2^h[z]\|_{L^\infty}\right)\left(1+\frac{\mathcal{K}}{\sqrt{2\pi}}\sup_{|\mathcal{K}|<1}\|G_{h_2,\mathcal{K}}\|_{L^1}\right). \end{equation*} Now it is enough to take \begin{equation}\label{IIeqK1} |\mathcal{K}_1(z(0),h_2)|<\frac{(C(h_2)8\partial_\alpha z_2(0)\|\partial_\alpha z_2\|_{L^\infty})^{-1}a^2}{\left(\|d_1^h[z]\|_{L^\infty}+\|d_2^h[z]\|_{L^\infty}\right)\left(1+\frac{1}{\sqrt{2\pi}}\sup_{|\mathcal{K}|<1}\|G_{h_2,\mathcal{K}}\|_{L^1}\right)}, \end{equation} to ensure that $\partial_\alpha v_1(0)<0$ for this curve $z(0)$ and any $|\mathcal{K}|<|\mathcal{K}_1(z(0),h_2)|$. \textbf{Showing the forward and backward solvability:} We define \begin{equation}\label{IIeq19c} d^-[z](\gamma,\beta)=\frac{\sinh^2(\beta/2)}{\cosh(z_1(\gamma)-z_1(\gamma-\beta))-\cos(z_2(\gamma)-z_2(\gamma-\beta))}, \end{equation} and \begin{equation}\label{IIeq19e} d^+[z](\gamma,\beta)=\frac{\cosh^2(\beta/2)}{\cosh(z_1(\gamma)-z_1(\gamma-\beta))+\cos(z_2(\gamma)-z_2(\gamma-\beta))}. \end{equation} Using the equations \eqref{IIeq19d},\eqref{IIeq19f},\eqref{IIeq19c} and \eqref{IIeq19e}, the proof of this step mimics the proof in Theorem \ref{IIteo2} and the proof in \cite{CGO} and so we only sketch it. As before, we consider curves $z$ satisfying the arc-chord condition and such that $$ \lim_{|\alpha|\rightarrow\infty}|z(\alpha)-(\alpha,0)|=0. $$ We define the complex strip $\mathbb B_r=\{\zeta+i\xi, \zeta\in\mathbb R, |\xi|<r\},$ and the spaces \eqref{IIXdef} with norm \eqref{IIXnorm} (see \cite{bakan2007hardy}). We define the set \begin{multline*} O_R=\{z\in X_r\text{ such that }\|z\|_r<R, \|d^-[z]\|_{L^\infty(\mathbb B_r)}<R, \|d^+[z]\|_{L^\infty(\mathbb B_r)}<R,\\ \|d_1^h[z]\|_{L^\infty(\mathbb B_r)}<R,\|d_2^h[z]\|_{L^\infty(\mathbb B_r)}<R\}, \end{multline*} where $d_i^h[z]$ and $d^\pm[z]$ are defined in \eqref{IIeq19d}, \eqref{IIeq19f}, \eqref{IIeq19c} and \eqref{IIeq19e}, respectively. As before, we have that, for $z,w\in O_R$, complex extension of \eqref{IIeqv}, $F:O_R\rightarrow X_{r'}$ is continuous and the following inequalities holds: \begin{eqnarray*}&&\|F[z]\|_{H^3(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z\|_r,\\ &&\|F[z]-F[w]\|_{H^3(\mathbb B_{r'})}\leq\frac{C_R}{r-r'}\|z-w\|_{H^3(\mathbb B_r)},\\ &&\sup_{\gamma\in \mathbb B_r,\beta\in\mathbb R}|F[z](\gamma)-F[z](\gamma-\beta)|\leq C_R|\beta|. \end{eqnarray*} We consider $$ z_{n+1}=z(0)+\int_0^tF[z_n]ds, \; z_0=z(0). $$ Using the previous properties of $F$ we obtain that, for $T=T(z(0),R)$ small enough, $z^{n+1}\in O_R,$ for all $n$. The rest of the proof follows in the same way as in \cite{nirenberg1972abstract, nishida1977note}. \end{proof} \bibliographystyle{abbrv}
\section{\@startsection {section}{1}{\z@}% {-3.5ex plus -1ex minus -.2ex}{2.3ex plus .2ex}{\large\bf}} \def\subsection{\@startsection{subsection}{2}% {\z@}{-3.25ex plus -1ex minus -.2ex}{1.5ex plus .2ex}{\normalsize\bf}} \def\@fnsymbol#1{\ensuremath{\ifcase#1\or *\or 1\or 2\or 3\or 4\or 5\or 6\or 7 \or 8\ or 9 \or 10\or 11 \else\@ctrerr\fi}} \makeatother \newcommand{\emphi}[1]{\emph{#1}} \newcommand{{et~al.}\xspace}{{et~al.}\xspace} \let\par\vspace{-\baselineskip}\relax \fi \pltopsep1ex \let\geq\geqslant \let\leq\leqslant \let\ge\geqslant \let\le\leqslant \newcommand{\symbol{'100}}{\symbol{'100}} \newcommand{\mathcal{P}}{\mathcal{P}} \newcommand{\mathcal{P}^{*}}{\mathcal{P}^{*}} \newcommand{T\'{o}th\xspace}{T\'{o}th\xspace} \newcommand{\si}[1]{#1} \newcommand{\face}{\psi}% \newcommand{\faceA}{\varphi}% \newcommand{{{e}}}{{{e}}} \newcommand{{{f}}}{{{f}}} \newcommand{{{g}}}{{{g}}} \newcommand{{v}}{{v}} \newcommand{{u}}{{u}} \newcommand{{G}}{{G}} \newcommand{{H}}{{H}} \newcommand{{Q}}{{Q}} \newcommand{{V}}{{V}} \newcommand{{\mathcal{F}}}{{\mathcal{F}}} \newcommand{{E}}{{E}} \newcommand{\pth}[2][\!]{#1\left({#2}\right)} \newcommand{\brc}[1]{\left\{ {#1} \right\}} \newcommand{\cardin}[1]{\left| {#1} \right|} \newcommand{L}{L} \renewcommand{\th}{th\xspace} \begin{document} \title{On the Number of Edges of Fan-Crossing Free Graphs% \thanks{OC and HSK were supported in part by NRF grant~2011-0016434 and in part by NRF grant 2011-0030044 (SRC-GAIA), both funded by the government of Korea. SHP was partially supported by NSF AF awards CCF-0915984 and CCF-1217462.}} \author{% Otfried Cheong% \thanks{Department of Computer Science, KAIST, Daejeon, Korea. \si{otfried}\symbol{'100}{}kaist.edu.}% \and Sariel Har-Peled% \thanks{% Department of Computer Science, University of Illinois, Urbana, USA. \si{sariel}\symbol{'100}{}\si{illinois.edu}.} \and Heuna Kim% \thanks{Freie \si{Universit\"at} Berlin, Berlin, Germany. heunak\symbol{'100}{}mi.fu-berlin.de.}% \and Hyo-Sil Kim% \thanks{Department of Computer Science and Engineering, POSTECH, Pohang, Korea. hyosil.kim@\symbol{'100}{}gmail.com.}} \maketitle \begin{abstract} A graph drawn in the plane with $n$ vertices is \emphi{$k$-fan-crossing free} for $k \geq 2$ if there are no $k+1$ edges $g,e_1,\dots, e_k$, such that $e_1,e_2,\dots,e_k$ have a common endpoint and $g$ crosses all~$e_i$. We prove a tight bound of $4n-8$ on the maximum number of edges of a $2$-fan-crossing free graph, and a tight $4n-9$ bound for a straight-edge drawing. For $k \geq 3$, we prove an upper bound of $3(k-1)(n-2)$ edges. We also discuss generalizations to monotone graph properties. \end{abstract} \section{Introduction} A \emphi{topological graph}~${G}$ is a graph drawn in the plane: vertices are points, and the edges of the graph are drawn as Jordan curves connecting the vertices. Edges are not allowed to pass through vertices other than their endpoints. We will assume the topological graph to be \emphi{simple}, that is, any pair of its edges have at most one point in common (so edges with a common endpoint do not cross, and edges cross at most once). Figure~\ref{fig:bad}~(a--b) shows configurations that are not allowed. If there are no crossings between edges, then the graph is planar, and Euler's formula implies that it has at most $3n-6$ edges, where $n$ is the number of vertices. What can be said if we relax this restriction---that is, we permit some edge crossings? For instance, a topological graph is called \emphi{$k$-planar} if each edge is crossed at most $k$~times. Pach and T\'{o}th\xspace~\cite{pt-gdfce-97} proved that a $k$-planar graph on~$n$ vertices has at most $(k+3)(n-2)$ edges for $0 \leq k \leq 4$, and at most~$4.108 \sqrt{k}n$ edges for general~$k$. The special case of $1$-planar graphs has recently received some attention, especially in the graph drawing community. Pach and T\'{o}th\xspace's bound is~$4n-8$, and this is tight: starting with a planar graph~$H$ where every face is a quadrilateral, and adding both diagonals results in a $1$-planar graph with~$4n-8$ edges. However, Didimo~\cite{d-dslop-13} showed that \emph{straight-line} $1$-planar graphs have at most~$4n-9$ edges, showing that F\'{a}{r{}y}'s theorem does not generalize to $1$-planar graphs. This bound is tight, as he constructed an infinite family of straight-line $1$-planar graphs with $4n-9$~edges. Hong {et~al.}\xspace~\cite{help-ftopg-12} characterize the $1$-planar graphs that can be drawn as straight-line $1$-planar graphs. Grigoriev and Bodlaender~\cite{gb-agefc-07} showed that testing if a given graph is $1$-planar is NP-hard. A topological graph is called \emphi{$k$-quasi planar} if it does not contain $k$~pairwise crossing edges. It is conjectured that for any fixed~$k$ the number of edges of a $k$-quasi planar graph is linear in the number of vertices~$n$. Agarwal {et~al.}\xspace~\cite{aapps-qpglne-1997} proved this for straight-line $3$-planar graphs, Pach {et~al.}\xspace~\cite{prt-rptg-02} for general $3$-planar graphs, Ackerman~\cite{a-mnetg-09} for~$4$-planar graphs, and Fox {et~al.}\xspace~\cite{fp-nekqp-13} prove a bound of the form~$O(n \log^{1+o(1)}n)$ for $k$-planar graphs. A different restriction on crossings arises in graph drawing: Humans have difficulty reading graph drawings where edges cross at acute angles, but graph drawings where edges cross at right angles are nearly as readable as planar ones. A \emph{right-angle crossing graph} (RAC graph) is a topological graph with straight edges where edges that cross must do so at right angle. Didimo {et~al.}\xspace~\cite{del-dgrac-11} showed that an RAC graph on $n$~vertices has at most~$4n-10$ edges. Testing whether a given graph is an RAC graph is NP-hard~\cite{abs-slrdp-12}. Eades and Liotta~\cite{el-racgo-13} showed that an extremal RAC graph, that is, an RAC graph with $n$ vertices and $4n - 10$ edges, is $1$-planar, and is the union of two maximal planar graphs sharing the same vertex set. A \emphi{radial $(p,q)$-grid} in a graph~${G}$ is a set of $p+q$ edges such that the first $p$~edges are all incident to a common vertex, and each of the first $p$~edges crosses each of the remaining $q$~edges. Pach {et~al.}\xspace \cite{ppst-tgnlg-05} proved that a graph without a radial $(p,q)$-grid, for $p,q \geq 1$, has at most $8 \cdot 24^{q} pn$ edges. We will call a radial $(k,1)$-grid a \emphi{$k$-fan crossing}. In other words, a fan crossing is formed by an edge~$g$ crossing $k$~edges~$e_1,\dots,e_k$ that are all incident to a common vertex, see Figure~\ref{fig:bad}~(c). A topological graph is \emphi{$k$-fan-crossing free} if it does not contain a $k$-fan crossing. We are particularly interested in the special case $k=2$. For brevity, let us call a $2$-fan crossing simply a \emphi{fan crossing} (shown in Figure~\ref{fig:bad}~(d)), and a $2$-fan-crossing free graph a \emphi{fan-crossing free graph}. \begin{figure}[t] \centerline{% \hfill% \includegraphics[page=1]{figs/not_allow}% \hfill \includegraphics[page=2]{figs/not_allow} \hfill \includegraphics[page=2]{figs/fan_crossing} \hfill \includegraphics[page=1]{figs/fan_crossing}% \hfill} \caption{(a) and (b) shows illegal embeddings of edges of a graph. (c) is a $4$-fan crossing, (d) a fan crossing.} \label{fig:bad} \end{figure} By Pach {et~al.}\xspace{}'s result, a $k$-fan-crossing free graph on $n$~vertices has at most $192kn$~edges, and a fan-crossing free graph has therefore at most $384n$~edges. We improve this bound by proving the following theorem. \begin{theorem}% \label{theo:main}% A fan-crossing free graph on~$n \geq 3$ vertices has at most~$4n-8$ edges. If the graph has straight edges, it has at most $4n-9$ edges. Both bounds are tight for~$n \geq 10$. \end{theorem} A $1$-planar graph is fan-crossing free, so Theorem~\ref{theo:main} generalizes both Pach and T\'{o}th\xspace's and Didimo's bound. We also extend their lower bounds by giving tight constructions for every value of~$n$. In an RAC graph all edges crossed by a given edge~$g$ are orthogonal to it and therefore parallel to each other, implying that an RAC graph is fan-crossing free. Our theorem, therefore, ``nearly'' implies Didimo {et~al.}\xspace's bound: a fan-crossing free graph has at most one edge more than an RAC graph. We can completely characterize extremal fan-crossing free graphs, that is, fan-crossing free graphs on $n$~vertices with $4n-8$ edges: Any such graph consists of a planar graph~$H$ where each face is a quadrilateral, together with both diagonals for each face. This implies the same properties obtained by Eades and Liotta for extremal RAC graphs: An extremal fan-crossing free graph is $1$-planar, and is the union of two maximal planar graphs. For $k$-fan-crossing free graphs with $k \geq 3$, we obtain the following result. \begin{theorem} \label{theo:k-main}% A $k$-fan-crossing free graph on~$n \geq 3$ vertices has at most~$3(k-1)(n-2)$ edges, for $k \geq 3$. \end{theorem} This bound is not tight, and the best lower-bound construction we are aware of has only about $kn$~edges. Most of the graph families discussed above have a common pattern: the subgraphs obtained by taking the edges crossed by a given edge~${{e}}$ may not contain some forbidden subgraph. We can formalize this notion as follows: For a topological graph~${G}$ and an edge ${{e}}$ of~${G}$, let~${G}_{{e}}$ denote the subgraph of ${G}$ containing exactly those edges that cross~${{e}}$. A graph property~$\mathcal{P}$ is called \emphi{monotone} if it is preserved under edge-deletions. In other words, if ${G}$ has $\mathcal{P}$ and ${G}'$ is obtained from ${G}$ by deleting edges, then ${G}'$ must have~$\mathcal{P}$. Given a monotone graph property~$\mathcal{P}$, we define a \emphi{derived graph property} $\mathcal{P}^{*}$ as follows: A topological graph ${G}$ has $\mathcal{P}^{*}$ if for every edge~${{e}}$ of~${G}$ the subgraph~${G}_{{e}}$ has~$\mathcal{P}$. Some examples are: \begin{compactitem} \item If $\mathcal{P}$ is the property that a graph does not contain a path of length two, then $\mathcal{P}^{*}$ is the property of being fan-crossing free; \item if $\mathcal{P}$ is the property of having at most $k$ edges, then $\mathcal{P}^{*}$ is $k$-planarity; \item if $\mathcal{P}$ is planarity, then $\mathcal{P}^{*}$ is 3-quasi-planarity. \end{compactitem} We can consider $\mathcal{P}^{*}$ for other interesting properties~$\mathcal{P}$, such as not containing a path of length~$k$, or not containing a~$K_{2,2}$. We prove the following very general theorem: \begin{theorem} \label{theo:gmain}% Let $\mathcal{P}$ be a monotone graph property such that any graph on $n$ vertices that has~$\mathcal{P}$ has at most $O(n^{1+\alpha})$ edges, for a constant $0 \leq \alpha \leq 1$. Let ${G}$ be a graph on $n$ vertices that has $\mathcal{P}^{*}$. If $\alpha > 0$, then ${G}$ has $O(n^{1+\alpha})$ edges. If $\alpha = 0$, then ${G}$ has $O(n \log^{2} n)$ edges. \end{theorem} This immediately covers many interesting cases. For instance, a graph where no edge crosses a path of length~$k$, for a constant~$k$, has at most $O(n \log^{2} n)$ edges. Graphs where no edge crosses a $K_{2,2}$ have at most $\Theta(n^{3/2})$ edges (and this is tight, as there are graphs with $\Theta(n^{3/2})$ edges that do not contain a~$K_{2,2}$, implying that no edge can cross a~$K_{2,2}$). \paragraph{Paper organization.} Section~\ref{sec:combi} tackles the problem in the simplest settings involving a single simply-connected ``face'' of a fan-crossing free graph. Section~\ref{sec:upperbound} extends this argument to the whole graph. Section~\ref{sec:lowerbound} describes lower bound constructions, and the straight-line case. In Section~\ref{sec:k-geq-3}, the argument is extended to the $k$-fan case. Section~\ref{sec:general} proves the bound for the case where we consider the forbidden structure to be a hereditary property defined on the intersection graphs induced by the edges. Finally, Section~\ref{sec:conclusions} ends the paper with a discussion and some open problems. \section{A combinatorial puzzle: Arrows and fans} \label{sec:combi} At the core of our bound lies a combinatorial question that we can express as follows: An \emphi{$m$-star} is a regular $m$-gon~$\face$ with a set of \emphi{arrows}. An arrow is a ray starting at a vertex of~$\face$, pointing into the interior of~$\face$, and exiting through an edge of~$\face$. \begin{figure}[t] \centerline{\hfill\includegraphics{figs/mstar}\hfill \includegraphics{figs/k3}\hfill} \caption{Left: A $7$-star. Right: A $3$-star has at most one arrow.} \label{fig:mstar} \end{figure} We require the set of edges and arrows to be \emphi{fan-crossing free}---that is, no edge or arrow intersects two arrows or an edge and an arrow incident to the same vertex. The left side of Figure~\ref{fig:mstar} shows a $7$-star. The dashed arrows are impossible---each of them forms a fan crossing with the solid edges and arrows. The question is: \emphi{How many arrows can an $m$-star possess?} \begin{observation} \label{observation:3star} A $3$-star has at most one arrow. \end{observation} \begin{proof} An arrow from a vertex has to exit the triangle through the opposing edge, so no vertex has two arrows. But two arrows from different vertices will also form a fan crossing, see the right side of Figure~\ref{fig:mstar}. \end{proof} It is not difficult to see that a $4$-star possesses at most~$2$ arrows. The reader may enjoy constructing $m$-stars with $2m-6$~arrows, for $m \geq 4$. We conjecture that this bound is tight. In the following, we will only prove a weaker bound that is sufficient to obtain tight results for fan-crossing free graphs. While we have posed the question in a geometric setting, it is important to realize that it is a \emphi{purely combinatorial} question. We can represent the $m$-star by writing its sequence of vertices and indicating when an arrow exits~$\face$. Whether or not three edges/arrows form a fan crossing can be determined from the ordering of their endpoints along the boundary of~$\face$ alone. Let $C={v}_1, \ldots, {v}_m$ be the sequence of vertices of~$\face$ in counter-clockwise order, such that the $i$\th boundary edge of~$\face$ is ${{e}}_i = {v}_i {v}_{i+1}$ (all indices are modulo~$m$). Consider an arrow~${{e}}$ starting at~${v}_i$. It exits~$\face$ through some edge~${{e}}_j$, splitting~$\face$ into two chains ${v}_{i+1}\dots{v}_j$ and ${v}_{j+1}\dots{v}_{i-1}$. The \emphi{length} of ${{e}}$ is the number of vertices on the shorter chain. We will call an arrow \emphi{short} if it has length one. A \emphi{long arrow} is an arrow of length larger than one. \begin{lemma} \label{lemma:kgon} For $m \geq 4$, an $m$-star~$\face$ has at most $2m - 8$ long arrows. \end{lemma} \begin{proof} The proof is by induction over~$m$. Any arrow in a $4$-star partitions the boundary into chains of length one and length two, and so there are no long arrows, proving the claim for $m = 4$. We suppose now that $m > 4$ and that the claim holds for all $4 \leq m' < m$. We \emph{delete all short arrows}, and let $L$ denote the remaining set of arrows, all of which are now long arrows. Let ${{e}}$ be an arrow of \emph{shortest length}~$\ell$ in~$\face$. Without loss of generality, we assume that ${{e}}$ starts in~${v}_1$ and exits through edge $e_{\ell+1} = {v}_{\ell+1} {v}_{\ell+2}$. Then, the following properties hold (see~Figure~\ref{fig:kgon}): \begin{figure}[t] \centerline{% \hfill% \includegraphics{figs/kgon}% \hfill% \includegraphics{figs/k_con}% \hfill% } \caption{Left: properties~(A) to~(F), right: property~(G) for the proof of~Lemma~\ref{lemma:kgon}.} \label{fig:kgon} \end{figure} \begin{compactenum}[(A)] \item Every arrow starting in ${v}_2,\ldots, {v}_{\ell+1}$ must cross ${{e}}$, as otherwise it would be shorter than ${{e}}$. \item There is no arrow that starts in $v_{\ell+1}$. By~(A), such an arrow must cross~${{e}}$, and so it forms a fan crossing with ${{e}}$ and ${{e}}_{\ell+1}$. \item At most one arrow starts in ${v}_i$, for $i = 2, \ldots, \ell$. Indeed, two arrows starting in ${v}_i$, for $i=2, \ldots,\ell$, must cross ${{e}}$ by~(A), and so they form a fan crossing with ${{e}}$. \item No arrow starting in~${v}_{\ell+2}$ exits through ${{e}}_2, \ldots, {{e}}_\ell$, as then it would be shorter than ${{e}}$. \item An arrow starting in~${v}_{\ell+2}$ and exiting through~${{e}}_1$ cannot exist either, as it forms a fan crossing with ${{e}}$ and ${{e}}_1$. \item No arrow starting in~${v}_m$ crosses ${{e}}_1, \ldots, {{e}}_{\ell-1}$, as then it would be shorter than ${{e}}$. \item The following two arrows cannot both exist: An arrow ${{e}}'$ starting in~${v}_m$ and exiting through~${{e}}_\ell$, and an arrow~${{e}}''$ starting in~${v}_\ell$. Indeed, if both ${{e}}'$ and ${{e}}''$ are present, then either~$e''$ exits through~$e_{m}$ and forms a fan crossing with $e$ and~$e_{m}$, or $e''$ intersects~$e'$ and so $e'$, $e''$, and $e_{\ell}$ form a fan crossing (see the right side of~Figure~\ref{fig:kgon}). \end{compactenum} We now create an $(m-\ell+1)$-star $\faceA$ by removing the vertices ${v}_2\dots{v}_\ell$ with all their incident arrows from~$\face$, such that ${v}_1$ and ${v}_{\ell+1}$ are consecutive on the boundary of~$\faceA$. An arrow that exits~$\face$ through one of the edges ${{e}}_1\dots{{e}}_\ell$ exits~$\faceA$ through the new edge~$g = {v}_1{v}_{\ell+1}$. Let $L' \subset L$ be the set of arrows of~$\faceA$, that is, the arrows of~$\face$ that do not start from~${v}_2\dots{v}_\ell$. Among the arrows in~$L'$, there are one or two short arrows: the arrow~${{e}}$, and the arrow~${{e}}'$ starting in~${v}_m$ and exiting through~${{e}}_\ell$ in~$\face$ (and therefore through~$g$ in~$\faceA$) if it exists. We set $q = 1$ if ${{e}}'$ exists, and else $q=0$. We delete from~$\faceA$ those one or two short arrows, and claim that there is now no fan crossing in~$\faceA$. Indeed, a fan crossing would have to involve the new edge~$g = {v}_1{v}_{\ell+1}$. But any arrow that crosses~$g$ must also cross~${{e}}$, and there is no arrow starting in~${v}_{\ell+1}$ by~(B). Since $\ell \geq 2$, we have $m - \ell + 1 < m$, and so by the inductive assumption $\faceA$ has at most $2(m - \ell + 1) - 8 = 2m - 2\ell -6$ long arrows. Since there are $1+q$ short arrows in~$L'$, we have $|L'| \leq 2m - 2\ell - 5 +q $. By~(C) and~(G), we have $|L| - |L'| \leq \ell-1 - q$. It follows that \begin{align*} |L| & \leq |L'| + \ell - 1 - q \leq 2m - \ell - 6 \leq 2m - 8. \ifmmode\qed\else\hfill\proofSymbol\fi \end{align*} \par\vspace{-\baselineskip} \end{proof} \parpic[l]{\includegraphics{figs/len1}} It remains to count the short arrows. Let $e$ be a short arrow, say starting in~${v}_i$ and exiting through~${{e}}_{i+1}$. Let us call~${v}_{i+1}$ the \emphi{witness} of~${{e}}$. We observe that no arrow~${{e}}'$ can start in this witness---${{e}}'$ would form a fan crossing with~${{e}}$ and~${{e}}_{i+1}$. The vertex~${v}_{i+1}$ can serve as the witness of only one short arrow: The only other possible short arrow~${{e}}''$ with witness~${v}_{i+1}$ starts in~${v}_{i+2}$ and exits through~${{e}}_i$. However, ${{e}}$, ${{e}}''$, and~${{e}}_i$ form a fan crossing. \picskip{0} We can now bound the number of arrows of an~$m$-star. \begin{lemma} \label{lemma:mstar} For $m \geq 3$, an $m$-star~$\face$ has at most $3m - 8$ arrows. The bound is attained only for $m = 3$. \end{lemma} \begin{proof} By Observation~\ref{observation:3star}, the claim is true for $m = 3$. We consider $m > 3$. By Lemma~\ref{lemma:kgon}, there are at most~$2m-8$ long arrows. Each short arrow has a unique witness. If all vertices are witnesses then there is no arrow, and so we can assume that at most $m-1$ vertices serve as witnesses, and we have at most $m-1$ short arrows, for a total of~$3m-9$ arrows. \end{proof} \section{The upper bound for fan-crossing free graphs} \label{sec:upperbound} \parpic[l]{\includegraphics{figs/easy}} \noindent Let ${G}=({V},{E})$ be a fan-crossing free graph. We fix an arbitrary maximal planar subgraph ${H} = ({V},{E}')$ of ${G}$. Let~$K = {E} \setminus {E}'$ be the set of edges of~${G}$ that is not in~${H}$. Since ${H}$ is maximal, every edge in~$K$ must cross at least one edge of~${H}$. We will replace each edge of~$K$ by \emph{two arrows}. Let ${{e}}\in K$ be an edge connecting vertices~${v}$ and~${u}$. The initial segment of~${{e}}$ must lie inside a face~$\face$ of~${H}$ incident to~${v}$, the final segment must lie inside a face~$\faceA$ of~${H}$ incident to~${u}$. It is possible that $\face = \faceA$, but in that case the edge~${{e}}$ does not entirely lie in the face, as ${H}$ is maximal. We replace~${{e}}$ by two arrows: one arrow starting in~${v}$ and passing through~$\face$ until it exits~$\face$ through some edge; another arrow starting in~${u}$ and passing through~$\faceA$ until it exits~$\faceA$ through some edge. In this manner, we replace the set of edges~$K$ by a set of~$2|K|$ arrows. The result is a planar graph whose faces have been adorned with arrows. The collection of edges and arrows is fan-crossing free. Every edge of~${H}$ is incident to two faces of~${H}$, which can happen to be identical. If we distinguish the sides of an edge, the boundary of each face~$\face$ of ${H}$ consists of simple chains of edges. If $\face$ is bounded, one chain forms the outer boundary of~$\face$ that makes $\face$~bounded, while all other chains bound holes inside~$\face$; if $\face$ is unbounded, then all chains bound holes in~$\face$. \parpic[l]{\includegraphics{figs/connec}} If the graph~${H}$ is connected, then the boundary of each face consists of a single chain. Let $\face$ be such a face whose boundary chain consists of $m$~edges (where edges that bound~$\face$ on both sides are counted twice). Then $\face$ has at most $3m-8$ arrows. This follows immediately from Lemma~\ref{lemma:mstar}: Recall that $m$-stars can be defined purely combinatorially. Whether three edges form a fan crossing can be decided solely by the ordering of their endpoints along the boundary chain. The boundary of a simply connected face is a single closed chain, and so Lemma~\ref{lemma:mstar} applies to this setting, see the side figure. \picskip{0} Unfortunately, we cannot guarantee that ${H}$ is connected. The following lemma bounds the number of arrows of a face~$\face$ in terms of its complexity and its number of boundary chains. The complexity of a face is the total number of edges of all its boundary chains, where edges that are incident to the face on both sides are counted twice. \begin{lemma} \label{lemma:face} A face of~${H}$ of complexity~$m$ bounded by~$p$ boundary chains possesses at most $3m + 8p - 16$ arrows. The bound can be attained only when $m=3$ and~$p=1$. \end{lemma} We will prove the lemma below, but let us first observe how it implies the upper bound on the number of edges of fan-crossing free graphs. \begin{lemma} \label{lemma:upper} A fan-crossing free graph~${G}$ on~$n$ vertices has at most $4n-8$ edges. \end{lemma} \begin{proof} Let $m$ be the number of edges, let $r$ be the number of faces, and let $p$ be the number of connected components of~${H}$. Let ${\mathcal{F}}$ be the set of faces of~${H}$. For a face $\face \in {\mathcal{F}}$, let $m(\face)$ denote the complexity of~$\face$, let $p(\face)$ denote the number of boundary chains of~$\face$, and let~$a(\face)$ denote the number of arrows of~$\face$. We have $\sum_{\face \in {\mathcal{F}}}m(\face) = 2m$ and $\sum_{\face\in{\mathcal{F}}}(p(\face)-1) = p - 1$ (each component is counted in its unbounded face, except that we miss one hole in the global unbounded face). The graph~${G}$ has $|E| = m + |K|$ edges. Using Lemma~\ref{lemma:face} we have \begin{align*} 2|E| & = 2m + 2|K| = \sum_{\face \in {\mathcal{F}}}m(\face) + \sum_{\face \in {\mathcal{F}}} a(\face) \\ & \leq \sum_{\face \in {\mathcal{F}}} \big(4m(\face) + 8p(\face) - 16\big) \\ & = 4\sum_{\face\in{\mathcal{F}}}m(\face) + 8\sum_{\face\in{\mathcal{F}}} (p(\face) - 1) - 8r\\ & = 8m + 8p - 8 - 8r. \intertext{By Euler's formula, we have $n - m + r = 1 + p$, so $m - r = n - 1 - p$, and we have} 2|E| & \leq 8(m-r) + 8p - 8 = 8n - 8 - 8p + 8p - 8 = 8n - 16. \ifmmode\qed\else\hfill\proofSymbol\fi \end{align*} \par\vspace{-\baselineskip} \end{proof} It remains to fill in the missing proof. \begin{proof}[Proof of Lemma~\ref{lemma:face}] Let $\face$ be a face of~${H}$, and let $m = m(\face)$ and $p = p(\face)$ be its complexity and its number of boundary components. A boundary component is a chain of edges, and could possibly degenerate to a single isolated vertex. We say that two boundary chains $\xi$ and $\zeta$ are \emphi{related} if an arrow starting in a vertex of~$\xi$ ends in an edge of~$\zeta$, or vice versa. Consider the undirected graph whose nodes are the boundary chains of~$\face$ and whose arcs connect boundary chains that are related. If this graph has more than one connected component, we can bound the number of arrows separately for each component, and so in the following we can assume that all boundary chains are (directly or indirectly) related. \begin{figure}[ht] \centerline{\hfill\includegraphics[page=1]{figs/bridge}\hfill \includegraphics[page=2]{figs/bridge}\hfill} \caption{Building a bridge between~$\xi$ and~$\zeta$.} \label{fig:bridge} \end{figure} Consider two related boundary chains~$\xi$ and~$\zeta$. By assumption there must be an arrow~${{e}}$, starting at a vertex~${v} \in\xi$, and ending in an edge~${u}_1{u}_2$ of~$\zeta$. We create a new vertex~$z$ on~$\zeta$ at the intersection point of ${{e}}$ and~${u}_1{u}_2$, split the boundary edge ${u}_1{u}_2$ into two edges ${u}_1z$ and $z{u}_2$, and insert the two new boundary edges~${v} z$ and~$z {v}$, see Figure~\ref{fig:bridge}. This operation has increased the complexity of~$\face$ by three. Note that some arrows of $\face$ might be crossing the new boundary edges---these arrows will now be shortened, and end on the new boundary edge. The two boundary chains~$\xi$ and~$\zeta$ have now merged into a single boundary chain. In effect, we have turned an arrow into a ``bridge'' connecting two boundary chains. No fan crossing is created, since all edges and arrows already existed. We do create a new vertex~$z$, but no arrow starts in~$z$, and so this vertex cannot cause a fan crossing. We insert $p-1$ bridges in total and connect all~$p$ boundary chains. In this manner, we end up with a face~$\faceA$ whose boundary is a single chain consisting of $m' = m + 3(p - 1)$ edges. If $m' = 3$, then $\faceA$ has at most one arrow, by Observation~\ref{observation:3star}. This case happens only for $m = 3$ and $p = 1$, and is the only case where the bound is tight. If $m' > 3$, then we can apply~Lemma~\ref{lemma:kgon} to argue that~$\faceA$ has at most~$2m'- 8$ long arrows. To count the short arrows, we observe that the vertex~$z$ created in the bridge-building process cannot be the witness of a short arrow: such a short arrow would imply a fan crossing in the original face~$\face$. It is also not the starting point of any arrow. It follows that building a bridge increases the number of possible witnesses by only one (the vertex~${v}$ now appears twice on the boundary chain). There are thus at most $m + p - 1$ possible witnesses in~$\faceA$. However, if all of these vertices are witnesses, then there is no arrow at all, and so there are at most $m + p - 2$ short arrows. Finally, we converted $p - 1$ arrows of~$\face$ into bridges to create~$\faceA$. The total number of arrows of~$\face$ is therefore at most \begin{align*} 2m' - 8 + (m+p-2) + (p-1)% &=% 2(m + 3p - 3) - 8 + (m + p - 2) + (p - 1) \\ & =% 3m + 8p - 17. \ifmmode\qed\else\hfill\proofSymbol\fi \end{align*} \par\vspace{-\baselineskip} \par\vspace{-\baselineskip} \end{proof} \section{Lower bounds and the straight-line case} \label{sec:lowerbound} Consider a fan-crossing free graph~${G}$ with $4n-8$~edges. This means that the inequality in the proof of Lemma~\ref{lemma:upper} must be an equality. In particular, every face~$\face$ of~${H}$ must have exactly~$3m(\face) + 8p(\face) - 16$ arrows. By Lemma~\ref{lemma:face}, this is only possible if $\face$ is a triangle, and so we have proven \begin{lemma} \label{lemma:maximal} A fan-crossing free graph~${G}$ with $4n-8$ edges contains a planar triangulation~${H}$ of its vertex set. Each triangle of~${H}$ possesses exactly one arrow. \end{lemma} Note that the arrow must necessarily connect a vertex of the triangle with the opposite vertex in the triangle adjacent along the opposite edge, as otherwise, it forms a fan crossing, and so we get the following second characterization of extremal fan-crossing free graphs: \begin{lemma} \label{lemma:maximal-quad} A fan-crossing free graph~${G}$ with $4n-8$ edges contains a planar graph~$Q$ on its vertex set, where each face of~$Q$ is a quadrilateral. ${G}$ is obtained from~$Q$ by adding both diagonals for each face of~$Q$. \end{lemma} By Euler's formula, a planar graph~$Q$ on~$n$ vertices whose faces are quadrilaterals has $2n-4$ edges and $n-2$ faces. Adding both diagonals to each face of $Q$, we obtain a fan-crossing free graph~${G}$ with $4n-8$ edges. However, it turns out that this construction needs to be done carefully: Otherwise, diagonals of two distinct faces of~$Q$ can connect the same two vertices, and the result is not a simple graph. And indeed, it turns out that for $n\in \{7, 9\}$, no (simple) fan-crossing free graph with $4n-8$ edges exists! When $n \geq 8$ is a multiple of four, Figure~\ref{fig:quadrilaterals} shows planar graphs~$Q$ where every bounded face is a convex quadrilateral. Since all their diagonals are straight, no multiple edges can arise. Only the two diagonals of the unbounded face are not straight and need to be checked individually. We will return to other values of~$n$ below. \begin{figure}[ht] \centerline{\includegraphics[page=1]{figs/quadrilaterals}} \caption{Planar graphs with convex quadrilateral faces for $n \in \{8, 12, 16\ldots\}$.} \label{fig:quadrilaterals} \end{figure} Lemma~\ref{lemma:maximal} implies immediately that a fan-crossing free graph with $4n-8$ edges cannot exist if \emph{all} edges are straight: Since the unbounded face of~${H}$ is a triangle, it cannot contain a straight arrow, and so any fan-crossing free graph drawn with straight edges has at most $4n-9$ edges. This bound is tight: for any~$n \geq 6$, we can construct a planar graph~$Q$ such that two faces of~$Q$ are triangles, and all other faces are convex quadrilaterals. Euler's formula implies that $Q$ has $2n-3$ edges and $n-1$ faces, and adding both diagonals to each quadrilateral face results in a graph with $2n-3 + 2(n-3) = 4n-9$ edges. The construction of~$Q$ is shown in Figure~\ref{fig:triangles}. The upper row shows the construction when $n \geq 6$ is a multiple of three. If $n \equiv 1 \pmod{3}$, we replace the two innermost triangles as shown in the lower left of the figure. If $n \equiv 2 \pmod{3}$, the two innermost triangles are replaced as in the lower right figure. \begin{lemma} \label{lemma:straight} A fan-crossing free graph drawn with straight edges has at most $4n-9$ edges. This bound is tight for $n\geq6$. \end{lemma} \begin{figure}[ht] \centerline{\includegraphics[page=2]{figs/quadrilaterals}} \caption{Adding both diagonals for each quadrilateral face results in a straight-line fan-crossing free graph with $4n-9$ edges.} \label{fig:triangles} \end{figure} We observe that the extremal fan-crossing free graph constructed for~$n=6$ has $15$~edges and is therefore the complete graph~$K_6$. It follows that the complete graph~$K_n$ is fan-crossing free for~$n \leq 6$, and it remains to discuss extremal fan-crossing free graphs for $n \geq 7$ when the edges are not necessarily straight. \begin{lemma} Extremal fan-crossing free graphs with $4n-8$ edges exist for $n = 8$ and all $n \geq 10$. For $n \in \{7, 9\}$, extremal fan-crossing free graphs have $4n-9$ edges. \end{lemma} \begin{proof} Let $n \in \{7, 9\}$, and assume ${G}$ is a fan-crossing free graph on~$n$ vertices with~$4n-8$ edges. By Lemma~\ref{lemma:maximal-quad}, ${G}$ contains a planar graph~$Q$ whose faces are quadrilaterals. We claim that $Q$ has a vertex~$v$ of degree two. This implies that the two quadrilaterals incident to~$v$ have diagonals connecting the same two vertices, a contradiction. It follows that a fan-crossing free graph with $4n-8$ edges does not exist. By Lemma~\ref{lemma:straight}, a fan-crossing free graph with $4n-9$ edges does exist. For $n = 7$, the claim is immediate: $Q$ has $2n-4=10$ edges, and so its average degree is $20/7 < 3$, and a vertex of degree two must exist. For $n = 9$, the total degree of~$Q$ is $2(2n-4) = 28$. We assume there is no vertex of degree two. Since nine vertices of degree three already contribute~$27$ to the total degree, it follows that there is one vertex~$w$ of degree four, while the other eight vertices have degree three. Since $Q$ contains only even cycles, it is a bipartite graph, and the vertices can be partitioned into two classes~$A$ and~$B$. Let $w \in A$, and let $k = |A| - 1$. The total degree of vertices in~$A$ and in~$B$ is identical, so we have $4 + 3k = 3(8-k)$, or $6k=20$, a contradiction. The case $n = 8$ was already handled in Figure~\ref{fig:quadrilaterals}, so consider $n \geq 10$. We will again start with a planar graph~$Q$ with quadrilateral faces. To avoid diagonals connecting identical vertices, we would like to make \emph{all} faces convex. This is obviously not possible when drawing the graph in the plane, and so we will draw the graph on the sphere, such that all faces are spherically-convex quadrilaterals. For even~$n$, we place a vertex at the North pole and at the South pole each. The remaining $n-2$ vertices form a zig-zag chain near the equator, distributing the points equally on two circles of equal latitude, see left-hand side of Figure~\ref{fig:spherical} (for~$n=10$). For odd~$n$, we place two vertices close to the North pole, one vertex at the South pole, and let the remaining~$n-3$ vertices again form a zig-zag chain near the equator, see the right-hand side of Figure~\ref{fig:spherical} (for $n = 11$). One of the resulting quadrilaterals is long and skinny, but it is still spherically convex. \begin{figure}[ht] \centerline{\includegraphics{figs/spherical}} \caption{Spherically-convex quadrilateralization for even~$n$ and odd~$n$.} \label{fig:spherical} \end{figure} \end{proof} \section[k-fan-crossing free graphs]{$k$-fan-crossing free graphs for $k \geq 3$} \label{sec:k-geq-3} Our proof of Theorem~\ref{theo:k-main} has the same structure as for the case~$k = 2$: Let $G = (V, E)$ be a $k$-fan-crossing free graph for~$k \geq 3$. We fix an arbitrary maximal planar subgraph~$H = (V, E')$ of~$G$, and let $K = E \setminus E'$. We replace each edge in~$K$ by two arrows, so we end up with a planar graph~$H$ whose faces have been adorned with $2|K|$~arrows in total. \begin{lemma} \label{lemma:k-face} A $k$-fan-crossing free face of~$H$ of complexity~$m \geq 3$ bounded by~$p$ boundary chains possesses at most $3(k-1)(m + 2p - 4) - 2m + 3$ arrows, for $k \geq 3$. \end{lemma} As in the case~$k=2$, we prove Lemma~\ref{lemma:k-face} by converting arrows connecting different boundary components of the face into bridges, until we have obtained a single boundary chain. The technical details are more complicated, and so we first discuss how Lemma~\ref{lemma:k-face} implies Theorem~\ref{theo:k-main}. \begin{proof}[Proof of Theorem~\ref{theo:k-main}] Let $m$ be the number of edges, let $r$ be the number of faces, and let $p$ be the number of connected components of~${H}$. Let ${\mathcal{F}}$ be the set of faces of~${H}$. For a face $\face \in {\mathcal{F}}$, let $m(\face)$ denote the complexity of~$\face$, let $p(\face)$ denote the number of boundary chains of~$\face$, and let~$a(\face)$ denote the number of arrows of~$\face$. As observed before, we have $\sum_{\face \in {\mathcal{F}}}m(\face) = 2m$ and $\sum_{\face\in{\mathcal{F}}}(p(\face)-1) = p - 1$. The graph~${G}$ has $|E| = m + |K|$ edges. Since $m(\face) \geq 3$, Lemma~\ref{lemma:k-face} implies $m(\face) + a(\face) \leq 3(k-1)(m(\face) + 2p(\face)-4)$, and so \begin{align*} 2|E| & = 2m + 2|K| = \sum_{\face \in {\mathcal{F}}} m(\face) + \sum_{\face \in {\mathcal{F}}} a(\face) \\\ & \leq \sum_{\face \in {\mathcal{F}}} \Big((3k-3)\big(m(\face) + 2(p(\face) - 1) - 2\big)\Big) \\ & = 3(k-1)\Big(\sum_{\face\in{\mathcal{F}}}m(\face) + 2\sum_{\face\in{\mathcal{F}}} (p(\face) - 1)- 2r\Big)\\ & = 6(k-1)(m + (p-1) - r) = 6(k-1)(n-2). \end{align*} In the last line we used Euler's formula $n - m + r = 1 + p$ for~$H$. \end{proof} To prove Lemma~\ref{lemma:k-face}, we need a $k$-fan equivalent of Lemma~\ref{lemma:mstar}. One additional difficulty is now keeping the contribution of the additional vertices formed by creating bridges low. We need a few new definitions. Again, an \emphi{$m$-star} is a regular $m$-gon~$\face$ with a set of arrows. We require the set of edges and arrows to be $k$-fan-crossing free. We number the vertices counter-clockwise as~$v_1, v_2, \dots, v_m$, and denote the edge between~$v_i$ and~$v_{i+1}$ as~$e_i$ (where index arithmetic is again modulo~$m$). We let $a_{ij}$ denote the number of arrows starting in vertex~$v_i$ and exiting through edge~$e_j$. The \emphi{degree}~$a_i$ of vertex~$v_i$ is the total number of arrows starting in~$v_i$. We define the length of an arrow and the notions of short and long arrows as before. We need to distinguish different kinds of vertices. If vertex~$v_i$ has degree~$a_i = 0$, we call it \emphi{void}. If, in addition, its left neighbor~$v_{i-1}$ has no short arrow passing over~$v_i$, that is, if $a_i = 0$ and $a_{i-1,i} = 0$, then we call it \emphi{left-light}. Similarly, if $a_i = 0$ and $a_{i+1,i-1} = 0$, then we call~$v_i$ \emphi{right-light}. A vertex is \emphi{light} when it is left-light or right-light. A vertex is \emphi{heavy} when its degree is non-zero. Finally, we do not allow a left-light vertex to be adjacent to a right-light vertex. What this means is that if a sequence of consecutive vertices, say~$v_1,v_2,\dots,v_t$ is a maximal sequence of zero-degree vertices, then either they are all left-light (if $a_{m,1} = 0$), or they are all right-light (if~$a_{t+1,t-1} = 0$), or only~$t-1$ of them are light and the last one is counted as void. We can now formulate our lemma. \begin{lemma} \label{lemma:k-mstar} For $m \geq 3$, $k \geq 3$, and $h\geq 2$, an $m$-star~$\face$ with $h$~heavy vertices, $\lambda$~light vertices, and~$\nu$~void (but not light) vertices has at most $(3k-5)h + k\lambda + (2k-3)\nu - (6k-9)$ arrows. \end{lemma} Before we prove Lemma~\ref{lemma:k-mstar}, let us first see how it implies Lemma~\ref{lemma:k-face}. \begin{proof}[Proof of Lemma~\ref{lemma:k-face}] We consider a face~$\face$ of~${H}$ of complexity~$m$ bounded by $p$~boundary components. As in the proof of Lemma~\ref{lemma:face}, we can assume that all boundary components are related. Thus there must be boundary components~$\xi, \zeta$ and an arrow starting at a vertex~$v \in \xi$ and ending in an edge~$u_1u_2$ of~$\zeta$. More precisely, there could be at most~$k-1$ such arrows starting in~$v$ and ending in~$u_1u_2$. We pick the two extreme arrows, that is, the one closest to~$u_1$ and the one closest to~$u_2$, and convert them to bridges, creating two new vertices~$z_1, z_2$ on $u_1u_2$, see Figure~\ref{fig:bridge_radial}. \begin{figure}[ht] \centerline{\hfill\includegraphics[page=1]{figs/bridge_radial}\hfill\ \includegraphics[page=2]{figs/bridge_radial}\hfill} \caption{Building a bridge between~$\xi$ and~$\zeta$.} \label{fig:bridge_radial} \end{figure} Inserting these two bridges increases the complexity of the face by three. However, $z_1$ and~$z_2$ are by construction light vertices (in Figure~\ref{fig:bridge_radial}, $z_1$~is right-light, $z_2$~is left-light). There are at most $k-1$ arrows from~$v$ to~$u_1u_2$ that are deleted from the face. We continue in this manner until the face is bounded by a single boundary chain. Every bridge adds one heavy vertex and two light vertices to the complexity of the face. (Note that later bridges could possibly create vertices on bridges built earlier, but that doesn't stop the vertices from being light.) We finally obtain a face with $m + p-1 \geq 2$ heavy vertices and $2(p-1)$ light vertices, and during the process we deleted at most $(k-1)(p-1)$ arrows. Applying Lemma~\ref{lemma:k-mstar}, the total number of arrows of~$\face$ is at most \begin{align*} & \hspace{-1cm}% (3k-5)(m + p - 1) + 2k(p-1) - (6k-9) + (k-1)(p-1) \\ & = (3k-5)m + (3k - 5 + 2k + k - 1)(p-1) - (6k-9) \\ & = (3k-3)m - 2m + (6k-6)(p-1) - (6k-6) + 3 \\ & = 3(k-1)(m + 2p - 4) - 2m + 3. \ifmmode\qed\else\hfill\proofSymbol\fi \end{align*} \par\vspace{-\baselineskip} \end{proof} Finally, it remains to prove the bound on the number of arrows of a $k$-fan-crossing free $m$-star. \begin{proof}[Proof of Lemma~\ref{lemma:k-mstar}] Let $A(h, \lambda, \nu)$ denote the maximum number of arrows of an $m$-star with $h$~heavy vertices, $\lambda$~light vertices, and $\nu$~void but not light vertices, where $m = h + \lambda + \nu$. We set \[ B(h, \lambda, \nu) = (3k-5)h + k\lambda + (2k-3)\nu - (6k-9), \] and show by induction over~$m$ that $A(h,\lambda, \nu) \leq B(h, \lambda, \nu)$ under the assumption that $m \geq 3$, $k \geq 3$, $h \geq 2$. We have several base cases. The reader may enjoy verifying that the claim holds for the triangle and the quadrilateral: \begin{align*} A(3, 0, 0) & = 3k-6 = B(3, 0, 0), \\ A(2, 0, 1) & = 2k -4 = B(2, 0, 1), \\ A(2, 1, 0) & = k - 1 = B(2, 1, 0), \\ A(4, 0, 0) & = 5k - 9 \leq 6k - 11 = B(4, 0, 0), \\ A(3, 0, 1) & = 4k - 6 \leq 5k - 9 = B(3, 0, 1), \\ A(3, 1, 0) & = 3k - 5 \leq 4k - 6 = B(3, 1, 0), \\ A(2, 0, 2) & = 4k-8 \leq 4k-7 = B(2,0,2), \\ A(2, 1, 1) & = 3k-5 \leq 3k-4 = B(2,1,1), \text{ and } \\ A(2, 2, 0) & = 2k-2 \leq 2k-1 = B(2,2,0). \end{align*} The second base case is when $m > 4$ and all vertex degrees are at most~$k-2$. In this case we have $A(h, \lambda, \nu) \leq (k-2)h$. If $h \geq 3$ then \begin{align*} (3k-5)h - (6k-9) - (k-2)h = (2k-3)(h-3) \geq 0, \end{align*} and the claim holds. If $h = 2$ then $\lambda+\nu > 2$ and so \begin{align*} A(2, \lambda, \nu) & \leq 2k - 4 \leq 2k-1 = B(2, 2, 0) \leq B(2, \lambda, \nu). \end{align*} The third base case is when $m > 4$ and $h = 2$, and the two heavy vertices are adjacent. Let $v_1$ and $v_2$ denote the two heavy vertices. No arrows start from the other vertices~$v_3,\ldots,v_m$. If no arrows starting from $v_1$ and $v_2$ intersect each other, then \begin{align*} A(2,\lambda,\nu) & \leq (k-1)(m-2) = (k-1)\lambda+(k-1)\nu \leq B(2,\lambda,\nu). \end{align*} Assume now that there is at least one arrow from $v_1$ that intersects an arrow from~$v_2$. Let $x_1$ denote the rightmost arrow from~$v_1$ and let $x_i$ (for $i > 1$) denote the $i$\th arrow from $x_1$ in counter-clockwise order. Similarly, let $y_1$ denote the leftmost arrow from~$v_2$ and let $y_i$ (for $i > 1$) denote the $i$\th arrow from~$v_2$ in clockwise order. We claim that only the arrows $x_1,\ldots,x_{k-1}$ from $v_1$ can intersect the arrows $y_1,\ldots,y_{k-1}$ from $v_2$. Indeed, $x_1$ cannot intersect more than $k-1$ arrows starting from $v_2$, since otherwise, it forms a $k$-fan crossing. If $x_1$ intersects $k-1$ arrows from $v_2$, it has to intersect $y_1,\ldots,y_{k-1}$. Since $x_1$ is the rightmost arrow from $v_1$, $y_j$ for $j \geq k$ must lie on the right side of~$x_1$, which implies that $y_j$ cannot intersect any arrow from~$v_1$. Similarly, since $y_1$ is the leftmost arrow from $v_2$, $x_j$ for $j \geq k$ cannot intersect any arrow from $v_2$, proving the claim. This implies that there can exist at most $(k-1)(m-2) + (k-1) = (k-1)(m-1)$ arrows in total, and so \begin{align*} & \hspace{-1cm}% [(3k-5)(2+\nu+\lambda)+(k-1)\nu+(2k-3)\lambda-(6k-9)] - (k-1)(1+\nu+\lambda) \\ & = (2k-4)(1+\nu+\lambda) + (k-1)\nu+(2k-3)\lambda - (3k-4) \geq 0, \end{align*} for $k \geq 3$ and $\nu+\lambda \geq 3$, and thus $A(2,\nu,\lambda) \leq B(2,\nu,\lambda)$. We now turn to the inductive step. Consider an $m$-star~$\face$ with $A(h, \lambda, \nu)$ arrows, where $m > 4$, $h \geq 2$, at least one vertex has degree at least~$k-1$, and if $h = 2$ then the two heavy vertices are not adjacent. We consider first the case where there are no long arrows: all arrows are short. By renumbering, we can assume that $v_2$ has degree $a_2 \geq k-1$. This implies $a_{13} = 0$ and $a_{31} = 0$, see Figure~\ref{fig:radial_free_short}. \begin{figure}[t] \centerline{\includegraphics{figs/radial_free_short}} \caption{When all arrows are short.} \label{fig:radial_free_short} \end{figure} We now construct an $(m-1)$-star~$\faceA$ by deleting~$v_2$ and its arrows, and inserting the edge~$g = v_1v_3$. Since $\face$ has only short arrows, the only arrows intersecting $g$ are~$a_{m1} \leq k-1$ arrows starting in~$v_m$ and $a_{42} \leq k-1$ arrows starting in~$v_4$. Since $m > 4$ the vertices $v_m$ and $v_4$ are distinct, and this implies that $\faceA$ has no $k$-fan-crossing. The vertex~$v_2$ is obviously heavy and has $a_2 \leq 2(k-1)$. If $v_3$ is left-light in~$\face$, it is still left-light in~$\faceA$ since~$a_{13} = 0$. Similarly, if $v_1$ is right-light in~$\face$, it remains so in~$\faceA$. (It cannot happen that $v_1$ is right-light and $v_3$~is left-light in~$\face$, as then $a_2 = 0 < k-1$.) We thus have \begin{align*} A(h, \lambda, \nu) & \leq 2k-2 + B(h-1, \lambda, \nu) \leq B(h, \lambda, \nu). \end{align*} It remains to consider the case where $\face$ has long arrows. Let $e$ denote the \emph{shortest} long arrow in~$\face$ and let $\ell$ denote its length. Renumbering the vertices we can assume that $e$~starts in~$v_1$ and exits through edge~$e_{\ell+1} = v_{\ell+1}v_{\ell+2}$. See Figure~\ref{fig:radial_free_long}. \begin{figure}[t] \centerline{\includegraphics{figs/radial_free_long}} \caption{$e$ is the shortest long arrow.} \label{fig:radial_free_long} \end{figure} The following property holds: \begin{compactenum} \item[(A)] Every long arrow starting in $v_2,\ldots,v_{\ell+1}$ must cross $e$, as otherwise it would be shorter than $e$. \end{compactenum} Let $a_{2e}$ denote the number of arrows starting in~$v_2$ that cross~$e$. These arrows cannot form a $k$-fan, so we have $a_{2e} \leq k-1$. An arrow starting in~$v_2$ that does not cross $e$ is short by~(A) and must exit through~$e_3$. This implies that $a_{2} = a_{2e} + a_{23} \leq 2(k-1)$. We now create an $(m-1)$-star~$\faceA$ by deleting~$v_2$ and all its arrows, deleting the $a_{12}$ arrows starting in~$v_1$ and exiting through~$e_3$, deleting the $a_{31}$ arrows starting in~$v_3$ and exiting through~$e_1$, and inserting the new edge~$g=v_1v_3$. If $\faceA$ has a $k$-fan crossing, it must involve the new edge~$g$. There are three ways in which this could happen: \begin{compactenum}[\qquad \bf {Case}~(a):] \item $g$ and $k-1$ arrows starting in $v_1$ are intersected by an arrow~$x$ that also intersects~$e_2$. See left side of Figure~\ref{fig:create-fan}. \item $g$ and $k-1$ arrows starting in $v_3$ are intersected by an arrow~$y$ that also intersects $e_1$. See right side of Figure~\ref{fig:create-fan}. \item $g$ intersects $k$ arrows starting in the same vertex. \end{compactenum} \begin{figure}[t] \centerline{% \hfill% \includegraphics{figs/radial_free_long_x}% \hfill% \includegraphics{figs/radial_free_long_y}% \hfill% } \caption{Two cases where $g$ creates a $k$-fan in~$\faceA$.} \label{fig:create-fan} \end{figure} We first observe that case~(c) cannot happen: If the $k$ arrows intersecting $g$ also intersect~$e$, they form a $k$-fan crossing in $\face$. If there is an arrow that does not intersect~$e$, it must start in one of $v_4,\ldots,v_{\ell+1}$, and so it is short by~(A). To intersect $g$, these arrows must start in $v_4$ and exit through $e_2$, and there are at most $(k-1)$ such arrows. If case~(a) happens, we delete from~$\faceA$ one more arrow, namely, the ``lowest'' arrow starting in~$v_1$. If case~(b) happens, we take out the ``lowest'' arrow starting in $v_3$. This ensures that $\faceA$ has no $k$-fan crossing. The inductive assumption holds for~$\faceA$, since it has $m-1$~vertices and at least two heavy vertices. The latter follows since $v_1$ has at least the arrow~$e$ and is therefore heavy, so $v_1$ and~$v_2$ cannot be the two only heavy vertices of~$\face$. We set $t_x = 1$ if case~(a) happens, and $t_x = 0$ otherwise; similarly we set $t_y = 1$ if case~(b) happens and $t_y = 0$ otherwise. Or original face~$\face$ has \[ \Delta := a_{2e} + a_{23} + a_{12} + a_{31} + t_x + t_y \] arrows more than the new face~$\faceA$. We collect a few more properties: \begin{compactenum} \item[(B)] If $a_{23} = k-1$ then case~(a) cannot happen: Since $x$ exits through~$e_2$, it intersects the $k-1$ arrows starting in~$v_2$ and exiting through~$e_3$, and in addition the edge~$e_2$ incident to~$v_2$. \item[(C)] If $a_{23} \geq 1$ then case~(b) cannot happen: An arrow starting in~$v_2$ and exiting through~$e_3$ intersects all arrows starting in~$v_3$ as well as the edge~$e_3$, and so $a_3 < k-1$. But $v_3$ needs to have~$k-1$ arrows for case~(b) to occur. \item[(D)] If $a_{31}=k-1$ then case~(a) cannot happen: Since $x$ exits through~$e_2$, it intersects the $k-1$ arrows starting in~$v_3$ and exiting through~$e_1$, and in addition the edge~$e_2$ incident to~$v_3$. \item[(E)] If $a_{12} \geq k-2$ then case~(b) cannot happen: If $y$ intersects~$e$, then it intersects~$e$, $e_1$, and $k-2$ arrows from~$v_1$ to~$e_2$, a $k$-fan crossing. If $y$ does not intersect~$e$, then $y$ must start in $v_2, \ldots, v_{\ell+1}$, and so by~(A), $y$~is short. To cross~$e_1$, $y$~must therefore start in~$v_3$, and so~$y$ does not cross~$g$. \end{compactenum} We now distinguish several cases to bound~$\Delta$: \paragraph{Case 1:} $a_{2} =a_{2e}+a_{23} \geq k-1$. In this case, $a_{12} = a_{31} = 0$, and so $\Delta = a_{2} + t_x + t_y$. Since $a_{2e} \leq k-1$, we have by~(B) and~(C): \[ \begin{array}{lccccccccl} & & & a_{2} & & t_x & & t_y & & \\ \textrm{If $a_{23} = k-1$:\qquad} & \Delta & \leq & 2(k-1) & + & 0 & + & 0 & = & 2k-2. \\ \textrm{If $1 \leq a_{23} \leq k-2$:\qquad} & \Delta & \leq & (2k-3) &+& 1 &+& 0 &=& 2k-2. \\ \textrm{If $a_{23} = 0$:\qquad} & \Delta & \leq &(k-1) &+& 1 &+& 1 &=& k+1 \leq 2k-2. \end{array} \] \paragraph{Case 2:} $a_{2} \leq k-2$ and $\max(a_{12}, a_{31})=k-1$. If $a_{12} = k-1$ then $a_{31} = 0 $, and if $a_{31}=k-1$ then $a_{12}=0$. By (D) and (E), we have \begin{align*} \Delta & = a_{2} + (a_{12}+a_{31}) + (t_x + t_y) \leq (k-2) + (k-1) + 1 = 2k-2. \end{align*} \paragraph{Case 3:} $\max\{a_{2},a_{12},a_{31}\} \leq k-2$. By~(E), $a_{12} = k-2$ implies $t_y = 0$, and so \[ \begin{array}{lccccccccccccl} & & & a_{2} & & a_{12} & & a_{31} & & t_x & & t_y & & \\ \text{If $a_{12} = k-2$:} & \Delta & \leq & k-2 &+& k-2 &+& k-2 &+& 1 &+& 0 &=& 3k-5. \\ \text{If $a_{12} < k-2$:} & \Delta & \leq & k-2 &+& k-3 &+& k-2 &+& 1 &+& 1 &=& 3k-5. \end{array} \] We now have all the ingredients to complete the inductive argument. Clearly $v_1$ is heavy because of the arrow~$e$. We distinguish the types of~$v_2$ and~$v_3$ (note that case~1 can occur only when~$v_2$ is heavy): \begin{itemize} \item If $v_2$ is heavy and $v_3$ is not left-light, then we have \begin{align*} A(h, \lambda, \nu) & \leq \max(2k-2, 3k-5) + B(h-1, \lambda, \nu) = B(h, \lambda, \nu). \end{align*} \item If $v_2$ is heavy and $v_3$ is left-light, then $v_3$ might become void in~$\faceA$. In this case~$a_{31} = a_{23} = 0$. The bound in case~3 improves to~$\Delta \leq 2k-3 \leq 2k-2$: \begin{align*} A(h, \lambda, \nu) & \leq 2k-2 + B(h-1, \lambda-1, \nu+1) = B(h, \lambda, \nu). \end{align*} \item If $v_2$ is right-light, then $a_2 = a_{31} = 0$. If $v_2$ is left-light and $v_3$ is not left-light, then $a_2 = a_{12} = 0$. Either way, in case~2 the bound improves to $\Delta \leq k$, and in case~3 it improves to $\Delta \leq k-1$. We have \begin{align*} A(h, \lambda, \nu) & \leq k + B(h, \lambda-1, \nu) = B(h, \lambda, \nu). \end{align*} \item If $v_2$ and $v_3$ are both left-light, then $v_3$ might become void in~$\faceA$. We have $a_2 = a_3 = a_{12} = 0$. We are thus in case~3 and have the improved bound $\Delta \leq 2$. This gives \begin{align*} A(h, \lambda, \nu) & \leq 2 + B(h, \lambda-2, \nu + 1) < B(h, \lambda, \nu). \end{align*} \item If $v_2$ is void and $v_3$ is not left-light, then $a_2 = 0$. In case~2 this implies~$\Delta \leq k$, in case~3 it implies $\Delta \leq 2k-3$. Since $k \leq 2k-3$ we have \begin{align*} A(h, \lambda, \nu) & \leq 2k - 3 + B(h, \lambda, \nu-1) = B(h, \lambda, \nu). \end{align*} \item Finally, if $v_2$ is void and $v_3$ is left-light, then $v_3$ might become void in~$\faceA$. We have $a_2 = a_3 = 0$. In case~2 this implies~$\Delta \leq k$, in case~3 it implies $\Delta \leq k-1$. We have \begin{align*} A(h, \lambda, \nu) & \leq k + B(h, \lambda-1, \nu) = B(h, \lambda, \nu). \end{align*} \end{itemize} This completes the inductive step. \end{proof} \section{The general bound} \label{sec:general} We now prove Theorem~\ref{theo:gmain}. The proof makes use of the following lemma by Pach {et~al.}\xspace \cite{pss-acn-94}: \begin{lemma}[{{\cite[Theorem~2.1]{pss-acn-94}}}] \label{lemma:pach}% Let ${G}$ be a graph with $n$ vertices of degree~$d_1,\dots,d_n$ and crossing number~$\chi$. Then there is a subset $E$ of $b$ edges of ${G}$ such that removing $E$ from ${G}$ creates components of size at most~$2n/3$, and \begin{align*} b^{2} \leq (1.58)^{2}\big(16\chi + \sum_{i = 1}^{n} d_{i}^{2}\big). \end{align*} \end{lemma} \begin{proof}[Proof of Theorem~\ref{theo:gmain}] Let ${G}$ be a graph on~$n$ vertices with $m$ edges having property~$\mathcal{P}^{*}$. Since each edge~$e$ crosses a graph that has property~$\mathcal{P}$, the crossing number of~${G}$ is at most $\chi\leq O(mn^{1+\alpha})$. The degree of any vertex is bounded by~$n-1$, and so we have $d_{i}^{2} \leq n \cdot d_i$. It follows using Lemma~\ref{lemma:pach} that there exists a set~$E$ of~$b$ edges in~${G}$ such that \begin{align*} b^{2} & \leq O(\chi + \sum_{i=1}^{n} d_i^{2}) \leq O(mn^{1+\alpha} + n \sum_{i=1}^{n}d_{i}) \leq O(mn^{1+\alpha} + mn) \leq O(mn^{1+\alpha}), \end{align*} and removing $E$ from ${G}$ results in components of size at most~$2n/3$. We recursively subdivide ${G}$. Level~$0$ of the subdivision is ${G}$ itself. We obtain level~$i+1$ from level~$i$ by decomposing each component of level~$i$ using Lemma~\ref{lemma:pach}. Consider a level~$i$. It consists of $k$~components $G_{1},\dots,G_{k}$. Component~$G_{j}$ has $n_{j}$ vertices and $m_{j}$ edges, where $n_{j}\leq (\frac 23)^{i}n$. The total number of edges at level~$i$ is $r = \sum_{j=1}^{k} m_{j}$. Using the Cauchy-Schwarz inequality for the vectors $\sqrt{m_j}$, $\sqrt{n_{j}}$, we have \begin{align*} \sum_{j=1}^{k} \sqrt{m_jn_{j}} \leq \sqrt{\sum_{j=1}^{k} m_{j}} \sqrt{\sum_{j=1}^{k} n_{j}} = \sqrt{r n} \leq \sqrt{m n}. \end{align*} We first consider the case~$\alpha > 0$. The number of edges needed to subdivide $G_{j}$ is $O(\sqrt{m_{j}n_{j}^{1+\alpha}})$. We bound this using $n_{j} \leq (\frac 23)^{i}n$ as $O(\sqrt{m_{j}n_{j}}((\frac 23)^{i}n)^{\alpha/2})$, and we obtain that the total number of edges removed between levels~$i$ and~$i+1$ is bounded by $O(\sqrt{mn}((\frac 23)^{i}n)^{\alpha/2})$. Since $(\frac 23)^{\alpha/2} < 1$, summing over all levels results in a geometric series, and so the total number of edges removed is $O(\sqrt{mn^{1+\alpha}})$. But this implies that the total number of edges in the graph is bounded as \begin{align*} m \leq O(\sqrt{mn^{1+\alpha}}), \end{align*} and squaring both sides and dividing by~$m$ results in \begin{align*} m \leq O(n^{1+\alpha}). \end{align*} Next, consider the case $\alpha = 0$. The number of edges removed between levels~$i$ and~$i+1$ is bounded by $O(\sqrt{m n})$. Adding over all $O(\log n)$ levels shows that \begin{align*} m & \leq O(\sqrt{mn} \log n). \end{align*} Again, squaring and dividing by~$m$ leads to \begin{align*} m & \leq O(n \log^{2} n). \ifmmode\qed\else\hfill\proofSymbol\fi \end{align*} \par\vspace{-\baselineskip} \end{proof} \section{Conclusions} \label{sec:conclusions} We have proven bounds on the number of edges of $k$-fan-crossing free graphs. For $k = 2$ our bound is tight, and we could even characterize the extremal graphs. In comparison, the bound we obtain for $k > 2$ is much weaker. For $k = 3$, for instance, our bound is~$6n-12$. The best lower bound we are aware of is the construction of Pach and T\'{o}th\xspace~\cite{pt-gdfce-97} of a 2-planar graph with $5n-10$~edges. For general~$k \geq 2$, a lower bound of $n(k-1)/2$ follows by considering a $(k-1)$-regular graph, that is a graph where every vertex has degree~$k-1$. Clearly it is $k$-fan-crossing free and has $n(k-1)/2$ edges. For $k \ll n$ this can be improved by a factor two: Consider a $\sqrt{n} \times \sqrt{n}$ integer grid of vertices. Connect every vertex to $2(k-1)$ neighbors within a $O(\sqrt{k}) \times O(\sqrt{k})$ grid in a symmetric way---that is, whenever we connect $(x,y)$ with $(x+s, y + t)$, we also connect with $(x-s, y-t)$. Then no line can intersect more than $k-1$ edges incident to a common vertex, and so the graph is $k$-fan-crossing free. Except for vertices near the boundary of the grid, every vertex has degree~$2(k-1)$, so the number of edges is $(k-1)(n - O(\sqrt{nk})) \approx kn$. By contrast, our upper bound is $\approx 3kn$. The weakness in our technique is that it analyzes arrows separately for every face of~$H$. When $H$ is a triangulation, it has $3n-6$ edges and $2n-4$ triangles. Each triangle can have $3k-6$ arrows, so we could have $(2n-4)(3k-6)$ arrows, implying $3n-6 + (2n-4)(3k-6)/2 = 3(k-1)(n-2)$ edges. Our bound is thus the best bound that can be obtained with this technique. For $k = 3$, it is possible to improve it slightly by observing that two adjacent triangles of~$H$ can only have four arrows in total. A $(k-1)$-planar graph is automatically $k$-fan-crossing free. Lemma~\ref{lemma:maximal-quad} implies that an extremal $2$-fan-crossing free graph is $1$-planar. Is the same statement true for $k=3$? It certainly doesn't hold for large~$k$, as Pach and T\'{o}th\xspace's bound on the number of edges of $k$-planar graphs is only~$O(\sqrt{k}n)$~\cite{pt-gdfce-97}. Already for $k=4$, $3$-planarity is a stronger condition than absence of a $4$-fan crossing: Pach {et~al.}\xspace~\cite{prtt-iclfm-06} showed that $3$-planar graphs have at most $5.5(n-2)$ edges. A $4$-fan-crossing free graph with~$6n-12$ edges can be constructed by starting with a triangulation, and adding the ``dual'' of every edge: for every pair of adjacent triangles, connect the two vertices not shared between the triangles. The crossing number of a $1$-planar graph on $n$~vertices is at most $n-2$ (this implies the bound $4n-8$ on the number of edges). In contrast, the crossing number of a fan-crossing free graph can be quadratic. For instance, start with the complete graph~$K_q$ on $q$~vertices. It has $q \choose 2$ edges and crossing number~$\Omega(q^{4})$. Now subdivide every edge into three edges by inserting two vertices very close to the original vertices. The resulting graph is fan-crossing free and has $n = q + 2{q \choose 2}$ vertices. Since any drawing of this graph can be converted into a drawing of~$K_q$ with the same number of crossings, it has crossing number $\Omega(q^{4}) = \Omega(n^{2})$. (The same construction could be done with any graph whose crossing number is quadratic in the number of edges, such as an expander graph or even a random graph.) A natural next question to ask is if our techniques can be used for graphs that do not contain a radial $(p,q)$-grid for $q > 1$, and if we can find tighter bounds than Pach {et~al.}\xspace~\cite{ppst-tgnlg-05}. In Theorem~\ref{theo:gmain}, we have given a rather general bound on the number of edges of graphs that exclude certain crossing patterns. The theorem shows that for graph properties~$\mathcal{P}$ that imply that the number of edges grows as $\Theta(n^{1+\alpha})$, for $\alpha > 0$, the size of the entire graph is bounded by the same function. For the interesting case $\alpha = 0$, which arises for instance for fan-crossing free graphs, our bound includes an extra $\log^{2} n$-term. Is this term an artifact of our proof technique, or are the examples of graph properties where $\mathcal{P}$ implies a linear number of edges, but graphs with~$\mathcal{P}^{*}$ and a superlinear number of edges exist? \paragraph{Acknowledgments.} For helpful discussions, we thank David Eppstein, J\'anos Pach, Antoine Vigneron, Yoshio Okamoto, and Shin-ichi Tanigawa, as well as the other participants of the Korean Workshop on Computational Geometry~2011 in Otaru, Japan. \bibliographystyle{alpha}
\section{Introduction} \label{sec:introduction} In recent years, spin qubits hosted in solid state nanostructures have been under extensive research due to their possible applications in quantum information processing and computation. Among the host materials of spin qubits, quite different approaches can be found, for instance, III-V-semiconductor and carbon nanotube quantum dots \cite{Hanson2007} (QD) as well as nitrogen vacancies in diamond \cite{Doherty2013}. These host materials show promising prospects but, unfortunately, also come with certain drawbacks. A precise control of the qubit state is the major advantage of III-V-semiconductor QDs based on Al(Ga)As heterostructures. Preparation and readout of the qubit with high fidelity via electrostatical gates has been demonstrated in many ground-breaking experiments\cite{Petta2005,Koppens2008,Greilich2009,Eble2009,Foletti2009,Ladd2010,Press2010,Bluhm2010,Bluhm2010a,Shulman2012,Dial2013}. However, the disadvantage of this material system is the presence of many nuclear spins inherent to the atoms of group III and group V elements of the periodic table. These nuclear spins give rise to a fast decoherence of the electron spin\cite{Schliemann2003,Coish2009}. Elaborate design of experiments including pulse sequences and methods to polarize the nuclear spins\cite{Reilly2008,Xu2009,Latta2009,Vink2009,Issler2010,Carter2011} such as dynamic nuclear polarization may help to compensate the effect of the spin bath. Nevertheless, it seems desirable to reduce the number of nuclear spins which can be achieved on the basis of other host materials. Obvious candidates are carbon and silicon, since their spin carrying isotopes have only very low natural abundances of about $1\%$ and $5\%$, respectively. In silicon, the qubits can be fabricated\cite{Morello2010,Morton2011} either using donor impurities or by confining a single electron via electrostatical gates. However, a controlled localization of the donor impurities is still a challenging task and electrostatically confined Si QDs often involve nanostructures with other materials like Ge, which potentially introduce additional nuclear spins\cite{Morton2011}. Carbon based QDs can be realized by confining an electron spin in carbon nanotubes\cite{Kuemmeth2008,Churchill2009,Churchill2009a,Steele2009,Monge2011,Chorley2011,Eichler2011,Lai2012} (CNT) via electrical gates allowing for a good control in the few electron regime. However, the curvature of the CNTs gives rise to a sizable spin-orbit coupling, yet another intrinsic source of decoherence to the electron spin. A different approach to a carbon based QD is the use of nitrogen vacancies in diamonds\cite{Dutt2007,Neumann2008,Togan2010,Togan2011,Doherty2013}, which show tremendously long coherence times. Unfortunately, control and readout of the qubit have to be done optically, which is disadvantageous for the realization of future on-chip electric circuits. These examples illustrate a more general issue of designing qubits, where an easy (electric) control and scalability seem to compete with noiseless environments and, hence, long decoherence times. A system potentially providing the best of both worlds is a graphene QD \cite{Trauzettel2007,Recher2010}, which offers very interesting electronic properties\cite{CastroNeto2009} and a small spin-orbit coupling\cite{Huertas-Hernando2006,Min2006,Yao2007,Gmitra2009}, as well as the possibility to control the number of nuclear spins by isotopic purification\cite{Banholzer1992,Simon2005,Churchill2009}. Moreover, the hyperfine interaction between the remaining nuclear spins and the electron spin is much smaller than in GaAs or Si. Additionally, the hyperfine interaction in graphene is anisotropic which could provide interesting applications as we discuss at the end of this article. Experimentally, graphene QDs are, for instance, realized by confining electrons with gates in bilayer graphene\cite{Goossens2012,Allen2012} and graphene nanoribbons\cite{Liu2009,Liu2010a}, respectively, or by etching the QD structure out of graphene flakes\cite{Stampfer2008,Ponomarenko2008,Schnez2009,Molitor2009,Guttinger2009,Molitor2010,Wang2010,Guttinger2010,Guttinger2011,Fringes2011,Engels2013}. Typical diameters are of the order of tens to hundreds of nanometers resulting in $K=15$ to $K=1500$ nuclear spins assuming a natural abundance of spin carrying $^{13}{\rm C}$ of $1\%$. Thus reducing the abundance of $^{13}{\rm C}$ by only two orders of magnitude leads to very small spin baths even in the case of rather large QDs. Recently, ultra small graphene QDs with diameters in the $1\,\mathrm{nm}$ range were made by electroburning\cite{Barreiro2012}. Altogether, these considerations show that the study of few nuclear spin models with $K<10$ as considered in this work is highly relevant for future research in the field. In this paper, we aim to set the basis for forthcoming investigations of the spin dynamics in graphene nanostructures. Besides quantum information theory, especially ongoing research on magnetism on edges\cite{Luitz2011,Schmidt2012,Schmidt2013,Golor2013,Golor2013a,Golor2013b} and vacancies\cite{Chen2011,Nair2012} in graphene can benefit from a detailed knowledge of the properties of the anisotropic hyperfine interaction (AHI). Moreover, we intend to complement our previous analytic study\cite{Fuchs2012} of the electron spin dynamics. Considering a large nuclear spin bath, we investigated the coherence of the electron spin in a non-Markovian approach using a generalized master equation. In this work, however, we were limited to large external magnetic fields in order to justify the perturbative treatment of the hyperfine interaction. Since we restrict ourselves to less than ten nuclear spins in the present work, we can apply exact diagonalization to the hyperfine Hamiltonian, which offers a powerful tool to investigate the dynamics of the electron spin for a wide parameter regime\cite{Schliemann2002,Schliemann2003,Dobrovitski2003,Shenvi2005,Zhang2007,Zhang2008,Sarkka2008,Cywinski2010,Erbe2012}. In particular, we analyze the role of the number of nuclear spins $K$, their position within the QD, as well as their initial spin state. Thereby, we use the long-time average $\av{S_{z}}_{T}$ of the longitudinal electron spin and its decoherence time $T_{D}$ to quantify the influence of these different aspects. Moreover, we investigate the dependence of $\av{S_{z}}_{T}$ and $T_{D}$ on the orientation of the spin quantization axis with respect to the graphene plane. For the long-time average, we find a continuous crossover from a initial state dominated regime for $K<5$ to a regime more affected by the configuration of the nuclear spins for $K>6$ where the relative positions of the nuclear spins with respect to each other matter. As we will show below, this behavior can be understood by an analysis of the Hilbert space dimensions as well as of the spatial distribution of the nuclear spins in the QD. Besides this regime change, a growing nuclear spin bath suppresses fluctuations around the long-time average more and more effectively, while the average itself is almost constant for all $K<9$ with $\av{S_{z}}_{T}\approx \hbar/4$ in the out-of-plane orientation and $\av{S_{z}}_{T}\approx0$ in the in-plane case. By resolving the orientation dependence in more detail, we find good agreement with $\av{S_{z}}_{T}(\beta)=\av{S_{z}}_{T}(0)\cdot\cos(\beta)^{2}$, where $\beta=0$ and $\beta=\pi/2$ correspond to the out-of-plane and in-plane orientation, respectively, {\it cf.} Fig. \ref{fig:QD} (a). Evidently, the decoherence time $T_{D}$ strongly depends on the number of nuclear spins $K$. We observe that the configuration of the nuclear spins is decisive even for very small numbers of nuclear spins, whereas the initial states play only a minor role. Depending on the relative positions of the nuclei, the decoherence times may deviate over several orders of magnitude. This behavior can be traced back to changes of the spectrum of eigenvalues of the full Hamiltonian. Moreover, the decoherence times significantly differ between out-of-plane and in-plane orientation. For $K=3$ and $\beta=0$, the majority of investigated configurations show very long decoherence times above $10\,\mathrm{ms}$ where, in many cases, even no decoherence at all was found. For $\beta=\pi/2$, in contrast, we always find decoherence, which predominantly occurs within $500\,\mu\mathrm{s}$. With increasing bath size, the decoherence times decrease for both orientations of the quantization axis. Then, decoherence times below $500\,\mu\mathrm{s}$ are most common. The article is organized as follows. In Sec.~\ref{sec:model}, we explain our model of the QD and discuss all relevant interactions of the spins with each other. Subsequently, in Sec.~\ref{sec:method}, we present the method used to obtain both the long-time average of the electron spin and its decoherence times. All results are shown and analyzed in Sec.~\ref{sec:results}. Based on a summary in Sec.~\ref{sec:conclusion}, we give an outlook on possible applications of few nuclear spin graphene QDs and on interesting future projects in this field. \section{Model} \label{sec:model} \begin{center} \begin{figure*} \centering \includegraphics[width=\textwidth,type=eps,ext=.eps,read=.eps]{fig1} \caption{(Color online) \textbf{(a):} The graphene ($(x_{1},x_{2},x_{3})$) and quantization axis ($(x,y,z)$) reference frames. \textbf{(b):} A graphene QD (red sites) for a Gaussian envelope function with $K=10$ uniformly random distributed $^{13}{\rm C}$ atoms (blue squares) carrying a nuclear spin $1/2$. The extent of the dot over the graphene lattice is defined via the electron envelope function, where all sites within the dot obey the cut-off relation defined in Eq. (\ref{eq:QD_def}). \textbf{(c):} The envelope function for fixed $x_{1}=0$ and $x_{1}=22\,a_{NN}$, respectively. The dashed line indicates the cut-off defined in Eq. (\ref{eq:QD_def}).} \label{fig:QD} \end{figure*} \end{center} We study the spin dynamics in a graphene quantum dot, where one electron spin is in contact with a bath of nuclear spins hosted by the $^{13}{\rm C}$ atoms. Due to the confinement, the electron can occupy a discrete spectrum of bound states, with an energy splitting between different states\cite{Silvestrov2007,DeMartino2007,Trauzettel2007,Matulis2008,Recher2009,Titov2010}. If the temperature is small compared to the level spacing $\Delta E$ of these bound-state energies, the electron resides in the ground state, which we describe by an envelope function $\phi(\vec{r})$. Hence, the probability to find the electron in a certain region of the QD can be described by its absolute square $|\phi(\vec{r})|^2$. In this paper we define the \textit{``center``} of the dot as the region around $\vec{r}_{max}$, where the envelope function is maximal $|\phi(\vec{r}_{max})|^2$. Far away from this center, the envelope function has to vanish: \begin{equation} |\phi(\vec{r}_{max}+\Delta\vec{r})|^2\rightarrow0 \;\; \mathrm{for} \;\; |\Delta\vec{r}| \rightarrow \infty \,. \label{eq:integrability} \end{equation} In this work, we model a graphene QD by the set of atomic sites $\lbrace\vec{r}_{k}\rbrace_{i=1}^{N_{sites}}$ obeying \begin{equation} \frac{|\phi(\vec{r}_{i})|^2}{|\phi(\vec{r}_{max})|^2}>C \,, \label{eq:QD_def} \end{equation} where $C=10^{-6}$ is a constant cut-off. A plot of a QD realized in this way is shown in Fig. \ref{fig:QD}. The choice of this finite system of discrete sites $\vec{r}_{i}$ imposes a normalization condition \begin{equation} \sum\limits_{i=1}^{N_{sites}} |\phi(\vec{r}_{i})|^2 = 1 \,, \label{eq:QD_norm} \end{equation} since we want to find the electron with probability $1$ somewhere within the dot. Effectively, we ignore everything outside the barrier defined by the cut-off, which is justified by the vanishing probability to find the electron there. This choice will become more clear when we discuss the hyperfine interaction between the electron and a single nuclear spin below. A possible choice for the envelope function is a Gaussian \begin{equation} \phi\left(\vec{r}\right) = \phi_{0}\exp\left[-\frac{1}{2}\left( \frac{r}{R}\right)^{2}\right] = \phi\left(r\right) \,, \label{eq:envfunc} \end{equation} where $r=\left|\vec{r}\right|$ is the absolute value of the electron position and the norm $\phi_{0}$ is chosen to satisfy the normalization condition in Eq. (\ref{eq:QD_norm}). This assumption is also in agreement with a recent experiment investigating the wave function of a graphene QD with soft confinement\cite{Subramaniam2012}. Note that the envelope function is not the exact electron wave function, but should give a good approximation to the precise solution. This can be seen, for instance, in graphene QDs based on semi-conducting armchair nano-ribbons\cite{Trauzettel2007}. The most important aspects, which are captured by this specific choice, are the absence of nodes in the ground state, a peak of the wave function in the center as well as a strong decay at the edges of the dot, which is illustrated by Fig. \ref{fig:QD} (c). Having defined the shape of our dot, we are now able to introduce the nuclear spins present in the system. Since we do not have further knowledge about the distribution of $^{13}{\rm C}$ within the dot, we randomly place the nuclear spins on the sites defined by Eq. (\ref{eq:QD_def}), where each site is chosen with equal probability. An example of a configuration of ten nuclear spins is shown in Fig. \ref{fig:QD}. We now proceed to the interactions between the nuclear spins and the electron spin and between themselves. The most relevant spin-spin interaction in our system is the hyperfine interaction between the electron spin $\vec{S}$ and $K$ nuclear spins $\vec{I}_{k}$ located at sites $\vec{r}_{k}$, \begin{equation} \hat{H}_{HI} = A_{HI} \sum\limits_{k=1}^{K} \sum\limits_{\mu,\nu} \overleftrightarrow{A}_{\mu\nu}\, |\phi(\vec{r}_{k})|^{2}\,\hat{S}_{\mu}\hat{I}_{k,\nu} \,, \label{eq:ahi} \end{equation} where the indices $\mu$ and $\nu$ run over spatial coordinates $x,y,z$. The energy scale of this interaction is given by $A_{HI}=0.6\,\mu\mathrm{eV}$ and $\overleftrightarrow{A}_{\mu\nu}$ is a spherical tensor\cite{Sakurai2010} of rank $2$, which takes into account the anisotropy of the hyperfine interaction in graphene\cite{Fischer2009a}. Remarkably, this interaction is strongly modulated by the envelope function, a fact which arises from the on-site nature of the hyperfine interaction. Thus, making the boundary of the dot smooth by taking a very small cutoff $C\rightarrow0$ in Eq. (\ref{eq:QD_def}), would only add vanishingly small contributions to the interaction in Eq. (\ref{eq:ahi}). Therefore we choose a small, but finite cutoff for simplicity. Besides the anisotropic hyperfine interaction (AHI), there is also a dipole-dipole interaction between pairs of nuclear spins. In the parameter regime considered in this work, this interaction, however, is about five orders of magnitude smaller than the AHI and, thus, neglected. We also proved its irrelevance by a numerical study, which we will not present here. Since external magnetic fields allow to manipulate spins experimentally, we include a Zeeman-Hamiltonian to account for this: \begin{equation} \hat{H}_{ZE} = \hbar \gamma_{S} B_{z} \hat{S}_{z} + \hbar \gamma_{^{13C}} B_{z} \sum\limits_{k=1}^{K} \hat{I}_{k,z} \approx A_{ZE} \hat{S}_{z} \,, \label{eq:aeb} \end{equation} where we used the fact that the electron gyromagnetic ratio $\gamma_{S}=1.76\times10^{11}\mathrm{s}^{-1}\mathrm{T}^{-1}$ is much larger than the gyromagnetic ratio of the nuclear spins $\gamma_{^{13}{\rm C}}=6.73\times10^{7}\mathrm{s}^{-1}\mathrm{T}^{-1}$ to justify the right-hand side of Eq. (\ref{eq:aeb}). In the presence of an external magnetic field, an interplay of the spin orbit coupling with acoustic phonons can lead to spin relaxation times $T_{1}$ ranging from milliseconds to seconds\cite{Struck2010,Droth2011,Droth2013,Hachiya2013} for small external magnetic fields. The exact value of $T_{1}$ significantly varies with the spectrum of the phonons which depends on the details of the dot nanostructure. Providing that the graphene flake is flat throughout the spatial extent of the QD, however, the spin orbit interaction should be small. Thus, this assumption justifies to neglect the influence of the spin orbit interaction on our problem. In the following, we aim to simulate a model experiment consisting of a preparation of the spins and the actual measurement of the spin dynamics. For the preparation, one can think of two different scenarios. First, the states of both the electron spin and the nuclear spins can be prepared in the presence of a strong external magnetic field $B_{0}\gg K \cdot A_{HI}|\phi(r_{max})|^2/\hbar\gamma_{S}=K\cdot|\phi(r_{max})|^2\cdot5.7\,\mathrm{mT}$, which imprints a well defined quantization axis. At the beginning of the actual measurement, this field is turned off or reduced to a finite value and the time evolution of the quantity of interest is recorded. However, since the tuning of magnetic fields is typically slow, this preparation scheme might not be adequate for a real experiment and we have to look for other solutions. In this case, we can think of injecting a spin polarized electron via a spin dependent tunneling process from a normal lead or via a spin conserving tunneling process from a spin polarized lead. Anyhow, in both considered scenarios our system features two natural reference frames, the first defined by the graphene plane and the second one by the quantization axis of the electron as depicted in Fig. \ref{fig:QD} (a). In order to clarify the notation, the graphene coordinate system (GCS) is written as $\vec{\tilde{v}}=(\tilde{v}_{x_{1}},\tilde{v}_{x_{2}},\tilde{v}_{x_{3}})$ with all objects being marked with a tilde, whereas a vector in the quantization axis coordinate system (QCS) is labeled by $\vec{v}=(v_{x},v_{y},v_{z})$ \textit{cf.} Fig. \ref{fig:QD}. If one chooses, without loss of generality, the $x_{2}$- and the $y$-axis to coincide, both coordinate systems are connected via a rotation $\hat{D}(\beta)$ by an angle $\beta$ around this common axis. Due to the symmetries of the carbon orbitals \cite{Fischer2009a}, the spherical tensor $\overleftrightarrow{A}_{\mu\nu}$ of the AHI in Eq. (\ref{eq:ahi}) takes its simplest form in the GCS, namely \begin{equation} \overleftrightarrow{\tilde{A}} = \left( \begin{array}{ccc} -\frac{1}{2} & 0 & 0 \\ 0 & -\frac{1}{2} & 0 \\ 0 & 0 & 1 \end{array} \right) \,, \label{eq:spherical_tensor} \end{equation} while the spin operators, $\hat{S}_{\mu}$ and $\hat{I}_{k,\nu}$, and the spin states are most conveniently defined in the QCS. In the main part of this work, we are interested in the time-dependent expectation value of an arbitrary operator $\hat{O}$ \begin{equation} \av{O}(t) = \bra{\psi_{0}}\,\hat{U}^{\dagger}(t)\,\hat{O}\,\hat{U}(t)\,\ket{\psi_{0}} \,, \label{eq:av_operator} \end{equation} where $\ket{\psi_{0}}$ describes the initial state of the total spin system and $\hat{U}(t)=\exp[-\mathrm{i}\hbar^{-1}t\,\hat{H}]$ is the time evolution operator determined by the total Hamiltonian $\hat{H}=\hat{H}_{HI}+\hat{H}_{ZE}$. Note, that this is the total Hamiltonian with respect to the QCS, where the Zeeman Hamiltonian is always diagonal and the AHI Hamiltonian is obtained from its simple GCS form $\hat{\tilde{H}}_{HI}$ by \begin{equation} \hat{H}_{HI} = \hat{D}(\beta)\hat{\tilde{H}}_{HI}\hat{D}^{\dagger}(\beta) \,. \label{eq:h_hi} \end{equation} Likewise, we could keep the Hamiltonian fixed in its GCS form and instead transform the operators $\hat{O}$ and $\hat{H}_{ZE}$ as well as the initial state $\ket{\psi_{0}}$ from the QCS to the GCS. For technical reasons, however, we choose to transform the Hamiltonian, while keeping the operator and the initial state fixed for arbitrary $\beta$. \section{Method} \label{sec:method} In order to numerically compute the time evolution in Eq. (\ref{eq:av_operator}), we need a basis to represent the state of our system and the operators acting on it. A natural choice for $N=1+K$ spins is given by the tensor product states of the electron spin and the nuclear spin eigenstates \begin{equation} \ket{n} = \ket{m^{n}_{S}}\otimes\bigotimes_{k=1}^{K}\ket{m^{n}_{k}} = \ket{\Downarrow\uparrow\downarrow\da\uparrow\dots} \,, \label{eq:Hilbert_basis} \end{equation} where the electron spin is represented by $\ket{m^{n}_{S}}$, $m^{n}_{S}=\Downarrow,\Uparrow$ and the nuclear spin states by $\ket{m^{n}_{k}}$, $m^{n}_{k}=\downarrow,\uparrow$. For convenience, we have ordered the nuclear spins $\ket{m^{n}_{S}m^{n}_{K}m^{n}_{K-1}\dots}$ according to the value of the envelope function at the corresponding site: \begin{equation} |\phi(r_{K})|^{2} \geq |\phi(r_{K-1})|^{2} \geq \dots \end{equation} Within this basis, an arbitrary state is given by a linear superposition of these $2^{N}$ states \begin{equation} \ket{\psi} = \sum\limits_{n=0}^{2^{N}-1} \alpha_{n} \ket{n} \,, \qquad \sum\limits_{n=0}^{2^{N}-1} |\alpha_{n}|^{2} = 1 \label{eq:state_superpos} \end{equation} with complex coefficients $\alpha_{n}$, while all operators are represented by $2^{N}\times2^{N}$ matrices. By diagonalizing the total Hamiltonian $\hat{M}\hat{H}\hat{M}^{\dagger}=\mathrm{diag}(\lambda_{0},\lambda_{1},\dots,\lambda_{2^{N}-1})$ we are able to re-express the time evolution operator \begin{equation} \hat{V}(t) = \hat{M}\hat{U}(t)\hat{M}^{\dagger} = \mathrm{diag}(\exp[-\mathrm{i}\hbar^{-1}\lambda_{0}\,t],\dots) \,, \label{eq:V_t} \end{equation} where $\hat{M}$ is an unitary operator formed by the eigenvectors of $\hat{H}$ and the $\lambda_{n}$ are the corresponding eigenvalues. Finally, we re-write Eq. (\ref{eq:av_operator}) and find \begin{equation} \av{O}(t) = \bra{\psi_{0}}\hat{M}^{\dagger}\;\;\hat{V}^{\dagger}(t)\;\;\hat{M}\,\hat{O}\,\hat{M}^{\dagger}\;\;\hat{V}(t)\;\;\hat{M}\ket{\psi_{0}} \,, \label{eq:av_operator_final} \end{equation} which we will evaluate for different parameter regimes in the following section. The numerical diagonalization is performed using the {\small\textrm{EIGEN}}\cite{Guennebaud2010} package for {\small\textrm{C++}}. As one can notice from the explanations above, we deal with a quite big parameter space in which we can analyze the outcome of Eq. (\ref{eq:av_operator_final}). First, we control the shape of the dot by means of the envelope function $|\phi(r)|^{2}$, secondly the number of nuclear spins $K$ is variable and finally these spins can have different positions or configurations $C$ within the dot. All of these parameters change the AHI Hamiltonian in Eq. (\ref{eq:ahi}). Moreover, we will investigate different initial states $\ket{\psi_{0}}$ of the electron and the nuclear spins affecting Eq. (\ref{eq:av_operator_final}). Additionally, the eigenvector matrix $\hat{M}$, appearing in this equation is a function of the twisting angle $\beta$ between the GCS and the QCS. Note that the spectrum of eigenvalues $\lambda_{n}$ is unaffected by a change of $\beta$. Finally, we can also modify the absolute value of the external magnetic field, which we will parametrize by the resulting Zeeman energy of the electron $A_{ZE}$. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig2} \caption{(Color online) Exemplary time evolution of the longitudinal electron spin component $\av{S_{z}}(t)$ as a function of time for out-of-plane orientation $\beta=0$, cf. Fig. \ref{fig:QD} (a). For a certain range of time $[T_{min},T_{max}]$ using a resolution $\Delta T$, we calculate the long-time average $\av{S_{z}}_{T}$ and the standard deviation $\sigma_{S_{z}}$ and find the maximal deviation $\Delta S_{z}$ around this value. These quantities as well as the details of the oscillations including the beating structure depend on the choice of the parameters. The decoherence time $T_{D}$ is determined by a constant threshold $C_{S}$.} \label{fig:Sz_example} \end{figure} \end{center} \section{Results} \label{sec:results} In this section, we present our findings for the model system defined above. All calculations were carried out using an envelope function of the Gaussian type in Eq. (\ref{eq:envfunc}) with $R=7\;a_{NN}$ and a cut-off $C=10^{-6}$. This corresponds to a dot with diameter $D\approx7.2\,\mathrm{nm}$ containing $N_{sites}\approx10^3$ carbon atoms, such that $K=9$ atoms correspond to the natural abundance $n_{I}=0.01$ of $^{13}{\rm C}$. In order to investigate the impact of different initial states, we choose random complex (RC) initial states\cite{Schliemann2002,Schliemann2003}. These states were created by drawing complex coefficients $\alpha_{n}$ from $\mathrm{Re}[\alpha_{n}],\mathrm{Im}[\alpha_{n}]\in[-1,1]$ with equal probability and normalizing them according to Eq. (\ref{eq:state_superpos}). Moreover, we choose the electron spin always to point down resulting in initial states consisting of $\ket{\Downarrow\dots}$ states only, which means that $\alpha_{n}=0$ for $n\geq2^{K}$. In order to determine qualitatively and quantitatively the impacts of the parameters, we investigate the time dependent expectation value $\av{S_{z}}(t)$ of the longitudinal electron spin component, which is calculated using Eq. (\ref{eq:av_operator_final}). A typical time evolution of $\av{S_{z}}(t)$ is plotted in Fig. \ref{fig:Sz_example}. Within the decoherence time $T_{D}$, the initial amplitude of the electron spin of $\hbar/2$ decays to its long-time average value, where still finite oscillations and beatings occur. This can be traced back to the finite size of the spin bath considered here. Its long-time average value is calculated by \begin{equation} \av{S_{z}}_{T} = \frac{1}{N_{T}}\sum\limits_{s=0}^{N_{T}}\av{S_{z}}(T_{min}+s\cdot\Delta T) \,, \label{eq:Sz_longtime} \end{equation} where we average over $N_{T}=(T_{max}-T_{min})/\Delta T$ time steps of width $\Delta T$. In order to investigate the oscillations of $\av{S_{z}}(t)$ quantitatively, we consider the standard deviation \begin{equation} \sigma_{S_{z}} = \sqrt{\frac{1}{N_{T}}\sum\limits_{s=0}^{N_{T}} \Big(\av{S_{z}}(T_{min}+s\cdot\Delta T) - \av{S_{z}}_{T}\Big)^{2}} \label{eq:Sz_stdv} \end{equation} as well as the sample range \begin{equation} \Delta S_{z} = \max_{t\in[T_{min},T_{max}]}[\av{S_{z}}(t)] - \min_{t\in[T_{min},T_{max}]}[\av{S_{z}}(t)] \,. \label{eq:Sz_delta} \end{equation} The latter quantity is a measure for the occurrence of oscillations with a big amplitude which originate from either recurrences of the signal, beatings or an entire lack of decoherence. While for beatings one expects rather small sample ranges $\Delta S_{z}<S_{z}(0)$, the former two cases should give values on the order of the initial amplitude, $\Delta S_{z}\sim O(S_{z}(0))$ in the out-of-plane case $\beta=0$ and $\Delta S_{z}\sim 2 \cdot O(S_{z}(0))$ in the in-plane case $\beta=\pi/2$. Besides these quantities characterizing the long time average of the electron spin, we are also interested in the amount of time it takes to decohere the system. In order to be independent from specific models of the decay, such as exponential or power-law decoherence, and to account for the characteristics of the numerics, we find this decoherence time $T_{D}$ by the first minimum exceeding a certain threshold $C_{S}$. For clarity, $C_{S}$ is also illustrated in Fig. \ref{fig:Sz_example}. This approach is similar to the one used in Ref. \onlinecite{Erbe2012} to find the decoherence times. Of course, the choice of this constant $C_{S}$ changes the value of $T_{D}$. However, its order of magnitude and its dependence on the different parameters is rather independent from a specific choice as long as $C_{S}$ is not too close to $\av{S_{z}}_{T}$, which we confirmed for different values of $C_{S}$. In the following, we analyze both the decoherence time and the long-time average of the longitudinal electron spin component for different parameter sets. For each number $K$ of nuclear spins many initial states and configurations are created and labeled by numbers $0,1,2\dots$ for later comparison of the results. Note, that for different nuclear spin numbers $K$ these labels describe different initial states and configurations. Moreover, we concentrate on two orientations of the quantization axis, namely out-of-plane orientation for $\beta=0$ and in-plane orientation with $\beta=\frac{\pi}{2}$. We investigated the effect of finite magnetic fields for exemplary initial states, configurations and $K=2,4,6$ nuclear spins, where we varied the resulting Zeeman constant from $A_{ZE}/A_{HI}\ll1$ to $A_{ZE}/A_{HI}\gg1$. For increasing $A_{ZE}$, we find a continuous crossover to a perfect alignment of the electron spin in the case of a very strong magnetic field. In the following, we put $A_{ZE}=0$ because we would like to better understand the low-magnetic field behavior of the spin dynamics in the presence of the AHI. \subsection{Dependence of the long-time average on different initial states, configurations, and the number of nuclear spins} \label{sec:sub:inst_config} \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig3} \caption{(Color online) Plot of the long-time average $\av{S_{z}}_{T}$ for out-of-plane orientation $\beta=0$ and $K=3$ and $K=6$ nuclear spins without an external magnetic field. We considered $51$ different RC initial states and $51$ random configurations. For both numbers of nuclear spins, the electron spin looses roughly one half of its amplitude to $\av{S_{z}}_{T}\approx-0.22\,\hbar$. The horizontal stripes for $K=3$ indicate the importance of the initial states for few nuclear spins, while configurations seem less relevant signaled by weaker vertical structures. This changes for $K=6$ nuclear spins, where the configurations dominate over initial states indicated by the vertical structures in the right plot.} \label{fig:Sz_mean__rd_050__c_050__A_+0.00e+00} \end{figure} \end{center} \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig4} \caption{(Color online) Same plot as in Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00} but for in-plane orientation $\beta=\pi/2$. For all parameters the long-time average of the longitudinal electron spin component sharply saturates at $\av{S_{z}}_{T}\approx0$. In contrast to the out-of-plane case, the results seem independent of the initial state even in the few spin regime. For certain configurations, however, we find a non negligible dependence on the initial state.} \label{fig:Sz_mean__rd_050__c_050__A_+1.57e+00} \end{figure} \end{center} \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig5} \caption{(Color online) Plot of the long-time average $\av{S_{z}}_{T}$, the standard deviation $\sigma_{S_{z}}$ and, the sample range $\Delta S_{z}$ as a function of the number of nuclear spins $K$ for in-plane ($\beta=\pi/2$, red) and out-of-plane orientation ($\beta=0$, black). The values are obtained by averaging over $51$ RC initial states and $51$ different configurations, see Figs. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00} and \ref{fig:Sz_mean__rd_050__c_050__A_+1.57e+00}. Error bars are given by the standard deviation with respect to averaging over all $51\times51$ results. While the long-time average is almost constant, the averaged standard deviation $\sigma_{S_{z}}$ as well as the averaged sample range $\Delta S_{z}$ strongly decrease for larger $K$ indicating the reduction of fluctuations and of the occurrence of beating or recurrence events.} \label{fig:rd_C__C__rd_mean_sample_range} \end{figure} \end{center} First, we investigate the consequences of both different RC initial states and different configurations of the nuclei within the dot. We calculated $\av{S_{z}}_{T}$, $\sigma_{S_{z}}$ and, $\Delta S_{z}$ for different parameter sets and found stable results for $T_{min}=0.5\times10^{9}\tau_{HI}$, $T_{max}=1.5\times10^{9}\tau_{HI}$, and $\Delta T=10^{4}\tau_{HI}$ with $\tau_{HI}=\hbar/A_{HI}\approx1\,\mathrm{ns}$. In Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00}, we plot the long-time average $\av{S_{z}}_{T}$ as a function of different RC states and configurations for $K=3$ and $K=6$, respectively, in out-of-plane orientation. The color map in Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+1.57e+00} was created for the same parameters with in-plane orientation. For a small number of nuclear spins $K=3$ and $\beta=0$, we observe strong fluctuations for both different $RC$ states and different configurations around an average value of $\av{S_{z}}_{T}\approx-0.22\,\hbar$ as depicted in the color map of Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00}. The horizontal stripes dominate over the vertical structures indicating, that the choice of the RC initial states has a greater influence on the results than the spatial configuration of the nuclear spins within the dot. Moreover, we find large oscillations around this long-time average value for many configurations and initial states. This results in both sizable sample ranges $\Delta S_{z}$ and standard deviations $\sigma_{S_{z}}$. By averaging over all $51\times51$ results, we find $\av{\av{S_{z}}_{T}}=(-0.22\pm0.06)\,\hbar$, $\av{\sigma_{S_{z}}}=(0.13\pm0.04)\,\hbar$, and $\av{\Delta S_{z}}=(0.52\pm0.11)\,\hbar$, which is also shown in Fig. \ref{fig:rd_C__C__rd_mean_sample_range}. The large average value of the sample range $\av{\Delta S_{z}}$ shows that for most cases analyzed, there was at least one big change in amplitude. However, no total spin flip to $+\hbar/2$ is achieved. The occurrence of sizable standard deviations indicates that there are on average many of these events. Thus in the few nuclear spin regime, coherent oscillations of the electron spin are the dominant dynamics, where recurrences of the initial value take place with a period $T_{P}=\hbar/\max_{i}(|\lambda_{i}|)\sim \hbar/(A_{HI}\cdot|\phi(r_{K})|^{2})\geq100\,\mathrm{ns}$. If we consider a larger environment of nuclear spins as presented in Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00} with $K=6$, the behavior of the long-time average changes. First of all, the result is much more uniform with respect to both the RC initial states and the configurations. In addition, the remaining differences in $\av{S_{z}}_{T}$ depend on the configurations rather than on the initial states, which is obvious from the vertical lines present in this color map. Averaging over all $51\times51$ results gives $\av{\av{S_{z}}_{T}}=(-0.22\pm0.02)\,\hbar$, which is essentially the same as for $K=3$. However, the standard deviation $\av{\sigma_{S_{z}}}=(0.06\pm0.03)\,\hbar$ and the sample range $\Delta S_{z}=(0.37\pm0.06)\,\hbar$ clearly decrease. We confirmed this trend of decreasing fluctuations by repeating the above averaging procedure for other numbers of nuclear spins. These results are presented as a function of $K$ in Fig. \ref{fig:rd_C__C__rd_mean_sample_range}. While the long-time average value is constant, both the standard deviation and the sample range become smaller. Especially, the pronounced decay of the sample range clearly indicates that recurrences occur much less and, hence, that the corresponding recurrence times are increasing with more nuclear spins. Thus, the major effect of an increased number of nuclear spins is to suppress the oscillations around the long-time average and changing the system from an initial state dominated regime to a regime where the configuration of the nuclear spins is important. This behavior can be understood by analyzing the impact of the nuclear spin number on the dimension of the Hilbert space and on the strength of the hyperfine interaction. For a small number of nuclear spins, the dimension of the corresponding Hilbert space $D=2^{K+1}$ is small and, hence, we draw our RC initial states from a rather limited set, where individual single product states $\ket{n}$ lead to very different dynamics of the electron spin. Due to the combination of only $2^{K}$ states $\ket{n}$ to a RC initial state, it is not unlikely that one of these states dominates over the rest leading to rather diverse results. By increasing K and, thus, the Hilbert space dimension this situation is changed. Since the individual state $\ket{\Downarrow\dots\uparrow\downarrow}$ of nuclear spins at the border of the dot is almost irrelevant due to a small $|\phi(\vec{r}_{k})|^{2}$, groups of effectively equivalent states are superposed. Thus, a more effective averaging is achieved suppressing the dependence on a specific initial state. As a consequence, it is very unlikely for a single state to dominate over the rest. The coupling strength of nuclei is the key in understanding the dependence of the results on the configuration. Its energy scale is given by the product $A_{HI}\,|\phi(r_{K})|^{2}$ of the hyperfine coupling constant and the maximal value of the envelope function at the sites of the nuclear spins. For a small number $K$, the probability to find two or more nuclear spins, which couple almost equally with the electron spin, is low due to the large gradient of the envelope function. Hence, effectively only one nuclear spin strongly interacts with the electron leading to simple oscillations. This behavior can be also easily derived by diagonalizing the resulting, effective $4\times4$ matrix of the AHI Hamiltonian given in Eq. (\ref{eq:ahi}). In doing so, one finds a discrete spectrum of frequencies given by the degenerate eigenvalues $\lbrace\lambda_{i}\rbrace=\lbrace-1/2,0,1/4,1/4\rbrace\cdot A_{HI}|\,\phi(r_{K})|^{2}$. This fact is responsible for the rather uniform dynamics with respect to different configurations in a small $K$ regime. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig6} \caption{(Color online) $(a)$: Eigenvalues $\lambda_{i}$ of the AHI Hamiltonian for for $K=6$ nuclear spins and configuration $C=10$ in normalized units of $A_{HI}\,|\phi(r_{K})|^{2}$. $(b)$: Eigenvalues $\lambda_{i}$ for $K=6$ and $C=3$. $(c)$: Number of distinct eigenvalues $\lambda_{i}$ as a function of the relative probabilities $|\phi(r_{K-1})|^{2}/|\phi(r_{K})|^{2}$ and $|\phi(r_{K-2})|^{2}/|\phi(r_{K})|^{2}$ for $K=6$ nuclear spins. If both $|\phi(r_{K-1})|^{2}/\phi(r_{K})|^{2}\approx1$ and $|\phi(r_{K-2})|^{2}/\phi(r_{K})|^{2}\approx1$ at least three nuclear spins are strongly interacting with the electron spin causing a spectrum with many different eigenvalues as depicted in $(b)$. If only the most central nuclear spin couples strongly to the electron spin (lower left part), the spectrum is highly degenerate showing only three different eigenvalues as shown in $(a)$. The upper limit for the number of $15$ has no deeper meaning besides distinguishing both types of spectra.} \label{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} \end{figure} \end{center} This situation can of course also occur for larger nuclear spin environments, as shown in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} $(a)$ for $K=6$. It is, however, rather the exception from the more probable case of several nuclei coupling comparably to the electron, where a almost continuous spectrum is found as depicted in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} $(b)$. If we characterize these spectra quantitatively by counting the number of distinct eigenvalues, i.e., eigenvalues which differ significantly, we can map the configuration of the nuclei to the spectra as depicted in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} $(c)$. For the in-plane case, our findings are quite different from the former ones. The electron spin saturates around $\av{S_{z}}_{T}=0$ for both $K=3$ and $K=6$ as shown in Fig. \ref{fig:Sz_mean__rd_050__c_050__A_+1.57e+00}. Interestingly, we find already for $K=3$, that this average is reached very precisely with smaller fluctuations than in the out-of-plane case. This fact becomes also clear from averaging the longitudinal electron spin $\av{\av{S_{z}}_{T}}=(0.000\pm0.004)\,\hbar$ over all results. Moreover, the results are independent from the choice of the RC initial state. Some single configurations, however, give rise to deviations from this, where also a dependence on the initial state is restored. It seems, that this is the case, where several nuclear spins couple comparably to the electron spin explaining the sensitivity on initial states. The size of the fluctuations is on average given by $\av{\sigma_{S_{z}}}=(0.15\pm0.02)\,\hbar$. The mean value of the sample range of $\av{\Delta S_{z}}=(0.92\pm0.07)\,\hbar$ close to $1$ indicates, that in most cases, the electron spin is at least once almost completely flipped to $+\frac{1}{2}\,\hbar$ in contrast to the out-of-plane orientation. The $K=6$ study shows qualitatively the same result with $\av{\av{S_{z}}_{T}}=(0.000\pm0.001)\,\hbar$, where the fluctuations $\av{\sigma_{S_{z}}}=(0.057\pm0.004)\,\hbar$ are further suppressed. Moreover, the appearance of recurrences and total spin flips is also strongly decreased for $K=6$ as is clear from the sample range $\av{\Delta S_{z}}=(0.51\pm0.09)\,\hbar$. Analyzing this observable as a function of the number of nuclear spins, we observe again a prominent suppression of the fluctuations for growing $K$ as is apparent in Fig. \ref{fig:rd_C__C__rd_mean_sample_range}. In order to understand the differences between the in-plane and out-of-plane dynamics of the electron spin in more detail, an analytic analysis of the dynamics in the case of only one nuclear spin is very useful. Calculating the long-time average analytically for $K=1$ yields \begin{align} & \av{S_{z}}_{T}(\beta) = \lim_{\Delta T\rightarrow\infty}\frac{1}{2\Delta T}\int\limits_{T+\Delta T}^{T-\Delta T} \av{S_{z}}(t,\beta) \nonumber\\ & = -\frac{\hbar}{4}\cos(\beta)\Big[2\rho_{\downarrow\da}\cos(\beta)+(\rho_{\uparrow\downarrow}+\rho_{\downarrow\uparrow})\sin(\beta)\Big] \,, \label{eq:szlta_analytically} \end{align} where the initial density matrix \begin{equation} \rho_{0} =\ket{\psi_{0}}\bra{\psi_{0}} = \left( \begin{array}{cccc} \rho_{\downarrow\da} & \rho_{\downarrow\uparrow} & 0 & 0 \\ \rho_{\uparrow\downarrow} & \rho_{\uparrow\ua} & 0 & 0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ \end{array} \right) \label{eq:rho_0} \end{equation} is only non-zero for the electron spin pointing down as for the RC initial states. For more nuclear spins involved, the resulting equations become much more complicated. However, for the special case of only one strongly coupling nuclear spin, the structure of the AHI Hamiltonian remains the same and Eq. (\ref{eq:szlta_analytically}) still holds. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig7} \caption{(Color online) Dependence of the long-time average $\av{S_{z}}_{T}(\beta)$ on the orientation of the quantization axis with respect to the graphene plane for $K=6$ nuclear spins. The example shown here was calculated for the configuration $C=3$, whose continuous spectrum is presented in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} (b). The only parameter used to fit the numerical values to the analytic curve $\av{S_{z}}_{T}(\beta)=\av{S_{z}}_{T}(0)\cdot\cos^{2}(\beta)$ is the out-of-plane value $\av{S_{z}}_{T}(0)$.} \label{fig:beta_example} \end{figure} \end{center} We investigated the $\beta$ dependence of the long-time average numerically for some configurations and initial states and $K=2,4,6$ and $9$ nuclear spins, where we find good agreement of our results with $\av{S_{z}}_{T}(\beta)=\av{S_{z}}_{T}(0)\cos^{2}(\beta)$ with increasing $K$. Particularly, we observed this behavior also for configurations with several nuclear spins coupling almost equally to the electron spin. As an example, we plot in Fig. \ref{fig:beta_example} the $\beta$ dependence of $\av{S_{z}}_{T}$ for $K=6$ and configuration $C=3$. Its spectrum is shown in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} (b). This fact is also supported by our results presented in Figs. \ref{fig:Sz_mean__rd_050__c_050__A_+0.00e+00} and \ref{fig:Sz_mean__rd_050__c_050__A_+1.57e+00}, where we find on average $\av{S_{z}}_{T}\approx-0.22\,\hbar$ for $\beta=0$ and $\av{S_{z}}_{T}\approx0$ for $\beta=\pi/2$. The deviation of $\av{S_{z}}_{T}(\beta=0)$ from $-\hbar/4$ originates from the finite time window $[T_{min},T_{max}]$ used in the numerical calculations, which misses recurrences of the full initial value of $\av{S_{z}}(t=0)=-\hbar/2$. From this numerical findings and Eqs. (\ref{eq:szlta_analytically}) and (\ref{eq:rho_0}), we suppose that contributions from the off diagonal parts cancel each other almost completely and that the elements of the diagonal parts of the density matrix $\rho_{\downarrow\da},\rho_{\uparrow\ua}$ have approximately equal weight of $1/2^{K}$, which seems reasonable for random complex initial states. \subsection{Decoherence times} \label{sec:sub:decoherence_times} In this section, we want to investigate the decoherence times of the longitudinal electron spin $S_{z}$ for different initial states and different configurations. We chose the threshold to be always about $0.1\,\hbar$ below the obtained long-time average, which gives $C_{S}=-0.325\,\hbar$ for the out-of-plane case $\beta=0$ and $C_{S}=-0.1\,\hbar$ for the in-plane case $\beta=\pi/2$. Moreover, we used exactly the same initial states and configurations for all $K$ as for the calculation of the long-time average. The decoherence times were estimated for times up to $10^{7}\,\tau_{HI}\approx 10\,\mathrm{ms}$ with a time resolution $\Delta T = 10^{2}\,\tau_{HI}$, which yields at least $P=2\pi/(\Delta T \cdot \lambda_{max})\approx 20$ points per period of the highest absolute frequency $\max_{i}(|\lambda_{i}|)$. For $K=6$ we extended the investigated time regime to $10^{8}\,\tau_{HI}\approx 100\,\mathrm{ms}$ using the same time resolution $\Delta T$. As it turns out, the decoherence times obtained by this method are rather independent from the initial states. Several factors are important for this fact. First of all, for larger numbers of nuclei of course the same arguments concerning the Hilbert space dimensions as for the long-time average hold. However, we also find for small $K$ only little dependence on the initial states. One reason for this is probably, that our method is robust against small changes of the longitudinal electron spin caused by different initial states, since we measure when the signal is above a certain threshold, but not how much. Finally, as we show below, the decoherence seems strongly related to the presence of many incommensurate frequencies. These frequencies are proportional to the eigenvalues of the hyperfine Hamiltonian and, hence, independent from the initial state. Therefore, we focus in the following on the consequences of different configurations on the decoherence times for different numbers of nuclear spins. In principle, there are two relevant aspects concerning the positions of the nuclei, the absolute value of the envelope function $|\phi(\vec{r}_{K})|^{2}$ at the site of the strongest coupling nuclear spin and the relative position of the nuclei with respect to each other. The importance of the former is obvious, since the envelope function sets the maximal energy scale of the AHI in Eq. (\ref{eq:h_hi}) to $A_{HI}\cdot |\phi(r_{K})|^{2}$ and, consequently, rescales all times by a factor $|\phi(r_{K})|^{-2}$. Therefore, if we want to analyze the influence of the relative positions, we have normalize the decoherence times according to $T_{D}\rightarrow T_{D}\cdot|\phi(r_{K})|^{2}$. We begin our discussion with investigating these normalized decoherence times for $K=6$ nuclei in more detail and then turn to absolute decoherence times as a function of $K$ afterwards. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig8} \caption{(Color online) Normalized decoherence time $T_{D}\cdot|\phi(r_{K})|^{2}$ as a function of $51$ RC initial states and $51$ random configurations for $K=6$ nuclear spins in in-plane and out-of-plane orientation. While the decoherence time is almost the same for different initial states, it strongly depends on the configurations showing deviations over several orders of magnitude. White spaces indicate the total lack of decoherence up to absolute times of $0.1\,\mathrm{s}$ given a threshold of $C_{S}=-0.325\,\hbar$. For special configurations, $C=10,32$ and $35$, and $\beta=0$ there is no decoherence at all, but coherent oscillations of the electron spin.} \label{fig:tr__N_007__A_+0.00e+00__THD_-3.25e-01__rd_map} \end{figure} \end{center} A color map of the normalized decoherence times for $51$ initial states and $51$ configurations is shown in Fig. \ref{fig:tr__N_007__A_+0.00e+00__THD_-3.25e-01__rd_map}. For the out-of-plane case, we find that the decoherence times are almost independent of the initial state, but vary over several orders of magnitude for different configurations. If we plot the normalized times as a function of the number of distinct eigenvalues, cf. Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num}, we find a direct connection between these times and the configuration of the nuclei in the dot. As is clear from Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__A_+0.00e+00__THD_-3.25e-01_tr_bin_num_map}, long decoherence times can be only found for the discrete spectra, which are realized if only one nuclear spin strongly interacts with the electron. The configurations without any decoherence, which are indicated by white spaces in Fig. \ref{fig:tr__N_007__A_+0.00e+00__THD_-3.25e-01__rd_map}, exhibit discrete spectra with the minimal number of distinct eigenvalues of $3$. An example of such a spectrum is shown in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} (a). In these cases the dynamics of the longitudinal electron spin are coherent oscillations, where recurrences appear with a period of $T_{P}=|\phi(r_{K})|^{-2}\cdot\tau_{HI}\geq100\,\mathrm{ns}$. In contrast to this, short normalized decoherence times are a consequence of continuous spectra as presented in Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} (b). Thus, by the configurations studied, we can proof a direct relation between the relative positions of the nuclear spins and their relative coupling strengths, respectively, and the order of magnitude of the decoherence times. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig9} \caption{(Color online) Normalized decoherence time $T_{D}\cdot|\phi(r_{K})|^{2}$ as a function of the number of distinct eigenvalues of the AHI Hamiltonian for $K=6$ nuclear spins in out-of-plane orientation. For this plot all out-of-plane results presented in Fig. \ref{fig:tr__N_007__A_+0.00e+00__THD_-3.25e-01__rd_map} are considered. Obviously, long decoherence times occur only for a small number of distinct eigenvalues. Together with the results of Fig. \ref{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__phi1_phi2_ev_bin_num} the importance of the relative coupling strengths $\propto|\phi(r_{K-1})|^{2}/|(\phi(r_{K}))|^{2}\;,\;\dots$ and, hence, of the relative position of different nuclei becomes evident.} \label{fig:N_007__zero_thd_+1.00e-04__jump_thd_+1.00e-01__A_+0.00e+00__THD_-3.25e-01_tr_bin_num_map} \end{figure} \end{center} For the in-plane case, the qualitative picture is similar, however, with shorter normalized decoherence times over all, such that we find decoherence within the investigated times for all configurations. In contrast to the out-of-plane case, also discrete spectra can show rather short decoherence times for specific configurations. Altogether, this demonstrates a much faster decoherence due to the broken symmetry in the in-plane orientation. \begin{center} \begin{figure} \centering \includegraphics[width=0.45\textwidth,type=eps,ext=.eps,read=.eps]{fig10} \caption{(Color online) Relative number of absolute decoherence times $T_{D}$ falling in a certain time interval $[T_{min},T_{max}[$ for different numbers $K$ of nuclei in in-plane and out-of-plane orientation. For each $K$ $51$ RC initial states and $51$ configurations were considered leading to $N_{calc}=2601$ calculations in total. $(a)$: In the out-of-plane case, long decoherence times are clearly dominating for few nuclear spins. Increasing $K$ leads to a quick decay of the decoherence times such that short times $T_{D}$ are common. Very short decoherence times start to become relevant for $K>6$. $(b)$: In the in-plane case, even for few nuclear spins short relaxation times are the rule. For increasing $K$, the percentage of short decoherence times is growing further.} \label{fig:gauss-----__tr_bins_oop} \end{figure} \end{center} Turning from normalized times to absolute decoherence times, the value of the envelope function $|\phi(r_{K})|^{2}$ at the site of the strongest coupling nuclear spins additionally becomes relevant, since it sets the order of magnitude of all times. Putting a larger and larger number of nuclear spins on a QD of constant area increases the average value of $|\phi(r_{K})|^{2}$, since it is more likely to find a spin very close to the center. Moreover, as we discussed above, an increased $K$ makes it much more probable to have several nuclear spins coupling almost equally to the electron spin. Altogether, this lets us expect a prominent decay of long decoherence times as a function of growing $K$, which is confirmed by Fig. \ref{fig:gauss-----__tr_bins_oop}. For $\beta=0$ and very few nuclear spins $K=3$, we find that the majority of decoherence times is longer than $10\,\mathrm{ms}$, whereas very short $T_{D}$ are almost completely irrelevant. For $K=8$ the percentages of short and long times are inverse. Now, only less than $7\%$ of the decoherence times are longer than $10\,\mathrm{ms}$, while most of the decay of the electron spins takes place within $500\,\mu\mathrm{s}$. However, for $K=6$, surprisingly, still about one-fifth of the cases shows ultra long decoherence times. In the in-plane orientation, long decoherence times make up only a small fraction even for few nuclear spins. Short decoherence times in the range of $5\,\mu\mathrm{s}$ to $500\,\mu\mathrm{s}$ are significantly increasing for more spins. Notably, ultra short times below $5\,\mu\mathrm{s}$ do not become much more important. In summary, typical decoherence times are on the order of $\mathrm{ms}$ under ideal conditions of small nuclear spin numbers and out-of-plane orientation. In the case of such long decoherence times, of course, other effects like spin orbit coupling could become relevant. In the presence of acoustic phonons and small external magnetic fields, this spin orbit coupling\cite{Struck2010,Hachiya2013} can lead to spin relaxation times of $T_{1}\sim1\,\mathrm{ms}$ below the decoherence times found here. For larger numbers of nuclear spins and, generally, for in-plane orientation, decoherence times are smaller, but still above $5\,\mu\mathrm{s}$. Typical decoherence times of $GaAs$ QDs under spin echo\cite{Koppens2008} lie in the $T_{2,echo}\sim1\mu\mathrm{s}$ regime, whereas the current record of $T_{2,CPMG}\approx200\,\mu\mathrm{s}$ was measured using the Carr-Purcell-Meiboom-Gill (CPMG) pulse sequence\cite{Bluhm2010}. Pure dephasing times $T^{*}_{2}$ are below $50\,\mathrm{ns}$ for $GaAs$. Although all our estimates for the decoherence times are done for a model without any effort to improve the coherence of the electron spin like pulse sequences or strong magnetic fields, in almost all considered cases, we are above the $GaAs$ spin echo time $T_{2,echo}$. For smaller nuclear spin numbers, graphene even outperforms the CPMG time, which lets us expect very long decoherence times in graphene QDs when using pulse sequences. \section{Summary and Conclusion} \label{sec:conclusion} Starting from a generic model of a graphene QD, we studied the dynamics of the electron spin caused by the hyperfine interaction with the nuclear spins present in the dot. The number of nuclei was varied from $K=2$ to $K=9$, where the upper limit corresponds to the natural abundance of spin carrying $^{13}{\rm C}$ for the dot size considered in this article. Besides the role of the number of nuclei, we also investigated the influence of the initial conditions as well as the impact of different configurations of the nuclei in the dot. Moreover, we explored the consequences of the orientation of the spin quantization axis with respect to the graphene plane. In order to characterize and quantify these effects, we analyzed both the long-time average $\av{S_{z}}_{T}$ of the longitudinal electron spin component and its decoherence time $T_{D}$. Since nuclear spins are usually very hard to control in the envisioned experiments, we chose the initial states to be random complex (RC) superpositions of single product states. For this class of initial states, we found an appreciable effect on the long-time average only in the case of very few nuclear spins with $K<5$. Upon increasing the number of nuclear spins the effects of quantum parallelism and amplitude averaging\cite{Schliemann2002,Schliemann2003} reduce the differences between individual RC initial states more and more effectively. In this parameter regime, the results are dominated by the configuration of the nuclear spins within the dot, i.e., by their relative positions with respect to each other. For different configurations, the spectrum of eigenvalues of the hyperfine interaction varies from a highly discrete one with many degenerate eigenvalues to a continuous spectrum with many incommensurate frequencies. For all $K$, a pronounced dependence of the long-time average on the orientation angle $\beta$ between the spin quantization axis and the normal vector of the graphene plane was found. It saturates at approximately one-half of its initial value of $\av{S_{z}}_{T}\approx-\hbar/4$ for $\beta=0$ and at $\av{S_{z}}_{T}\approx0$ in the in-plane case with $\beta=\pi/2$. While the long-time average of the electron spin is surprisingly almost constant with respect to $K$, we observed a strong reduction of fluctuations around it for larger nuclear spin baths. In contrast to the long-time average, the decoherence times $T_{D}$ never showed a recognizable dependence on the initial states. Instead, the decoherence times depended decisively on the configuration of the nuclear spins in the dot. Long decoherence times were observed for only one nuclear spin strongly interacting with the electron spin, while several almost equally coupled nuclei lead to a very fast decoherence. Moreover, the decoherence times showed a strong dependence on the number of nuclear spins as well as on the orientation of the quantization axis. In the out-of-plane case, about $75\%$ of our results experienced decoherence times longer than $T_{max}=10\mathrm{ms}$ for $K=3$. For $K=8$, instead, less than $10\%$ showed no decoherence within this time frame while, in most cases, the electron spin decayed in less than $500\,\mu\mathrm{s}$. Considering the in-plane orientation, already for $K=3$ the majority of investigated initial state / configuration sets decohere within $500\,\mu\mathrm{s}$. Although our results were obtained for a specific model of the graphene quantum dot using a Gaussian envelope function, they could be generalized quite naturally. In our model, the QD was comparably small with a sharp boundary. This choice resulted in a steep envelope function. Physically, this situation corresponds approximately to an etched QD. Thinking of larger QDs with smoother boundaries, we expect a flatter envelope function which gives rise to more nuclear spins interacting comparably with the electron spin. Consequently, it becomes more likely to end up with rather low fluctuations around the long-time average and to find quite short decoherence times. In contrast, the realization of even smaller dots\cite{Barreiro2012} with diameters of about $1\,\mathrm{nm}$ causes a very steep envelope function. This case should result in, at most, one nuclear spin interacting with the electron spin. Both scenarios seem experimentally interesting in order to engineer QDs for different applications. A $^{13}{\rm C}$ enriched QD could potentially be used to prepare the electron spin very precisely in a certain superposition of spin up and down for subsequent experiments. A very small QD, in contrast, could serve as a storage for the electron spin where very long decoherence times are to be expected. Besides these technical points, graphene QDs could also serve as a rich playground to test fundamental aspects of quantum mechanics and quantum information theory in an interesting system-bath setup. If we consider the hosted electron spin as the system, we are able to control both its spatial size and its state electrostatically. Moreover, the electron spin can be straightforwardly addressed via external magnetic fields or in an optical way. In contrast, a direct preparation of the state of the nuclear spin bath seems challenging. However, the size of the nuclear spin bath can be modified systematically by isotopic purification of either $^{12}{\rm C}$ or $^{13}{\rm C}$. Finally, as we argued above, the design of the QD enables the experimentalist to manipulate the strength as well as the nature of the system-bath interaction. Thus, these considerations render graphene QDs to be a flexible system-bath realization offering a controllable, fermionic bath of spin $1/2$ nuclei. Given these opportunities, our setup not only seems very promising for studying a quantum to classical crossover as a function of the bath size in a fermionic environment, but also for more advanced concepts of quantum information theory such as quantum Darwinism\cite{Zurek2009}. With the notion of ``quantum Darwinism,'' W. H. Zurek summarizes his ideas of emergent classicality in a pure quantum universe, where the formation of classical and quantum system bath correlations, as well as the accompanied information exchange matter. Since measurements are most often indirect, relying on the environment as a mediator, it seems to be evident that the environment plays the crucial role in the creation of objective properties. However, as far as we know, this theory is up to now tested only in few experiments\cite{Burke2010,Cornelio2012} in a rather indirect way. In our opinion, taking advantage of the controllable spin bath in graphene QDs, this system offers a unique opportunity to reveal new insights in the role of the environment in the classical world we experience. \section{Acknowledgements} \label{sec:aknowledgement} We would like to thank Manuel Schmidt for valuable discussions. M.F. and B.T. acknowledge support from the Priority Program 1459 ``Graphene'' of the DFG and from the Euro-GRAPHENE Program of the ESF. The work of J.S. was supported by DFG via Collaborative Research Center 631.
\section{Introduction and Motivations} The precise definition and quantitative description of memory effects (or non-Markovianity) have become a central issue in the theory of open quantum systems \cite{uno,blp,Chruscinski2010,rhp,sun,erica,vasile,luo,due,spettro,rev}, and have been the subject of recent experimental efforts \cite{experiments}. It has been argued that non-Markovianity has the potential of being exploited to pursue new quantum technologies \cite{due}, and that it can be thought as a resource in quantum metrology \cite{tre}, for the generation of entangled states \cite{entgeneration} and for quantum key distribution \cite{quattro}. Various characterizations of the non-Markovian behavior have been given, which capture different aspects of the decohering dynamics of an open system. They include the lack of divisibility of the map describing the time evolution of the system of interest \cite{uno,rhp}, or the back-flow of the information that the system itself had previously lost, described either in terms of the distinguishability of the evolved states \cite{blp} or of the quantum Fisher information \cite{sun}. Further proposals have been put forward, based on the decay rates entering the master equation \cite{erica}, on the use of the quantum mutual information \cite{luo} or of channel capacities \cite{due}, on spectral considerations \cite{spettro}, and on the temporary expansion of the volume of the states accessible through the reduced dynamics \cite{det}. Moreover, many other properties, related to the locality \cite{sei}, to the complexity \cite{znidaric} or to the size of the environment \cite{lorenzoPRA}, have been investigated in different physical settings, ranging from spin systems \cite{apollaro11,haikkagold} to Bose-Einstein condensates of ultra-cold atoms \cite{haikkabec}. Despite the conceptual differences, several quantifiers of the amount of non-Markovianity give similar qualitative (and sometimes also quantitative) descriptions when applied to the dynamics of simple quantum systems such as a qubit\cite{altricin,haikkagold}. In particular, this holds true for the specific case of a purely dephasing dynamics where the system loses coherence due to its interaction with the environment, without any energy exchange. In this case, indeed, the open system evolution is completely characterized by a so called decoherence factor, which is the only ingredient necessary to evaluate the amount of non-Markovianity. Specifically, let us consider a two-level system (a qubit, with energy eigenstates $\ket{\alpha}$, $\alpha=0,1$) interacting with its environment in such a way as to preserve its energy. This implies that the interaction Hamiltonian commutes with that of the qubit and, as a result, the qubit state can be written as \be \rho(t) = \begin{pmatrix}\rho_{00}(0) & \nu(t) \rho_{01} (0) \\ \nu^*(t) \, \rho_{10}(0) & \rho_{11}(0) \end{pmatrix}, \quad \nu(t)=Tr_{env} \{e^{i H_{1} t} \, e^{-i H_{0} t} \, \rho_{env}\} \, ,\label{tev} \ee where $\rho_{env}$ is the initial state of the environment, while $H_{\alpha} = H_{env} + \bra{\alpha} H_{int} \ket{\alpha}$ are effective Hamiltonians for the environment, conditioned on the state of the qubit ($\alpha =0,1$). If the initial state of the environment is pure, $\rho_{env} = \ket{\phi}\bra{\phi}$, then the decoherence factor (whose square is known as the Loschmidt echo, $L(t) = \modulo{\nu(t)}^2$ \cite{losch}) is given by the overlap $\nu(t) = \braket{\phi_1(t)}{\phi_0(t)}$, where $\ket{\phi_{\alpha}} = e^{-i H_{\alpha} t} \ket{\phi}$. The quantity $\nu(t)$ can be used to characterize the environment itself, and, in particular, it gives a very peculiar behavior for fermionic environment. Indeed, as first pointed out by P. W. Anderson over 40 years ago, such an overlap of the two many-body wavefunctions, describing deformed and un-deformed Fermi seas, respectively, scales with the size of the environment and vanishes in the thermodynamic limit, giving rise to an `orthogonality catastrophe' \cite{anderson}. The dynamic counterpart of Anderson's theory was investigated a few years later with the prediction of a universal absorption-edge singularity in the X-ray spectrum of simple metals, which has become known as the `Fermi-edge singularity' \cite{mnd}. Mahan, Nozieres and De Dominicis (MND) obtained an expression for the function $\nu(t)$ describing the response of a Fermi gas to the sudden switching of a local perturbation, i.e. a core-hole induced by the X-ray, which gives rise to a deformation (or shake up) of the many body state of the gas. It is our aim in this paper to study the analogous of such a phenomenon for a trapped gas of ultra-cold Fermi atoms in which the very fast excitation of an impurity atom (e.g. by a focused laser pulse) produces a sudden local perturbation \cite{catastro1,catastro2,knap}. The time response of the gas is directly related to the decoherence of the impurity, which experiences a purely dephasing dynamics. As mentioned above, for such a case the non-Markovianity of the map is strictly connected to the decoherence factor. As a result, with our theoretical construction we are able to explore the link between the non-Markovianity and the orthogonality catastrophe occurring within the environment. We will adopt the geometric measure recently put forward by some of us in Ref. \cite{det}, which allows for an intuitive visualization of the information exchange between system and environment. This is briefly recalled in Sect. \ref{secuno}. Then, after the model for the fermionic bath is explicitly described in Sec. \ref{secdue}, in Sec. \ref{sectre} we use the geometric measure to discuss the decoherence of the impurity. Some final remarks are given in Sec. \ref{secquattro}. \section{Geometric description of non-Markovianity} \label{secuno} In this section we briefly review the definition and meaning of the measure of non-Markovianity introduced in Ref. \cite{det}, adapting it to the case of a qubit undergoing a purely dephasing dynamics. The basic idea is that the time evolution of the density matrix of a qubit can always be recast into the form of an affine transformation for the Bloch vector $\vett b = \mbox{Tr} \{ \vett{\sigma} \}$ (where $\vett{\sigma}$ is the vector of Pauli matrices), which can be contracted, rotated and translated by a given amount. In particular, for the time evolution given in Eq. (\ref{tev}), the translation term is absent and we have \be \vett b(t) = A_t \, \vett b(0) \, ,\ee with \begin{equation} A_t=\left(\begin{matrix} \mathfrak{Re} \nu(t) & \mathfrak{Im} \nu(t) & 0 \\ - \mathfrak{Im} \nu(t) & \mathfrak{Re} \nu(t) & 0 \\ 0 & 0 & 1 \end{matrix}\right) \, . \end{equation} The generic initial state for the qubit corresponds to a Bloch vector lying within the unit (Bloch) sphere. The set of accessible states changes as a function of time, being contracted (with respect to the initial sphere) in the equatorial plane for the case of a purely dephasing evolution. In Fig. \ref{detpal} we provide a representation of this set as a function of time for two specific situations which will be described in more detail in the next section. \begin{figure}[t] \centerline{\scalebox{0.3}{\includegraphics{detpal_new.eps}}} \caption{Behavior of the determinant $|A_t|$, of the dynamical map governing the time evolution of the two-level impurity atom, for the case of an environment with $N_{F}=200$ fermions and for $\beta=0.05$ , $\alpha=0.001$ (left plot) and $\beta=3$ , $\alpha=0.1$ (right plot). (Energies are in units of $\hbar \omega$).} \label{detpal} \end{figure} In the general case, the absolute value of the determinant of the matrix describing the dynamical map, $||A_t||$, gives the ratio between the volume of the set of states (or, more precisely, the set of $\vett b$ vectors) accessible by the system at time $t$ and the volume of the initial set of all possible Bloch vectors, which is the entire Bloch sphere. A dynamical map described by a Lindblad-like master equation gives rise to a non-increasing volume \cite{uno,det}. This is true even if the master equation has time decay coefficients, provided that the latter are strictly positive quantities \cite{piilo}. Hence, it is natural to call a process Markovian if the determinant does not increase in time, and non-Markovian if an initial volume contraction is followed by a temporary inflation, giving rise to a determinant that has a positive time derivative in some specific interval. The two examples reported in Fig. \ref{detpal} explicitly depict a Markovian and a non-Markovian evolution in terms of the determinant. Using the same method adopted in Ref. \cite{blp} to single out the intervals in which the determinant increases in time, we define \cite{det} \begin{equation} \mathcal{N}_V=\int_{\frac{d\left||A_t|\right|}{dt}>0}\frac{d\left||A_t|\right|}{dt} \, dt \label{NMnostra} \end{equation} as a non-Markovianity measure. For the map corresponding to a dephasing evolution, the determinant is related to the decoherence factor. Explicitly, $\left||A_t|\right|=|\nu(t)|^2$. In Section \ref{sectre}, we will discuss the behavior of ${\cal N}_V$ for a qubit in a fermionic environment as a function of the coupling strength and of the temperature of the environment. \section{Impurity in a Fermionic environment} \label{secdue} The explicit model for the environment that we discuss in this paper is given by a gas of ultra-cold non-interacting fermionic atoms, trapped in an harmonic potential of frequency $\omega$. This is described by the Hamiltonian $$\hat{H}_{env} = \sum_{n} \varepsilon_{n} {c}_{n}^{\dagger} {c}_{n} \, ,$$ with $c_{n}$ being the annihilation operator for the $n$-th single particle level of energy $\varepsilon_{n} = \hbar \omega (n+1/2)$. $\hat{H}_{env}$, together with the number operator $\hat N = \sum_{n} {c}_{n}^{\dagger} {c}_{n}$, also sets the initial equilibrium state of the gas, $\rho_{env} = \exp \{ - \beta ( \hat H_{env} - \mu \hat N)\} /Z$, where the chemical potential $\mu$ is fixed by the requirement that the gas contains (on average) $N_F$ fermions, while $\beta$ is the inverse temperature. We consider a two-level impurity, trapped in an auxiliary potential and brought in contact with the Fermi gas. We assume that when the impurity is in the state $|0\rangle$, it has a negligible scattering interaction with the gas. On the other hand, if the impurity is in the state $\ket 1$, the gas feels a localized perturbation $\hat V$, describing a neutral $s$-wave like interaction that we treat in the pseudo-potential approximation: $V(x)=\pi V_{0}x_{0}\delta (x)$ (the trap length $x_{0}$ as well as the factor $\pi$ are put in the definition for future convenience only). The interaction Hamiltonian, then, has the form \be H_{int} = \hat{V} \otimes \ket{e}\bra{e} \, . \ee As mentioned in the introduction, the key quantity for our discussion is the decoherence factor \be \nu(t) = \left\langle e^{\frac{i}{\hbar }\hat{H}_{env}t}{\,}e^{-\frac{i}{\hbar} (\hat{H}_{env}+\hat{V}) t}\right\rangle \text{,} \label{nu} \end{equation} where the braket symbol is a short-hand notation for the thermal equilibrium average over the unperturbed environment. An analytic estimate can be given for this quantity both at zero and at finite temperatures \cite{catastro2}. Here, we will proceed to a numerical evaluation of the decoherence factor (and, in particular, of its modulus) using the linked cluster theorem and up to two-vertices connected Feynman diagrams, which amount to a partial re-summation of a perturbative expansion in the ratio of the interaction strength $V_0$ with the Fermi energy $\varepsilon_F = \hbar \omega (N_F + \frac{1}{2})$. The details of this kind of evaluation are given in Ref. \cite{catastro2}, where it is shown that a good indicator of the effect of the perturbation induced on the Fermi gas is the a-dimensional parameter \be \alpha =\frac{V_{0}^{2}}{\hbar \omega \, \varepsilon_F} \, .\ee This coincides with the critical parameter of the MND theory, which is obtained as the limiting case of a free Fermi gas, $\omega \rightarrow 0$. In the following, we will consider spin $1/2$ fermions, fix their number to $N_F = 200$, and discuss the behavior of the measure of non-Markovianity ${\cal N}_V$ as a function of $\alpha$ and of the inverse temperature $\beta = \frac{1}{KT}$. \section{Non-Markovianity and its relation to the shake-up of the Fermi gas} \label{sectre} The parameter $\alpha$ introduced in the previous section is a measure of the strength of the perturbation due to the switching impurity. One could naively expect that by increasing $\alpha$ the amount of non-Markovianity in the dynamics of the qubit should increase. However, this is not the case as Fig. \ref{noncre} clearly shows. In particular, ${\cal N}_V$ increases for a very small $\alpha$, reaching a maximum value for $\alpha \leq 0.2$, which depends on the chosen value of $\beta$. After such a maximum, ${\cal N}_V$ decreases with increasing the interaction strength. \begin{figure}[t] \centerline{\scalebox{0.3}{\includegraphics{noncre2_new.eps}}} \caption{Non-Markovianity measure ${\cal N}_V$ as function of the effective coupling strength $\alpha$, for various values of the inverse thermal energy $\beta$ expressed in units of $1/\hbar \omega$.} \label{noncre} \end{figure} This is due to the peculiar way a fermionic system responds to the perturbation, especially at small temperatures. For very small $\alpha$, an almost linear increase with the intensity of the perturbation is expected from simple second order perturbation theory. Indeed, to first order in $V_0$ one would obtain only a shift of the energy levels, resulting in a purely oscillating $\nu(t)$; the second order correction in $V_0$ (which are linear in $\alpha$), instead, introduces a distortion of the single-particle energy eigenstates. A small $\alpha$ implies that the Fermi surface is not substantially modified and that only those fermions whose energy is close to the Fermi energy are excited. This, in turns, means that only a few fermionic modes (and, thus, few almost undistorted frequencies) enter the dynamics, and give rise to a quasi-periodicity of the function $\nu(t)$ with frequency $\omega$. As a result, every half a period the derivative of the determinant changes sign and a contribution is given to the integral in Eq. (\ref{NMnostra}). The accumulation of such positive contributions gives rise to an ${\cal N}_V$ that grows with $\alpha$. Notice, however, that a true periodicity is quickly lost with increasing $\alpha$, especially if the temperature is not kept low. From Fig. \ref{detpal}, one can see that only a vague periodicity survives already at $\alpha=0.1$ if $KT = \hbar \omega/3$. This increase of ${\cal N}_V$ with increasing $\alpha$ is counteracted by an effective suppression of the oscillations in $\nu(t)$ which occurs when the single-particle energies become more and more distorted (so that they are not anymore multiple of $\omega$) and when more and more transitions are induced by the increasing-in-strength of the perturbation. For large values of $\alpha$, indeed, the entire Fermi sea responds to the perturbation and the non-markovianity decreases. Such a behavior can be also interpreted in a complementary way. For small $\alpha$'s the effective environment felt by the impurity has a very prominent spectral structure, given by the Fermi edge. On the other hand, with increasing $\alpha$, the effective environmental frequency spectrum becomes more and more flat, giving rise to an effective Markovian dynamics for the qubit. This line of reasoning is confirmed by the fact that the amount of non-Markovianity decreases with increasing the temperature due to the fact that the Fermi edge is more and more blurred for a smaller and smaller $\beta$. Fig. \ref{omeg} gives the behavior of the amount of non-Markovianity as function of the trap frequency, to explicitly confirm that ${\cal N}_V$ increases with $\omega$ as a result of the fact that the periods in which the determinant grows become closer to each other in time. \begin{figure}[t] \centerline{\scalebox{0.6}{\includegraphics{omeg_new.eps}}} \caption{Non-Markovianity measure as function of the trap frequency $\omega$ for different temperatures.} \label{omeg} \end{figure} In particular, for the free gas originally treated by MND, which is obtained when $\omega\rightarrow 0$, the dynamics of the impurity is fully markovian due to the absence of any oscillations in the decoherence factor. \subsection{Build-up in time of the non-Markovianity} The non-Markovianity measure ${\cal N}_V$ gives an integral characterization of the whole dynamics. More details on the time development of the memory effects during the system-environment information-exchange process can be obtained if we consider separately the time intervals in which the determinant increases and decreases, up to a given time $t$. To this end, it is useful to define \begin{equation} \mathcal{N}_{\pm}(t)= \pm \int_{\pm \frac{d\left||A_{\tau}|\right|}{d \tau}>0}\frac{d\left||A_{\tau}|\right|}{d \tau} \, \theta(t-\tau) \, d\tau \label{pm} \, , \end{equation} which give, respectively, (the sum of ) the amount of expansion/contraction of the volume of the set of accessible states within the time $t$. In particular, it is meaningful to consider the ratio \be {\cal R}(t) = \frac{N_+(t)}{ N_-(t)} \, ,\ee which expresses the fraction of the volume which is recovered in the expansions with respect to the one lost due to the previous contractions. The function ${\cal R} (t)$ is strictly zero until the volume starts increasing and then it grows/diminishes with the volume. Every time interval in which ${\cal R}$ grows corresponds to a positive contribution to the build-up of the integral measure ${\cal N}_V$. \begin{figure}[t] \centerline{\scalebox{0.3}{\includegraphics{ratio_new.eps}}} \caption{Ratio expansion over contraction factors of the volume of states accessible through the dynamical map after time $t$. The left panel reports ${\cal R}(t)$ at fixed $\alpha=0.01$ for various temperatures, while the right panel shows the same quantity for different values of $\alpha$ and a fixed $KT = \hbar \omega/3$.} \label{ratio} \end{figure} Fig. \ref{ratio} confirms the interpretation that the non-Markovianity is due to quasi-periodic oscillations of the volume of the set of accessible states, occurring with a period $2 \pi/\omega$, which is more and more suppressed with increasing temperature \section{Concluding remarks} \label{secquattro} In this paper we studied the dynamics of a two-level impurity encapsulated in a trapped fermion environment. Due to the presence of the Fermi edge, at low temperatures such an environment enjoys the presence of a spectral structure which induces a non-Markovian behavior in the time evolution of the impurity. This is due to the oscillating response of the trapped Fermi gas in which transition are induced by the perturbing local impurity. Such a behavior is found to occur provided the interaction strength is not too large (with respect to the relevant energy scale of the gas, given by $\sqrt{\varepsilon_F \hbar \omega}$). On the other hand, a stronger interaction tends to smoothen the environmental spectral density giving rise to a Markovian dynamics for the impurity. Markovianity is generically obtained also if the temperature is larger than the scale set up by the trapping frequency and, more generally, whenever the discrete structure of the single-particle energy levels of the gas can be confused with a continuum, e. g. in the absence of the trap. The effects that we reported can be studied with a system of trapped cold fermionic atoms. Indeed, in this realm, many experiments have recently dealt with the effect of impurities within the gas~\cite{fermionexp}, which can be easily tested both as function of the interaction strength (changeable via the phenomenon of Feshbach resonance) and of the trapping frequency.
\section{Introduction} Localization plays a crucial role for understanding the exact quantization of the quantum Hall effect. In strong magnetic fields, electronic transport can be described in terms of narrow edge channels, leading to the well-known observations of $R_\mathrm{{xx}}\approx0$ and $R_\mathrm{{xy}}=h/(e^2\nu_\mathrm{bulk})$. In this case, varying the magnetic field only (de)populates localized states in the bulk, which do not affect transport because backscattering of the chiral edge states across the wide bulk region is negligible. Alternative pictures, where the current is believed to flow in the bulk, exist (see for example Ref. \onlinecite{komiyama_edge_1998} for an overview). Also in this case, the (de)population of localized states plays the key role for the conductance quantization. These theoretically predicted localized states have been investigated in various experiments using spatially resolved imaging techniques \cite{tessmer_subsurface_1998,ilani_microscopic_2004, martin_localization_2004, steele_imaging_2005,hashimoto_quantum_2008} or single-electron transistors fabricated on top of a two-dimensional electron gas (2DEG) \cite{wei_edge_1998}. In a pioneering work by \textit{Ilani et al.} \cite{ilani_microscopic_2004}, the equilibrium properties of such localized states have been investigated in the bulk of a 2DEG with scanning single-electron transistor (SET) techniques. The behavior of the bulk localizations was shown to be dominated by Coulomb blockade physics. This picture cannot be explained as a single-electron effect, but requires the formation of compressible and incompressible regions in the bulk, in analogy to an edge state picture that takes self-consistent screening into account \cite{chklovskii_electrostatics_1992, efros_homogeneous_1992, cooper_coulomb_1993}. Here, the system is decomposed into compressible regions, in which potential fluctuations are screened and the density varies, and incompressible regions of constant density but varying background potential. Apart from the aforementioned experiments which probe localizations on a very local scale in the bulk, conductance fluctuations in the quantum Hall regime, believed to be related to localized states, have been studied in direct transport experiments \cite{simmons_resistance_1989,simmons_resistance_1991,cobden_fluctuations_1999,machida_resistance_2001,couturaud_local_2009,peled_near-perfect_2003,staring_coulomb-blockade_1992,main_resistance_1994,timp_quantum_1987}. They have been investigated for example in Si-MOSFETs \cite{cobden_fluctuations_1999}, Graphene \cite{velasco_probing_2010,martin_nature_2009,branchaud_transport_2010}, InGaAs quantum wells \cite{granger_few-electron_2011} and in narrow AlGaAs/GaAs heterostructures \cite{simmons_resistance_1989,simmons_resistance_1991}, where localized states couple to the edge and thus become accessible. In the latter experiments \cite{simmons_resistance_1989,simmons_resistance_1991}, resistance fluctuations have been interpreted as magnetically bound states. As pointed out later \cite{lee_comment_1990,rosenow_influence_2007,halperin_theory_2011}, Coulomb blockade effects are of great importance for such experiments and have to be taken into account for the interpretation of $B$-field and gate-voltage periodicities. In the work of \textit{Cobden et al.} \cite{cobden_fluctuations_1999}, conductance fluctuations in the quantum Hall regime span a distinct pattern in the density vs. magnetic field plane, with resonances parallel to neighboring conductance plateaus, similar to the phase diagram obtained by \textit{Ilani et al.} This has been interpreted as Coulomb charging of localized states in the bulk of the employed small structures. The absence of a clear periodicity suggests either the contribution of many localized states or the validity of other interpretations\cite{machida_resistance_2001}, which are based on the presence of a network of compressible stripes.\\ More recently, scanning gate experiments have tried to combine spatial resolution with transport \cite{hackens_imaging_2010,martins_scanning_2013,martins_coherent_2013}. \textit{Hackens et al.} have investigated Coulomb dominated islands inside quantum Hall interferometers\cite{hackens_imaging_2010}. Modulations of transport, due to the coupling of the localized islands to the edge states were found. In contrast to this behavior dominated by Coulomb charging, recent experiments \cite{martins_coherent_2013} report phase coherent tunneling across constrictions in the quantum Hall regime.\\ Quantum point contacts are one of the conceptually most simple, though interesting systems studied in mesoscopic physics. Despite their simplicity, complex many-particle phenomena like g-factor enhancement \cite{martin_enhanced_2008,yoon_nonlocal_2009,rossler_transport_2011} or the 0.7 anomaly \cite{thomas_possible_1996,kristensen_bias_2000,cronenwett_low-temperature_2002,komijani_evidence_2010,micolich_what_2011,komijani_origins_2013} are observed. Furthermore, they offer the possibility of locally probing a physical system, which is for example used in charge detection experiments \cite{gustavsson_counting_2006-1,gustavsson_frequency-selective_2007,rossler_tunable_2013,baer_cyclic_2013} or tunneling experiments in the quantum Hall regime \cite{milliken_indications_1996,chang_observation_1996,roddaro_interedge_2004}. This allows us to employ QPCs in the quantum Hall regime for investigating the influence of disorder-induced localizations on transport. The influence of localized states on (fractional) quantum Hall states confined in QPCs is not fully understood. Furthermore, the influence of individual localizations on non-equilibrium transport was not accessible in the mentioned transport experiments. In contrast, scanning SET experiments provided information about individual localizations, but not about their influence on transport. For the interpretation of interference experiments in the quantum Hall regime \cite{zhang_distinct_2009,kou_coulomb_2012,camino_quantum_2007,mcclure_fabry-perot_2012,willett_measurement_2009}, a detailed understanding of the transmission properties of single QPCs is necessary. Even in 2DEGs with the highest mobilities technologically achievable at the moment, disorder significantly influences transport through the QPC, as soon as a perpendicular magnetic field is applied. We show that even in simple QPCs complicated behavior can be observed, which is interpreted in terms of single- and many-electron physics of individual disorder-induced localizations. We argue that the influence of localizations can be minimized by employing growth and gating techniques, which result in a very steep QPC confinement potential (perpendicular to transport direction) and low disorder in the channel. By this, the $\nu=5/2$ state can be confined to a QPC without noticeable backscattering and preserving the bulk properties in an unprecedented quality, giving a good starting position for tunneling- and interference experiments in the second Landau level.\\ \section{Experimental details} The QPCs used in this study are defined by electron-beam lithography and subsequent Ti/Au evaporation on photo-lithographically patterned high-mobility wafers. Constrictions with different geometries have been studied for this manuscript (see Tab.\ref{TabelleSamples} for an overview). For the 250 nm wide QPC I.a and the 500 nm wide QPC I.b, a 30 nm wide quantum well with a carrier sheet density $n_\mathrm{s}\approx3.04\times 10^{11}~$cm$^{-2}$ and a mobility $\mu\approx13\times 10^6~$cm$^{2}$/Vs has been used. In this structure the 165 nm deep quantum well is neighbored by two $\delta$-Si doped GaAs layers, enclosed in 2 nm thick layers of AlAs. These screening layers reside 70 nm below and above the 2DEG. The electrons in the AlAs wells populate the X-band and provide additional low-mobility electron layers, which screen the $\Gamma$-electrons in the 2DEG from remote ionized impurities. Due to the screening layers, hysteretic and time-dependent processes make gating difficult. The gating properties of these wafers have been studied earlier \cite{rossler_gating_2010}. The 1.2 $\mathrm{\mu}$m wide QPC III.a has been fabricated on a wafer which employs a similar growth technique ($\mu \approx 17.8\times 10^6~$cm$^2$/Vs, $n_\mathrm{s} \approx 2.13\times 10^{11}~$cm$^{-2}$, 250 nm deep, 30 nm wide quantum well, screening layers 100 nm below and above the 2DEG). The high mobility structures used for QPCs I.a,b and III.a are optimized for the $\nu=5/2$ state without the requirement of prior LED illumination \cite{reichl_increasing_2014}. QPC II.a, QPC II.b (both 700 nm wide), QPC II.c (900 nm wide) and QPC II.d (800 nm wide) were fabricated on a single side doped GaAs/Al$_\mathrm{x}$Ga$_\mathrm{1-x}$As heterostructure with a mobility of approximately $8\times10^6$ cm$^2$/Vs and a 320 nm deep 2DEG with an electron density of approximately $1.5\times10^{11}$cm$^{-2}$. Hysteresis effects are much less pronounced in these structures which employ a reduced proportion of Al in the spacer layer between the doping plane and the 2DEG (x=0.24 compared to typically x=0.30$-$0.33). Long-range scattering is therefore reduced, thus facilitating the formation of the $\nu$=5/2 state and other fragile FQH states \cite{pan_impact_2011,gamez_5/2_2013}. The measurements have been conducted in a dilution refrigerator at a base temperature of approximately 85 mK and in magnetic fields up to 13 T. Measurements of QPC III.a have been performed in a dry dilution refrigerator with an electronic temperature of approximately 12 mK, achieved by massive filtering and thermal anchoring at every temperature stage. Standard four-terminal lock-in measurement techniques have been used to measure $R_\mathrm{xx}$ and $R_\mathrm{xy}$ of the bulk 2DEG and the differential conductance the QPC, $G=\partial I_\mathrm{AC}/\partial V_\mathrm{diag}$, which gives access to the effective QPC filling factor $\nu_\mathrm{QPC}$\cite{beenakker_quantum_2004}. Here, the voltage drop $V_\mathrm{diag}$ is measured diagonally across the QPC. \begin{table*} \begin{tabular}{ccccc} \toprule QPC & $w$ (nm) & Heterostructure & $n_s$ (10$^{11}$ cm$^{-2}$) & $\mu$ (10$^{6}$ cm$^{2}$/Vs) \\ \toprule QPCI.a & 250 & 30 nm QW, $\delta$-Si screening & 3.0 & 13.0 \\ \cline{1-2} QPCI.b & 500 & & & \\\hline QPCII.a & 700 & & 1.5 & 8.0 \\\cline{1-2} QPCII.b & 700 & Single-side doped & & \\\cline{1-2} QPCII.c & 900 & GaAs/Al$_{0.24}$Ga$_{0.76}$As heterostructure & & \\\cline{1-2} QPCII.d & 800 & & & \\ \hline QPCIII.a & 1200 & 30 nm QW, $\delta$-Si screening & 2.1 & 17.8 \\ \hline \end{tabular} \caption{Overview of the different samples used in this study. Channel width $w$, electron sheet density $n_s$ and mobility $\mu$ are indicated for the different QPCs.} \label{TabelleSamples} \end{table*} \section{Results and discussion} The main part of this article will be organized as follows: First, an exemplary quantum Hall phase diagram will be discussed (\ref{phase}). The influence of different QPC geometries on the width of the incompressible region separating the edge states and the density distribution is discussed in section \ref{geometry}. In the main part of our article, section \ref{main}, QPC resonances are characterized and explained via a microscopic model. A short summary of this central chapter is given afterwards. The resonances' dependence on the spatial position of the conducting channel inside the QPC is investigated in the following (\ref{AsymmetrieChap}). At the end, methods for confining the most fragile fractional quantum Hall states are discussed (\ref{Fragile}). \subsection{Quantum Hall phase diagram of a QPC} \label{phase} \begin{figure} \begin{center} \includegraphics[width=8cm]{FIG-Schmetterling.pdf} \end{center} \caption{(Color online) \textbf{a:} Transconductance of QPC I.a as a function of the QPC voltage and the magnetic field. Conductance plateaus at multiples of $e^2/h$ can be seen as black regions. Resonances, bending in the $B-V_\mathrm{QPC}$-plane are indicated by white dashed lines. The 1/$B$ periodic kinks (white arrows) are believed to originate from a change of the filling factor in the bulk. A possible combination of bulk filling factors is indicated. Inset: full B-field and voltage dependence of the system. Here, the voltage was swept from -1.3 V $\rightarrow$ -2.7 V $\rightarrow$ -1.3 V repeatedly and the $B$-field was stepped from 0 T $\rightarrow$ 13 T $\rightarrow$ 0 T. \textbf{b:} The $B$=0 QPC conductance plateaus at multiples of $2\times e^2/h$ spin-split for increasing magnetic fields (magnetic field from 0 T to 3 T). For $0<G<2\times e^2/h$ and $2\times e^2/h<G<4\times e^2/h$, local minima in the slope of the conductance are marked by green diamonds or red circles. In contrast to the second and third subband, the spin-splitting of the lowest subband starts at conductance values of approx. $0.7\times2\times e^2/h$ and approaches $1\times e^2/h$ as the magnetic field strength is increased \cite{wharam_one-dimensional_1988, van_wees_quantized_1988-1}. \textbf{c:} At strong magnetic fields (with bulk filling factors $\nu_{bulk}$), conductance plateaus corresponding to different integer and fractional filling factors in the QPC can be observed.} \label{Schmetterling} \end{figure} Fig. \ref{Schmetterling}.a shows the differential conductance $G$ (plotted: numerical derivative $\partial G/\partial V_\mathrm{QPC}$ in colorscale) of QPC I.a as a function of the voltage applied to the QPC gates ($V_\mathrm{QPC}$) and a perpendicular magnetic field $B$. At zero magnetic field, the well-known QPC conductance quantization in multiples of $2\times e^2/h$ is found. As the magnetic field is increased, conductance steps (or plateaus), seen as maxima (or black areas) of $\partial G/\partial V_\mathrm{QPC}$, bend to more positive QPC voltages, due to magneto-electric depopulation of the QPC channel \cite{buttiker_quantized_1990,beenakker_quantum_2004}. The quantized conductance plateaus successively develop into regions of constant effective filling factor of the QPC ($\nu_\mathrm{QPC}$) with a diagonal resistance $R_\mathrm{diag}=h/(e^2\nu_\mathrm{QPC})$. In this regime the spin splitting is sufficiently strong to also observe conductance plateaus at $G=1,3,5,...~\times e^2/h$ \cite{wharam_one-dimensional_1988, van_wees_quantized_1988-1}. The low-field behavior of the spin-splitting is shown in Fig. \ref{Schmetterling}.b. Numerically extracted local minima of the slope of the conductance curve have been marked with (green) diamonds / (red) circles. As $B~\rightarrow~0$, the $G=1\times e^2/h$ plateau seems to join the $0.7\times2\times e^2/h$ anomaly\cite{graham_interaction_2003}. No similar behavior can be observed at $G=3\times e^2/h$ and $G=5\times e^2/h$. In the quantum Hall regime, conductance curves of the QPC (Fig \ref{Schmetterling}.c) show fractional effective filling factors at $\nu_\mathrm{QPC}=2/3$ and $\nu_\mathrm{QPC}=4/3$ for different integer and fractional filling factors of the bulk ($\nu_\mathrm{bulk})$. The shape of the boundary of the $G=0$ region of Fig. \ref{Schmetterling}.a is determined by different effects: first, an increasing magnetic field leads to magneto-electric depopulation due to an increase of the single-particle electron energy, thus moving the pinch-off region to less negative gate voltages. In addition, time-dependent and hysteretic processes of the X-electron screening layers lead to an additional drift of the pinch-off line towards less negative QPC voltages. The small inset shows the $B$-field and voltage dependence of the system. Here, the voltage was swept from -1.3 V $\rightarrow$ -2.7 V $\rightarrow$ -1.3 V repeatedly (horizontal axis) and the $B$-field was stepped from 0 T $\rightarrow$ 13 T $\rightarrow$ 0 T (vertical axis). Upper and lower part of the inset are not mirror symmetric - over the time of the measurement, the pinch-off-line drifts towards less negative voltages, indicating a time-dependence of the system. Furthermore, changes of the filling factor in the bulk can lead to an abrupt decrease of the Fermi energy of the system as observed in quantum dots (QDs) \cite{ciorga_addition_2000}. This effect is believed to cause the $1/B$-periodic kinks in the pinch-off line. When increasing the $B$-field across the kinks, the pinch-off line suddenly moves towards more positive QPC voltages (marked by white arrows in Fig. \ref{Schmetterling}.a), though an assignment to the individual filling factors in the bulk is not uniquely possible, probably due to a reduced density in the bulk near the QPC, which governs the local coupling of bulk states into the QPC. In the regions of Fig. \ref{Schmetterling}.a where the QPC filling factor is changing, the QPC conductance does not vary monotonically. For $G>1\times e^2/h$, resonances which are parallel to the boundary of one of the neighboring conductance plateaus are observed (green (gray) arrows). In contrast, the region $G<1\times e^2/h$ shows resonant features without any preferred slope, or even with varying slope at different $B$-fields (a set of bending resonances is indicated by the white dashed line). Very similar resonances have been found in several QPCs in different cooldowns. The origin of these resonances will be discussed later in the framework of single- and many-electron physics. \subsection{Influence of QPC geometry on incompressible separating region and density distribution} \label{geometry} \begin{figure} \begin{center} \includegraphics[width=8cm]{FIG-DensityCuts.pdf} \end{center} \caption{(Color online) Density distributions in the channels of QPCII.a (\textbf{a}) and QPCII.b (\textbf{b}). The original zero magnetic field density (dashed (blue), blow-up in second row) has been altered by the formation of an incompressible region in the center of the constriction at $B=1.71$ T ($\nu_\mathrm{QPC}=2$), resulting in a region of constant density $n_0$. The width $\Delta$a of the incompressible region is indicated for different filling factors $\nu_\mathrm{QPC}$ in both QPCs.} \label{Simulation} \end{figure} To be able to understand the mechanisms behind the resonances in more detail, we have investigated two different QPC designs, fabricated on a 2DEG of lower density. QPC II.a is 700 nm wide with a top-gate above the conducting channel (see inset Fig. \ref{Simulation}.a). QPCs II.b (inset Fig. \ref{Simulation}.b) and II.c are standard 700 nm / 800 nm wide QPCs. Density profiles in the y-direction (along the lateral confinement potential) for the two QPC designs (at $B$ = 0) have been obtained from a self-consistent bandstructure calculation using next\textbf{nano}. The doping concentration was adjusted to account for surface charges and to reproduce the gate pinch-off voltages correctly. The applied voltages to the gates were chosen such that the calculated density at $B$ = 0 corresponds to the density necessary for $\nu_\mathrm{QPC}$ = 2 at $B$ = 1.71 T, as in the measurements. Using the calculated density profile at $B$ = 0, we have calculated the altered density profile when a compressible region is formed in the center of the QPC (Fig. \ref{Simulation}) and the width $\Delta$a of the incompressible region for different QPC filling factors, using the electrostatic model of \textit{Chklovskii et al.}\cite{chklovskii_ballistic_1993}. In this model, perfect metallic screening in the compressible regions is assumed. For the gap energies, $\hbar \omega_c$ and $\tilde{g}\mu_BB$ with an exchange enhanced $\tilde{g}\approx 4~$\cite{rossler_transport_2011} have been used as estimates for $\nu=2$ and $\nu=1$. The energy gap at $\nu=1/3$ has been measured (see appendix). In Fig. \ref{Simulation}, the resulting self-consistent densities (for $\nu_\mathrm{QPC}$ = 2) are shown as solid lines. The original density at $B$ = 0 (dashed (blue) line, second row) is modified by the formation of an incompressible stripe (constant density $n_0$) in the center of the channel. When a negative voltage is applied to the channel top-gate (CTG) of QPC II.a, the subband minimum is lifted. In this situation, the curvature of the electron density in the center of the constriction is small. When QPC II.b is tuned to a similar density in the constriction, the density curvature in the center is much greater, leading to a narrower compressible region (Fig.\ref{Simulation}.b). Comparing $\Delta$a of the two QPCs, we conclude that for QPC II.a, a significantly wider incompressible region is expected according to the model of \textit{Chklovskii et al.} \cite{chklovskii_ballistic_1993}. The widths $\Delta$a range from approximately 20 nm to 90 nm. Disorder potential fluctuations have typical length scales of the order of 100 nm \cite{sohrmann_compressibility_2007,ilani_microscopic_2004}. If the amplitude of such a disorder potential fluctuation in the incompressible region in the center of the QPC is large enough to create an intersection of the Landau level with the Fermi energy (Fig. \ref{Schema}.B,C, left column), compressible regions of enhanced or reduced density (Fig. \ref{Schema}.B,C, middle column) are formed. Thus, the small width of the incompressible region in QPCII.b (and hence in the QPCII.c with similar geometry) makes it less likely that a disorder potential fluctuation leads to the formation of a localization in the constriction. Furthermore, the coupling to such a localization is strongly varied as the width of the separating incompressible region changes, making the observation of periodic charging of a single localization impossible. To observe periodicities and study resonances in more detail, we now investigate electronic transport in QPC II.a, where a much stronger influence of disorder-induced localizations is expected. Here, a periodic behavior is expected over a larger parameter range, as the width of the incompressible regions separating edge and localizations is sufficiently wide. \subsection{Characterization of QPC resonances and microscopic model} \label{main} \subsubsection{Periodic conductance oscillations in QPCs of different geometries} \begin{figure*} \begin{center} \includegraphics[width=16cm]{FIG-Faecher.pdf} \end{center} \caption{(Color online) \textbf{a}: Transconductance of QPC II.a as a function of the voltage $V_{\mathrm{CTG}}$ and magnetic field $B$. Here, the density is tuned roughly linearly by the gate voltage. \textbf{b}: Transconductance of QPC II.c as a function of $V_\mathrm{QPC}$. Pronounced fractional and integer filling factors are observed (black regions A.1, A.2, etc.). Apart from these regions of nearly-perfect transmission, disorder modulates transport in other regions: for small transmission (C.0, C.1), small backscattering (B.1) and at the low density, low $B$-field end of the conductance plateaus (D.1). A similar behavior is found when one underlying edge state is perfectly transmitted (regions B.2 - D.2). \textbf{c:} Transconductance of QPC II.a, when -400 mV are applied to the CTG. \textbf{d:} Zoom of Fig. \ref{Faecher}.c: transition from $\nu_\mathrm{QPC}=2$ to $\nu_\mathrm{QPC}=1$. Two distinct slopes (green solid / white dashed lines), parallel to the boundary of the neighboring conductance plateaus, are observed. \textbf{e,f:} Close-ups of the conductance oscillations for $\nu_\mathrm{QPC}=1$ and $\nu_\mathrm{QPC}=1/3$ (enframed areas in \textbf{c}).} \label{Faecher} \end{figure*} The filling factor spectra of QPC II.a and QPC II.c are investigated similarly to the measurement of Fig. \ref{Schmetterling}.a, by varying the QPC gate voltage versus the magnetic field $B$. First, the channel top-gate voltage $V_\mathrm{CTG}$ has been varied (Fig. \ref{Faecher}.a). This gate varies the density of the channel roughly linearly with applied voltage (neglecting filling-factor dependent capacitances), as seen from the slope $dB/dV_\mathrm{CTG}\varpropto 1/\nu_\mathrm{QPC}$ of the conductance plateaus, which show up as black areas of quantized conductance. In addition to the full series of integer filling factors, fractional states at $\nu_\mathrm{QPC}=1/3$, 2/3, 4/3 and 5/3 can be observed. Close to the low- and high density edges of the conductance plateaus, sets of conductance oscillations with a slope parallel to the boundaries are observed, similar to the ones observed in small Hall-bars \cite{cobden_fluctuations_1999,machida_resistance_2001}. The slope and number of these resonances are independent of density and magnetic field strength.\\ A qualitatively similar behavior can be found for QPC II.c as $V_\mathrm{QPC}$ is varied (Fig.\ref{Faecher}.b). Here, regions of perfect transmission have been marked (A.1). Modulations occur at the low-density side (B.1) and high-density side of the conductance plateaus (C.1) or pinch-off (C.0). Furthermore, resonances at the low-density and low-$B$-field end of conductance plateaus are observed (D.1). These resonances disappear as the density and $B$-field strength increase. Similar regions can be attributed to higher filling factors, for which underlying edge states are perfectly transmitted (A.2-D.2). \\ As mentioned above, a much stronger influence of a disorder-induced localization is expected for QPC II.a, as the wider incompressible region is much more likely to accommodate one or several extrema of the disorder potential. Here, quasi-periodic conductance modulations should occur over a larger parameter range, as the width of the incompressible regions separating edge and localizations is sufficiently wide. In order to verify this expectation, Fig. \ref{Faecher}.c shows the transconductance of QPC II.a, obtained by keeping $V_\mathrm{CTG}$ fixed while varying $V_\mathrm{QPC}$ and $B$. Compared to Fig. \ref{Faecher}.b, more pronounced conductance oscillations are observed (red empty arrow). Especially gate voltage regions close to pinch-off are now dominated by equidistant conductance peaks parallel to the magnetic field axis. Regions C and D overlap, which can be seen from the coexistence of two different distinguishable slopes (indicated by green solid arrows). In the integer quantum Hall regime (for example in Fig. \ref{Faecher}.d), conductance oscillations with distinct slopes are observed between neighboring conductance plateaus. The resonances are parallel to either of the two neighboring plateau boundaries (Fig. \ref{Faecher}.d, green solid / white dashed line). Close to $\nu_\mathrm{QPC}=1/3$, strong resonances, parallel to the conductance plateau occur (Fig. \ref{Faecher}.f). At lower B-Fields, between $\nu_\mathrm{QPC}=1/3$ and $\nu_\mathrm{QPC}=1$, weak modulations with an intermediate slope are observed. \begin{figure*} \begin{center} \includegraphics[width=16cm]{FIG-Schema.pdf} \end{center} \caption{(Color online) Schematic guiding center energies of extended states and densities within a QPC for the different transmission situations A-D, as indicated in Fig. \ref{Faecher}.b. Empty / filled circles symbolize empty or occupied states. In the case of perfect transmission (\textbf{A}), adding an exemplary disorder potential does not alter the density distribution within the channel. As density fluctuations in the transmitted (\textbf{B}) or energetically lowest reflected (\textbf{C}) Landau level lead to partially occupied states at the Fermi energy, compressible regions of enhanced (\textbf{C}, green (dark grey) area) or reduced density (\textbf{B}, red striped area) are formed within the incompressible region. This gives rise to a quantized charge on the compressible regions of enhanced or reduced density formed in the incompressible region which separates the edge states. For smaller magnetic fields or stronger disorder fluctuations, wide compressible regions are absent and only states below the Fermi energy are occupied. Here, compressible regions of enhanced and reduced density modulate the transport in the constriction at the same time (\textbf{D}). As wide incompressible regions are absent, the compressible regions of enhanced or reduced density are no longer governed by Coulomb dominated physics. Here, single-electron resonances arise from localized states, encircling a certain number of magnetic flux quanta. In contrast to the Coulomb dominated mechanism, such single-electron resonances give rise to a dependence in the $B-V_\mathrm{QPC}$ plane which may differ from the slope of the conductance plateaus.} \label{Schema} \end{figure*} \subsubsection{Screening and localization model} The mechanism which gives rise to the different resonances in regions A-D can be understood in terms of an edge-state picture which takes non-linear screening of potential fluctuations into account (Fig. \ref{Schema}). Similar models have been employed to understand bulk localizations in scanning SET and scanning capacitance experiments \cite{ilani_microscopic_2004,martin_localization_2004,steele_imaging_2005,steele_imaging_2006}. Regions of locally enhanced or reduced density are formed on top of the background density, associated with different extended quantum Hall states. These localizations in the constriction couple to the edge states and give rise to conductance oscillations. In Fig. \ref{Schema}, the guiding center energies of two extended quantum Hall states are shown as a function of the spatial direction $y$, intersecting the QPC channel (Fig. \ref{Schema}.A, left column). Empty / filled circles symbolize empty or occupied states. The extended states could for example be associated with Landau levels (in this case $\Delta_\mathrm{ext}=\hbar \omega_c$), spin-split Landau levels ($\Delta_\mathrm{ext}=\tilde{g}\mu_B B$ with an exchange enhanced $\tilde{g}$), or $\Lambda$-levels of composite fermions, corresponding to a FQH state at $\nu$=1/$m$ \cite{macdonald_edge_1990,brey_edge_1994,sim_composite-fermion_1999}($\Delta_\mathrm{ext}=\Delta_{1/m}$ is the energy gap of the FQH state). For simplicity, we will constrain the discussion in the following to the situation, where extended states arise from a Landau level splitting. If spin-split Landau levels or $\Lambda$-levels are considered, an analog picture can be constructed.\\ In Fig. \ref{Schema}.A, energies of the second Landau level are far above the Fermi energy. In the most simple edge state picture \cite{halperin_quantized_1982,buttiker_absence_1988}, Landau level energies are bent up by the confinement potential of the QPC, giving rise to chiral edge states at the intersections with the Fermi energy, thus leading to a step-wise density increase towards the bulk of the sample. However, self-consistency of the Poisson and Schr\"odinger equations at a smooth, electrostatically defined edge \cite{chklovskii_electrostatics_1992} leads to a screened potential and smooth density variations in compressible regions of finite width (\ref{Schema}.A, middle column). In this compressible region, partially filled states (half-filled circles) reside at the Fermi energy. The density of electrons in the lowest Landau level is constrained via $n\le n_0=2 eB/h$, due to its finite degeneracy. Where the Landau level energy lies below the Fermi energy, all states are occupied (filled circles) and this maximum density is reached. Potential fluctuations can no longer be screened, in contrast to the ideally perfect screening in compressible regions where the potential is flat. In this picture, compressible regions in between regions of constant filling factors $\nu_1$ and $\nu_2$ contribute $G=e^2/h\times\Delta\nu$ to the conductance, where $\Delta\nu=\nu_2-\nu_1$ \cite{beenakker_theory_1991}. Alternate models exist, where the current is flowing in the bulk (see for example Ref. \onlinecite{komiyama_edge_1998} for an overview). In our case however, the details of the current distribution in the QPC are not important, as only the total conductance through the QPC can be measured. A schematic spatial density distribution within the QPC is shown in the right column of Fig. \ref{Schema}.A. Here, the boundaries between compressible and incompressible regions are indicated as black arrows. For simplicity, these will be referred to as 'edge states' from now on. In this picture, the edge state is perfectly transmitted through the QPC constriction (between black polygons) and both counter-propagating directions are separated by a wide incompressible region \cite{chklovskii_ballistic_1993} (white), yielding a quantized QPC conductance. Far away from the QPC, additional Landau levels eventually fall below the Fermi energy, leading to additional compressible regions where the density increases towards its bulk value (green (dark grey) area). Adding schematic potential fluctuations (\ref{Schema}.A, left column) does not change the overall situation, as long as no states in the second Landau level become occupied. This is the analog situation in the regions A.1 and A.2 of Fig. \ref{Faecher}.b. As the magnetic field strength is increased, Landau levels are lifted in energy, leading to a narrower incompressible region in the center of the QPC between the edge states. The density is locally reduced (\ref{Schema}.B, middle column) where maxima of the potential fluctuations intersect the Fermi energy (\ref{Schema}.B, left column). This leads to the formation of a compressible region of reduced density (red striped) that is separated from the edge states via incompressible stripes.\\ For an increasing disorder amplitude or decreasing Landau level splitting, compressible regions of enhanced or reduced density can occur in the constriction at the same time, explaining the simultaneous visibility of resonances with a different slope in Fig. \ref{Faecher}.c (indicated by solid green arrows). When disorder dominates over the Landau level splitting, i.e. when the the gradient of the background potential $\partial V/\partial y$ becomes comparable to $E_\mathrm{gap}/l_\mathrm{B}$\cite{beenakker_edge_1990}, where $l_\mathrm{B}$ is the magnetic length, the system is no longer described by a many-electron picture with screening via compressible and incompressible regions (\ref{Schema}.D). In that confinement-dominated case, single-electron states localized around a potential minimum or maximum in the constriction enclose a fixed number of flux quanta \cite{jain_quantum_1988} (Fig. \ref{Schema}D, solid blue line). As the area of the localized state is tuned non-linearly with the QPC gate voltage, resonances with varying slope in the $B-V_\mathrm{QPC}$ plane are expected \cite{ilani_microscopic_2004}. \\ Alternate tunneling paths that lead to a qualitatively similar behavior have been proposed \cite{van_loosdrecht_aharonov-bohm_1988}. Here, non-adiabaticity of the QPC potential leads to enhanced tunneling between the edge channels at the entrance and exit of the constriction (Fig. \ref{Schema}D, red dashed lines). In contrast, the situation of Fig. \ref{Schema}.B is described by Coulomb dominated physics of the compressible region of reduced density inside the constriction. Here, electron-electron interactions lead to a potential with compressible and incompressible regions. The charge of the compressible region of reduced density is quantized, leading to a certain slope in the $B-V_\mathrm{QPC}$ plane \cite{ilani_microscopic_2004}, whenever an electron is added or removed from the compressible region of reduced density. The slope is uniquely determined by the filling factor of the incompressible region in which the compressible region of enhanced or reduced density is formed and equals the slope of the corresponding conductance plateaus in the $V_\mathrm{QPC}$-$B$-field plane. This explains why resonances only occur with one of the slopes of the neighboring conductance plateaus (Fig. \ref{Faecher}.d). Conductance resonances are only visible in the transport data when the incompressible region between the edge states and the compressible region of reduced density is sufficiently small, allowing for resonant backscattering across the constriction. This is the case as the conductance starts to decrease below the plateau value, as in Fig. \ref{Faecher}.b B.1 and B.2. Similarly, potential minima of the second Landau level fall below the Fermi energy, as the magnetic field strength is decreased (\ref{Schema}.C, left column), leading to compressible region of enhanced density within the incompressible region separating the edge states. As additional transmission sets in (Fig. \ref{Faecher}.b, C.1 and C.2), the coupling of these compressible region of enhanced density leads to a periodic modulation of the transmission.\\ In this discussion, the additional complication of possible edge reconstruction of integer quantum Hall (IQH) edge states \cite{venkatachalam_local_2012} has not been taken into account. Furthermore, we observe faint conductance plateaus at $G=2/3 \times e^2/h$ in the QPC. This state is clearly visible in the QPCII.c (Fig. \ref{Faecher}.b) and in QPCII.a when the voltage applied to the CTG is swept (Fig. \ref{Faecher}.a). Surprisingly, the $\nu$=2/3 state is not observed, when the QPC voltage of QPCII.a is swept while a constant voltage is applied to the CTG (Fig. \ref{Faecher}.c). The edge structure of the $\nu=2/3$ state is still not understood in detail. Theory and experimental findings suggest, that this state may consist of a $\delta\nu$=1 IQH edge state and a counter-propagating $\delta\nu$=-1/3 edge state of holes which are equilibrated by interaction, resulting in a single chiral charged mode and a counter-propagating neutral mode \cite{girvin_particle-hole_1984,macdonald_edge_1990,wen_electrodynamical_1990,bid_observation_2010}. Even more advanced theoretical proposals exist \cite{wang_edge_2013-1}, which can explain the experimental findings of these states. How to interpret localizations in the case of such a complicated edge structure remains an open question. The weak visibility of the $\nu$=2/3 state could be due to this complicated edge structure and suggests a smaller energy gap than observed for the $\nu=1/3$ state.\\ As mentioned before, resonances with bending slopes in the $B$-field$-V_\mathrm{QPC}$ plane are expected for single-electron resonances \cite{ilani_microscopic_2004,jain_quantum_1988}. The detailed behavior of the slope depends on the disorder potential intersecting the Fermi energy. This suggests that the resonances in the FQH regime of Fig. \ref{Schmetterling}.a (marked by white dashed line) could be interpreted as single-electron effects. In Ref. \onlinecite{simmons_resistance_1991}, a model for a disorder potential maximum in a constriction is proposed, leading to magnetically bound states which could qualitatively reproduce the bending of the resonances. In this situation, disorder dominates over the smaller FQH gaps and the formation of wide compressible and incompressible regions in the constriction is no longer possible (Fig. \ref{Schema}.D). Thus, the slope in the $B$-field - $V_\mathrm{QPC}$ plane depends on the influence of the QPC voltage on the enclosed area, which depends on the shape of the disorder potential maximum.\\ \subsubsection{B-field and voltage periodicities} Within this framework, we may now investigate the periodicities of the resonances in Fig. \ref{Faecher}.c-\ref{Faecher}.f. For a Coulomb dominated quantum dot, a distinct behavior of the periodicities $\Delta B(\nu_\mathrm{QPC})$ and $\Delta V_\mathrm{QPC}(\nu_\mathrm{QPC})$ is expected. These periodicities depend on the filling factor of the incompressible region, in which the Coulomb dominated region is formed, in our case this is $\nu_\mathrm{QPC}$. From theoretical models \cite{rosenow_influence_2007, halperin_theory_2011} for Coulomb dominated Fabry-P\'erot interferometers it is expected that $\Delta B(\nu_\mathrm{QPC}=1)\approx2 \Delta B(\nu_\mathrm{QPC}=2)\approx\Delta B(\nu_\mathrm{QPC}=1/3)$ and $\Delta V_\mathrm{QPC}(\nu_\mathrm{QPC}=1)\approx\Delta V_\mathrm{QPC}(\nu_\mathrm{QPC}=2)\approx3 \Delta V_\mathrm{QPC}(\nu_\mathrm{QPC}=1/3)$ \footnote{For the voltage periodicity at $\nu_\mathrm{QPC}$=1/3, the gating effect of the background electrons has to be taken into account, as described in Ref. \onlinecite{mcclure_fabry-perot_2012}. In our case however, this only gives a negligible correction from $\Delta V_\mathrm{QPC}(\nu_\mathrm{QPC}=1)\approx3 \Delta V_\mathrm{QPC}(\nu_\mathrm{QPC}=1/3)$.}, which has been observed in lithographically defined quantum dots\cite{mcclure_fabry-perot_2012}. In the IQH regime, our periodicities for $\nu_\mathrm{QPC}=2$ ($\Delta B\approx 30~\mathrm{mT},~\Delta V_\mathrm{QPC}\approx 62~mV$) and $\nu_\mathrm{QPC}=1$ ($\Delta B\approx 55~\mathrm{mT},~\Delta V_\mathrm{QPC}\approx 60~mV$) are in good agreement with these predictions. Periodicities for $\nu_\mathrm{QPC}=1/3$ ($\Delta B\approx 73~\mathrm{mT},~\Delta V_\mathrm{QPC}\approx 24~mV$) are at least compatible with a Coulomb dominated localization of fractional $e/3$ charges. The area which can be extracted from these periodicities ($A\approx0.075~\mu$m$^2$) is compatible with a localization in the channel of the QPC. However, it should be noted that the geometry of the compressible region of enhanced or reduced density within the constriction might change as the $B$-field is varied, because it is not lithographically defined but might change self-consistently. A different behavior is observed in the low-$n$/low-$B$-field end of conductance plateaus ('D' in Fig. \ref{Schema}), where single-electron physics is expected to dominate. In the measurement of Fig. \ref{Faecher}.c (regions encircled by white dashed line), periodicities for $\nu_\mathrm{QPC}=1$ ($\Delta B\approx 200~\mathrm{mT},~\Delta V_\mathrm{QPC}\approx 53~mV$) and $\nu_\mathrm{QPC}=1/3$ ($\Delta B\approx 360~\mathrm{mT},~\Delta V_\mathrm{QPC}\approx 48~mV$) are incompatible with a Coulomb dominated mechanism and indicate single-electron behavior. Similar enhancements of $\Delta B$ for $\nu=1/3$ have been interpreted as magnetically bound states in earlier experiments \cite{simmons_resistance_1989}. However in this interpretation, finite temperature effects or an interplay with Coulomb blockade mechanisms might have to be taken into account \cite{lee_comment_1990}.\\ \subsubsection{Summary} To summarize, the most important findings of this section are: periodic conductance oscillations with a slope, parallel to either of the neighboring conductance plateaus were observed. They were interpreted to originate from the Coulomb dominated charging of compressible region of enhanced or reduced density, formed in a constant filling factor background. This filling factor determines the slope. $B$-field and gate voltage periodicities agree with expectations for a Coulomb dominated Fabry-P\'erot interferometer. At low densities and in weak magnetic fields, disorder prevents the formation of compressible and incompressible regions. Here, resonances are interpreted as single-electron effects, where electronic states are dominated by confinement and encircle a local potential maximum and enclose a certain number of flux quanta. In the fractional quantum Hall regime where energy gaps are smaller than in the integer quantum Hall regime, an influence of both mechanisms can be seen. At the plateau boundaries of the $\nu_\mathrm{QPC}$=1/3 state, conductance oscillations, compatible with Coulomb dominated charging of fractionally charged quasiparticles, are observed. For 1/3 $<$ $\nu_\mathrm{QPC}$ $<$ 1, modulations of the conductance with an intermediate slope (in-between slopes of the $\nu_\mathrm{QPC}$=1 and $\nu_\mathrm{QPC}$=1/3 plateaus) are observed. These slopes move with the local filling factor of the QPC, i.e. correspond to a certain number of flux quanta per electron. This indicates the importance of single electron interference, where resonances are expected to emanate from the $B$ = 0, $n$ = 0 origin of the Landau fan \cite{sohrmann_compressibility_2007}. \subsection{Spatial dependence of QPC resonances} \label{AsymmetrieChap} \begin{figure*} \begin{center} \includegraphics[width=16cm]{FIG-Asymmetrie.pdf} \end{center} \caption{(Color online) \textbf{a-d}: Transconductance (numerical derivative in diagonal direction) of QPCs on a high-density sample (QPC I.a, \textbf{a},\textbf{b}) and a low-density sample (QPC II.d, \textbf{c}, QPC II.a, \textbf{d}) as the voltages of left and right QPC gates $V_\mathrm{l}$ and $V_\mathrm{r}$ are varied. White arrows mark resonances that are believed to be due to single-electron interference. These resonances move with a more complicated dependence as the asymmetry is varied, in contrast to many-electron resonances which bend parallel to the conductance plateaus.} \label{Asymmetrie} \end{figure*} By applying different voltages to the two different QPC gates, it is possible to laterally shift the QPC channel in the lithographically defined constriction (this technique was for example used in references \onlinecite{sarkozy_zero-bias_2009,rossler_transport_2011,komijani_origins_2013}). For QPCs similar to QPC I.a, this shift was found to be of the order of the lithographic QPC width \cite{rossler_transport_2011,schnez_imaging_2011}. Fig. \ref{Asymmetrie} shows the numerical derivative of $G$ in diagonal direction (transconductance $\partial G/\partial V_\mathrm{l\&r}$), as the voltages $V_\mathrm{l}$ and $V_\mathrm{r}$ of the left and right QPC gate are varied. In these measurements, the 2DEG far away from the QPC (bulk) is tuned to a fixed filling factor $\nu_\mathrm{bulk}$ with $R_\mathrm{xx}\approx 0$. Regions of constant conductance and pinch-off show up as black areas, bright regions of increasing conductance bend around the pinch-off region. Fig. \ref{Asymmetrie}.a and \ref{Asymmetrie}.c show the asymmetry-dependence of resonances (the diagonal of Fig. \ref{Asymmetrie}.a is a cut across the resonances of Fig. \ref{Schmetterling}.a indicated by the dashed line) in the low-density low-$B$-field end of the $\nu_\mathrm{QPC}=1$ plateau for three different QPCs (Fig. \ref{Asymmetrie}.a, \ref{Asymmetrie}.b: QPC I.a, Fig. \ref{Asymmetrie}.c: QPC II.d, \ref{Asymmetrie}.d: QPC II.a) on 2DEGs of different density. Resonances believed to originate from single-electron effects (indicated by white arrows), are observed at the low-density end of the $G=1\times e^2/h$ conductance plateau. The resonances show up as two or three parallel lines with a varying slope clearly different from the conductance plateaus' slope and sit deep in the $G=1\times e^2/h$ conductance plateau. Such resonances, occurring mainly in symmetric configurations, have been observed in most of the QPCs in study. Additional modulations of the conductance can be observed between the conductance plateaus. These many-electron resonances bend roughly in the same way as the pinch-off line but vary in intensity, as the asymmetry is varied. \\ Fig. \ref{Asymmetrie}.b and \ref{Asymmetrie}.d show the asymmetry behavior in strong magnetic fields. For a bulk filling factor $\nu_\mathrm{bulk}=1$, conductance plateaus in the QPC at $G=1/3\times e^2/h$ are observed. In Fig. \ref{Asymmetrie}.b, mainly resonances bending with the pinch-off line are observed. In contrast, in Fig. \ref{Asymmetrie}.d, non-regular resonances without any preferred slope are observed.\\ With our model (Fig. \ref{Schema}) we can now try to distinguish the asymmetry-behavior of the two different types of resonances: on the one hand, the confinement dominated resonances (Fig. \ref{Schema}.D) for the situation where compressible and incompressible regions are absent and the system is described by single-electron physics, on the other hand the many-electron resonances (Fig. \ref{Schema}.B, \ref{Schema}.C) where a compressible region of enhanced or reduced density, situated in an incompressible region, is charged. \\ Confinement-dominated single-electron resonances are expected to occur as a result of a localized state at a certain position in the channel, to which both edges couple. As the asymmetry and thus the background potential is varied, single-particle energy levels are shifted in energy, which changes the position in gate voltages of the resonance relative to pinch-off. Thus, single-electron resonances are expected to possess a dependence on gate voltage which is not parallel to the respective conductance plateau as the asymmetry is varied. They should disappear, as soon as the coupling to one of the edges is lost. Here, the gate voltage dependence is influenced by the details of the confinement and disorder potential. The proposed gate voltage dependence of the single-electron resonances (which causes a bending not necessarily parallel to the pinch-off line) and the disappearance of the resonances with increasing asymmetry are indeed observed (Fig. \ref{Asymmetrie}.a, \ref{Asymmetrie}.b, white arrows). A similar behavior might be expected from an Aharonov-Bohm mechanism, where non-adiabaticity of the QPC saddle-point potential leads to enhanced tunneling between the edge channels at the entrance and exit of the constriction and thus defines a QPC-voltage dependent area \cite{van_loosdrecht_aharonov-bohm_1988}.\\ In the ideal model of many-electron resonances, the charge of the compressible island of reduced or enhanced density is quantized and changes when the total density in the constriction is varied (i.e. when moving perpendicular to pinch-off in the $V_\mathrm{l}-V_\mathrm{r}$ plane). When the asymmetry is varied parallel to pinch-off, we expect to change mainly the width of the incompressible regions separating the compressible island of reduced or enhanced density from the edge. Thereby the resonance amplitude which highly depends on the width of the incompressible region \cite{baer_cyclic_2013} is changed. At the same time, the occupation of the compressible region of reduced or enhanced density is expected to be approximately constant, as long as the picture of compressible and incompressible regions does not break down. In this scenario, resonances are thus expected to run parallel to the conductance plateau edges, as observed in the measurements (Figs. \ref{Asymmetrie}.a-c). \\ Because the conductance varies strongly in-between the plateaus, resonances cannot be attributed to individual localizations as it was possible for example in Fig. \ref{Faecher}.c. Thus, in a yet different scenario, conductance oscillations could also originate from single-electron effects, where we only probe localizations that couple to both edges for a given voltage asymmetry. At this asymmetry, they possess a local gate voltage dependence, shifting them parallel to the conductance plateaus. The overall behavior of the resonances could result from averaging the contributions of many single-electron resonances.\\ Summarizing, we may state that the bending resonances of Fig. \ref{Asymmetrie}.a and \ref{Asymmetrie}.c (marked by white arrows) are compatible with a confinement dominated single-electron effect, whereas resonances parallel to the conductance plateaus (Figs. \ref{Asymmetrie}.a-.c) are compatible with a many-electron effect. However, other mechanisms leading to similar observations cannot be excluded. The fact that in Fig. \ref{Asymmetrie}.d no resonances bending with the conductance plateaus are observed may indicate that in Fig. \ref{Asymmetrie}.d transport is dominated by single-electron physics, while many-electron effects dominate in Fig. \ref{Asymmetrie}.b, where the applied magnetic field is much stronger and the disorder potential is smaller due to a higher mobility 2DEG. \subsection{Fragile fractional quantum Hall states in QPCs} \label{Fragile} Fig. \ref{Cold}.a shows the transmission of QPC I.b (light / dark blue) and QPC II.d (red) as a function of applied QPC voltage. The conductance of QPC II.d as a function of $V_\mathrm{QPC}$ (red) shows conductance oscillations on the low-density side of the $\nu_\mathrm{QPC}=1$ plateau. These are those resonances of Fig.\ref{Asymmetrie}.c, which were interpreted as single-electron effects.\\ QPC I.b exhibits conductance plateaus at 2/3, 3/5, 2/5 and $1/3\times e^2/h$ in strong magnetic fields ($B = 13~\mathrm{T}$, $\nu_\mathrm{bulk}=1$). Close to pinch-off, the conductance strongly fluctuates. Unfortunately the observation of $\nu_\mathrm{QPC}=2/3,~3/5,~2/5$ and 1/3 does not allow to draw conclusions about the edge reconstruction of the $\nu_\mathrm{bulk}=1$ edge state. Over the whole QPC voltage range, not only the transmission, but also the channel density and the shape of the QPC confinement potential strongly vary \cite{rossler_transport_2011}. The measurement in dark blue shows the first $V_\mathrm{QPC}$ sweep after the cool-down. When closing the channel for a second time (light blue (gray)), fractional filling factors are still visible, but a more negative gate voltage has to be applied to pinch off the channel. As the QPC is subsequently opened again, a pronounced hysteresis is visible and the more fragile conductance plateaus at $G=2/5\times e^2/h$, $G=3/5\times e^2/h$, $G=1/3\times e^2/h$ and $G=2/3\times e^2/h$ disappear. This behavior can be understood considering the time- and voltage-dependent density in the X-electron screening layers. After the screening layers have been depleted, the density only relaxes with long time constants. The electron density of the 2DEG is inversely proportional to the charge carrier density in the X-electron bands due to capacitive coupling. Thus, depleted screening layers lead to a increased 2DEG electron density at the same QPC voltage, explaining why the QPC conductance is higher for opening the QPC than for closing it. \begin{figure} \begin{center} \includegraphics[width=8cm]{FIG-Cold.pdf} \end{center} \caption{(Color online) \textbf{a}: In strong magnetic fields ($B$ = 13 T), the transmission of QPC I.b close to pinch-off is strongly fluctuating (light / dark blue (gray)). The dashed (red) curve depicts a situation in which a transmitted edge state is weakly backscattered in QPC II.d (see inset). \textbf{b}: Transmission of QPC III.a for 2 $\le$ $\nu_\mathrm{bulk}$ $\le$ 3. In the bulk, $\nu_\mathrm{bulk}=$ 7/3, 8/3 and 5/2 are fully quantized with a strong minimum in $R_{\mathrm{xx}}$ (solid black line) and a plateau in $R_{\mathrm{xy}}$ (dashed (green) line). In addition, pronounced reentrant integer quantum Hall (RIQH) states are observed. The diagonal resistance across the QPC, $R_\mathrm{diag}$ (solid blue (gray) line), shows a plateau at $\nu=5/2$, indicating nearly perfect transmission through the QPC. The density within the constriction is very similar to the bulk density.} \label{Cold} \end{figure} Fig. \ref{Cold}.a, demonstrates that many different fractional filling factors $\nu_\mathrm{QPC}$ can be transmitted by applying an appropriate QPC voltage and keeping the magnetic field fixed. However, relaxation of the barely mobile X-band screening layer electrons makes the observation of the most fragile fractional quantum Hall states difficult. To overcome this limitation, the fact that the X-band screening layers become mobile for temperatures above approximately 1 K can be used \cite{rossler_gating_2010,radu_quasi-particle_2008}. By applying top-gate voltages at higher temperatures, the screening layer density can relax in a steady state and density fluctuations in the constriction are avoided. By this relaxation, additional screening is provided, which is believed to result in a much steeper QPC confinement potential. The density relaxation is extremely slow at dilution refrigerator temperatures. At $T$ $\approx$ 1 K, the density already saturates within minutes. At $T$ $\approx$ 4 K, relaxation to lower screening electron density is nearly instantaneous. To allow a full relaxation of the screening layers, the system is kept at $T$ $\approx$ 4 K for several hours. If in contrast the same gate voltage is applied at mK temperatures, the channel density is strongly reduced due to the vicinity of negatively charged screening electrons. Simultaneously to the slow depletion of the screening layers, the electron density in the channel rises. However, this process leads to strong fluctuations in the conductance which destroy the quantization of fragile fractional quantum Hall states.\\ Fig. \ref{Cold}.b shows the diagonal resistance of the 1.2 $\mu$m wide QPC III.a for 2 $\le$ $\nu_\mathrm{bulk}$ $\le$ 3. Here, -4 V have been applied to the QPC gates at $T$ $\approx$ 4 K. The electron gas below the metallic top-gates is depleted at approx. -3.2 V. At a base temperature of 9 mK (electronic temperature $\approx$ 12 mK), the filling factors 7/3, 8/3 and 5/2 are fully quantized in the bulk, with a strong minimum in $R_{\mathrm{xx}}$ and a plateau in $R_{\mathrm{xy}}$. In addition, pronounced reentrant integer quantum Hall states are observed. The density in the constriction is nearly identical to the bulk density, as seen from the overlap of different filling factors. At a magnetic field of approx. 3.6 T, the plateau in $R_\mathrm{diag}$ shows, that the $\nu=5/2$ state is nearly perfectly transmitted through the QPC, without significant backscattering. Here, the applied QPC voltage of -4 V has been kept fixed while cooling down to the base temperature. The deviation of $R_\mathrm{diag}$ and $R_\mathrm{xy}$ at $B$ $\approx$ 3.4 T originates from a small longitudinal component in $R_\mathrm{diag}$ due to an asymmetry of the sample geometry. The optimized growth and gating procedure allow the definition of a QPC without decreasing the density in the constriction and without destroying the quantization of the $\nu$=5/2 state, which is otherwise not possible. Interference experiments at $\nu$=5/2 \cite{stern_proposed_2006,bonderson_detecting_2006,bonderson_probing_2006,ilan_signatures_2011,ilan_coulomb_2008,willett_measurement_2009,willett_magnetic-field-tuned_2013} require a filling factor $\nu$=5/2 in the center of the employed QD, while edge states are only partially transmitted. Here, the diameter of the QD is constrained to a few $\mu$m (due to the finite quasiparticle coherence length \cite{bishara_interferometric_2009}), thus making the conservation of the bulk density and $\nu$=5/2 quantization on a $\mu$m length-scale crucial. The steep confinement potential of QPCIII.a leads to a decreased width of the compressible regions in the QPC and a wider separating incompressible region, thus reducing backscattering across. The anticipated complex edge structure of the $\nu$ = 5/2 state (which was experimentally found to occur only in QPCs of rather large width \cite{miller_fractional_2007}) might facilitate its formation in a steeper confinement potential. Furthermore, the additional screening of the disorder in the constriction via X-band electrons reduces the amplitude of the disorder potential fluctuations. Hence, the influence of conductance oscillations as discussed in section \ref{main} is expected to be reduced. The main drawback of the utilized gating method is the low tuneability of gate voltages at mK temperatures. Here, the gate voltages have to remain in very small range around the voltage that has been applied at $T$=4 K. Otherwise, slow relaxation processes of the X-band screening layers destroy the quantization of the $\nu$=5/2 state. Growth methods which utilize conventional DX-doping and a reduced Al molar fraction might help to overcome this problem, while still providing a good quantization of $\nu$=5/2 \cite{reichl_increasing_2014,pan_impact_2011,gamez_5/2_2013}. \\ Having demonstrated that we can confine a fully gapped $\nu$=5/2 state to a QPC, we are at a good starting point for conducting tunneling and interference experiments with the fragile fractional quantum Hall states at $\nu$=7/3 and 5/2. \section{Conclusion} In conclusion, we have investigated the interplay of electronic transport and localization in quantum point contacts of different geometries and based on 2DEGs utilizing different growth techniques. In these systems, various integer and fractional quantum Hall states were observed. Using a QPC with a top-gate, we were able to investigate conductance resonances in greater detail. In this sample, edge states are separated by a wide incompressible region thus leading to a significant influence of localizations due to disorder potential fluctuations. Regions of perfect QPC transmission are surrounded by periodic conductance oscillations with an identical slope in the $V_\mathrm{QPC}-B$-field plane. Within a many-electron picture, the resonances on the high (low) density end of the plateau can be interpreted as regions of enhanced or reduced density formed within incompressible regions between the counter-propagating edge states. As the charge of these regions is conserved, changing the density or magnetic field leads to periodic conductance oscillations, whenever an electron is added or removed. $B$-field and $V_\mathrm{QPC}$-periodicities agree with expectations for a Coulomb dominated quantum dot in strong magnetic fields and are determined by the filling factor background in which the compressible region of enhanced or reduced density is formed. At low densities and in weaker magnetic fields, resonances within the conductance plateau occur. In this regime, disorder broadening becomes comparable to the Landau level separation, thus compressible regions of reduced and enhanced density, situated in different Landau levels modulate transport at the same time. Here, the many-electron picture is not valid anymore and resonances with a dependence in the $B-V_\mathrm{QPC}$ plane, not necessarily equal to the conductance plateaus' dependence, are observed. These resonances are interpreted as confinement dominated single-electron interference effects. In the fractional quantum Hall regime, the behavior of the system seems to be influenced by both, single- and many-electron physics. Due to the much smaller gaps of the FQH states, disorder becomes more important. Close to perfect transmission, resonances similar to those associated with compressible regions of reduced or enhanced density in a many-electron picture can be observed. Periodicities at $\nu_\mathrm{QPC}=1/3$ are compatible with the localization of fractionally charged quasiparticles in a Coulomb dominated quantum dot. However, for intermediate transmissions, weak resonances with a slope different from the slopes of the neighboring conductance plateaus are observed, indicating the importance of single-electron physics where the formation of compressible and incompressible regions breaks down. Single-electron resonances have been studied as a function of the position of the conducting channel in the constriction. In contrast to many-electron resonances, single-electron resonances are expected to possess slopes in the gate-voltage plane, not necessarily parallel to the conductance plateaus. Here, the slope depends on the details of the disorder potential. Using optimized growth techniques and gating procedures, we are able to form QPC constrictions with extremely weak backscattering and a density equal to the bulk density. This allows us to observe the $\nu=5/2$ state in the QPC with a fully developed plateau. The bulk properties, like the reentrant integer quantum Hall states, are fully conserved in the QPC, making this system promising for future tunneling and interference experiments at $\nu$=5/2. \section{Acknowledgments} We gratefully acknowledge discussions with Bernd Rosenow, Yigal Meir, Yuval Gefen, Ferdinand Kuemmeth and Hiske Overweg. We acknowledge the support of the ETH FIRST laboratory and financial support of the Swiss Science Foundation (Schweizerischer Nationalfonds, NCCR 'Quantum Science and Technology').
\section{Introduction}\label{int} Current cosmological observations, such as the cosmic microwave background (CMB) measurements of temperature anisotropies and polarization at high redshift $z\sim 1090$ and the redshift-distance measurements of supernovae (SNIa) at $z < 2$, have demonstrated that the universe is now undergoing an accelerated phase of expansion. The nature of dark energy, the mysterious power to drive the expansion, is among the biggest problems in modern physics and has been studied widely. The simplest candidate of dark energy is the cosmological constant, whose equation of state (EoS) $w$ always remains $-1$. Although this model is compatible with the current observational data \cite{planck_fit}, it suffers from the well-known fine-tuning and coincidence problems \cite{ccproblem1,ccproblem2,ccproblem3}. In order to lift these severe problems, many alternative dynamical dark energy models, such as quintessence \cite{quint1,quint2,quint3,quint4}, phantom \cite{phantom}, k-essence \cite{kessence1,kessence2}, and quintom \cite{quintom1,quintom2,cai}, have been proposed. Interestingly, the dynamical dark energy component is naturally expected to interact with the other components, such as the cold dark matter \cite{copeland98,amendola00} or massive neutrinos \cite{gu03,fardon04}, in the field theory framework. If these interactions really exist, it would open up the possibility of detecting the dark energy non-gravitationally. In the coupled dark energy models, the quintessence scalar field could nontrivially couple to the cold dark matter component. The presence of the interaction clearly modifies the cosmological background evolutions. The evolution of cold dark matter energy density is dependent on the quintessence scalar field \cite{cde_xia}. The energy density can be transferred between the cold dark matter and the dark energy. On the other hand, the interaction between dark sectors will also affect the evolution of cosmological perturbations (see ref. \cite{cde_xia}, and references therein). The non-zero coupling could shift the matter-radiation equality scale factor, and affect the locations and amplitudes of acoustic peaks of CMB temperature anisotropies and the turnover scales of large scale structure (LSS) matter power spectrum. Furthermore, the coupling affects the dynamics of the gravitational potential, and also affect the late integrated Sachs-Wolfe (ISW) effect \cite{isw}. Therefore, it is of great interest to investigate the non-minimally coupled dark energy models from the current observational data, such as the CMB measurements \cite{Amendola:2002bs,Amendola:2003eq,Bean:2008ac,LaVacca:2009yp,He:2010im,Pettorino:2012ts,Salvatelli2013,Pettorino:2013oxa}, the LSS clustering \cite{Baldi:2010ks,Baldi:2010td}, the weak lensing \cite{DeBernardis:2011iw,Amendola:2011ie}, the cross-correlation between CMB and LSS \cite{cde_xia,Mainini:2010ng,Bertacca:2011in}, and the low redshift observations \cite{Honorez:2010rr}. Since the Planck collaboration has released the first cosmological papers providing the high resolution, full sky, CMB maps \cite{planck_map}, it is important to study the coupled dark energy models and revisit the constraint on parameters from the latest cosmological probes. In this paper we investigate this kind of model and present the tight constraints from the latest Planck and WMAP9 data, the baryon acoustic oscillations (BAO) measurements from several large scale structure (LSS) surveys, the ``Union2.1'' compilation which includes 580 supernovae, and the CMB lensing data from the Planck measurement. More interestingly, we find a non-zero value for the interaction parameter from the ``SNLS'' supernovae sample and the direct measurement on the Hubble constant from the Hubble Space Telescope (HST). The structure of the paper is as follows: in Sec.II we show the basic equations of background evolution and linear perturbations of the coupled dark energy model. In Sec.III we present the current observational datasets we used. Sec.IV contains our main global fitting results from the current observations, while Sec.V is dedicated to the summary. \section{Coupled Dark Energy Model}\label{theory} In this section we briefly review the basic equations for the coupled dark energy model. We refer the reader to refs. \cite{cde_xia,Pettorino:2012ts,Pourtsidou:2013nha} for a very detailed description of all equations involved and effects on the CMB and LSS measurements. We assume a flat universe described by the Friedmann-Robertson-Walker metric and the adiabatic initial conditions for all components in our analyses. When including the interactions, the conservation of energy momentum for each component becomes \cite{brookfield08}: \begin{equation} T^{\mu}_{\nu;\mu}=\beta\phi_{,\nu}T^{\alpha}_{\alpha}~,\label{tmunu} \end{equation} where $T^{\mu}_{\nu}$ is the energy-momentum tensors and $\phi$ is the quintessence scalar field. In our analysis we only consider the interaction between cold dark matter and dark energy, the energy conservation equations of cold dark matter and dark energy will be violated: \begin{eqnarray} \dot{\rho}^{}_{\rm c}+3\mathcal{H}{\rho}^{}_{\rm c}&=&Q~,\\ \dot{\rho}^{}_{\phi}+3\mathcal{H}{\rho}^{}_{\phi}(1+w^{}_{\phi})&=&-Q~,\label{decoveq} \end{eqnarray} where $\rho_{\rm c}$ and $\rho_\phi$ denote the energy densities of cold dark matter and dark energy, $\mathcal{H}=\dot{a}/a$ (the dot refers to the derivative with respect to the conformal time $\eta$) and $Q$ is the interaction energy exchange. Here, we consider the exponential form $\rho^{}_{\rm c}(\phi)=\rho^{\ast}_{\rm c}e^{\beta\phi}$ as the interaction form between quintessence and cold dark matter, where $\rho^{\ast}_{\rm c}$ is the bare energy density of cold dark matter and $\beta$ is the strength of interaction. Then, the energy exchange $Q$ can be written as \cite{cde_xia,Pourtsidou:2013nha}: \begin{equation} Q = \beta\dot{\phi}\rho_{\rm c}~. \end{equation} When $Q<0$ (or $Q>0$), the energy of cold dark matter (or dark energy) transfers to dark energy (or cold dark matter). Equivalently, the quintessence scalar field $\phi$ evolves according to the Klein-Gordon equation: \begin{equation} \ddot{\phi}+2\mathcal{H}\dot{\phi}+a^2V'(\phi)=-a^2\beta\rho^{}_{\rm c}~\label{quinKGeq}. \end{equation} In this paper we choose the typical exponential form as the quintessence potential: $V(\phi)=V_{0}e^{-\lambda\phi}$, and the prime denotes the derivative with respect to the quintessence scalar field $\phi\,$: $V'(\phi)\equiv\partial V(\phi)/\partial\phi\,$. We modify these background equations of the coupled dark energy model in the CAMB code \cite{camb} to calculate the cosmological distance information for the SNIa and BAO measurements. In the synchronous gauge, we could calculate the evolution equations for the perturbation of cold dark matter in the linear regime \cite{cde_xia}: \begin{eqnarray} \dot{\delta}_{\rm c}&=&-\theta_{\rm c}-\dot{h}/2+\beta\delta\dot{\phi}~,\\ \dot{\theta}_{\rm c}&=&-\mathcal{H}\theta_{\rm c}-\beta\dot{\phi}\theta_{\rm c}+k^2\beta\delta{\phi}~, \end{eqnarray} where $\delta_{\rm c}$ and $\theta_{\rm c}$ are the density perturbation and the gradient of the velocity of the cold dark matter, $\delta\phi$ is the perturbation of dark energy scalar field, and $h$ is the usual synchronous gauge metric perturbation. In the presence of interaction, $\theta_{\rm c}$ will evolve to be nonzero, even if its initial value is zero. Therefore, we compute the perturbation equations in an arbitrary synchronous gauge, instead of the cold dark matter rest frame \cite{cde_xia,Bean:2008ac}. On the other hand, we get the perturbed Klein-Gordon equation in the non-minimally coupled system: \begin{equation} \delta\ddot{\phi}+2\mathcal{H}\delta\dot{\phi}+k^2\delta\phi+a^2V''\delta\phi+\dot{h}\dot{\phi}/2=-a^2\beta\rho_{\rm c}\delta_{\rm c}~. \end{equation} We include these evolution equations for the perturbations of the coupled dark energy model in the CAMB code to compute the theoretical prediction of the CMB temperature power spectrum. \begin{figure}[t] \centering \includegraphics[scale=0.5]{figure/cl.eps} \caption{The CMB temperature anisotropies for two different models: $\lambda=1.22$, $\beta=0$ (black solid lines) and $\lambda=1.22$, $\beta=0.15$ (red dashed lines).}\label{cmb_eff} \end{figure} Here, we use the best fit model of Planck data \cite{planck_like}: $\Omega_bh^2 = 0.02203$, $\Omega_ch^2 = 0.1204$, $\tau = 0.0925$, $h = 0.6704$, $n_s = 0.9619$ and $A_s = 2.215 \times 10^{-9}$ at $k = 0.05$ ${\rm Mpc}^{-1}$. In figure \ref{cmb_eff} we plot the CMB temperature power spectra for two different models: the uncoupled model $\lambda=1.22$, $\beta=0$ (black solid lines) and the coupled system $\lambda=1.22$, $\beta=0.15$ (red dashed lines). We can see that the amplitude on the small scales CMB temperature power spectrum decreases with increasing coupling, due to the earlier epoch of matter-radiation equality $a_{\rm eq}$. And the coupling also shifts locations of the CMB peaks towards smaller scales and suppresses the CMB temperature anisotropies on very large scales (the late-time ISW effect) \cite{cde_xia}. \section{Data}\label{data} In our analyses, we consider the following cosmological probes: i) CMB power spectra; ii) the BAO signal in the galaxy power spectra; iii) direct measurement of the current Hubble constant; iv) luminosity distances of type Ia supernovae. For the Planck data from the 1-year data release \cite{planck_fit}, we use the low-$\ell$ and high-$\ell$ CMB temperature power spectrum data from Planck with the low-$\ell$ WMAP9 polarization data (Planck+WP). We marginalize over the nuisance parameters that model the unresolved foregrounds with wide priors \cite{planck_like}. We also consider the CMB lensing data obtained from Planck \cite{planck_lens} seperately. For comparison, we also use the WMAP9 CMB temperature and polarization power spectra \cite{wmap9} in our calculations. BAO provides an efficient method for measuring the expansion history by using features in the clustering of galaxies within large scale surveys as a ruler with which to measure the distance-redshift relation. Since the current BAO data are not accurate enough, one can only determine an effective distance \cite{baosdss}: \begin{equation} D_V(z)=[(1+z)^2D_A^2(z)cz/H(z)]^{1/3}~. \end{equation} Following the Planck analysis \cite{planck_fit}, in this paper we use the BAO measurement from the 6dF Galaxy Redshift Survey (6dFGRS) at a low redshift ($r_s/D_V (z = 0.106) = 0.336\pm0.015$) \cite{6dfgrs}, and the measurement of the BAO scale based on a re-analysis of the Luminous Red Galaxies (LRG) sample from Sloan Digital Sky Survey (SDSS) Data Release 7 at the median redshift ($r_s/D_V (z = 0.35) = 0.1126\pm0.0022$) \cite{sdssdr7}, and the BAO signal from BOSS CMASS DR9 data at ($r_s/D_V (z = 0.57) = 0.0732\pm0.0012$) \cite{sdssdr9}. We also add a gaussian prior on the current Hubble constant given by ref. \cite{hst_riess}; $H_0 = 73.8 \pm 2.4$ ${\rm km\,s^{-1}\,Mpc^{-1}}$ (68\% C.L.). The quoted error includes both statistical and systematic errors. This measurement of $H_0$ is obtained from the magnitude-redshift relation of 240 low-z Type Ia supernovae at $z < 0.1$ by the Near Infrared Camera and Multi-Object Spectrometer Camera 2 of HST. Finally, we include data from Type Ia supernovae, which consists of luminosity distance measurements as a function of redshift, $D_L(z)$. In this paper we consider two SNIa samples: the ``Union2.1'' compilation with 580 samples \cite{union2} and the ``SNLS'' compilation with 473 supernovae reprocessed by ref. \cite{snls}. However, we do not combine them together in the numerical analysis, since we find that these two SNIa samples give quite different constraints on the coupled dark energy model. When calculating the likelihood, we marginalize the nuisance parameters, like the absolute magnitude $M$ and the parameters $\alpha$ and $\beta$, as explained by Ref. \cite{planck_fit}. \section{Numerical Results}\label{results} We perform a global fitting of cosmological parameters using the CosmoMC package \cite{cosmomc}, a Markov Chain Monte Carlo (MCMC) code. We assume purely adiabatic initial conditions and neglect the primordial tensor fluctuations. The basic six cosmological parameters are allowed to vary with top-hat priors: the cold dark matter energy density parameter $\Omega_c h^2 \in [0.01, 0.99]$, the baryon energy density parameter $\Omega_b h^2 \in [0.005, 0.1]$, the scalar spectral index $n_s \in [0.5, 1.5]$, the primordial amplitude $\ln[10^{10}A_s] \in [2.7, 4.0]$, the ratio (multiplied by 100) of the sound horizon at decoupling to the angular diameter distance to the last scattering surface $100\Theta_s \in [0.5, 10]$, and the optical depth to re-ionization $\tau \in [0.01, 0.8]$. The pivot scale is set at $k_{s0} = 0.05$ ${\rm Mpc}^{-1}$. There are two more parameters: $\lambda$ and $\beta$ in the potential and coupling forms of the coupled dark energy model. In addition, CosmoMC imposes a weak prior on the Hubble parameter: $h \in [0.4,1.0]$. \begin{table} \caption{The median values and $1\,\sigma$ error bars on some cosmological parameters obtained from different data combinations in the coupled dark energy model. For the interaction strength $\beta$, we quote the $95\%$ upper limits instead.}\label{table1} \begin{center} \begin{tabular}{l|c|c|c|c|c|c} \hline\hline &WMAP9&Planck+WP&Planck+WP+BAO&Planck+WP+Union&Planck+WP+Lens&Normal Data\\ \hline $\Omega_bh^2$&$0.02282\pm0.00054$&$0.02196\pm0.00029$&$0.02199\pm0.00026$&$0.02200\pm0.00028$&$0.02210\pm0.00027$&$0.02210\pm0.00025$\\ $\Omega_ch^2$&$0.1066\pm0.0080$&$0.1170\pm0.0044$&$0.1182\pm0.0023$&$0.1159\pm0.0043$&$0.1160\pm0.0037$&$0.1175\pm0.0018$\\ $\Omega_m$&$0.2599\pm0.0616$&$0.3093\pm0.0470$&$0.3097\pm0.0189$&$0.2823\pm0.0336$&$0.3030\pm0.0425$&$0.2997\pm0.0126$\\ $\sigma_8$&$0.8331\pm0.0772$&$0.8410\pm0.0480$&$0.8373\pm0.0262$&$0.8648\pm0.0393$&$0.8337\pm0.0435$&$0.8353\pm0.0172$\\ $H_0$&$72.04\pm7.99$&$67.59\pm4.68$&$67.35\pm1.82$&$70.21\pm3.47$&$67.97\pm4.25$&$68.29\pm1.19$\\ $\beta$&$<0.1446$&$<0.1021$&$<0.0636$&$<0.1002$&$<0.0943$&$<0.0522$\\ \hline\hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \centering \includegraphics[scale=0.45]{figure/cmb_alone.eps} \caption{Marginalized two-dimensional likelihood (1, $2\,\sigma$ contours) constraints on the parameters $\beta$ and $\sigma_8$ in the coupled dark energy model from the Planck+WP (red) and WMAP9 (blue) data, respectively.}\label{cmb_alone} \end{figure} \subsection{Tight Constraints on $\beta$} Firstly, we consider the constraints on the coupled dark energy model from the Planck+WP and WMAP9 data alone. In table \ref{table1} we list the constraints on some cosmological parameters in the coupled dark energy model from different data combinations. As we know, the CMB anisotropies mainly contain the information about the high-redshift universe, but it is not directly sensitive to phenomena which affect the lower redshift Universe, such as the nature of dark energy. Thus, the WMAP9 data alone can not constrain the parameter $\beta$ of the coupled dark energy model very well. We only obtain the upper limit on the strength of interaction, namely the $95\%$ C.L. constraint is $\beta < 0.145$, which is consistent with previous works \cite{cde_xia,Bean:2008ac}. When we use the more accurate Planck+WP data, the constraint becomes tighter \begin{equation} \beta<0.102~~(95\%~{\rm C.L.})~. \end{equation} As we mentioned before, the non-minimal coupling shifts the acoustic peaks of CMB temperature anisotropies on the small scales. The small-scale CMB measurements should improve the constraints on the coupling strength. However, since the new Planck data have measured the small-scale CMB power spectrum with very high precision, adding other small-scale CMB data, like Atacama Cosmology Telescope (ACT) \cite{act} and South Pole Telescope (SPT) \cite{spt}, the constraint on $\beta$ is only slightly improved \cite{cde_xia,Pettorino:2013oxa}. In order to save some CPU time for running MCMC, in this work we do not include the ACT and SPT data, since they include many nuisance parameters which significantly slow down our calculations. In figure \ref{cmb_alone}, we show the two-dimensional contours in the ($\beta$,$\sigma_8$) panel. We can see that $\beta$ is correlated with the $\sigma_8$. In the coupled dark energy model, a positive coupling between cold dark matter and dark energy leads to a high value of $\sigma_8$. That is because we have an energy transfer from cold dark matter to dark energy, which means that there is more dark matter at early times. In this case, the epoch of matter-radiation equality occurs earlier in the coupled model relative to the uncoupled system. Only the very small scale modes could enter the horizon and grow during the radiation dominated era. The growth of perturbations in the coupled model under consideration is enhanced, small-scale power is increased and the value of $\sigma_8$ is larger \cite{cde_xia,Pourtsidou:2013nha}. Using the Planck+WP data alone, we obtain the limit on $\sigma_8$ today of $\sigma_8=0.841\pm0.048$ ($68\%$ C.L.), which is obviously higher than one obtained in the standard $\Lambda$CDM model: $\sigma_8=0.829\pm0.013$ ($68\%$ C.L.) \cite{planck_fit}. In the standard flat $\Lambda$CDM framework, the constraint on the Hubble constant $H_0$ is significantly improved by the new Planck data, $H_0=67.4\pm1.4$ ${\rm km\,s^{-1}\,Mpc^{-1}}$ at $68\%$ confidence level \cite{planck_fit}. However, this result is obviously in tension with that measured by various lower-redshift methods, such as the direct $H_0$ probe from HST \cite{hst_riess}. The Planck team argued that the local measurements are more likely to be affected by some unknown systematics, such as the effect of a local underdensity, which might lead to this tension \cite{Amendola2013,Verde2013}. Recently, ref. \cite{xia2013} found that if the unknown systematic is not an issue, this $H_0$ tension actually implies that the Planck data favor the dynamical dark energy model, especially one with the EoS $w<-1$. When the model with $w<-1$ is allowed in the analysis, the constraint on $H_0$ from Planck+WP data is consistent with the HST $H_0$ prior. However, this $H_0$ tension cannot be eased in the coupled dark energy model. Planck+WP data yield the $68\%$ C.L. constraint on the Hubble constant of $H_0=67.6\pm4.7$ ${\rm km\,s^{-1}\,Mpc^{-1}}$, which is still significantly lower than the HST measurement. This is because we use the quintensense dark energy model to couple with the cold dark matter. The effective equation of state of dark energy sector cannot be smaller than $-1$ \cite{cde_xia}, therefore, the tension between constraints on $H_0$ still exists. Different from the quintessence scalar field, there are some coupling models where the dark energy component is modeled as a fluid with constant equation of state parameter $w$. It would be very interesting whether the $H_0$ tension can be relaxed when $w$ allowed to be smaller than $-1$ in these models \cite{xia_new}. \begin{figure}[t] \centering \includegraphics[scale=0.35]{figure/beta_1d.eps} \caption{One-dimensional posterior distributions of the strength of interaction $\beta$ from various data combinations: Planck+WP+BAO (red dashed line), Planck+WP+Union (blue dotted line), Planck+WP+Lens (magenta dash-dotted line), all ``Normal Data'' together (black solid line).}\label{beta_1d} \end{figure} Since the CMB data alone cannot very well constrain the strength of coupling $\beta$ in the coupled dark energy model, we need to add some extra information from the low-redshift probes to break the degeneracy. In our calculations, we consider three kinds of measurements: the BAO signal, the ``Union2.1'' compilation of SNIa and the CMB lensing data from Planck measurement. We also combine these datasets together with the Planck+WP data and call this combined dataset the ``Normal'' data. In figure \ref{beta_1d} we show the one-dimensional posterior distributions of $\beta$ from different data combinations. The BAO measurements give the tight constraint on the matter density $\Omega_m$. When using Planck+WP and BAO data, the constraint on the strength of coupling is significantly improved, namely $\beta<0.064$ at $95\%$ confidence level. However, the constraining powers of ``Union2.1'' compilation of SNIa and the CMB lensing data are not strong enough for the coupled dark energy model. Planck+WP+Union and Planck+WP+Lens data yield the $95\%$ upper limits on the strength of $\beta<0.100$ and $\beta<0.094$, respectively. When we combine these low redshift probes together, the ``Normal'' data give very stringent upper limit on the strength of interaction between cold dark matter and dark energy: \begin{equation} \beta<0.052~~(95\%~{\rm C.L.})~. \end{equation} The minimally coupled system is still consistent with these observational datasets. We do not find the evidence for the non-minimally coupled dark energy model for these data combinations. \subsection{Non-zero Coupling} \begin{table} \caption{The median values and $1\,\sigma$ error bars on some cosmological parameters obtained from the Planck+WP, HST and SNLS data in the coupled dark energy model.}\label{table2} \begin{center} \begin{tabular}{l|c|c|c|c} \hline\hline &Planck+WP&Planck+WP+HST&Planck+WP+SNLS&Tension Data\\ \hline $\Omega_bh^2$&$0.02196\pm0.00029$&$0.02202\pm0.00028$&$0.02200\pm0.00028$&$0.02203\pm0.00028$\\ $\Omega_ch^2$&$0.1170\pm0.0044$&$0.1125\pm0.0037$&$0.1117\pm0.0049$&$0.1112\pm0.0033$\\ $\Omega_m$&$0.3093\pm0.0470$&$0.2533\pm0.0270$&$0.2453\pm0.0371$&$0.2398\pm0.0229$\\ $\sigma_8$&$0.8410\pm0.0480$&$0.8919\pm0.385$&$0.9031\pm0.0523$&$0.9055\pm0.0385$\\ $H_0$&$67.59\pm4.68$&$73.13\pm3.07$&$74.38\pm4.56$&$74.75\pm2.75$\\ $\beta$&$<0.1021$&$0.0728\pm0.0265$&$0.0735\pm0.0311$&$0.0782\pm0.0217$\\ \hline\hline \end{tabular} \end{center} \end{table} \begin{figure}[t] \centering \includegraphics[scale=0.3]{figure/beta.eps} \includegraphics[scale=0.3]{figure/h0.eps} \caption{One-dimensional posterior distributions of the strength of interaction $\beta$ and the Hubble constant $H_0$ from various data combinations: Planck+WP (red dotted line), Planck+WP+HST (blue dashed line), Planck+WP+SNLS (magenta dash-dotted line), all ``Tension Data'' together (black solid line). The vertical blue dashed line in the right panel denotes the HST prior on the Hubble constant.}\label{tension_1d} \end{figure} Besides the ``Normal'' data mentioned above, we also have some other low redshift measurements, like the direct probe on $H_0$ from HST and the ``SNLS'' compilation of SNIa. Interestingly, these two datasets strongly favor a high value of the Hubble constant, which is significantly different from that obtained from the ``Normal'' data \cite{xia2013}. Therefore, we combine these two datasets together to constrain the coupled dark energy model and call it the ``Tension'' data. In table \ref{table2} we list the constraints on some parameters from the ``Tension'' data combinations. The interaction between the quintessence scalar field and the cold dark matter significantly modifies the evolutions of their energy density \cite{cde_xia}. For a positive coupling, the energy of cold dark matter transfers to dark energy. Consequently, the effective EoS of dark energy becomes smaller than that in the uncoupled system. The current Hubble constant must be increased correspondingly in order to produce the same expansion rate. Therefore, the strength of coupling $\beta$ is correlated with the Hubble constant $H_0$. Including the measurements on $H_0$ could be helpful to break the degeneracy. So we first add the HST gaussian prior on $H_0$ in the calculations. Due to the large discrepancy on $H_0$ between Planck+WP and HST prior, adding the HST prior to the Planck+WP data forces the obtained median value of $H_0$ towards to the higher one, namely $H_0=73.1\pm3.1$ ${\rm km\,s^{-1}\,Mpc^{-1}}$ ($68\%$ C.L.). As a consequence, the degeneracy between $\beta$ and $H_0$ induces a likelihood peak for the strength of coupling: \begin{equation} \beta = 0.0728\pm0.0265~~(68\%~{\rm C.L.})~, \end{equation} which apparently departs from the minimally coupled system, which is shown in figure \ref{tension_1d} (blue dashed lines). We also include the ``SNLS'' supernovae sample into the calculations and find the similar conclusion with that obtained by adding the HST prior. Planck+WP+SNLS data favor a high value of Hubble constant $H_0=74.4\pm4.6$ ${\rm km\,s^{-1}\,Mpc^{-1}}$ ($68\%$ C.L.) and a non-zero coupling parameter $\beta=0.0735\pm0.0311$ ($68\%$ C.L.). Finally, we combine the SNLS and the HST prior together and the situation becomes worse, shown in figure \ref{tension_1d} (black solid lines). The ``Tension'' data combination yields the tight constraints on the Hubble constant $H_0=74.7\pm2.8$ ${\rm km\,s^{-1}\,Mpc^{-1}}$ ($68\%$ C.L.) and the strength of coupling parameter: \begin{equation} \beta = 0.0782\pm0.0217~~(68\%~{\rm C.L.})~. \end{equation} The preference for a non-zero coupling increases. In figure \ref{tension_2d} we show the two-dimensional contours in the ($\beta$,$H_0$) panel from different data combinations, which is clearly shown that $\beta$ and $H_0$ are correlated. When comparing with the contour obtained from Planck+WP data, the ``Normal'' data give consistent constraint on the Hubble constant. Therefore, this data combination only reduces the correlation between $\beta$ and $H_0$ to give better constraint on $\beta$. However, the ``Tension'' data favor a high value of the Hubble constant. Using this data combination shifts the two-dimensional contour towards high values of $H_0$. Consequently, a non-zero coupling parameter is slightly favored by this data. But it is not strong enough to claim a deviation from the standard $\Lambda$CDM model. \begin{figure}[t] \centering \includegraphics[scale=0.45]{figure/new_tension.eps} \caption{Marginalized two-dimensional likelihood (1, $2\,\sigma$ contours) constraints on the parameters $\beta$ and $H_0$ in the coupled dark energy model from Planck+WP alone (green), ``Normal'' data (red) and ``Tension'' data (blue), respectively.}\label{tension_2d} \end{figure} \section{Summary}\label{sum} In this paper we have presented the latest cosmological constraints on the coupled dark energy models, in which the quintessence scalar field non-minimally couples to the cold dark matter, from the recent Planck measurements. Since the CMB anisotropies mainly contain the information about the high-redshift universe, Planck+WP data alone cannot constrain the strength of coupling strongly, namely the $95\%$ C.L. upper limit is $\beta<0.102$. When we combine the Planck+WP with the BAO, SNIa ``Union2.1'' compilation and the CMB lensing data from Planck, this ``Normal'' data yield a very tight constraint on the strength, $\beta<0.052$ at $95\%$ confidence level. Different from the dark energy model with a constant equation of state parameter $w$, the tension between constraints on $H_0$ from the Planck+WP and the HST $H_0$ prior cannot be eased in the coupled dark energy models, since in these models the effective EoS of dark energy is always larger than $-1$. Finally, we use the HST prior and the SNIa ``SNLS'' sample to break the degeneracy between $\beta$ and $H_0$. Since these two data strongly favor a high value of the Hubble constant, we find an interesting preference for the non-zero coupling: $\beta = 0.0782\pm0.0217$ ($68\%$ C.L.) from the ``Tension'' data combination. This result mainly comes from a tension between constraints on the Hubble constant from the Planck measurement and the local direct $H_0$ probes and needs to be clarified further. \section*{Acknowledgements} We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA). Support for LAMBDA is provided by the NASA Office of Space Science. JX is supported by the National Youth Thousand Talents Program and the grants No. Y25155E0U1 and No. Y3291740S3.
\section{Introduction} Many physical systems exhibit symmetries. A number of techniques have been developed during the last two centuries to take advantage of the conservation laws that are usually associated to these invariance properties to simplify or {\it reduce} those systems in order to make easier the computation of their solutions. The presence of symmetries also creates natural dynamical features that generalize distinguished solutions of their non-symmetric counterparts like the so called {\it relative equilibria} or {\it relative periodic orbits}; relative equilibria are solutions of a symmetric system that coincide with one-parameter group orbits of the action that leaves that system invariant. The justification of this denomination lies in the fact that relative equilibria are equilibria for the reduced Hamiltonian system~\cite{mwr} constructed with the momentum map associated to the action, provided that this object exists. Regarding the stability of these solutions, the degeneracies caused by the presence of symmetries in a system cause drift phenomena that make non-evident the selection of a stability definition. A very reasonable choice is the concept of stability relative to a subgroup introduced in~\cite{smodg, thesis:patrick} for which a number of energy-momentum based sufficient conditions have been formulated in the literature under different assumptions and levels of generality on the group actions involved and the momentum values at which the relative equilibrium in question takes place~\cite{smodg, drift, thesis:patrick, mrs, mre, persistence:periodic, molecules, Ortega1999a, singreleq, po, pos, thesis, Roberts2002, patrick:roberts:wulff:2002, Ortega2004, MontaldiOlmos2011}. These methods have been used, for example, in the study of the stability of relative equilibria present in different configurations of rigid bodies~\cite{lrsm, llt, thesis:patrick}, Riemann ellipsoids~\cite{Fasso2001, Rodriguez-Olmos2008}, underwater vehicles~\cite{leonard:1997, leonard:marsden:1997}, vortices~\cite{Pekarsky1998, laurent:nonlinearity, vortices:lim, Laurent-Polz2011}, and molecules~\cite{molecules}. {\it In this paper we use these methods to establish sufficient conditions for the stability of various branches of relative equilibria present in a mechanical system consisting of a small axisymmetric magnetic body-dipole in an also axisymmetric external magnetic field that additionally exhibits a mirror symmetry}. When the external field is created by two magnetic poles modeled by two distant ``charges"~\cite{Smythe1939} we call this system the {\bfseries\itshape standard orbitron}; the setup involving arbitrary external fields exhibiting the above mentioned symmetries will be referred to as the {\bfseries\itshape generalized orbitron}. The generic term {\bfseries\itshape orbitron} will refer simultaneously to both the standard and the generalized orbitrons. This problem has been studied for a long time already: the model was introduced in the 1970s, and the first theoretical and experimental results were presented in~\cite{Kozorez1974, Kozorez1981}; numerical simulations were carried out later on in~\cite{sszub08, Grygoryeva2009, sszub10}. Unfortunately, these works do not contain a complete mathematical proof of nonlinear stability due to the limitations of the classical Lyapunov type approach that was followed. In the following pages we will see how in this case, the methods of geometric mechanics and symmetry-based stability analysis are capable of providing sets of sufficient conditions that complement the partial results already existing in the literature and that ensure nonlinear stability. Geometric mechanical methods have already been applied in the context of two systems involving spatially extended magnetic bodies, namely the {\it levitron} and the {\it magnetic dumbbells}. The levitron~\cite{Harrigan} is a magnetic spinning top in the presence of gravitation that can levitate in the air repelled by a base magnet. The stability of this dynamical phenomenon has been explored with the tools of geometric mechanics in~\cite{Dullin1999, dullin:2004, levitronMarsden}. Unfortunately, in this system there are not sufficient conserved quantities available to conclude nonlinear stability using energy-momentum methods and only linear stability estimates are available. The magnetic dumbbells~\cite{Kozorez1974} are two axisymmetric magnetic rigid bodies in space interacting contactlessly with each other; this system exhibits stable regular relative equilibria for which stability conditions have been found using the energy-momentum method in~\cite{sszub12}. The results obtained in this paper are also of much interest at the time of clarifying several misconceptions in the physics literature that incorrectly state that purely magnetic systems cannot exhibit stable behavior in the absence of other long-range forces. This belief goes back to Earnshaw's Global Instability Theorem~\cite{Earnshaw1842}, one of the most profusely cited results in the physics literature concerning stability in magnetic systems~\cite{Bassani2006}. Earnshaw's theory concerns mainly point particles and it was generalized during the last 170 years to a large variety of systems dealing with pure/combined confinement \cite{Simon2001} exhibiting both static and dynamic~\cite{Ginzburg1947, Tamm1979} solutions. Such extensions were in many occasions not rigorously proved and hence experimental results in the last eighty years by Meissner~\cite{Meissner1933}, Braunbeck~\cite{Braunbek1939}, Arkadiev-Kapitsa~\cite{Arkadiev1947} (levitation with Type I superconductors), Brandt~\cite{Brandt1989}, \cite{Brandt1990} (levitation with Type II superconductors), or Harrigan~\cite{Harrigan} (levitron), raised questions as to the universal applicability of Earnshaw's theory. The positive stability results obtained in this paper for dynamic solutions of the orbitron lead us to believe that other similar configurations that have been experimentally observed to be stable could be rigorously proved to have this property despite widespread beliefs in the opposite direction. An important example are the 1941--1947 results by Tamm and Ginzburg that claim that in the case of two interacting magnetic dipoles, orbital motion is impossible using both classical and quantum mechanical descriptions~\cite{Ginzburg1947}; nevertheless, there exist experimental prototypes where a small permanent magnet exhibited quasiorbital motion around another fixed permanent magnet for up to six minutes~\cite{Kozorez1974}. We plan to tackle these questions with methods similar to those put at work in this paper for the orbitron in a forthcoming publication. The paper is organized as follows: in Section~\ref{The orbitron section} we present the Hamiltonian description of the orbitron by including a detailed geometric description of its phase space, equations of motion, symmetries, and associated momentum map. Section~\ref{Relative equilibria of the orbitron} contains a characterization of the relative equilibria of the orbitron that is obtained out of the critical points of the augmented Hamiltonian, constructed using the momentum map associated to the toral symmetry of this system spelled out in the preceding section. Section~\ref{Stability analysis of the relative equilibria of the orbitron} is dedicated to the stability analysis of two branches of equatorial relative equilibria introduced in Section~\ref{Relative equilibria of the orbitron}. One of these branches is singular, in the sense that it exhibits nontrivial isotropy group, and the other one is regular. The stability study is carried out for both the standard and the generalized orbitrons using the energy--momentum method, which yields in this case a set of conditions whose joint satisfaction is sufficient for the toral stability of the regular relative equilibria. Concerning the singular relative equilibria, none of these solutions can be proved to be stable using the energy--momentum method for the standard orbitron, while in the generalized case we are able to specify sufficient conditions involving both the design parameters of the external magnetic field and the dynamical features of the system that guarantee its nonlinear stability. In the second part of Section~\ref{Stability analysis of the relative equilibria of the orbitron} we introduce new linear methods to assess the sharpness of the stability conditions; more specifically, we show that the spectral instability of a natural linearized Hamiltonian vector field that can be associated to any relative equilibrium, ensures its nonlinear instability. This result is very instrumental in our setup since it allows us, for example, to prove the nonlinear instability of the singular branch of relative equilibria of the standard orbitron and the sharpness of some of the nonlinear stability conditions obtained in the regular case. In order to improve the readability of the paper, most proofs of the results in the paper and a number of technical details about the geometry of the system that are used in those proofs, have been included in appendices at the end of the paper (Section~\ref{Appendices}). \medskip \noindent {\bf Acknowledgments:} The authors thank the Fields Institute and the organizers of the Marsden Memorial Program on Geometry, Mechanics, and Dynamics that made possible the collaboration that lead to this work. LG acknowledges financial support from the Faculty for the Future Program of the Schlumberger Foundation. \section{The orbitron} \label{The orbitron section} The standard orbitron is a a small axisymmetric magnetized rigid body (for example a small permanent magnet or a current-carrying loop) with magnetic moment $\mathbf{\mu} $, in the permanent magnetic field created by two fixed magnetic poles modeled by opposite charges placed at distance $h$~\cite{Smythe1939} in the absence of gravity (see Figure \ref{fig:orbitronSketch}); in this definition the adjective ``small'' refers to the size of the body in comparison with the distance $2h$ between the magnetic poles. In this section we provide the Hamiltonian description of this physical system. \begin{figure}[h] \centering \includegraphics[scale=0.3]{orbitronSketch.png} \caption{Schematic representation of the standard orbitron. The magnetic rigid body interacts exclusively with the fixed magnetic poles represented in the picture. The opposite poles of each fixed magnet are assumed to be very distant in comparison with the dimensions of the small rigid body; therefore, their influence is negligible and they are hence not represented .} \label{fig:orbitronSketch} \end{figure} \medskip \noindent {\bf Phase space.} The configuration space of the orbitron is the special Euclidean group in three dimensions $SE(3) = SO(3) \times \mathbb{R}^3$. The $\mathbb{R}^3 $ factor of $SE(3) $ accounts for the position of the center of mass in space of the rigid body and $SO(3)$ specifies its orientation with respect to a fixed initial frame. The orbitron is a simple mechanical system in the sense that its Hamiltonian function is of the form kinetic plus potential energy and that its phase space is the cotangent bundle $T ^\ast SE (3) $ of its configuration space $SE(3) $ endowed with the canonical symplectic structure $\omega $ obtained as minus the differential of the corresponding Liouville one form. As the cotangent bundle of any Lie group, $T ^\ast SE (3) $ can be right or left trivialized in order to obtain the so called space or body coordinates, respectively (see Appendix~\ref{details phase space}), of the phase space. These trivializations provide an identification of the bundle $T ^\ast SE (3) $ with the product $SE(3)\times \mathfrak{se}(3)^\ast $, where the symbol $\mathfrak{se}(3)^\ast $ stands for the dual of the Lie algebra $\mathfrak{se}(3)$ of $SE (3) $. In this paper we will work in body coordinates unless it is specified otherwise. Using this representation, we denote by $(A, \mathbf{x} )$ the elements of $SE(3)=SO(3) \times \mathbb{R}^3$ and by $((A, \mathbf{x} ),({\bf \Pi}, \mathbf{p} )) $ those of $T ^\ast SE(3) \simeq SE(3)\times \mathfrak{se}(3)^\ast $ using body coordinates. \medskip \noindent {\bf Equations of motion.} The Hamiltonian of the orbitron is given by the sum of the kinetic $T( \boldsymbol{\Pi} , \mathbf{p} ) $ and the potential $V \left(A, \mathbf{x} \right) $ energy, that is, \begin{equation} \label{Hamiltonian orbitron} h(\left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi},\mathbf{p}\right) ) = T( \boldsymbol{\Pi} , \mathbf{p} ) + V \left(A, \mathbf{x} \right). \end{equation} The expression of the kinetic energy is: \begin{equation} \label{potential and kinetic} T(\boldsymbol{\Pi}, {\bf p} ): = \frac{1}{2} \boldsymbol{\Pi}^T \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \frac{1}{2M} \| \mathbf{p} \|^2, \end{equation} where $M$ is the mass of the axisymmetric magnetic body and $\mathbb{I}_{ref}$ the reference inertia tensor $\mathbb{I}_{ref} = {\rm diag}(I_1,I_1,I_3)$. The coincidence between the first two principal moments of inertia is related to an axial symmetry with respect to the third coordinate that we assume in the body. The potential energy is given by \begin{equation} V \left(A, \mathbf{x} \right) := -\mu \langle \mathbf{B}(\mathbf{x}), A \mathbf{e}_3 \rangle, \end{equation} where $\mathbf{x} = (x, y, z)\in \mathbb{R} ^3$, $\mu $ is the magnetic moment of the axisymmetric rigid body/dipole, and $ \mathbf{B} (\mathbf{x} )$ is the strength of the magnetic field created by two magnetic poles/``charges" $\pm q$ placed at the points $(0,0,h)$ and $(0,0,-h)$, $h>0$, that is, \begin{equation}\label{magnetic_field} \mathbf{B} (\mathbf{x} )= \dfrac{\mu _0 q}{4\pi}\left( \dfrac{x}{D(\mathbf{x} )_+^{3/2}} - \dfrac{x}{D(\mathbf{x} )_-^{3/2}}, \dfrac{y}{D(\mathbf{x} )_+^{3/2}}-\dfrac{y}{D(\mathbf{x} )_-^{3/2}},\dfrac{z-h}{D(\mathbf{x} )_+^{3/2}}-\dfrac{z+h}{D(\mathbf{x} )_-^{3/2}}\right), \end{equation} with $D(\mathbf{x} )_+=x ^2 + y ^2 + (z-h) ^2$, $D(\mathbf{x} )_-=x ^2 + y ^2 + (z+h) ^2$, and $\mu _0$ the magnetic permeability of vacuum. A small axisymmetric magnetized rigid body subjected to a external magnetic field of the form specified in~(\ref{magnetic_field}) will be called a {\bfseries\itshape standard orbitron}. As we will see later on, most of the results that we present in this paper hold for systems with external magnetic fields that share the following symmetry properties presented by $\mathbf{B}$ in~(\ref{magnetic_field}), namely: \begin{description} \item [(i)] Equivariance with respect to rotations $R _{\theta_S} ^Z$ around the $OZ$ axis: \begin{equation} \label{mag_equivariant} \mathbf{B}(R_{\theta_S}^Z \mathbf{x} ) = R_{\theta_S}^Z\mathbf{B}(\mathbf{x}) \enspace {\rm for} \enspace \theta _S \in \mathbb{R}. \end{equation} \item [(ii)] Behavior with respect to the mirror transformation \begin{equation} \label{mirror transformation} (x,y,z) \longmapsto (x,y,-z) \end{equation} according to the prescription \begin{align}\label{Bx_transformation} &B _x (x,y,z)=-B _x (x,y,-z),\\ \label{By_transformation} &B _y (x,y,z)=-B _y (x,y,-z),\\ \label{Bz_transformation} &B _z (x,y,z)=B _z (x,y,-z). \end{align} \end{description} Consider an arbitrary magnetic field in the magnetostatic approximation in a domain free of other magnetic sources that satisfies these symmetry properties. A small axisymmetric magnetized rigid body subjected to the influence of such an external field will be called a {\bfseries\itshape generalized orbitron}. The generic term {\bfseries\itshape orbitron} will refer simultaneously to both the standard and the generalized orbitrons. The equations of motion of the orbitron are determined by Hamilton's equations: \begin{equation} \label{Hamiltonian equations} \mathbf{i}_{X _h} \omega= \mathbf{d} h, \end{equation} where $\mathbf{i}$ denotes the interior derivative, $\mathbf{d} $ is Cartan's exterior derivative, and $X _h \in \mathfrak{X}(T ^\ast SE (3))$ the Hamiltonian vector field associated to $h \in C^{\infty} (T ^\ast SE (3))$. It can be proved (see Appendix~\ref{Equations of motion of the orbitron}) that in body coordinates, Hamilton's equations~(\ref{Hamiltonian equations}) amount to the set of differential equations: \begin{align} &\dot{A} = A \widehat{ \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi}},\label{equations motion 1}\\ &\dot{\mathbf{x} } = \dfrac{1}{M}A {\bf p},\label{equations motion 2}\\ &\dot{\boldsymbol{\Pi} } = \boldsymbol{\Pi} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + A^{-1} \mathbf{B}(\mathbf{x} )\times \mathbf{e}_3, \label{equations motion 3}\\ &\dot{\mathbf{p} } = \mathbf{p} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \mu A^{-1} D\mathbf{B}( \mathbf{x} ) ^T A \mathbf{e}_3. \label{equations motion 4} \end{align} The symbol $\widehat{ \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi}} $ stands for the antisymmetric matrix associated to the vector $ \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} \in \mathbb{R} ^3$ via the Lie algebra isomorphism\ \ $\widehat{} : \left( \mathbb{R}^3, \times \right) \longrightarrow \left( \mathfrak{so}(3), [\cdot,\cdot] \right) $ introduced in Appendix~\ref{details phase space} and $D$ for the differential. \medskip \noindent {\bf Toral symmetry of the orbitron and associated momentum map.} The axial symmetry of the magnetic rigid body and the rotational spatial symmetry of the magnetic field created by the two poles with respect to rotations around the OZ axis endow this system with a toral symmetry which is obtained as the cotangent lift of the following action on the configuration space: \begin{equation} \label{toral action config space} \begin{array}{cccc} \Phi : &(\mathbb{T}^2=S^1\times S^1)\times SE(3)& \longrightarrow &SE(3)\\ &\left(\left( e^{i\theta_S},e^{i\theta_B}\right) ,(A,\mathbf{x} )\right)&\longmapsto& ({R}^{Z}_{\theta_S} A {R}^{Z}_{-\theta_B},{R}^{Z}_{\theta_S}\mathbf{x} ), \end{array} \end{equation} where $R _\theta ^Z$ denotes the rotation matrix around the third axis by an angle $\theta $. The first circle action involving ${R}^{Z}_{\theta_S} $ implies a spatial rotation of the center of mass of the body and the second one, given by ${R}^{Z}_{\theta_B}$, accounts for a rotation of the magnetic body around its symmetry axis. In Appendix~\ref{The toral action on phase space} we show that the cotangent lift, also denoted by $\Phi $, is a canonical symmetry given by \begin{equation} \label{toral action cotangent} \begin{array}{ccccc} &\Phi: & (\mathbb{T}^2=S^1\times S^1)\times T^*SE(3) & \longrightarrow & T^*SE(3) \\ &&\left(\left( e^{i\theta_S},e^{i\theta_B}\right) ,((A,\mathbf{x} ), (\boldsymbol{\Pi} , \mathbf{p} ))\right)&\longmapsto&\left( ( R_{\theta _S } A R_{-\theta _B }, R_{\theta _S } \mathbf{x} ) , ( R_{\theta _B } \boldsymbol{\Pi}, R_{\theta _B } \mathbf{p} ) \right). \end{array} \end{equation} that has an invariant momentum map associated $\boldsymbol{J} :T^*SE(3) \longrightarrow \mathfrak{t}^*$ given by: \begin{align} \label{momentum map toral action} &\boldsymbol{J} \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi},\mathbf{p}\right) \right) = \left( \langle A \boldsymbol{\Pi} + \mathbf{x} \times A \mathbf{p} , \mathbf{e}_3 \rangle , -\langle \boldsymbol{\Pi}, \mathbf{e}_3 \rangle \right). \end{align} A straightforward computation shows that the Hamiltonian of the orbitron is invariant with respect to the action~(\ref{toral action cotangent}), that is, \begin{equation*} h \circ \Phi_{\left(e^{i\theta_S},e^{i\theta_B}\right)}=h, \quad \mbox{for any} \quad \left(e^{i\theta_S},e^{i\theta_B}\right) \in \mathbb{T}^2, \end{equation*} which, by Noether's Theorem~\cite[Theorem 4.2.2]{Abraham1978}, allows us to conclude that the level sets of the momentum map~(\ref{momentum map toral action}) are preserved by the associated Hamiltonian dynamics, that is, if $F _t $ is the flow of the vector field $X _h$ then $ \mathbf{J} \circ F _t = \mathbf{J} $ for any $t$. The action~(\ref{toral action cotangent}) has two isotropy subgroups, namely, the identity $\{e\} $ and the diagonal circle $H:=\left\{\left(e^{i\theta},e^{i\theta}\right)\mid e^{i\theta} \in S ^1\right\} $. The orbit type submanifold $\left(T^*SE(3)\right) _H $ is given by \begin{equation} \label{singular orbit type} \left(T^*SE(3) \right)_H = \left\{ \left(\left(R _\theta^Z, a \mathbf{e} _3 \right),\left(b \mathbf{e} _3, c \mathbf{e} _3 \right) \right)\mid \theta, a, b, c \in \mathbb{R}\right\}, \end{equation} and $\left(T^*SE(3) \right)_{\{e\}} =T^*SE(3) \setminus \left(T^*SE(3)\right) _H$. The bifurcation lemma (see for instance~\cite[Proposition 4.5.12]{Ortega2004}) guarantees that the restriction of the momentum map to the regular isotropy type $T^*SE(3) _{\{e\}}$ is a submersion and that it has rank one at points in the isotropy type $T^*SE(3) _H $. \section{Relative equilibria of the orbitron} \label{Relative equilibria of the orbitron} In this section we specify the equations that characterize the relative equilibria of the orbitron with respect to its toral symmetry. \medskip \noindent {\bf Relative equilibria: setup and background.} Consider a vector field $X \in \mathfrak{X}(M) $ on a manifold $M$ that is equivariant with respect to action of a Lie group $G$ on it. We say that the point $m \in M $ is a relative equilibrium with velocity $\xi\in \mathfrak{g} $ if the value of vector field at that point coincides with the infinitesimal generator $\xi_M$ associated to $\xi $, that is, \begin{equation} \label{relative equilibrium condition} X _h (m)= \xi_M(m). \end{equation} The Lie algebra element $\xi\in \mathfrak{g} $ is called the velocity of the relative equilibrium. This defining property is equivalent to saying that the flow $F _t $ associated to the vector field $X$ at the point $ m\in M $ coincides with the one-parameter Lie subgroup of $G$ generated by $\xi\in \mathfrak{g}$, that is, \begin{equation} \label{relative equilibrium with flow} F _t (m)=\exp t \xi \cdot m, \end{equation} where $\exp $ is the Lie group exponential map $\exp : \mathfrak{g} \rightarrow G$. In the Hamiltonian setup, relative equilibria have a very convenient characterization that uses the critical points of a function instead of the equilibria of the vector field $X- \xi_M $, as in~(\ref{relative equilibrium condition}). Indeed, consider now a symmetric Hamiltonian system $(M, \omega, h, G, \mathbf{J}:M \rightarrow \mathfrak{g}^\ast) $ and assume that the momentum map $\mathbf{J} $ is coadjoint equivariant; it can be shown~\cite{Abraham1978} that the point $m \in M $ is a relative equilibrium of the Hamiltonian vector field $X _h $ with velocity $\xi\in \mathfrak{g} $ if and only if \begin{equation} \label{relative critical points} \mathbf{d} \left(h- \mathbf{J}^\xi\right)(m)=0, \end{equation} where $\mathbf{J}^\xi:= \langle\mathbf{J}, \xi\rangle$. The combination $h- \mathbf{J}^\xi$ is usually referred to as the {\bfseries\itshape augmented Hamiltonian}. If the relative equilibrium $m\in M$ is such that $\mathbf{J}(m)= \mu \in \mathfrak{g}^\ast $ and we denote its isotropy subgroup with respect to the $G$ action by $G _m$, the law of conservation of the isotropy~\cite{Ortega2004} and Noether's Theorem imply~\cite[Theorem 2.8]{Ortega1999a} that $\xi \in {\rm Lie} \left(N_{G _\mu}(G _m)\right) $, where $G _\mu $ is the coadjoint isotropy of $\mu \in \mathfrak{g}^\ast$ and $N_{G _\mu}(G _m) $ is the normalizer group of $G _m $ in $G _\mu $ (note that $G _m \subset G _\mu $ necessarily due to the equivariance of the momentum map). Finally, notice that the velocity of a relative equilibrium with nontrivial isotropy is not uniquely defined; indeed, it is clear in~(\ref{relative equilibrium with flow}) that if $\xi\in \mathfrak{g}$ is a velocity for the relative equilibrium $m $, then so is $\xi+ \eta $ for any $\eta \in {\rm Lie} \left(G _m\right) $. \medskip \noindent {\bf Relative equilibria equations of the orbitron.} The next proposition, proved in Appendix~\ref{proof proposition relative}, specifies the critical point equations~(\ref{relative critical points}) in the case of the orbitron and shows the existence of branches of relative equilibria whose stability we will study in the next section. \begin{proposition} \label{prop_rel_eq} Consider the orbitron system introduced in Section~\ref{The orbitron section} whose Hamiltonian function is given by~(\ref{Hamiltonian orbitron}) and let $\mathbf{z}=\left(\left(A, \mathbf{x}\right), \left(\boldsymbol{\Pi}, {\bf p}\right)\right) \in T ^\ast SE(3) $. Then: \begin{description} \item [(i)] The point $\mathbf{z}$ is a relative equilibrium of the orbitron with velocity $(\xi_1, \xi_2)\in \mathbb{R} ^2$ with respect to the toral symmetry introduced in Section~\ref{The orbitron section} if and only if the following identities are satisfied: \begin{align} \label{equationderivative1s} &\mu \left[ \mathbf{B} (\mathbf{x} ) \times A \mathbf{e}_3 \right] + \xi _1 \left[ A \mathbf{p} \times (\mathbf{x} \times \mathbf{e}_3 ) - A \boldsymbol{\Pi} \times \mathbf{e}_3 \right] = 0, \\ \label{equationderivative2s} &- \mu D\mathbf{B} (\mathbf{x} ) ^T (A \mathbf{e}_3) - \xi _1 \left( A \mathbf{p} \times \mathbf{e}_3 \right) = 0,\\ \label{equationderivative3s} &\mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \xi _2 \mathbf{e}_3 - \xi _1 A^{-1} \mathbf{e}_3 =0,\\ \label{equationderivative4s} & \frac{1}{M} \mathbf{p} - \xi _1 A^{-1} \left( \mathbf{e}_3 \times \mathbf{x} \right) = 0. \end{align} \item [(ii)] Consider now $A_0=R_{\theta_0}^Z$, $\mathbf{x} _0= \left( x, y, 0\right) $, $\boldsymbol{\Pi} _0 = I_3 \left( \xi _1 - \xi _2 \right) \mathbf{e}_3 $ and $\mathbf{p} _0=M \xi _1 A_0^{-1}\left( -y,x, 0\right) $. The point $\mathbf{z} _0 = \left( \left(A_0, \mathbf{x}_0 \right) , \left(\boldsymbol{\Pi}_0,\mathbf{p}_0\right) \right) $ is a relative equilibrium of the standard orbitron with velocity $ \left( \xi _1 , \xi _2 \right) $, where $\xi _2 $ is an arbitrary real number and $\xi _1 $ is either arbitrary when $\mathbf{x}_0 =\mathbf{0} $ or \begin{equation} \label{xx1 regular} \xi _1 = \pm \left( - \dfrac{3h\mu q \mu _0 }{2 \pi MD(\mathbf{x}_0 )^{5/2}}\right)^{1/2}, \end{equation} when $\mathbf{x}_0 \neq \mathbf{0}$. In view of the expression~(\ref{xx1 regular}) for the spatial velocity $\xi_1$, the existence of the latter relative equilibrium is only guaranteed when $\mu q <0 $. \item[(iii)] The conclusions in the previous part also hold for the generalized orbitron. In this situation $B _z (x,y,z) = f(x ^2 + y ^2,z )$ for some $f \in C^\infty(\mathbb{R} ^2 )$, and the spatial velocity $\xi _1 $ of the relative equilibria with $\mathbf{x}_0 \neq \mathbf{0} $ is given by \begin{equation} \label{xx1 generalized} \xi _1 = \pm \left( -\dfrac{2}{M} \mu f'_1\right) ^{1/2}, \end{equation} where $f'_1 = \left.\dfrac{\partial f(v,z )}{\partial v}\right|_{v=x ^2+ y ^2, z = 0}$. In view of the expression of the spatial velocity $\xi_1$ in~(\ref{xx1 generalized}), the existence of this relative equilibrium is only guaranteed when $\mu f _1 ' <0 $. \end{description} The relative equilibria in these statements for which $\mathbf{x}_0 \neq \mathbf{0} $ (respectively $\mathbf{x}_0 = \mathbf{0} $) have trivial (respectively nontrivial $H$) isotropy and hence belong to the orbit type $\left(T^*SE(3)\right) _{\{e\}} $ (respectively $\left(T^*SE(3)\right) _{H} $); we will refer to them as {\bfseries\itshape regular relative equilibria} (respectively {\bfseries\itshape singular relative equilibria}). \end{proposition} \begin{figure}[!htp] \centering \includegraphics[scale=0.3]{relEqulibriaSketch.png} \caption{Regular and singular relative equilibria of the standard orbitron. The symbols $r_{min} $ and $r_{max}$ represent the stability region in configuration space determined by the conditions that will be presented later on in part {\bf (i)} of Theorem~\ref{kozorez relations}.} \label{fig:relEqulibriaSketch} \end{figure} \section{Stability analysis of the relative equilibria of the orbitron} \label{Stability analysis of the relative equilibria of the orbitron} In this section we study the stability properties of the branches of relative equilibria of the orbitron introduced in the second and third parts of Proposition~\ref{prop_rel_eq}. \medskip \noindent {\bf The energy--momentum method.} As we already explained in the introduction, the degeneracies present in symmetric systems cause various drift phenomena that complicate the selection of a stability criterion. The most natural and fruitful choice is that of stability relative to a subgroup, introduced in~\cite{smodg} for relative equilibria and in~\cite{po} for relative periodic orbits. \begin{definition} \label{stability relative to a subgroup} Let $X\in\frak{X}(M)$ be a $G$--equivariant vector field on the $G$--manifold $M$ and let $G'$ be a subgroup of $G$. A relative equilibrium $m\in M$ of $X$, is called $G'${\bfseries\itshape--stable}, or {\bfseries\itshape stable modulo} $G'$, if for any $G'$--invariant open neighborhood $V$ of the orbit $G'\cdot m$, there is an open neighborhood $U\subset V$ of $m$, such that if $F_t$ is the flow of the vector field $X$ and $u\in U$, then $F_t (u)\in V$ for all $t\geq 0$. \end{definition} In the Hamiltonian setup there exists a variety of Dirichlet type results that provide sufficient conditions for the $G _\mu $--stability of a given relative equilibrium, where $\mu \in \mathfrak{g}^\ast $ is the momentum value in which it is sitting and $G _\mu $ is its coadjoint isotropy. The reason why the subgroup $G _\mu$ arises naturally is clear if we look at the stability problem from the symplectic reduction point of view; more explicitly, consider a symmetric Hamiltonian system $(M, \omega, h, G, \mathbf{J}: M \rightarrow \mathfrak{g}^\ast)$ that exhibits a relative equilibrium at the point $m \in M $ such that $\mathbf{J}(m)= \mu $. Suppose that the momentum map $\mathbf{J}$ is coadjoint equivariant and that the coadjoint isotropy $G _\mu$ acts freely and properly on the momentum fiber $\mathbf{J}^{-1}(\mu) $; in these conditions, the quotient space $ \mathbf{J}^{-1}(\mu) / G _\mu$ is naturally symplectic~\cite{mr,mwr} and the $G$-equivariant Hamiltonian vector field associated to $h$ projects onto another Hamiltonian vector field in which the relative equilibrium $m$ becomes a standard equilibrium. The importance of this construction in our context comes from the fact that the standard Lyapunov stability of the reduced equilibrium is equivalent to the $G _\mu $--stability of the relative equilibrium. The following result, known as the {\bfseries\itshape energy--momentum method}, provides a sufficient condition for the $G _\mu $--stability of a given relative equilibrium. This result has been introduced at different levels of generality in~\cite{smodg, Ortega1999a, patrick:roberts:wulff:2002, MontaldiOlmos2011}. \begin{theorem}[Energy-momentum method] \label{energy momentum statement} Let $(M,\omega, h)$ be a symplectic Hamiltonian system with a symmetry given by the Lie group $G$ acting properly on $M$ with an associated coadjoint equivariant momentum map $\mathbf{J} :M\rightarrow \mathfrak{g}^\ast$. Let $m\in M$ be a relative equilibrium such that $\mathbf{J}(m)=\mu\in\mathfrak{g}^\ast $ and assume that the coadjoint isotropy subgroup $G _\mu $ is compact. Let $\xi\in {\rm Lie}N_{G _\mu}(G _m)$ be a velocity of the relative equilibrium. If the quadratic form \begin{equation} \label{quadratic form stability} \mathbf{d}^2(h-\mathbf{J}^\xi)(m)|_{W\times W} \end{equation} is definite for some (and hence for any) subspace $W$ such that \[ \label{definitionofwre} \ker T_m \mathbf{J}= W\oplus T_m(G_\mu\cdot m), \] then $m$ is a $G_\mu$--stable relative equilibrium. If $\dim W=0$, then $m$ is always a $G_\mu$--stable relative equilibrium. The quadratic form $\mathbf{d}^2(h-\mathbf{J}^\xi)(m)|_{W\times W}$, will be called the {\bfseries\itshape stability form} of the relative equilibrium $m$ and $W$ a {\bfseries\itshape stability space}. \end{theorem} \begin{remark} \normalfont Even though we work exclusivey in the Hamiltonian setup, this criterion has elaborate counterparts in the Lagrangian side~\cite{slm, Rodriguez-Olmos2006}. \end{remark} \begin{remark} \normalfont The statement in Theorem~\ref{energy momentum statement} can be generalized to the context of Hamiltonian actions on Poisson manifolds and can be stated so that one can take advantage of existing Casimirs or other non-symmetry related conserved quantities in order to prove the stability of a given relative equilibrium~\cite[Theorem 4.8]{pos}. More explicitly, if in the conditions of Theorem~\ref{energy momentum statement} there exists a set of $G_\mu$--invariant conserved quantities $C_1,\ldots,\,C_n:M\rightarrow\mathbb{R}$ for which \[ \mathbf{d}(h-\mathbf{J}^\xi+C_1+\cdots+C_n)(m)=0, \] and \[ \mathbf{d}^2(h-\mathbf{J}^\xi+C_1+\cdots+C_n)(m)|_{W\times W} \] is definite for some (and hence for any) $W$ such that \[ \ker\mathbf{d} C^1(m)\cap\ldots\cap\ker\mathbf{d} C^n(m)\cap \ker T_m \mathbf{J}= W\oplus T_m(G_\mu\cdot m), \] then $m$ is a $G_\mu$--stable relative equilibrium. \end{remark} \medskip \noindent {\bf Nonlinear stability of the orbitron relative equilibria.} The application of the energy-momentum method to the relative equilibria of the orbitron introduced in Proposition~\ref{prop_rel_eq} makes possible the determination of sizeable regions in parameter space for which those solutions are $G _\mu $-stable ($\mathbb{T}^2 $-stable in this case). We spell this out in the statement of the following theorem whose proof is provided in the Appendix~\ref{proof kozorez}. \begin{theorem} \label{kozorez relations} Consider the relative equilibria introduced in Proposition~\ref{prop_rel_eq}. Then: \begin{description} \item[(i)] The regular relative equilibria of the standard orbitron in part {\bf (ii)} of Proposition~\ref{prop_rel_eq}, that is, those for which $\mathbf{x}_0 \neq \mathbf{0} $, are $\mathbb{T}^2 $--stable whenever the following three inequalities are satisfied: \begin{equation} \label{kozoriez} \dfrac{2}{3}<\dfrac{r^2}{h^2} < 4, \end{equation} \begin{equation}\label{signxi2} {\rm sign}(\xi _1^0 ) I_3 \xi _2<-\mid \xi_1^0 \mid \left( I_1-I_3+\dfrac{2}{3} M \dfrac{(r^2+h^2)h^2}{3 r^2-2 h^2}\right), \end{equation} where $r ^2=\|\mathbf{x}_0\| ^2 $, $\xi _1 ^0 = \pm \left( - \dfrac{3h\mu q \mu _0 }{2 \pi MD(\mathbf{x}_0 )^{5/2}}\right)^{1/2} $, and $\mu q <0$. The singular relative equilibria ($\mathbf{x}_0 = \mathbf{0} $) are always formally unstable, in the sense that the stability form~(\ref{quadratic form stability}) exhibits a nontrivial signature. \item[(ii)] \label{kozorez relations generalized} The regular relative equilibria of the generalized orbitron in part {\bf (iii)} of Proposition~\ref{prop_rel_eq} are $\mathbb{T}^2 $--stable whenever the following conditions hold: \begin{align} & \mu f'_1< 0,\label{f11}\\ &\mu\left(2 f_{1}'+r^2f_{1}''\right)<0,\label{kozorez generalized1}\\ &\mu f _2'' <0,\label{f22}\\ &{\rm sign}(\xi _1^0 ) I_3 \xi _2<- |\xi_1^0| \left( (I_1-I_3)+\dfrac{1}{2}M\left(\dfrac{f _0}{f'_1}+4 r ^2\dfrac{{f_{1}' }}{f_{2}'' } \right) \right),\label{cond3_xi2} \end{align} where $r ^2=\|\mathbf{x}_0\| ^2 $, $f\in C^{\infty}(\mathbb{R} ^2)$ is the function such that $B _z (x,y,z) = f(r ^2,z )$, $f _0= f (r ^2, 0)$, $f'_1 = \left. \dfrac{\partial f(v,z)}{\partial v}\right|_{v=r ^2, z= 0}$, $f''_1 = \left. \dfrac{\partial ^2 f(v,z)}{\partial v ^2}\right|_{v=r ^2, z= 0}$, $f''_2 = \left. \dfrac{\partial ^2 f(v,z)}{\partial z ^2}\right|_{v=r ^2, z= 0}$, and $\xi _1 ^0 = \pm{ \left( -\dfrac{2}{M} \mu f'_1\right)}^{1/2}$. The singular branch ($\mathbf{x}_0= {\bf 0}$) is $\mathbb{T} ^2$--stable if the following conditions are satisfied: \begin{align} &\mu f' _1<0,\label{f11_gen_singular_cond}\\ &\mu f'' _2< 0,\label{f22_gen_singular_cond}\\ & \xi_1^2< -\dfrac{2}{M} \mu f'_1,\label{xi1_gen_singular_cond}\\ & {\rm sign}(\xi_1)\Pi_0> \dfrac{I _1\xi_1^2- \mu f_0}{|\xi_1|},\label{Pi0_gen_singular_cond} \end{align} where $\Pi_0= I _3\left( \xi_1- \xi_2\right)$ and we use the same notation as above for $f _0$, $f' _1$, and $f'' _2$, replacing $v = r ^2$ by $v = 0$. When $\mu f _0<0$ and $\dfrac{f _0}{f' _1}<\dfrac{2}{M} I _1$, the conditions \eqref{xi1_gen_singular_cond} and \eqref{Pi0_gen_singular_cond} can be replaced by the following single $\xi_1$--independent optimal condition: \begin{equation} \label{Pi0_gen_singular_optimal_cond} |\Pi_0|>2 \sqrt{- \mu f_0 I _1}. \end{equation} This optimal condition is achieved by using the spatial velocities $\xi_1=\pm \left(-\dfrac{1}{I _1} \mu f_0\right)^{1/2}$; the positive (respectively negative) sign for the velocity corresponds to positive (respectively negative) values of $\Pi_0 $. \end{description} \end{theorem} \begin{remark} \normalfont The right inequality in~(\ref{kozoriez}) was already known by Kozorez~\cite{Kozorez1981} but it does not ensure by itself the nonlinear stability of this symmetric configuration. We will refer to this inequality as the {\bfseries\itshape Kozorez condition}. The extension of this inequality in the context of the generalized orbitron is given by \eqref{kozorez generalized1}. \end{remark} \begin{remark} \normalfont The formal instability of the singular branch of the standard orbitron is not informative about its actual nonlinear stability or instability. This point is determined via a complementary spectral stability analysis of the linearized system that we carry out later on in Theorem~\ref{unstable branch} and that allows us to conclude the nonlinear instability of this singular branch of relative equilibria. \end{remark} \begin{remark} \label{why gaussian elimination} \normalfont The proof of the theorem presented in Appendix~\ref{proof kozorez} consists of studying the definiteness of the stability form~(\ref{quadratic form stability}) introduced in Theorem~\ref{energy momentum statement}. A quick dimension count shows that the stability spaces corresponding to the regular and singular branches of relative equilibria are eight and ten dimensional, respectively. The need of determining the sign of the eigenvalues of stability forms in high dimensions like ours has motivated the introduction in the literature of various block diagonalizations for it based on arguments of dynamic~\cite{slm, Rodriguez-Olmos2006} or kinematic~\cite{Ortega1999a} nature. An elementary but important observation that we point out in the proof of this theorem is that in order to ensure the stability of the relative equilibrium in question there is no need to compute the eigenvalues of the stability form but only to determine its signature; the relevance of this statement lies in the fact that by Sylvester's Law of Inertia, the signature is invariant by conjugation with respect to invertible matrices and hence can be read out of the pivots of the matrix obtained by performing Gaussian elimination on the stability form. Unlike the situation faced when computing eigenvalues, Gaussian elimination can be carried out formally and not just numerically in virtually any dimension. This remark is of much importance for non-simple mechanical systems for which dynamic block diagonalizations similar to those cited above are rarely available. \end{remark} \begin{remark} \normalfont Conditions \eqref{f11_gen_singular_cond}--\eqref{Pi0_gen_singular_cond} can be used in the design of magnetic fields capable of confining magnetic rigid bodies that do not exhibit spatial rotation. This is the working principle of devices such as magnetic contactless flywheels or levitrons. In the case of flywheels, up until now only actively controlled versions have been developed; as to the levitron, the potentials that have been considered so far~\cite{Dullin1999, dullin:2004, levitronMarsden} do not allow to conclude nonlinear stability using the methods put at work in Theorem~\ref{kozorez relations} and only the spectral stability of the corresponding linearized systems has been considered. We plan to explore in detail these systems in a future publication. \end{remark} \medskip \noindent {\bf Linear stability and instability analysis tools for relative equilibria.} The use of the energy-momentum method provides sufficient but not necessary nonlinear stability conditions. More specifically, there is no guarantee that the stability regions determined by the inequalities in the statement of Theorem~\ref{kozorez relations} are optimal in the sense that as soon as those conditions are violated stability disappears. In the context of stability studies for standard equilibria one usually proceeds by examining the spectral stability of the linearization at the equilibrium of the vector field in question, that is, when the sufficient stability conditions obtained via a Dirichlet type criterion are violated, one looks for eigenvalues of the linearization that exhibit a nonzero real part, whose existence would imply the nonlinear instability of the equilibrium of the original vector field. This way to proceed can be extended in the context of regular relative equilibria by looking at the spectral stability of the linearization of the reduced Hamiltonian vector field at the equilibrium corresponding to the relative equilibrium in the symplectic Marsden--Weinstein reduced space~\cite{mwr}. Even though in the singular case, there exist reduced spaces that generalize the Marsden--Weinstein reduced space~\cite{sl, or2006a, or2006}, the equivalence between $G _\mu $-stability of a relative equilibrium and standard nonlinear stability of the corresponding reduced equilibrium does not hold anymore, which makes necessary the formulation of a criterion that, as the energy-momentum method in Theorem~\ref{energy momentum statement}, provides a linear stability analysis tool for relative equilibria whose formulation does not need reduction; such a statement is provided in the next proposition, whose proof can be found in the appendix, and we will apply it later on to the branches introduced in Proposition~\ref{prop_rel_eq} whose nonlinear stability was studied in Theorem~\ref{kozorez relations}. In order to fix the notation and to make the presentation self contained, we start by recalling the notion of linearization of a vector field at an equilibrium point. \begin{definition} Let $X \in \mathfrak{X}(M)$ be a vector field on the manifold $M$ and let $m_0 \in M$ be an equilibrium point, that is, $X(m_0) = 0$. The {\bfseries\itshape linearization} $X'$ of $X$ at the point $m_0$ is a vector field $X' \in \mathfrak{X}(T_{m_0}M)$ on the vector space $T_{m_0}M $, defined by \begin{align*} \begin{array}{ccc} X': T_{m_0}M & \longrightarrow & T_{m_0} M \times T_{m_0} M\\ v &\longmapsto&\left( v, \left. \frac{d}{d\lambda } \right|_{\lambda = 0} T_{m_0} F_{\lambda } \cdot v\right), \end{array} \end{align*} where $F_{\lambda } $ is the flow of $X$. The eigenvalues of the linear map $\Pi _2 \circ X': T_{m_0}M \longmapsto T_{m_0}M$ are called the {\bfseries\itshape characteristic exponents} of $X$ at $m_0$. The map $\Pi _2: T_{m_0} M \times T_{m_0} M \rightarrow T_{m_0} M $ is the projection onto the second factor. \end{definition} \begin{proposition} \label{linear tools for instability} Let $G$ be a Lie group acting canonically and properly on the symplectic manifold $(M , \omega ) $ and suppose that there exists a coadjoint equivariant momentum map $ \mathbf{J}:M \rightarrow \mathfrak{g}^\ast$ that can be associated to it. Let $h \in C^\infty(M)^{G} $ be a $G$--invariant Hamiltonian and let $m \in M $ be a relative equilibrium of the corresponding $G$--equivariant Hamiltonian vector field $X _h $ with velocity $\xi\in \mathfrak{g}$. Consider a $G _m$--invariant stability space $W$ such that \begin{equation*} \ker T _m\mathbf{J}=W\oplus T _m \left(G _\mu \cdot m\right), \end{equation*} with $\mu := \mathbf{J} (m) $ and $G _\mu \subset G $ the coadjoint isotropy of $\mu\in \mathfrak{g}^\ast$. Then: \begin{description} \item [(i)] $ \left(W, \omega _W\right) $ with $\omega_W:= \omega(m)| _W $ is a symplectic vector subspace of $\left(T _mM, \omega (m)\right) $. \item [(ii)] There exists a symplectic slice $ \left(S, \omega_S\right) $ at $m\in M$ such that $\left(T _mS, \omega_S (m)\right)= \left(W, \omega _W\right) $. \item [(iii)] The Hamiltonian vector field $X_{h _S ^\xi}\in \mathfrak{X}(S) $ in $S$ associated to the Hamiltonian function $h _S^\xi:= \left.\left(h- \mathbf{J}^\xi\right)\right | _S $ exhibits an equilibrium at the point $m \in S \subset M $. \item [(iv)] The linearization $X'_{h _S ^\xi}\in \mathfrak{X}(T _m S) = \mathfrak{X} (W)$ of $X_{h _S ^\xi} $ at $m\in S$ coincides with the linear Hamiltonian vector field $X _Q $ on $(W, \omega _W)$ that has as Hamiltonian vector field the stability form \begin{equation*} Q (w):= \mathbf{d}^2\left(h- \mathbf{J}^\xi\right)(m)(w,w), \qquad w \in W. \end{equation*} \item [(v)] Suppose that the two tangent spaces $T _m \left( G_{\mu } \cdot m\right) $ and $T _m \left( G \cdot m\right) $ coincide. Then \begin{equation} \label{TmM} T _m M = W \oplus W ^\omega. \end{equation} Additionally, let $h ^\xi:=h -\mathbf{J} ^\xi \in C^{\infty}(M) $ be the augmented Hamiltonian and let $X_{h ^\xi}' \in \mathfrak{X}(T _m M ) $ be the linearization of the Hamiltonian vector field $X_{h ^\xi}$ at $m$. Then \begin{equation} \label{XQ} X_Q = \mathbb{P}_W X_{h^ \xi }' \boldsymbol{i} _W, \end{equation} where $\boldsymbol{i}_W: W \hookrightarrow T _m M$ is the inclusion, $\mathbb{P}_W: T_mM \longrightarrow W$ is the projection according to \eqref{TmM}, and $X_{h^ \xi }'$ is the linearization of $X_{h^ \xi }$ at $m$. \item [(vi)] If the linear vector field $X _Q $ is spectrally unstable in the sense that it exhibits eigenvalues with a nontrivial real part, then the relative equilibrium $m \in M $ of $X _h $ is nonlinearly $K $--unstable, for any subgroup $K \subset G $. \end{description} \end{proposition} We now provide a result that spells out how to compute the linearization of a Hamiltonian vector field at an equilibrium for systems whose phase space is the cotangent bundle of a Lie group. The following proposition expresses the linearization that we need in terms of a linear map on the Euclidean vector space formed by the direct product of the Lie algebra and its dual. \begin{proposition} \label{linearization for t stars g} Let $G$ be a Lie group with Lie algebra $ \mathfrak{g} $ and let $T ^\ast G $ be its cotangent bundle endowed with the canonical symplectic form. Consider now the body coordinates (left trivialized) expression $G \times \mathfrak{g}^\ast $ of $T ^\ast G $ and let $h \in C^{\infty}(G \times \mathfrak{g}^\ast) $ be a Hamiltonian function whose associated Hamiltonian vector field $X _h $ exhibits an equilibrium at the point $(g, \mu) \in G \times \mathfrak{g}^\ast $. Then: \begin{description} \item [(i)] Let $\varphi : G \times \left(G \times \mathfrak{g}^\ast\right) \rightarrow G \times \mathfrak{g}^\ast $ be the cotangent lift of the action by left translations of $G$ on $G$ expressed in body coordinates. Let $h ^g:=h \circ \varphi_g $; the Hamiltonian vector field $X_{h ^g} $ exhibits an equilibrium at the point $(e,\mu)$. \item [(ii)] Let $\Phi _g:= T_{(e, \mu)} \varphi_g: T_{(e, \mu)}\left(G \times \mathfrak{g}^\ast\right)\simeq \mathfrak{g}\times \mathfrak{g}^\ast \longrightarrow T_{(g, \mu)}\left(G \times \mathfrak{g}^\ast\right) $ and let $Q \in C^{\infty} \left(T_{(g, \mu)}\left(G \times \mathfrak{g}^\ast\right)\right)$ (respectively $Q ^g \in C^{\infty}(\mathfrak{g}\times \mathfrak{g}^\ast))$ be the quadratic form associated to the second derivative of $h$ at $(g, \mu) $ (respectively of $h ^g $ at $(e, \mu)$). Then \begin{equation} \label{relation quadratic forms} Q ^g=Q \circ \Phi_g \end{equation} and the associated linear Hamiltonian vector fields considered as linear maps satisfy: \begin{equation} \label{relation linearizations} \Phi_g\circ X_{Q ^g}= X _Q \circ \Phi_g. \end{equation} \item [(iii)] The linearization $X_{Q ^g}: \mathfrak{g}\times \mathfrak{g}^\ast \rightarrow \mathfrak{g}\times \mathfrak{g}^\ast $ is given by: \begin{equation} \label{expression linearization at e} X_{Q ^g}(\xi, \tau)= \left(\pi_{\mathfrak{g}^\ast} \left({\rm Hess}(\xi, \tau)\right), -\pi_{\mathfrak{g}} \left({\rm Hess}(\xi, \tau)\right)+ {\rm ad}^\ast _{\pi_{\mathfrak{g}^\ast } {\rm Hess}(\xi, \tau)}\mu\right), \quad \mbox{for any} \quad \left(\xi, \tau\right)\in \mathfrak{g}\times \mathfrak{g}^\ast, \end{equation} where $\pi_{\mathfrak{g}} : \mathfrak{g} \times \mathfrak{g}^\ast \rightarrow \mathfrak{g} $, $\pi_{\mathfrak{g}^\ast } : \mathfrak{g} \times \mathfrak{g}^\ast \rightarrow \mathfrak{g}^\ast $ are the canonical projections and $ {\rm Hess}: \mathfrak{g} \times \mathfrak{g}^\ast \rightarrow \mathfrak{g} \times \mathfrak{g}^\ast $ is the linear map associated to the Hessian of $h ^g $ at $(e, \mu) $ by the relation \begin{equation*} \langle{\rm Hess}(\xi, \tau), (\eta, \rho)\rangle= \mathbf{d}^2h ^g(e, \mu)((\xi, \tau), (\eta, \rho)), \quad (\xi, \tau), (\eta, \rho) \in \mathfrak{g}\times \mathfrak{g}^\ast. \end{equation*} \end{description} \end{proposition} We now implement the expression for the linearization of a Hamiltonian vector field obtained in this proposition, in the particular case of the cotangent bundle $T ^\ast SE(3)$. Let $h \in C^{\infty}(T ^\ast (SE(3))) $ be a Hamiltonian function and let $X _h $ be the corresponding Hamiltonian vector field that we assume has an equilibrium at the point $\mathbf{z} _0= \left((A _0, \mathbf{x}_0), (\boldsymbol{\Pi}_0, {\bf p} _0)\right)$, that is, $\mathbf{d} h(\mathbf{z}_0) =0$. Let $g= (A _0, \mathbf{x} _0)\in SE(3) $ and let $\mathbf{z}= \left((I,{\bf 0}),( \boldsymbol{\Pi}_0, {\bf p} _0)\right)$; using the notation in the previous proposition, it is clear that $\mathbf{z} _0= \varphi_g (\mathbf{z}) $. Let ${\rm Hess}(\mathbf{z} ): \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast \rightarrow \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast$ be the linear map associated to the Hessian of $h \circ \varphi_g $ at $\mathbf{z}$, that is, for any $\mathbf{v}, \mathbf{w} \in T_{\mathbf{z}} \left( T ^\ast SE(3) \right) \simeq \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast$, \begin{equation*} \langle \mathbf{v}, {\rm Hess} (\mathbf{z}) \mathbf{w}\rangle = \mathbf{d} ^ 2 (h\circ \varphi _g) (\mathbf{z} ) (\mathbf{v},\mathbf{w}). \end{equation*} Now, given $\mathbf{v} =\left(\delta A , \boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right) \in \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast$, define the projection (also available also for the $\boldsymbol{\delta}\mathbf{x} $, $\boldsymbol{\delta}\boldsymbol{\Pi} $, $\boldsymbol{\delta}\mathbf{p}$ components): \begin{equation} \begin{array}{ccc} \boldsymbol{\pi}_{\delta A }: \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast&\longrightarrow&\mathbb{R}^{3}\\ \left( \delta A , \boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right)&\longmapsto& \delta A \end{array} \end{equation} By relations~(\ref{relation linearizations}) and~(\ref{expression linearization at e}) in Proposition~\ref{linearization for t stars g}, and the expression~(\ref{ad star}), the linearization $X'_h$ of $X _h $ at $\mathbf{z} _0 $ is given by \begin{equation} \label{expression of linearization 1} X _h'= \Phi_g \circ X _{h ^g}' \circ \Phi_{g ^{-1}}, \end{equation} where $X _{h ^g}':\mathfrak{se}(3) \times \mathfrak{se}(3)^\ast\simeq \mathbb{R}^{12} \rightarrow \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast\simeq \mathbb{R}^{12} $ is the linear map determined by the twelve by twelve matrix \begin{equation} \label{expression of linearization} X'_{h^g} = \left( \begin{array}{c} \boldsymbol{\pi} _{\boldsymbol{\delta}\boldsymbol{\Pi} } {\rm Hess} (\mathbf{z}_0 ) \\ \boldsymbol{\pi} _{\boldsymbol{\delta}\mathbf{p} } {\rm Hess}(\mathbf{z}_0 ) \\ - \boldsymbol{\pi} _{\delta A } {\rm Hess} (\mathbf{z}_0 ) + \widehat{ \boldsymbol{\Pi} }_0 \boldsymbol{\pi} _{\boldsymbol{\delta}\boldsymbol{\Pi}} {\rm Hess}(\mathbf{z}_0 ) + \widehat{ \mathbf{p} }_0 \boldsymbol{\pi} _{\boldsymbol{\delta}\mathbf{p} } {\rm Hess}(\mathbf{z}_0 ) \\ -\boldsymbol{\pi} _{\boldsymbol{\delta}\mathbf{x} } {\rm Hess}(\mathbf{z}_0 )+ \widehat{ \mathbf{p} }_0 \boldsymbol{\pi} _{\boldsymbol{\delta}\mathbf{\Pi} } {\rm Hess}(\mathbf{z}_0 ) \end{array} \right). \end{equation} This expression should be understood as a vertical concatenation of four matrices with three rows and twelve columns each. More explicitly, given that for any $\mathbf{v} =\left(\widehat{\delta A} , \boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right) \in \mathfrak{se}(3) \times \mathfrak{se}(3)^\ast$, $\Phi _g (\mathbf{v})=\left(A _0\widehat{\delta A} , A _0\boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right)\in T _{\mathbf{z} _0} \left(T ^\ast SE(3)\right)$, we can write \begin{equation*} X _h'\left(A _0\widehat{\delta A} , A _0\boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right)= \left(A _0 X _A, A _0 X _{\mathbf{x}}, X_{\boldsymbol{\Pi}}, X_{\boldsymbol{p}}\right), \end{equation*} where $\left( X _A, X _{\mathbf{x}}, X_{\boldsymbol{\Pi}}, X_{\boldsymbol{p}}\right) $ is the image by~(\ref{expression of linearization}) of the vector $\left(\delta A , \boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right) $. \medskip \noindent {\bf Linear stability and instability of the orbitron relative equilibria.} The results presented in the previous paragraph provide all the necessary tools to carry out the linear stability analysis of the relative equilibria of the standard and generalized orbitron introduced in the parts {\bf (ii)} and {\bf (iii)} of Proposition~\ref{prop_rel_eq}. We will proceed by using expressions~(\ref{expression of linearization 1}) and~(\ref{expression of linearization}) in order to compute the linearization at the relative equilibria of the Hamiltonian vector fields associated to the augmented Hamiltonians constructed with the appropriate relative equilibrium velocities. We subsequently use part {\bf (v)} of Proposition~\ref{linear tools for instability} in order to write down the linearization of the restriction of this vector field to the tangent space to a symplectic slice (equivalently, a stability space); finally, we use the last part of this result in order to search for instability regions by looking for eigenvalues of this linearization that exhibit a nontrivial real part and determine how sharp the nonlinear sufficient stability conditions in Theorem~\ref{kozorez relations} are; more specifically, we will see that there might exist relative equilibria that are nonlinearly stable even though the conditions in Theorem~\ref{kozorez relations} are not satisfied. A detailed description of this implementation is provided in Appendix~\ref{Linear stability and instability of the orbitron relative equilibria}. The following result, formulated using the terminology introduced in Proposition~\ref{prop_rel_eq}, summarizes the results of the linear analysis. \begin{theorem} \label{unstable branch} Consider the relative equilibria introduced in Proposition~\ref{prop_rel_eq}. Then: \begin{description} \item [(i)] Regarding the relative equilibria of the standard orbitron in part {\bf (ii)} of the proposition: \begin{description} \item [(a)] The regular relative equilibria that do not satisfy the Kozorez relation ($r^2/h^2 < 4$) are unstable and this stability condition is consequently sharp. The other two stability conditions in~(\ref{kozoriez}) and~(\ref{signxi2}) are not sharp, that is, there are regions in parameter space that do not satisfy them and where the linearized system is spectrally stable. \item [(b)] The singular relative equilibria of the standard orbitron are nonlinearly unstable. \end{description} \item [(ii)] Regarding the relative equilibria of the generalized orbitron in part {\bf (iii)} of the proposition: \begin{description} \item [(a)] The regular relative equilibria that do not satisfy the generalized Kozorez relation~(\ref{kozorez generalized1}), namely, $\mu \left(2f _1 ' + r ^2f _2''\right)<0$, are unstable and this stability condition is consequently sharp. The remaining stability conditions~(\ref{f11}),~(\ref{f22}), and~(\ref{cond3_xi2}) are not sharp, that is, there are regions in parameter space that do not satisfy them and where the linearized system is spectrally stable. \item [(b)] The spectral stability of the singular relative equilibria of the generalized orbitron is equivalent to the following three conditions: \begin{align} & \mu f _1'<0,\label{sing_gen_f11}\\ &\mu f _2''<0,\label{sing_gen_f22}\\ &\Pi_0^2> -4 \mu f _0I _1,\label{sing_gen_Pi0} \end{align} where $\Pi _0=I _3(\xi_1- \xi_2)$. This statement implies that the nonlinear stability conditions~(\ref{f11_gen_singular_cond}) and~(\ref{f22_gen_singular_cond}) are sharp and that the remaining conditions are not. \end{description} \end{description} \end{theorem} \noindent\textbf{Proof.\ \ (i) Part {\bf (a)}} The linearization $X _Q $ at the regular relative equilibria of the Hamiltonian vector field in the stability space associated to the augmented Hamiltonian is provided in the expression~(\ref{linearization regular}). This matrix is block diagonal and the top two by two block has as eigenvalues \begin{equation*} \lambda_{\pm}=\pm \xi_1^0\sqrt{\frac{r^2-4h^2}{r^2+h^2}}, \end{equation*} which are real whenever $r^2/h^2 >4$, that is, when the Kozorez relation is violated. In conclusion, the part {\bf (vi)} of Proposition~\ref{linear tools for instability} ensures that as soon as the Kozorez relation is violated the relative equilibria cease to be stable. The lack of sharpness of the two other stability conditions in~(\ref{kozoriez}) and~(\ref{signxi2}) is observed by studying the spectrum of the remaining six by six block of the linearization $X _Q$ which may be purely imaginary in regions of the parameter space in which those conditions are violated. The expressions corresponding to those six eigenvalues are very convoluted and we therefore do not include them in the paper; in turn, we illustrate this phenomenon in Figure~\ref{fig:stabilityGaps}, in which we plot the maximum absolute value of the real part of the eigenvalues of the linearization versus the radius of spatial rotation $r$ and the body rotation velocity $\xi_2$, respectively, when all the system parameters specified in the caption remain constant. The graph on the left hand side shows that when the radius goes beyond the critical value stipulated by the left inequality in~(\ref{kozoriez}) the spectrum of the linearization remains purely imaginary for a while and the system is hence potentially stable; it is also visible that, as we proved above, the system becomes spectrally unstable as soon as the Kozorez relation ceases to be satisfied. The lack of sharpness of the condition~(\ref{signxi2}) is illustrated in the right hand side graph and is of a slight different nature; indeed, as soon as the condition is not satisfied, spectral instability appears but if the body rotation velocity is sufficiently decreased the system becomes again spectrally stable in some interval of the $\xi_2$ parameter space. \medskip \noindent\textbf{(i) Part {\bf (b)}} The corresponding linearization $X_Q$ at the singular relative equilibria is described in~(\ref{xq singular}). Its spectrum includes the two following eigenvalues: \begin{eqnarray*} \lambda _1 &=& \dfrac{1}{h ^2 }\sqrt{-\dfrac{3 \mu _0 \mu q }{M \pi }},\\ \lambda _2 &=& \sqrt{ - \left( \xi _1 - \dfrac{1}{h ^2 }\sqrt{-\dfrac{3\mu _0 \mu q}{2 M \pi }} \right) ^2 }. \end{eqnarray*} The eigenvalue $\lambda _1 $ can only be purely imaginary when $\mu q > 0$. This in turns implies that the term $\sqrt{-\dfrac{3 \mu _0 \mu q}{2M \pi }}$ in $\lambda _2 $ is purely imaginary and prevents the eigenvalue to be purely imaginary unless ${-\dfrac{3 \mu _0 \mu q}{2M \pi }}$ is zero. \medskip \noindent {\bf (ii) Part (a)} Analogously to the situation in the proof of {\bf (i)}\ Part {\bf (a)}, the linearization $X _Q$ at the regular relative equilibria of the generalized orbitron exhibits the following two eigenvalues: \begin{equation*} \lambda_{\pm}= \pm 2\sqrt{\dfrac{1}{M}\mu\left(2 f _1'+ r ^2 f _1 ''\right)}, \end{equation*} which are obviously purely imaginary if and only if the generalized Kozorez relation~(\ref{kozorez generalized1}) holds. The lack of sharpness in the remaining relations follows from the fact that they contain as particular cases the stability conditions for the standard orbitron that, as we illustrated in Figure~\ref{fig:stabilityGaps}, are not necessary for the spectral stability of $X _Q $. \medskip \noindent {\bf (ii) Part (b)} The linearization $X _Q $ at the singular relative equilibria of the generalized orbitron is provided in~(\ref{xq singular generalized}) and its spectrum is made up by the following ten eigenvalues: \begin{align} \lambda_1 ^{\pm} &= \pm \sqrt{\dfrac{1}{M} \mu f ''_2},\label{sing_gen_lambda1}\\ \lambda _{2,\pm} ^{\pm} &= \pm \sqrt{-\dfrac{1}{M} \left( \xi _1 \sqrt{M} \pm \sqrt{-2 \mu f ' _1 }\right) ^2 },\label{sing_gen_lambda2}\\ \lambda _{3,\pm} ^{\pm} &= \pm \dfrac{1}{2}\sqrt{-\dfrac{1}{I _1 } \left( \left( 2 \xi _1 I _1 - \Pi _0 \right) \pm \sqrt{4 \mu f _0 I _1 + \Pi _0 ^2 } \right) ^2 }.\label{sing_gen_lambda3} \end{align} The eigenvalues $\lambda_1 ^{\pm} $ can be purely imaginary only when $\mu f '' _2 <0$. In order for the four eigenvalues $\lambda _{2,\pm} ^{\pm}$ to have the same property, the term $\sqrt{-2 \mu f ' _1 }$ has to be necessarily a real number, which yields the condition $\mu f' _1 <0$. These two relations obviously imply that the nonlinear stability conditions~(\ref{f11_gen_singular_cond}) and~(\ref{f22_gen_singular_cond}) are sharp. Finally, the remaining four eigenvalues $\lambda _{3,\pm} ^{\pm} $ are purely imaginary whenever the term $\sqrt{4 \mu f _0 I _1 + \Pi _0 ^2 }$ is real, which requires in turn that the relation $\Pi _0 ^2 >4 \mu f _0 I _1$ is satisfied. We note that this relation may hold without~(\ref{xi1_gen_singular_cond}) and~(\ref{Pi0_gen_singular_cond}) or~(\ref{Pi0_gen_singular_optimal_cond}) being satisfied. Indeed, take for example a system for which $\mu f _0<0 $; in that situation, the relation~(\ref{sing_gen_Pi0}) does not impose any constraint on $\Pi_0 $ an hence it is easy to find values for this variable that violate ~(\ref{xi1_gen_singular_cond}) and~(\ref{Pi0_gen_singular_cond}) or~(\ref{Pi0_gen_singular_optimal_cond}). \quad $\blacksquare$ \begin{figure}[!htp] \includegraphics[scale=0.3]{stabilityGaps3.png} \caption{Spectral stability study for the relative equilibria of a standard orbitron with $h = 0.05$ m, $M = 0.0068$ kg, $\mu_0 = 4\pi\cdot10^{-7}$~N$\cdot$A$^{-2}$, $\mu = -0.18375$ A$\cdot$m$^2$, $q = 17.58$ A$\cdot$m, $I_1 = 0.17\cdot10^{-6}$ kg$\cdot$m$^2$, $I_3=0.1\cdot10^{-6}$ kg$\cdot$m$^2$. The position of the red bullets indicates the critical values of $r$ (m) and $\xi_2 $ (rad$\cdot $s$^{-1}$) determined by the stability conditions~(\ref{kozoriez}) and~(\ref{signxi2}), respectively. The grey bands correspond to the stability gaps discussed in the proof of Theorem~\ref{unstable branch}, part {\bf (i)} in which the system is spectrally stable while the stability form exhibits a nontrivial signature.} \label{fig:stabilityGaps} \end{figure} \section{Appendices} \label{Appendices} \subsection{The geometry of the phase space of the orbitron $(T ^\ast SE (3), \omega) $} \label{details phase space} \noindent {\bf Lie group and Lie algebra structure of the configuration space.} The configuration space of the orbitron is the Lie group $SE(3) = SO(3) \times \mathbb{R}^3$ endowed with the semidirect product structure associated to the composition rule \begin{equation} \label{Lie product} \begin{array}{cccc} \Psi:& SE(3)\times SE(3)&\longrightarrow &SE(3) \\ &(({A}_1,\mathbf{x}_1),({A}_2,\mathbf{x}_2)) &\longmapsto &({A}_1 {A}_2, {A}_1 \mathbf{x}_2 + \mathbf{x}_1), \end{array} \end{equation} for which $e=(I,0)$ and $({A},\mathbf{x})^{-1}= ({A}^{-1},-{A}^{-1}\mathbf{x})$. In order to spell out the Lie algebra structure associated to the Lie product~(\ref{Lie product}) we start by recalling the Lie algebra isomorphism \ \ $\widehat{} : \left( \mathbb{R}^3, \times \right) \longrightarrow \left( \mathfrak{so}(3), [\cdot,\cdot] \right) $ between the Lie algebra $(\mathfrak{so}(3),[\cdot ,\cdot ])$ of $SO(3)$ and $(\mathbb{R} ^3, \times) $ endowed with the standard cross product, given by the assignment \begin{equation*} \mathbf{x} = (x_1,x_2,x_3)\in \mathbb{R} ^3 \longmapsto \widehat{\mathbf{x}}:=\begin{pmatrix} 0 & -x_3 & x_2 \\ x_3 & 0 & -x_1 \\ -x_2 & x_1 & 0 \end{pmatrix}. \end{equation*} We recall that isomorphism $\enspace \widehat{}\enspace$ satisfies that $\widehat{\mathbf{x}} \mathbf{w} = \mathbf{x} \times \mathbf{w}$ and that for any $A \in SO(3)$ and $\mathbf{x} \in \mathbb{R}^3$ \begin{align} & T_IL_{A} \cdot \widehat{\mathbf{x}} = A \widehat{\mathbf{x}},\label{elementary relation 1}\\ & \mathrm{Ad}_{A} \widehat{\mathbf{x}} = A \widehat{\mathbf{x}} A^{-1} = \widehat{A\mathbf{x}},\label{elementary relation 2}\\ & \mathrm{Ad}_{A} \widehat{\mathbf{x}} = T_I \left( L _A \circ R_{A^{-1}} \right) \widehat{\mathbf{x}} = \left. \frac{d}{dt} \right|_0 A \exp t \widehat{\mathbf{x}} A^{-1}=A \widehat{ \mathbf{x} } A^{-1}= \widehat{ A \mathbf{x} },\label{elementary relation 3} \end{align} where $L _A: SO(3) \rightarrow SO(3) $ (respectively $R _A $) denotes left (respectively right) translations and $ \mbox{\rm Ad} _A : \mathfrak{so}(3) \rightarrow \mathfrak{so}(3) $ is the adjoint representation. The $\enspace \widehat{} \enspace$ isomorphism induces another one \begin{align*} \enspace \widehat{}\enspace :\ &\mathbb{R}^3 \longrightarrow \mathfrak{so}(3)^*\nonumber \\ &\boldsymbol{\pi} \longmapsto \widehat{\boldsymbol{\pi} } \end{align*} uniquely determined by the relation $\langle \widehat{\boldsymbol{\pi} }, \widehat{ \mathbf{x}} \rangle :=\langle \boldsymbol{\pi} , \mathbf{x} \rangle _{\mathbb{R}^3} $, with $\langle \cdot , \cdot \rangle _{\mathbb{R}^3} $ the Euclidean inner product in $\mathbb{R} ^3$. Using this isomorphism, we have \begin{equation} \label{coadj} \mathrm{Ad}^*_A \widehat{ \boldsymbol{\pi} } = \widehat{ A ^{-1} \boldsymbol{\pi} }. \end{equation} Using this notation, the Lie algebra structure of $ \mathfrak{se}(3) = \mathfrak{so}(3)\times \mathbb{R}^3$ is given by the bracket \begin{equation} \left[ \left( \widehat{ \boldsymbol{\rho}_1 }, \boldsymbol{\tau} _1 \right) , \left( \widehat{\boldsymbol{\rho}_2},\boldsymbol{ \tau}_2 \right) \right] := \left( \widehat{\boldsymbol{ \rho}_1 \times \boldsymbol{\rho}_2} , \boldsymbol{\rho}_1 \times \boldsymbol{\tau}_2 - \boldsymbol{\rho}_2 \times \boldsymbol{\tau}_1 \right). \end{equation} Additionally, for any $(A, \mathbf{x})\in SE(3) $, $\left( \widehat{\boldsymbol{\rho}}, \boldsymbol{\tau} \right), \left( \widehat{\boldsymbol{\rho}} _1 , \boldsymbol{\tau} _1 \right), \left( \widehat{\boldsymbol{\rho} } _2 , \boldsymbol{\tau} _2 \right) \in \mathfrak{se}(3)$, $\left(\widehat{ \boldsymbol{\mu}} , \boldsymbol{\alpha} \right)\in \mathfrak{se}(3)^\ast $, $\boldsymbol{\beta}, \boldsymbol{\gamma} \in \mathbb{R} ^3$, the following relations that we use later on in the paper hold \begin{align} T_{\left( I, 0\right)}L_{\left( A, \mathbf{x} \right) } \cdot \left( \widehat{\boldsymbol{\rho}} , \boldsymbol{\tau} \right) & = \left( A \widehat{\boldsymbol{\rho}}, A \boldsymbol{\tau} \right) \label{leftact}\\ T_{\left( I , 0\right) }R_{\left( A, \mathbf{x} \right) } \cdot \left( \widehat{\boldsymbol{\rho}} , \boldsymbol{\tau} \right) &= \left( \widehat{\boldsymbol{\rho} } A, \boldsymbol{\rho} \times \mathbf{x} + \boldsymbol{\tau} \right) \label{rightact}\\ {\rm ad} _{\left( \widehat{\boldsymbol{\rho} _1} , \boldsymbol{\tau}_1 \right) }\left( \widehat{\boldsymbol{\rho} } _2 , \boldsymbol{\tau} _2 \right) &= \left( \widehat{ \boldsymbol{\rho} _1 \times \boldsymbol{\rho} _2 }, \boldsymbol{\rho} _1 \times \boldsymbol{\tau} _2 - \boldsymbol{\rho} _2 \times \boldsymbol{\tau} _1 \right), \\ {\rm ad}^* _{\left( \widehat{ \boldsymbol{\rho} } , \boldsymbol{\tau} \right) } \left(\widehat{\boldsymbol{\mu}} , \boldsymbol{\alpha} \right) &= \left( \widehat{ \boldsymbol{\mu} \times \boldsymbol{\rho} } + \widehat{ \boldsymbol{\alpha} \times \boldsymbol{\tau} }, \boldsymbol{\alpha} \times \boldsymbol{\rho} \right), \label{ad star} \\ T^*_{\left( I,0\right) } R_{\left( A, \mathbf{x} \right) } \left( \widehat{ \boldsymbol{\beta}} A,\boldsymbol{\gamma}\right) &= \left( \boldsymbol{\beta}+ \mathbf{x} \times \boldsymbol{\gamma}, \boldsymbol{\gamma}\right), \\ T^*_{\left( I,0\right) } L_{\left( A, \mathbf{x} \right) } \left( A \widehat{ \boldsymbol{\beta}}, \boldsymbol{\gamma}\right) &= \left( \boldsymbol{\beta}, A^{-1}\boldsymbol{\gamma}\right). \label{t star} \end{align} In the last two expressions we have identified $T ^\ast _{(A, \mathbf{x})}SE(3) $ with $T _{(A, \mathbf{x})}SE(3) $ using the Frobenius norm in the $SO(3) $ part and the Euclidean norm in the $\mathbb{R}^3 $ part. Using these equalities, it is easy to see that the adjoint and coadjoint actions of $SE(3)$ on its algebra $\mathfrak{se}(3)$ and its dual $\mathfrak{se}(3)^\ast $ are determined by: \begin{align} {\rm Ad}_{(A, \mathbf{x}) } \left( {\widehat{ \boldsymbol{\rho} }, \boldsymbol{\tau} } \right) &= \left( {\rm Ad}_A \widehat{\boldsymbol{\rho}}, - ({\rm Ad } _A \widehat{\boldsymbol{\rho} }) \mathbf{x} + A \boldsymbol{\tau} \right) = \left( \widehat{ A \boldsymbol{\rho}}, \mathbf{x} \times A \boldsymbol{\rho} + A \boldsymbol{\tau} \right), \\ {\rm Ad}^*_{(A, \mathbf{x} )} \left( {\widehat{\boldsymbol{\mu} }, \boldsymbol{\alpha} } \right) &= \left( {\rm Ad}^*_A \widehat{\boldsymbol{\mu} } - \widehat{\left( A^{-1} \left( \mathbf{x} \times \boldsymbol{\alpha }\right) \right)} , A^{-1} \boldsymbol{\alpha} \right) = \left( \left( \widehat{ A^{-1} ( \boldsymbol{\mu} - ( \mathbf{x} \times \boldsymbol{\alpha} ))} \right), A^{-1} \boldsymbol{\alpha} \right).\label{coadjaction} \end{align} \medskip \noindent {\bf Body and space coordinates for $T ^\ast SE (3) $.} Given an arbitrary Lie group $G$ with Lie algebra $\mathfrak{g} $, we recall (see for example~\cite{Abraham1978}) that the maps \begin{equation} \label{maps space coordinates} \begin{array}{cccc} \varrho_1:& TG &\longrightarrow &G \times g \\ &u_g &\longmapsto &(g,T_g R_{g^{-1}}\cdot u_g) \\ &T_e R_g\cdot \xi &\longmapsto &(g,\xi). \end{array}\qquad \mbox{and} \qquad \begin{array}{cccc} \varrho_2: &T^*G &\longrightarrow &G \times g^* \\ &\alpha_g &\longmapsto &(g,T_e^* R_{g}\cdot \alpha_g) \\ &T^*_g R_{g^{-1}}\cdot \mu &\longmapsto &(g,\mu) \end{array} \end{equation} define trivializations of the tangent $TG$ and cotangent bundles $T ^\ast G $, respectively, that are usually referred to as {\bfseries\itshape space coordinates} of these bundles. Notice that if $\varrho _1 \left( u _g \right) = \left( g, \xi \right)$, $\varrho _2 \left( \alpha _g \right) = \left( g, \mu \right)$, then $\left\langle \alpha _g , u _g \right\rangle = \left\langle \mu , \xi \right\rangle $. Analogously, the trivializations obtained using left translations instead via the maps \begin{equation} \label{maps body coordinates} \begin{array}{cccc} \lambda_1:& TG &\longrightarrow &G \times g \\ &u_g &\longmapsto &(g,T_g L_{g^{-1}}\cdot u_g) \\ &T_e L_g\cdot \xi &\longmapsto &(g,\xi). \end{array}\qquad \mbox{and} \qquad \begin{array}{cccc} \lambda_2: &T^*G &\longrightarrow &G \times g^* \\ &\alpha_g &\longmapsto &(g,T_e^* L_{g}\cdot \alpha_g) \\ &T^*_g L_{g^{-1}}\cdot \mu &\longmapsto &(g,\mu) \end{array} \end{equation} are usually referred to as {\bfseries\itshape body coordinates}. Notice that if $\lambda _1 \left( u _g \right) = \left( g, \xi \right)$, $\lambda _2 \left( \alpha _g \right) = \left( g, \mu \right)$, then $\left\langle \alpha _g , u _g \right\rangle = \left\langle \mu , \xi \right\rangle $. We now use these maps to establish the relation between the space and body coordinates $\left( (A, \mathbf{x} ), ( \boldsymbol{\Pi}_S, \mathbf{p} _S) \right) $ and $\left( (A, \mathbf{x} ), ( \boldsymbol{\Pi}_B, \mathbf{p} _B) \right) $, respectively, of an arbitrary point in $T^*SE(3)$. Indeed, using~(\ref{maps body coordinates}),~(\ref{maps space coordinates}), and \eqref{coadjaction}, we have that \begin{align*} \left( (A, \mathbf{x} ), (\boldsymbol{\Pi} _B, \mathbf{p} _B)\right)& = \lambda _2 \left( T^*_{(A, \mathbf{x} )} R_{(A, \mathbf{x} )^{-1}} \cdot (\boldsymbol{\Pi} _S, \mathbf{p} _S)\right) = \left( (A, \mathbf{x} ), {\rm Ad}^* _{(A, \mathbf{x} )} (\boldsymbol{\Pi} _S, \mathbf{p} _S)\right) \\ &= \left( (A, \mathbf{x} ), (A^{-1} (\boldsymbol{\Pi} _S - \mathbf{x} \times \mathbf{p} _S), A^{-1} \mathbf{p} _S)\right). \end{align*} Consequently, \begin{align} \label{spacetobody} &\boldsymbol{\Pi} _B = A^{-1} \left( \boldsymbol{\Pi} _S - \mathbf{x} \times \mathbf{p} _S\right),\nonumber \\ & \mathbf{p} _B = A^{-1} \mathbf{p} _S. \end{align} Conversely, \begin{align} \label{bodytospace} &\boldsymbol{\Pi} _S = A \boldsymbol{\Pi} _B + \mathbf{x} \times A \mathbf{p} _B, \nonumber \\ & \mathbf{p} _S = A \mathbf{p} _B. \end{align} \subsection{Equations of motion of the orbitron} \label{Equations of motion of the orbitron} In this section we obtain the equations of motion~(\ref{equations motion 1})-(\ref{equations motion 4}) of the orbitron using body coordinates. We will proceed by writing down first the differential equations that define a Hamiltonian vector field on the left trivialized cotangent bundle $G \times \mathfrak{g}^\ast$ of an arbitrary Lie group $G$ with Lie algebra $\mathfrak{g}$. \begin{proposition} Let $G$ be a Lie group with Lie algebra $\mathfrak{g} $ and let $T ^\ast G $ be its cotangent bundle endowed with the canonical symplectic form. Let $\omega _B $ be the corresponding symplectic form on the trivial bundle $G \times \mathfrak{g}^\ast $ obtained out of $T ^\ast G $ by left trivialization (body coordinates) and let $h\in C^{\infty}(G \times \mathfrak{g}^\ast)$ be a Hamiltonian function. For any $(g, \mu) \in G \times \mathfrak{g}^\ast $, the Hamiltonian vector field $X _h\in \mathfrak{X}(G \times \mathfrak{g}^\ast)$ associated to $h$ is given by \begin{equation} \label{h from partial} X _h \left( g, \mu \right) = \left( T _I L _g \cdot X _G \left( g, \mu \right) , X _{\mathfrak{g} ^\ast } \left( g, \mu \right) \right), \end{equation} where $X_{G}(g, \mu) \in \mathfrak{g} $ and $ X _{\mathfrak{g} ^\ast } \left( g, \mu \right) \in \mathfrak{g}^\ast $ are determined by \begin{eqnarray} X_G \left( g, \mu \right) &=& D_{\mathfrak{g} ^\ast } h \left( g, \mu \right),\label{xG}\\ X_{\mathfrak{g} ^\ast } \left( g, \mu \right) &=& - T_I ^\ast L_g \cdot D_G h \left( g, \mu \right) + {\rm ad} ^\ast _{D_{\mathfrak{g} ^\ast } h \left( g, \mu \right)} \mu.\label{xgstar} \end{eqnarray} \end{proposition} \noindent\textbf{Proof.\ \ } Using the expression of the canonical symplectic form $\omega_B$ of $T ^\ast G $ in body coordinates (see for instance~\cite[Expression (6.2.5)]{Ortega2004}) it is easy to see that $X _G$, $X _{\mathfrak{g}^\ast} $, and hence $X _h $, are determined by the relation \begin{multline*} \omega_B(g, \mu)\left(X _h \left( g, \mu \right), \left(T _I L _g \cdot \xi _G, \beta\right)\right)=\langle \beta , X _G \left( g, \mu \right) \rangle - \langle X_{g ^\ast } \left( g, \mu \right), \xi _G \rangle \\ + \langle \mu , \left[ X _G \left( g, \mu \right), \xi _G\right] \rangle = D_G h \left( g, \mu \right) \cdot T _I L _g \cdot \xi _G + D_{\mathfrak{g} ^\ast } h \left( g, \mu \right) \cdot \beta, \end{multline*} where $\xi _G \in \mathfrak{g} $ and $\beta \in \mathfrak{g} ^\ast $ are arbitrary and $D_G h $ and $D_{\mathfrak{g}^\ast}h $ are the partial derivatives of $h$ with respect to $G$ and $\mathfrak{g}^\ast $, respectively. Equivalently, \begin{eqnarray*} X_G \left( g, \mu \right) &=& D_{g ^\ast } h \left( g, \mu \right),\\ X_{g ^\ast } \left( g, \mu \right) &=& - T_I ^\ast L_g \cdot D_G h \left( g, \mu \right) + {\rm ad} ^\ast _{D_{g ^\ast } h \left( g, \mu \right)} \mu, \end{eqnarray*} as required. \quad $\blacksquare$ \medskip We now consider the case we are interested in, that is, $G = SE(3) = SO(3) \times \mathbb{R} ^3 $ and \begin{equation} \label{hamiltonian function prop} h \left( (A, \mathbf{x} ), (\boldsymbol{\Pi} , \mathbf{p} )\right) = \dfrac{1}{2} \boldsymbol{\Pi} ^T\mathbb{I}_{ref} ^{-1} \boldsymbol{\Pi} + \dfrac{1}{2M} \mathbf{p}^T \mathbf{p} - \mu \langle \mathbf{B}(\mathbf{x} ), A \mathbf{e}_3 \rangle . \end{equation} Let \begin{equation*} v _{\left( A, \mathbf{x} \right)} = T_{\left( I, \mathbf{0} \right) } L _{\left( A, \mathbf{x} \right) }\cdot \left( \widehat{ \delta A}, \boldsymbol{\delta } \mathbf{x} \right) = \left( A \widehat{\delta A }, A \boldsymbol{\delta}\mathbf{x} \right) \end{equation*} be an arbitrary element of $T_{\left( A, \mathbf{x} \right) } SE(3)$ and $\beta = \left( \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right) \in \mathfrak{se}(3)^\ast $. Then, as \begin{multline*} {\bf d} h \left( (A, \mathbf{x}) , (\boldsymbol{\Pi} , \mathbf{p} )\right) \cdot \left( v_{\left( A, \mathbf{x} \right) }, \beta\right) = \left. \dfrac{d}{dt} \right|_0 h \left( \left( (A, \mathbf{x} ) \cdot (\exp t \widehat{ \delta A }, t \boldsymbol{\delta}\mathbf{x} )\right), (\boldsymbol{\Pi} +t\boldsymbol{\delta}\boldsymbol{\Pi}, \mathbf{p} + t \boldsymbol{\delta}\mathbf{p} ) \right) = \\ \langle \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} , \boldsymbol{\delta}\boldsymbol{\Pi} \rangle + \dfrac{1}{M} \langle \mathbf{p} , \boldsymbol{\delta}\mathbf{p} \rangle - \mu \langle DB (\mathbf{x} )^T A \mathbf{e}_3 , A \boldsymbol{\delta}\mathbf{x} \rangle + \langle A (\widehat{ \mathbf{e}_3 \times A^{-1} \mathbf{B}(\mathbf{x} ) }), A \widehat{ \delta A } \rangle, \end{multline*} we can conclude that \begin{align} D_G h \left( (A, \mathbf{x} ), (\boldsymbol{\Pi} , \mathbf{p} )\right) &=\left( \begin{array}{c} A \left[ \widehat{ \mathbf{e}_3 \times A^{-1} \mathbf{B} (\mathbf{x} ) }\right] \\ -\mu D\mathbf{B} (\mathbf{x} )^T A \mathbf{e}_3 \end{array} \right),\label{DG}\\ D_{g ^\ast } h \left( (A, \mathbf{x} ), (\boldsymbol{\Pi} , \mathbf{p} )\right) &=\left( \begin{array}{c} \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} \\ \frac{1}{M} \mathbf{p} \end{array} \right).\label{Dgstar} \end{align} Now using \eqref{xG} and \eqref{xgstar}, together with and \eqref{DG}, \eqref{Dgstar},~(\ref{ad star}), and ~(\ref{t star}), we obtain \begin{align*} X_{g ^\ast } \left( g, \mu \right) &=\left( \begin{array}{c} - \mathbf{e}_3 \times A^{-1} \mathbf{B}(\mathbf{x} ) + \boldsymbol{\Pi} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} \\ \mu A ^{-1} D\mathbf{B} (\mathbf{x} )^T A \mathbf{e}_3 + \mathbf{p} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} \end{array} \right), \\ X_G \left( g, \mu \right) &=\left( \begin{array}{c} \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} \\ \frac{1}{M} \mathbf{p} \end{array} \right). \end{align*} Consequently, by \eqref{h from partial} we conclude that the equations of motion associated to the Hamiltonian~(\ref{hamiltonian function prop}) are \begin{align*} &\dot{A} = A \widehat{ \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi}},\\ &\dot{\mathbf{x} } = \dfrac{1}{M}A {\bf p},\\ &\dot{\boldsymbol{\Pi} } = \boldsymbol{\Pi} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + A^{-1} \mathbf{B}(\mathbf{x} )\times \mathbf{e}_3, \\ &\dot{\mathbf{p} } = \mathbf{p} \times \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \mu A^{-1} D\mathbf{B}( \mathbf{x} ) ^T A \mathbf{e}_3. \end{align*} \subsection{The toral action on phase space $T^*SE(3)$ and the associated momentum map} \label{The toral action on phase space} \noindent {\bf The expression of the lifted action in body coordinates.} We start by proving that the cotangent lift of the toral action on $SE (3) $ in~(\ref{toral action config space}) is given by~(\ref{toral action cotangent}) when using body coordinates. Consider $H$ and $G$ two arbitrary Lie groups and let $\Phi : H \times G \longrightarrow G$ be an action of $H$ on $G$. We recall that the lift of this action to the cotangent bundle $T ^\ast G $ of $G$, also denoted by $\Phi$, is given by \begin{equation*} \begin{array}{cccc} \Phi : &H \times T^*G &\longrightarrow &T^*G\\ &\left( h, \alpha _g \right) & \longmapsto &T^*_{\Phi_h \left( g\right) } \Phi _{h^{-1}} \cdot \alpha _g. \end{array} \end{equation*} Using the maps introduced in~(\ref{maps body coordinates}), this action is expressed in body coordinates as: \begin{equation*} \Phi \left( h, \left( g, \mu \right) \right) := \lambda _2 \left( \Phi \left( h, \lambda _2 ^{-1} \left( g, \mu \right) \right) \right), \quad \mbox{for any} \quad h \in H, g \in G, \text{ and } \mu\in \mathfrak{g}^\ast, \end{equation*} or equivalently, \begin{equation*} \Phi _h \left( g, \mu \right) = \left( \Phi _h \left( g \right) , T _e ^* L _{\Phi _h \left( g \right) } \cdot T^* _{\Phi _h \left( g \right)} \left( L _{g^{-1}} \circ \Phi _{h^{-1}} \right) \mu \right) = \left( \Phi _h \left( g \right) , T _e ^* \left( L _{g^{-1}} \circ \Phi _{h^{-1}} \circ L _{\Phi _{h} \left( g\right)} \right) \mu \right). \end{equation*} In the particular case of $H=\mathbb{T}^2$, $G=SE(3)$, and the toral action introduced in~(\ref{toral action config space}), that is, \begin{equation*} \begin{array}{cccc} \Phi : &(\mathbb{T}^2=S^1\times S^1)\times SE(3)& \longrightarrow &SE(3)\\ &\left(\left( e^{i\theta_S},e^{i\theta_B}\right) ,(A,\mathbf{x} )\right)&\longmapsto& ({R}^{Z}_{\theta_S} A {R}^{Z}_{-\theta_B},{R}^{Z}_{\theta_S}\mathbf{x} ), \end{array} \end{equation*} we consider $g= \left( A, \mathbf{x} \right) \in SE(3)$, $\mu = \left( \widehat{\boldsymbol{\Pi}} , \mathbf{p} \right) \in \mathfrak{se}(3)^\ast $, and $h= \left( e^{i\theta_S},e^{i\theta_B}\right) \in \mathbb{T}^2 $. Then, \begin{equation} \label{cotangent lift generic} \Phi _h \left( g, \mu \right) = \left( \left( R_{\theta _S }^Z A R_{\theta _{-B} }^Z, R _{\theta _S }^Z \mathbf{x} \right), T _e ^* \left( L _{g^{-1}} \circ \Phi _{h^{-1}} \circ L _{\Phi _{h} \left( g\right)} \right) \mu \right). \end{equation} In order to compute the second part of this expression let $\xi = \left( \widehat{ \boldsymbol{\rho} }, \boldsymbol{\tau} \right) \in \mathfrak{se}(3)$. Then \begin{align*} &\left\langle T _e ^* \left( L _{g^{-1}} \circ \Phi _{h^{-1}} \circ L _{\Phi _{h} \left( g\right)} \right) \mu , \xi \right\rangle = \left\langle \mu , T _e \left( L _{g^{-1}} \circ \Phi _{h^{-1}} \circ L _{\Phi _{h} \left( g\right)} \right) \xi \right\rangle \\ &= \left. \frac{d}{dt} \right|_0 \left\langle \left( \widehat{\boldsymbol{\Pi}} , \mathbf{p} \right) , L_{\left( A^{-1}, -A^{-1} \mathbf{x} \right) } \circ \Phi _{\left( e^{-i\theta_S},e^{-i\theta_B}\right) } \circ L_{\left( R_{\theta _S }^Z A R_{-\theta _{B} }^Z, R _{\theta _S }^Z \mathbf{x} \right)} \left( \exp t \widehat{ \boldsymbol{\rho} }, t \boldsymbol{\tau} \right) \right\rangle \\ & =\left. \frac{d}{dt} \right|_0 \left\langle \left( \widehat{\boldsymbol{\Pi}} , \mathbf{p} \right) , \left( A^{-1} A R_{-\theta _{B}}^Z \exp t \widehat{ \boldsymbol{\rho} } R_{\theta _B }^Z, A^{-1} A R_{-\theta _{B}}^Z t \boldsymbol{\tau} +A^{-1} \mathbf{x} - A^{-1} \mathbf{x} \right) \right\rangle\\ & =\left. \frac{d}{dt} \right|_0 \left\langle \left( \widehat{\boldsymbol{\Pi}} , \mathbf{p} \right) , \left( R^Z_{-\theta _{B}} \exp t \widehat{ \boldsymbol{\rho} } R^Z_{\theta _B }, t R^Z_{-\theta _{B}} \boldsymbol{\tau} \right) \right\rangle= \left\langle \left( \widehat{\boldsymbol{\Pi}} , \mathbf{p} \right) , \left( \rm{Ad}_{R^Z_{- \theta _B }} \widehat{ \boldsymbol{\rho} }, R^Z_{- \theta _B } \boldsymbol{\tau} \right) \right\rangle \\ &= \left\langle \left( {\rm Ad}^*_{R^Z_{- \theta _B }} \widehat{\boldsymbol{\Pi}}, R^Z_{\theta _B } \mathbf{p} \right), \left( \widehat{ \boldsymbol{\rho} }, \boldsymbol{\tau} \right) \right\rangle. \end{align*} Given that by~(\ref{elementary relation 3}) ${\rm Ad^*}_{R^Z_{- \theta _B }} \widehat{\boldsymbol{\Pi}} = \widehat{ R^Z_{\theta _B} \boldsymbol{\Pi}} $, the last equality together with~(\ref{cotangent lift generic}) yield the expression~(\ref{toral action cotangent}) of the lifted action in body coordinates, that is, \begin{equation} \Phi _{\left( e^{i\theta_S},e^{i\theta_B}\right)} \left( (A, \mathbf{x} ) , ( \boldsymbol{\Pi} ,\mathbf{p} ) \right) = \left( ( R_{\theta _S } A R^Z_{-\theta _B }, R^Z_{\theta _S } \mathbf{x} ) , ( R^Z_{\theta _B } \boldsymbol{\Pi} , R^Z_{\theta _B } \mathbf{p} ) \right). \end{equation} \medskip \noindent {\bf The infinitesimal generators of the toral action.} We first show that for any Lie algebra element $\left( \xi _1 , \xi _2 \right) \in \mathbb{R}^2 = {\rm Lie} (\mathbb{T}^2)$ and $\left(A, \mathbf{x} \right) \in SE(3)$, \begin{align} (\xi_1,\xi_2)_{SE(3)}\left(A, \mathbf{x} \right) &= T_{\left( I, 0 \right) } R_{\left(A, \mathbf{x} \right) } \left( \widehat{\xi_1 \mathbf{e} _3 -A\xi_2 \mathbf{e} _3 }, A\xi_2 \mathbf{e} _3 \times \mathbf{x} \right)\label{infgenright}\\ &= T_{\left( I, 0 \right) } L_{ \left( A, \mathbf{x} \right)} \left( {\rm Ad}_{A^{-1}} \widehat{\xi_1 \mathbf{e} _3} -\widehat{ \xi_2 \mathbf{e} _3 }, A^{-1} \left( \xi_1 \mathbf{e} _3 \times \mathbf{x}\right) \right).\label{infgenleft1} \end{align} We start by proving the first equality \begin{align*} &(\xi_1,\xi_2)_{SE(3)}\left(A, \mathbf{x} \right) = \left. \frac{d}{dt} \right|_0 (\exp t\widehat{\xi_1 \mathbf{e} _3 } A \exp ( - t \widehat{\xi_2 \mathbf{e} _3 }) ,\exp t \widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} ) = (\widehat{\xi_1 \mathbf{e} _3 }A-A\widehat{\xi_2 \mathbf{e} _3 },\widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} ) \\ &= (\widehat{\xi_1 \mathbf{e} _3 }A-A\widehat{\xi_2 \mathbf{e} _3 }A^{-1}A,\xi_1 \mathbf{e} _3 \times \mathbf{x} ) = \left( (\widehat{\xi_1 \mathbf{e} _3 }-A\widehat{\xi_2 \mathbf{e} _3 }A^{-1}) A, (\widehat{\xi_1 \mathbf{e} _3 }-A\widehat{\xi_2 \mathbf{e} _3 }+ A\widehat{\xi_2 \mathbf{e} _3 } ) \times \mathbf{x} \right) \\ &= \left( (\widehat{\xi_1 \mathbf{e} _3 -A\xi_2 \mathbf{e} _3 })A, (\xi_1 \mathbf{e} _3 -A\xi_2 \mathbf{e} _3) \times \mathbf{x} + (A\xi_2 \mathbf{e} _3 \times \mathbf{x}) \right) = T_{\left( I, 0 \right) } R_{\left(A, \mathbf{x} \right) } \left( (\widehat{\xi_1 \mathbf{e} _3 -A\xi_2 \mathbf{e} _3 }), A\xi_2 \mathbf{e} _3 \times \mathbf{x} \right), \end{align*} where in the last equality we used \eqref{rightact}. Regarding~(\ref{infgenleft1}), note that \begin{align*} &(\xi_1,\xi_2)_{SE(3)}\left(A, \mathbf{x} \right) = \left. \frac{d}{dt} \right|_0 (\exp t\widehat{\xi_1 \mathbf{e} _3 } A \exp ( - t \widehat{\xi_2 \mathbf{e} _3 }) ,\exp t \widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} ) = (\widehat{\xi_1 \mathbf{e} _3 }A-A\widehat{\xi_2 \mathbf{e} _3 },\widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} ) \\ &= (A A^{-1} \widehat{\xi_1 \mathbf{e} _3 }A-A\widehat{\xi_2 \mathbf{e} _3 },\xi_1 \mathbf{e} _3 \times \mathbf{x} ) =\left( T_I L_A({\rm Ad}_{A^{-1}} \widehat{\xi_1 \mathbf{e} _3 }-\widehat{\xi_2 \mathbf{e} _3 }), (AA^{-1}(\xi_1 \mathbf{e} _3 \times \mathbf{x})) \right)\\ & = T_{\left( I, 0 \right) } L_{ \left( A, \mathbf{x} \right)} \left( {\rm Ad}_{A^{-1}} \widehat{\xi_1 \mathbf{e} _3} -\widehat{ \xi_2 \mathbf{e} _3 }, A^{-1} \left( \xi_1 \mathbf{e} _3 \times \mathbf{x}\right) \right), \end{align*} where we used \eqref{leftact}. The infinitesimal generator of the lifted $\mathbb{T}^2$-action on $T^*SE(3)$ in body coordinates is given by \begin{equation} \label{infinitgenBody} (\xi_1,\xi_2)_{T^*SE(3)}\left(A, \mathbf{x}, \boldsymbol{\Pi} , \mathbf{p} \right)= \left( A ({\rm Ad}_{A^{-1}} (\widehat{\xi_1 \mathbf{e} _3} )-\widehat{ \xi_2 \mathbf{e} _3 }), \widehat{ {\xi_1 \mathbf{e} _3 } } \mathbf{x} , \widehat{ {\xi_2 \mathbf{e} _3 }} \boldsymbol{\Pi} , \widehat{ {\xi_2 \mathbf{e} _3 }} \mathbf{p} \right) \end{equation} Indeed, \begin{align*} &(\xi_1,\xi_2)_{T^*SE(3)}\left(A, \mathbf{x}, \boldsymbol{\Pi} , \mathbf{p} \right) = \left. \frac{d}{dt} \right|_0 \exp t (\xi_1,\xi_2) \cdot \left(A, \mathbf{x}, \boldsymbol{\Pi} , \mathbf{p} \right) \\ &= \left. \frac{d}{dt} \right|_0 \left( \exp t\widehat{\xi_1 \mathbf{e} _3 } A \exp (-t\widehat{\xi_2 \mathbf{e} _3 }), \exp t\widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} , \exp t\widehat{\xi_2 \mathbf{e} _3 } \boldsymbol{\Pi} , \exp t\widehat{\xi_2 \mathbf{e} _3 } \mathbf{p} \right) \nonumber \\ &= \left( AA ^{-1}\widehat{\xi_1 \mathbf{e} _3}A -A\widehat{ \xi_2 \mathbf{e} _3 }, \widehat{\xi_1 \mathbf{e} _3 } \mathbf{x} , \widehat{\xi_2 \mathbf{e} _3 } \boldsymbol{\Pi} , \widehat{\xi_2 \mathbf{e} _3 } \mathbf{p} \right) = \left( A \left({\rm Ad}_{A^{-1}} \left(\widehat{\xi_1 \mathbf{e} _3}\right) -\widehat{ \xi_2 \mathbf{e} _3 }\right), \widehat{ {\xi_1 \mathbf{e} _3 } } \mathbf{x} , \widehat{ {\xi_2 \mathbf{e} _3 }} \boldsymbol{\Pi} , \widehat{ {\xi_2 \mathbf{e} _3 }} \mathbf{p} \right). \end{align*} \medskip \noindent {\bf The momentum map of the toral action} Given a lifted action of a Lie group $H$ on the cotangent bundle $T ^\ast G $ of a Lie group $G$ endowed with the canonical symplectic form, the map $\mathbf{J}: T ^\ast G\longrightarrow \mathfrak{g}^\ast $ defined by \begin{equation} \label{canonical momentum map} \langle \boldsymbol{J} (\alpha_g),\xi\rangle = \langle\alpha_g,\xi_G(g)\rangle \quad \mbox{for any} \quad g \in G,\, \alpha _g \in T^*G, \text{ and } \xi \in \mathfrak{h}, \end{equation} is a coadjoint equivariant momentum map for this canonical action (see~\cite[Corollary 4.2.11]{Abraham1978}). We now study the particular case we are interested in, that is, $H = \mathbb{T}^2$, $G=SE(3)$, and consider an arbitrary point $g=\left(A, \mathbf{x} \right) \in SE (3)$, $\mu= \left(\boldsymbol{\Pi},\mathbf{p}\right) \in \mathfrak{se}(3) ^\ast $ and $\alpha _g = T^* _g L _{g^{-1}} \cdot \mu \in T ^\ast SE(3)$ the covector that in body coordinates is expressed as $\left(\left(A, \mathbf{x} \right),\left(\boldsymbol{\Pi},\mathbf{p}\right)\right)$. With this notation, the expression in body coordinates of the momentum map $\mathbf{J}: SE (3) \times \mathfrak{se}(3)^\ast \longrightarrow \mathbb{R}^2 $ in~(\ref{canonical momentum map}) is given by \begin{equation} \label{momentum map in body coordinates} \boldsymbol{J} \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi} ,\mathbf{p} \right) \right) = \left( \langle A \boldsymbol{\Pi} + \mathbf{x} \times A \mathbf{p} , \mathbf{e}_3 \rangle , -\langle \boldsymbol{\Pi} , \mathbf{e}_3 \rangle \right). \end{equation} Indeed, for any $(\xi_1, \xi_2)\in \mathbb{R}^2 $, \begin{align*} &\left\langle\boldsymbol{J} \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi} ,\mathbf{p} \right) \right) , \left( \xi _1 , \xi _2 \right) \right\rangle = \left\langle T^*_{\left(A, \mathbf{x} \right) } L_{\left(A, \mathbf{x} \right) ^{-1} } \left(\boldsymbol{\Pi} ,\mathbf{p} \right) , T_{\left( I, 0 \right) } L_{\left(A, \mathbf{x} \right) } \left( (\widehat{A^{-1} \xi_1 \mathbf{e} _3 -\xi_2 \mathbf{e} _3 }), A^{-1} (\xi_1 \mathbf{e} _3 \times \mathbf{x} )\right) \right\rangle \\ &= \langle \boldsymbol{\Pi} , A^{-1}\xi_1 \mathbf{e} _3 -\xi_2 \mathbf{e} _3 \rangle + \langle \mathbf{p} , A^{-1}(\xi_1 \mathbf{e} _3 \times \mathbf{x}) \rangle = \langle \boldsymbol{\Pi} , A^{-1}\xi_1 \mathbf{e}_3 -\xi_2 \mathbf{e}_3 \rangle + \langle \mathbf{p} , A^{-1}(\xi_1 \mathbf{e}_3 \times \mathbf{x}) \rangle, \end{align*} which proves~(\ref{momentum map in body coordinates}) since $(\xi_1, \xi_2)\in \mathbb{R}^2 $ is arbitrary. \subsection{Proof of Proposition~\ref{prop_rel_eq}} \label{proof proposition relative} \medskip \noindent {\bf (i)} Using the statement preceeding~(\ref{relative critical points}) we will specify the relative equilibria of the orbitron by characterizing the points $\mathbf{z}=\left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi},\mathbf{p}\right)\right) \in T^*SE(3)$ for which \begin{equation} \mathbf{d} (h - \boldsymbol{J} ^{(\xi_1,\xi_2)}) \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi},\mathbf{p}\right) \right) =0 \end{equation} for some $(\xi_1,\xi_2) \in \mathbb{R}^2$. We start by computing the tangent of the momentum map and the differential of the Hamiltonian. Let $\mathbf{v}=\left( ( \widehat{\delta A } A , \boldsymbol{\delta}\mathbf{x} ) , ( \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} )\right) \in T_{\mathbf{z}} \left( T^*SE(3) \right) $ be an arbitrary vector at the point $\mathbf{z}$, then it is easy to check that \begin{align} \mathbf{d} T \left(\boldsymbol{\Pi},\mathbf{p}\right) \cdot \mathbf{v} &= \langle \boldsymbol{\Pi}, \mathbb{I}_{ref}^{-1} \boldsymbol{\delta}\boldsymbol{\Pi} \rangle + \frac{1}{M} \langle \mathbf{p}, \boldsymbol{\delta}\mathbf{p} \rangle,\label{diff t}\\ \mathbf{d} V \left(A, \mathbf{x} \right) \cdot \mathbf{v} &= -\mu \left[ \langle D\mathbf{B}(\mathbf{x})(\boldsymbol{\delta}\mathbf{x}), A \mathbf{e}_3 \rangle + \langle \mathbf{B} (\mathbf{x} ), {\delta A } \times A \mathbf{e}_3 \rangle \right].\label{diff v} \end{align} with $T$ and $V$ the kinetic and potential energies introduced in~(\ref{potential and kinetic}). Additionally, \begin{align} &T_{ \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi} ,\mathbf{p} \right) \right)} \boldsymbol{J} \cdot \left( (\widehat{\delta A } A , \boldsymbol{\delta}\mathbf{x} ) , (\boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} ) \right) \nonumber \\ &=\Big( \langle \delta A \times A\boldsymbol{\Pi} +A \boldsymbol{\delta \Pi} + \boldsymbol{\delta} \mathbf{x} \times A\mathbf{p} + \mathbf{x} \times ( \delta A \times A \mathbf{p} ) + \mathbf{x} \times A \boldsymbol{\delta }\mathbf{p} , \mathbf{e}_3 \rangle,- \langle \boldsymbol{\delta}\boldsymbol{\Pi} , \mathbf{e}_3 \rangle \Big).\label{differential momentum map} \end{align} Consequently, using~(\ref{diff t}),~(\ref{diff v}) and \eqref{differential momentum map} we have, for any $(\xi_1, \xi_2)\in \mathbb{R}^2$ \begin{multline} \label{critical 1} \mathbf{d} \left(h - \boldsymbol{J} ^{(\xi_1,\xi_2)}\right) (\mathbf{z}) \cdot \mathbf{v} = \boldsymbol{\Pi} ^T\mathbb{I}_{ref}^{-1} \boldsymbol{\delta}\boldsymbol{\Pi} + \frac{1}{M} \mathbf{p} \cdot \boldsymbol{\delta}\mathbf{p} -\mu \left[ \langle D\mathbf{B}(\mathbf{x})(\boldsymbol{\delta}\mathbf{x}), A \mathbf{e}_3 \rangle + \langle \mathbf{B} (\mathbf{x} ), {\delta A } \times A \mathbf{e}_3 \rangle \right] \\ + \xi _2 \boldsymbol{\delta}\boldsymbol{\Pi} \cdot \mathbf{e}_3 - \xi _1 \left( \delta A \times A\boldsymbol{\Pi}+A \boldsymbol{\delta}\boldsymbol{\Pi} + \boldsymbol{\delta}\mathbf{x} \times A\mathbf{p} +\mathbf{x} \times (\delta A \times A \mathbf{p} ) + \mathbf{x} \times A\boldsymbol{\delta}\mathbf{p} \right) \cdot \mathbf{e}_3. \end{multline} Therefore, as $\widehat{\delta A }$, $ \boldsymbol{\delta}\mathbf{x}$, $\boldsymbol{\delta}\boldsymbol{\Pi}$, and $\boldsymbol{\delta}\mathbf{p}$ in this expression are arbitrary, it can be checked that the points $\mathbf{z} \in T^*SE(3)$ for which $ \mathbf{d} (h - \boldsymbol{J} ^{(\xi_1,\xi_2)}) \left(\mathbf{z} \right) =0 $ are characterized by the equations: \begin{align} \label{equationderivative1} &\mu \left[ \mathbf{B} (\mathbf{x} ) \times A \mathbf{e}_3 \right] + \xi _1 \left[ A \mathbf{p} \times (\mathbf{x} \times \mathbf{e}_3 ) - A \boldsymbol{\Pi} \times \mathbf{e}_3 \right] = 0, \\ \label{equationderivative2} &- \mu D\mathbf{B} (\mathbf{x} ) ^T (A \mathbf{e}_3) - \xi _1 \left( A \mathbf{p} \times \mathbf{e}_3 \right)= 0,\\ \label{equationderivative3} &\mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \xi _2 \mathbf{e}_3 - \xi _1 A^{-1} \mathbf{e}_3 =0,\\ \label{equationderivative4} & \frac{1}{M} \mathbf{p} - \xi _1 A^{-1} \left( \mathbf{e}_3 \times \mathbf{x} \right) = 0, \end{align} as required. \medskip \noindent {\bf (ii)} We show that the points $\mathbf{z} _0 = \left( \left(A_0, \mathbf{x}_0 \right) , \left(\boldsymbol{\Pi}_0,\mathbf{p}_0\right) \right) $ of the form specified in the statement of the proposition satisfy equations \eqref{equationderivative1}--\eqref{equationderivative4} and hence constitute a branch of relative equilibria. We proceed by considering $A_0=R_{\theta_0}^Z$ and $\mathbf{x} _0= \left( x, y, 0\right) $ and using equations \eqref{equationderivative1}--\eqref{equationderivative4} to determine $\boldsymbol{\Pi}_0 $, $\mathbf{p}_0 $, and the velocity $ \boldsymbol{\xi } = \left( \xi _1 , \xi _2 \right) $ in the statement. Notice first that $A_0 \mathbf{e}_3 = \mathbf{e}_3 $, hence by \eqref{equationderivative4} we have that \begin{equation} \label{p0regular} \mathbf{p} _0=M \xi _1 A_0 ^{-1}\left( -y,x, 0\right), \end{equation} necessarily. Now by \eqref{equationderivative3} \\ \begin{equation} \boldsymbol{\Pi} _0 = \mathbb{I}_{ref}\left( \xi _1 - \xi _2 \right) \mathbf{e}_3 = I_3 \left( \xi _1 - \xi _2 \right) \mathbf{e}_3. \end{equation} In order to handle \eqref{equationderivative2} we note that $D\mathbf{B}(\mathbf{x})$ is given by the matrix whose components are \begin{align*} & \dfrac{\partial{B_x}}{\partial{x}} =k \left( \dfrac{D(\mathbf{x} )_+ - 3x^2}{D(\mathbf{x} )_+ ^{5/2}} - \dfrac{D(\mathbf{x} )_- - 3x^2}{D( \mathbf{x} )_- ^{5/2}} \right), \\ &\dfrac{\partial{B_x}}{\partial{y}} =k \left( \dfrac{- 3xy}{D(\mathbf{x} )_+ ^{5/2}} + \dfrac{3xy}{D(\mathbf{x} )_- ^{5/2}} \right), \\ &\dfrac{\partial{B_x}}{\partial{z}} = k \left( \dfrac{- 3x(z-h)}{D(\mathbf{x} )_+ ^{5/2}} + \dfrac{3x(z+h)}{D(\mathbf{x} )_- ^{5/2}} \right), \\ &\dfrac{\partial{B_y}}{\partial{x}} =k \left( \dfrac{- 3xy}{D(\mathbf{x} )_+ ^{5/2}} + \dfrac{3xy}{D( \mathbf{x} )_-^{5/2}}\right), \\ &\dfrac{\partial{B_y}}{\partial{y}} = k \left( \dfrac{D(\mathbf{x} )_+ - 3y^2}{D(\mathbf{x} )_+ ^{5/2}} - \dfrac{D( \mathbf{x} )_- - 3y^2}{D( \mathbf{x} )_- ^{5/2}}\right), \\ &\dfrac{\partial{B_y}}{\partial{z}} = k \left( \dfrac{- 3y(z-h)}{D(\mathbf{x} )_+^{5/2}} + \dfrac{3y(z+h)}{D( \mathbf{x} )_- ^{5/2}}\right), \end{align*} \begin{align*} &\dfrac{\partial{B_z}}{\partial{x}} = k \left( \dfrac{- 3x(z-h)}{D(\mathbf{x} )_+^{5/2}} + \dfrac{3x(z+h)}{D( \mathbf{x} )_- ^{5/2}}\right), \\ &\dfrac{\partial{B_z}}{\partial{y}} = k \left( \dfrac{- 3y(z-h)}{D(\mathbf{x} )_+ ^{5/2}} + \dfrac{3y(z+h)}{D( \mathbf{x} )_- ^{5/2}}\right), \\ & \dfrac{\partial{B_z}}{\partial{z}} = k \left( \dfrac{D(\mathbf{x} )_+ - 3(z-h)^2}{D(\mathbf{x} )_+^{5/2}} - \dfrac{D( \mathbf{x} )_- - 3(z+h)^2}{D( \mathbf{x} )_-^{5/2}}\right), \end{align*} where $D(\mathbf{x} )_+=x ^2 + y ^2 + (z-h) ^2$, $D(\mathbf{x} )_-=x ^2 + y ^2 + (z+h) ^2$ and $k = \dfrac{\mu _0 q}{4 \pi }$. Consequently, \begin{align*} D\mathbf{B}({\bf x}_0)=k \left( \begin{array}{ccc} 0 & 0 & \frac{6xh}{D(\mathbf{x} _0)^{5/2}} \\ 0 & 0 & \frac{6yh}{D(\mathbf{x} _0)^{5/2}}\\ \frac{6xh}{D(\mathbf{x} _0)^{5/2}} & \frac{6yh}{D(\mathbf{x} _0)^{5/2}} & 0 \end{array} \right), \end{align*} where $D(\mathbf{x} _0)=D(\mathbf{x} _0)_+ = D (\mathbf{x} _0)_-$. Hence \begin{equation} \label{DBx0regular} D\mathbf{B}\left( \mathbf{x}_0 \right)^T \left( A_0 \mathbf{e}_3 \right)= D\mathbf{B}\left( \mathbf{x}_0 \right)^T \mathbf{e}_3 = \dfrac{6kh}{D(\mathbf{x}_0)^{5/2}} \mathbf{x}_0. \end{equation} Note additionally that by \eqref{p0regular} \begin{equation} \label{p0timesAe3} A_0\mathbf{p}_0 \times \mathbf{e}_3 = M \xi _1 \mathbf{x}_0. \end{equation} Then by equalities \eqref{DBx0regular} and by \eqref{p0timesAe3}, equation \eqref{equationderivative2} holds whenever $\mathbf{x}_0= \boldsymbol{0} $ or when $\mathbf{x}_0 \neq \boldsymbol{0} $ and $\xi _1 ^2 = - \dfrac{3h\mu q \mu _0 }{2 \pi MD(\mathbf{x}_0 )^{5/2}} $; we note that in both situations, there are no restrictions on the second component of the velocity $\xi_2$. Finally, it can be readily verified that \eqref{equationderivative1} always holds at the point $\left( \left(A_0, \mathbf{x}_0 \right) , \left(\boldsymbol{\Pi}_0,\mathbf{p}_0\right) \right) $ by using that $\mathbf{B}(\mathbf{x} _0)=-\dfrac{\mu_0qh}{2\pi D(\mathbf{x} _0)^{3/2}} \mathbf{e}_3$. \medskip \noindent {\bf (iii)} Suppose that we are in the presence of a magnetic field $\mathbf{B}$ equivariant with respect to rotations around the $OZ$ axis and that behaves as indicated in \eqref{Bx_transformation}--\eqref{Bz_transformation} with respect to the mirror transformation~(\ref{mirror transformation}). Notice first that by \eqref{Bx_transformation} and \eqref{By_transformation} \begin{equation}\label{Bxy_zero} B _x (x,y,0) = B _y (x,y,0) \end{equation} and hence \begin{equation}\label{Bz_e3} \mathbf{B}(x,y,0)=B _z (x,y,0) \mathbf{e}_3. \end{equation} Additionally, by \eqref{mag_equivariant}, $B_z (x,y,0)$ is rotationally invariant with respect to rotations in the $OXY$ plane, hence \begin{equation}\label{Bz_func_sqr_xy} B _z (x,y,0) = f(x ^2 + y ^2 ), \enspace {\rm for \enspace some} \enspace f \in C^\infty(\mathbb{R} ^2 ). \end{equation} Conditions \eqref{equationderivative4s} and \eqref{equationderivative3s} show that if $A _0 = R_{\theta _0 } ^z $ and $\mathbf{x} _0 = (x,y,0)$, then $\boldsymbol{\Pi} _0 = I _3 \left( \xi _1 - \xi _2 \right) \mathbf{e}_3 $ and $\mathbf{p} _0 = M \xi _1 A _0 ^{-1} \left(-y,x,0\right) $ necessarily. If we use $\mathbf{z} _0 = \left( (A _0 , \mathbf{x} _0 ),(\boldsymbol{\Pi} _0 , \mathbf{p} _0 )\right) $ and \eqref{Bz_e3} in the expression \eqref{equationderivative1s}, it can be easily verified that this relation is automatically satisfied. In order to study the expression \eqref{equationderivative2s}, we take derivatives on both sides of \eqref{Bz_transformation} and obtain that \begin{equation*} \partial _z B _z (x,y,z) = - \partial _z B _z (x,y,-z) \end{equation*} which shows that \begin{equation}\label{deriv_Bz_zero} \partial _z B _z (x,y,0) = 0. \end{equation} Finally, by \eqref{Bz_e3} and \eqref{deriv_Bz_zero} the relation \eqref{equationderivative2s} amounts to \begin{equation*} - \mu (\partial _x B _z , \partial _y B _z, 0 )=M \xi _1 ^2 \mathbf{x} _0. \end{equation*} By \eqref{Bz_func_sqr_xy} this is equivalent to \begin{equation*} -2 \mu f'(x ^2 + y ^2 ) \mathbf{x} _0 = M \xi _1 ^2 \mathbf{x} _0, \end{equation*} which guarantees that \eqref{equationderivative2s} is satisfied provided that \begin{equation} \xi _1 = \pm \left( -\dfrac{2}{M} \mu f'( x ^2 + y ^2 )\right) ^{1/2}, \end{equation} as required. \quad $\blacksquare$ \subsection{Proof of Theorem~\ref{kozorez relations}} \label{proof kozorez} We will proceed by using Theorem~\ref{energy momentum statement} in order to determine the regions in parameter space for which the stability form~(\ref{quadratic form stability}) at the relative equilibria is definite, which in turn ensures $\mathbb{T}^2 $--stability. We start by denoting the augmented Hamiltonian as $h ^{\boldsymbol{\xi}} :=h- \boldsymbol{J}^{\boldsymbol{\xi}} $, for any $\boldsymbol{\xi}= \left( \xi _1 , \xi _2 \right)\in {\rm Lie} \left(\mathbb{T}^2\right)$. Let $\mathbf{z}= \left( \left(A, \mathbf{x} \right) , \left(\boldsymbol{\Pi},\mathbf{p}\right) \right) \in T ^\ast \left(SE (3)\right)$ expressed in body coordinates. As we saw in the proof of Proposition~\ref{prop_rel_eq} (see Appendix~\ref{proof proposition relative}), the partial derivatives of $h ^{\boldsymbol{\xi}} $ are given by: \begin{itemize} \item $ h_A^ {\boldsymbol{\xi}}:=D_A h ^{\boldsymbol{\xi}} (\mathbf{z}) = \mu \left[ \mathbf{B} (\mathbf{x} ) \times A \mathbf{e}_3 \right] + \xi _1 \left[ A \mathbf{p} \times (\mathbf{x} \times \mathbf{e}_3 ) - A \boldsymbol{\Pi} \times \mathbf{e}_3 \right] $, \item $h_{\mathbf{x}} ^{\boldsymbol{\xi}}:=D_{\mathbf{x} } h ^ {\boldsymbol{\xi}} (\mathbf{z}) = - \mu D\mathbf{B} (\mathbf{x} ) ^T (A \mathbf{e}_3) - \xi _1 \left( A \mathbf{p} \times \mathbf{e}_3 \right) $, \item $h _{\boldsymbol{\Pi} } ^{\boldsymbol{\xi}}:=D_{\boldsymbol{\Pi} } h ^{\boldsymbol{\xi}} (\mathbf{z}) = \mathbb{I}_{ref}^{-1} \boldsymbol{\Pi} + \xi _2 \mathbf{e}_3 - \xi _1 A^{-1} \mathbf{e}_3$, \item $h_{\mathbf{p} }^ {\boldsymbol{\xi}} :=D_{\mathbf{p} } h ^ {\boldsymbol{\xi}} (\mathbf{z}) = \frac{1}{M} \mathbf{p} - \xi _1 A^{-1} \left( \mathbf{e}_3 \times \mathbf{x} \right)$. \end{itemize} In order to compute the Hessian of the augmented Hamiltonian, we write down the derivatives of its partial derivatives in the direction given by the vector $\mathbf{v}=\left. \dfrac{d}{dt} \right|_0 \left( ( \exp t \widehat{\delta A } A , \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} ) , ( \boldsymbol{\Pi} + t \boldsymbol{\delta}\boldsymbol{\Pi} , \mathbf{p} + t \boldsymbol{\delta}\mathbf{p} )\right)$. A straightforward computation yields: \begin{itemize} \item $\mathbf{d} h_A ^ {\boldsymbol{\xi}} ( \mathbf{z}) \cdot\mathbf{v} = \mu \left[ (D\mathbf{B}(\mathbf{x} ) \boldsymbol{\delta}\mathbf{x}) \times A \mathbf{e}_3 + \mathbf{B} (\mathbf{x} ) \times ( \widehat{\delta A } A \mathbf{e}_3 )\right] + \xi _1 \Big[ ( \widehat{\delta A } A \mathbf{p} + A \boldsymbol{\delta}\mathbf{p})\times ( \mathbf{x} \times \mathbf{e}_3 ) + A \mathbf{p} \times (\boldsymbol{\delta}\mathbf{x} \times \mathbf{e}_3 ) - \widehat{\delta A } A \boldsymbol{\Pi} \times \mathbf{e}_3 - ( A \boldsymbol{\delta}\boldsymbol{\Pi} \times \mathbf{e}_3 ) \Big]$, \item $\mathbf{d} h_{\mathbf{x} } ^{\boldsymbol{\xi}} ( \mathbf{z} ) \cdot \mathbf{v} = -\mu \left( T_{\mathbf{x} } \mathbf{F} ( \boldsymbol{\delta}\mathbf{x} )\right) (A \mathbf{e}_3 )-\mu \mathbf{F} (\mathbf{x} ) \left( \delta A \times A \mathbf{e}_3\right) - \xi _1 ( \widehat{ \delta A } A \mathbf{p} \times \mathbf{e}_3 + A \boldsymbol{\delta}\mathbf{p} \times \mathbf{e}_3 )$, where \begin{align*} \mathbf{F} :\ & \mathbb{R}^3 \longrightarrow M_{3 \times 3}\\ & \mathbf{x} \longmapsto D\mathbf{B}(\mathbf{x} )^T, \end{align*} \item $\mathbf{d} h_{\boldsymbol{\Pi} } ^ {\boldsymbol{\xi}} (\mathbf{z}) \cdot \mathbf{v} = \mathbb{I}_{ref}^{-1} \boldsymbol{\delta}\boldsymbol{\Pi} + \xi _1 A^T \widehat{ \delta A} \mathbf{e}_3$, \item $\mathbf{d} h_{\mathbf{p} } ^ {\boldsymbol{\xi}} ( \mathbf{z}) \cdot \mathbf{v} = \dfrac{\boldsymbol{\delta}\mathbf{p} }{M} - \xi _1 A^T\widehat{ \delta A } ( \mathbf{x} \times \mathbf{e}_3)+ \xi _1 A^T (\boldsymbol{\delta}\mathbf{x} \times \mathbf{e}_3 ) $. \end{itemize} Consequently, the matrix expression associated to $\mathbf{d}^2 \left(h- \mathbf{J}^{\boldsymbol{\xi}}\right) (\mathbf{z}) $ is given by: \begin{equation} \label{hessian in general} \left( \begin{array}{cccc} - \mu \left[ \widehat{ \mathbf{B} (\mathbf{x} ) } \widehat{ A \mathbf{e}_3 }\right] + \xi _1 \Big[ \widehat{ \mathbf{x} \times \mathbf{e}_3 } \widehat{ A \mathbf{p}} - \widehat{ \mathbf{e}_3 } \widehat{ A \boldsymbol{\Pi} }\Big] & -\mu \widehat{ A \mathbf{e}_3 } \mathbf{F}(\mathbf{x} )^T - \xi _1 \widehat{ A \mathbf{p} } \widehat{ \mathbf{e}_3 } & \xi _1 \widehat{\mathbf{e}_3 }A& -\xi _1 \widehat{ \mathbf{x} \times \mathbf{e}_3 } A\\ \mu \left[ \mathbf{F} (\mathbf{x} ) \widehat{ A \mathbf{e}_3 }\right] - \xi _1 \widehat{ \mathbf{e}_3 } \widehat{ A \mathbf{p} } & -\mu {T} _{\mathbf{x} } \mathbf{F} (\cdot) (A \mathbf{e}_3 )& 0& \xi _1 \widehat{\mathbf{e}_3 }A\\ - \xi _1 A^T\widehat{ \mathbf{e}_3 } & 0 & \mathbb{I}_{ref}^{-1} & 0\\ \xi _1 A^T \widehat{ \mathbf{x} \times \mathbf{e}_3 } & - \xi _1 A^T \widehat{\mathbf{e}_3 } & 0& \dfrac{1}{M} \mathbb{I}_{id} \end{array} \right) \end{equation} We now compute the value of the Hessian~(\ref{hessian in general}) at the relative equilibria in the second and third parts of Proposition~\ref{prop_rel_eq}, that is, $\mathbf{z}_0= \left( \left(A_0, \mathbf{x}_0 \right) , \left(\boldsymbol{\Pi}_0,\mathbf{p}_0\right) \right) $ with $A_0=R_{\theta}^Z$, $\mathbf{x}_0= \left( x, y, 0\right)^T$, $\boldsymbol{\Pi}_0=I_3 \left( \xi _1^0 - \xi _2\right) \mathbf{e}_3 $, and $\mathbf{p}_0=M \xi _1 ^0A _0^{-1} \left( -y, x, 0\right)^T $. We start by noticing that \begin{equation}\label{TxF_is_Hess} T _{\mathbf{x} } \mathbf{F} (\cdot)(\mathbf{e}_3 ) = {\rm Hess} \left( {B_z} \right)(\mathbf{x} ). \end{equation} Indeed, for any $\boldsymbol{\delta}\mathbf{x} \in T_{\mathbf{x}}\mathbb{R}^3 $ \begin{multline}\label{TxF_is_Hess_proof} T _{\mathbf{x} }\mathbf{F} (\boldsymbol{\delta}\mathbf{x} )(\mathbf{e}_3 )= \left. \dfrac{d}{dt} \right|_0 \mathbf{F} \left( \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} \right) (\mathbf{e}_3 ) = \left. \dfrac{d}{dt} \right|_0 D\mathbf{B} \left( \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} \right)^T \mathbf{e}_3 \\ = \left. \dfrac{d}{dt} \right|_0 \left( \begin{array}{ccc} \dfrac{\partial{B_z}}{\partial{x}} \left( \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} \right)\\ \\\dfrac{\partial{B_y}}{\partial{z}}\left( \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} \right)\\\\ \dfrac{\partial{B_z}}{\partial{z}} \left( \mathbf{x} + t \boldsymbol{\delta}\mathbf{x} \right) \end{array} \right) = {\rm Hess} \left( {B_z} \right)(\mathbf{x} ) \cdot \boldsymbol{\delta}\mathbf{x}. \end{multline} Therefore, the matrix expression associated to $\mathbf{d}^2 \left(h- \mathbf{J}^{\boldsymbol{\xi}}\right) (\mathbf{z}_0) $ is given by: \begin{equation} \label{hessian in particular} \left( \begin{array}{cccc} - \mu \left[ \widehat{ \mathbf{B} (\mathbf{x}_0 ) } \widehat{\mathbf{e}_3 }\right] + \xi _1 \Big[ \widehat{ \mathbf{x} _0 \times \mathbf{e}_3 } \widehat{ \mathbf{p} _0 } - \widehat{\mathbf{e}_3} \widehat{ \boldsymbol{\Pi} _0 }\Big] & -\mu\widehat{\mathbf{e}_3 } \mathbf{F}(\mathbf{x}_0 )^T - \xi _1 \widehat{ \mathbf{p}_0 } \widehat{ \mathbf{e}_3} & \xi _1 \widehat{ \mathbf{e}_3 } & -\xi _1 (\widehat{ \mathbf{x}_0 \times \mathbf{e}_3})\\ \mu\mathbf{F} (\mathbf{x}_0 ) \widehat{ \mathbf{e}_3 }- \xi _1 \widehat{ \mathbf{e}_3} \widehat{ \mathbf{p}_0} &-\mu {\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 ) &0 &\xi _1 \widehat{\mathbf{e}_3 }\\ - \xi _1 \widehat{\mathbf{e}_3 } &0 &\mathbb{I}_{ref}^{-1} &0\\ \xi _1( \widehat{ \mathbf{x} _0 \times \mathbf{e}_3 }) &- \xi _1 \widehat{ \mathbf{e}_3 } &0 &\dfrac{1}{M} \mathbb{I}_{id} \end{array} \right). \end{equation} In order to construct stability forms for the regular and singular branches, we now determine stability spaces $W$ to which we will restrict the Hessian~(\ref{hessian in particular}). \medskip \noindent {\bf A stability space for the regular branch ($r>0 $).} In this case, the kernel of the derivative of the momentum map is given by: \begin{align*} \ker T_{{\bf z}_0} \boldsymbol{J}& = \left\{ v= \left( ( \widehat{\delta A } , \boldsymbol{\delta}\mathbf{x} ) , \left( \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right) \right) \in T_{\mathbf{z}_0} \left( SE(3) \times \mathfrak{se}(3)^*\right) \mid T_{\mathbf{z}_0 } \boldsymbol{J} \cdot v = 0 \right\}\\ &= \left\{ v \in T_{\mathbf{z}_0 } \left( SE(3) \times \mathfrak{se}(3)^*\right) \mid \boldsymbol{\delta}\boldsymbol{\Pi} \cdot \mathbf{e}_3 =0, \delta p _2 =- M \xi _1 ^0 \delta x_1 \right\}, \end{align*} and using~(\ref{infinitgenBody}), the tangent space $\mathfrak{t}^2 \cdot {\bf z}_0:=T_{{\bf z}_0} \left(\mathbb{T}^2\cdot {\bf z}_0\right)$ to the toral orbit that goes through the relative equilibrium ${\bf z}_0 $ can be characterized as: \begin{align*} \mathfrak{t}^2 \cdot {\bf z}_0&= \left\{ \left( \xi _1 , \xi _2 \right)_{T^*SE(3)}(\mathbf{z}_0 )\mid \xi_1, \xi_2 \in \mathbb{R} \right\}= \left\{ \left( \widehat{\left( \xi _1 \mathbf{e}_3 - \xi _2 \mathbf{e}_3 \right)}, \xi _1 \mathbf{e}_3 \times \mathbf{x}_0, \mathbf{0} , \xi _2 \mathbf{e}_3 \times \mathbf{p}_0 \right) \mid \xi _1 , \xi _2 \in \mathbb{R}^2\right\}\\ &= \left\{ \left( \left( \xi _1 - \xi _2 \right) \mathbf{e}_3 , \xi _1 r \mathbf{e}_2, \mathbf{0} , - \xi _2 M r\xi _1 ^0 \mathbf{e}_1 \right) \mid \xi _1 , \xi _2 \in \mathbb{R}^2 \right\}. \end{align*} Finally, it can be easily verified that the vector subspace $W\subset \ker T_{{\bf z}_0} \boldsymbol{J}$ \begin{align*} &W:= \Big\{ \left( \delta A , \boldsymbol{\delta}\mathbf{x} , \boldsymbol{\delta}\boldsymbol{\Pi} , \boldsymbol{\delta}\mathbf{p} \right)\mid \boldsymbol{\delta}\boldsymbol{\Pi} \cdot \mathbf{e}_3 =0, \boldsymbol{\delta}\mathbf{x} \cdot \mathbf{p} _0 =0, \boldsymbol{\delta}\mathbf{p} \cdot \mathbf{x} _0 =0, \delta p _2 =- M \xi _1 ^0 \delta x_1 \Big\} \end{align*} is such that \begin{equation} {\rm Ker} T_{{\bf z}_0} \boldsymbol{J} = W \oplus \mathfrak{t}^2 \cdot {\bf z}_0, \end{equation} and hence constitutes a stability space. Moreover, let $\mathbf{u}_1 = \left( \mathbf{0} , \mathbf{e}_1, \mathbf{0} , -M \xi _1 ^0 \mathbf{e}_2 \right)$, $\mathbf{u}_2 = \left( \mathbf{e}_3 ,\mathbf{0} , \mathbf{0} , \mathbf{0} \right)$, $\mathbf{u}_3 =\left( \mathbf{0} , \mathbf{0} , \mathbf{e}_1, \mathbf{0} \right)$, $\mathbf{u}_4 = \left( \mathbf{0} , \mathbf{0} , \mathbf{e}_2, \mathbf{0} \right)$, $\mathbf{u}_5= \left( \mathbf{0} , \mathbf{0} , \mathbf{0} , \mathbf{e}_3 \right)$, $\mathbf{u}_6= \left( \mathbf{0} , \mathbf{e}_3 , \mathbf{0} , \mathbf{0} \right)$, $\mathbf{u}_7=\left( \mathbf{e}_2, \mathbf{0} , \mathbf{0} , \mathbf{0} \right)$, and $\mathbf{u}_8= \left( \mathbf{e}_1, \mathbf{0} , \mathbf{0} , \mathbf{0} \right)$. It can be checked that \begin{align} W={\rm span} \Big\{\mathbf{u}_1,\mathbf{u}_2,\mathbf{u}_3,\mathbf{u}_4,\mathbf{u}_5,\mathbf{u}_6,\mathbf{u}_7,\mathbf{u}_8 \Big\}.\label{basis regular branch} \end{align} The set $\mathcal{B}= \Big\{\mathbf{u}_1,\mathbf{u}_2,\mathbf{u}_3,\mathbf{u}_4,\mathbf{u}_5,\mathbf{u}_6,\mathbf{u}_7,\mathbf{u}_8 \Big\} $ will be used as a basis of the stability space in order to obtain matrix expressions for the stability form $\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1^0 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W} $ corresponding to each part of Theorem~\ref{kozorez relations}. \medskip \noindent {\bf A stability space for the singular branch ($r=0 $).} Consider now the relative equlibrium ${z} _0 = \left( \left(A_0, \mathbf{x}_0 \right) , \left(\boldsymbol{\Pi}_0,\mathbf{p}_0\right) \right) $ with $A_0=R_{\theta_0}^Z$, $\mathbf{x} _0= \left( 0, 0, 0\right) $, $\boldsymbol{\Pi} _0 = I_3 \left( \xi _1 - \xi _2 \right) \mathbf{e}_3 $, and $\mathbf{p} _0={\bf 0} $, $\xi_1, \xi_2 \in \mathbb{R} $. In this case, the matrix expression~(\ref{hessian in general}) associated to $\mathbf{d}^2 \left(h- \mathbf{J}^{\boldsymbol{\xi}}\right) (\mathbf{z}) $ is given by: \begin{equation}\label{Hessian_singular} \left( \begin{array}{cccc} - \mu \widehat{ \mathbf{B} (\mathbf{x}_0 ) } \widehat{\mathbf{e}_3 } - \xi _1 \widehat{\mathbf{e}_3} \widehat{ \boldsymbol{\Pi_0} } & 0 & \xi _1 \widehat{ \mathbf{e}_3 }& 0\\ 0&-\mu {\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 ) &0&\xi _1 \widehat{\mathbf{e}_3 }\\ - \xi _1 \widehat{\mathbf{e}_3 }&0&\mathbb{I}_{ref} ^{-1}&0\\ 0 &- \xi _1 \widehat{ \mathbf{e}_3 }&0&\dfrac{1}{M} \mathbb{I}_{id} \end{array} \right). \end{equation} These relative equilibria lay on the singular isotropy type manifold~(\ref{singular orbit type}) and hence by the Bifurcation Lemma (see~\cite[Proposition 4.5.12]{Ortega2004}), the kernel of the derivative of the momentum map is necessarily of dimension eleven at those points. Indeed, it can be checked that: \begin{equation*} \ker T_{{\bf z}_0} \boldsymbol{J}= \left\{ v \in T_{\mathbf{z}_0 } \left( SE(3) \times \mathfrak{se}(3)^*\right) \mid \boldsymbol{\delta}\boldsymbol{\Pi} \cdot \mathbf{e}_3 =0 \right\}, \end{equation*} and using~(\ref{infinitgenBody}), the tangent space $\mathfrak{t}^2 \cdot {\bf z}_0:=T_{{\bf z}_0} \left(\mathbb{T}^2\cdot {\bf z}_0\right)$ to the toral orbit that goes through the singular relative equilibrium ${\bf z}_0 $ can be characterized as: \begin{equation*} \mathfrak{t}^2 \cdot {\bf z}_0= \left\{ \left( \left( \xi _1 - \xi _2 \right) \mathbf{e}_3 , {\bf 0}, \mathbf{0} , {\bf 0} \right) \mid \xi _1 , \xi _2 \in \mathbb{R}^2 \right\}= {\rm span} \{ \left(\mathbf{e}_3, {\bf 0}, {\bf 0}, {\bf 0}\right)\}. \end{equation*} Finally, it can be easily verified that the vector subspace $W\subset \ker T_{{\bf z}_0} \boldsymbol{J}$ given by \begin{equation} \label{stability space singular} W={\rm span} \Big\{\mathbf{u}_1,\mathbf{u}_2,\mathbf{u}_3,\mathbf{u}_4,\mathbf{u}_5,\mathbf{u}_6,\mathbf{u}_7,\mathbf{u}_8, \mathbf{u}_9, \mathbf{u}_{10} \Big\}, \end{equation} with $\mathbf{u}_1= \left( \mathbf{0} ,{\bf 0}, \mathbf{0} , \mathbf{e}_3 \right)$, $\mathbf{u}_2= \left( {\bf 0} ,\mathbf{e}_3, \mathbf{0} , \mathbf{0} \right)$, $\mathbf{u}_3 =\left( \mathbf{0} , \mathbf{0} , {\bf 0}, \mathbf{e}_2\right)$, $\mathbf{u}_4 = \left( \mathbf{0} , \mathbf{0} , {\bf 0},\mathbf{e}_1 \right)$, $\mathbf{u}_5= \left( \mathbf{0} , \mathbf{0} , \mathbf{e}_2 , {\bf 0} \right)$, $\mathbf{u}_6= \left( \mathbf{0} , {\bf 0} , \mathbf{e}_1 , \mathbf{0} \right)$, $\mathbf{u}_7=\left( {\bf 0}, \mathbf{e}_2 , \mathbf{0} , \mathbf{0} \right)$, $\mathbf{u}_8= \left( {\bf 0}, \mathbf{e}_1 , \mathbf{0} , \mathbf{0} \right)$, $\mathbf{u}_9= \left( \mathbf{e}_2, {\bf 0} , \mathbf{0} , \mathbf{0} \right)$, and $\mathbf{u}_{10}= \left(\mathbf{e}_1 , {\bf 0}, \mathbf{0} , \mathbf{0} \right)$ is a $H$--invariant stability space, that is, \begin{equation} {\rm Ker} T_{{\bf z}_0} \boldsymbol{J} = W \oplus \mathfrak{t}^2 \cdot {\bf z}_0, \end{equation} and hence constitutes a stability space. We recall that $H:=\left\{\left(e^{i\theta},e^{i\theta}\right)\mid e^{i\theta} \in S ^1\right\} = \mathbb{T}^2_{{\bf z}_0}$ is the isotropy subgroup of the relative equilibrium ${\bf z} _0 $. We will use the set $\mathcal{B}= \Big\{\mathbf{u}_1,\mathbf{u}_2,\mathbf{u}_3,\mathbf{u}_4,\mathbf{u}_5,\mathbf{u}_6,\mathbf{u}_7,\mathbf{u}_8,\mathbf{u}_9,\mathbf{u}_{10} \Big\} $ as a basis of the stability space in order to obtain a matrix expression for the stability form~$\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W} $ for the parts {\bf (i)} and {\bf (ii)} of Theorem~\ref{kozorez relations}. \medskip \noindent {\bf Proof of part (i) of the theorem.} \medskip \noindent {\bf Stability study for the regular branch.} We start by noting that the stability of a relative equilibrium can be determined by using any of the the points that constitute its trajectory in phase space. Hence we can, without loss of generality, use the relative equilibrium point ${\bf z} _0 $ of the form ${\bf z} _0 = \left( (\mathbb{I}_{id}, r \mathbf{e}_1), (I _3 \left( \xi _1 ^0 - \xi _2 \right) \mathbf{e}_3 , M r \xi _1 ^0 \mathbf{e_2} )\right)$. We recall that the regular relative equilibria are those for which $r> 0 $ and \begin{equation*} \xi _1^0 = \pm \left( - \dfrac{3h\mu q \mu _0 }{2 \pi MD(\mathbf{x}_0 )^{5/2}}\right)^{1/2}. \end{equation*} We now provide the expression of ${\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 )$ using the same notation as in~(\ref{DBx0regular}) and conclude that \[ {\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 ) = \dfrac{6kh}{D(\mathbf{x}_0 )^{7/2}} \left( \begin{array}{ccc} D(\mathbf{x}_0 )-5x^2& -5xy&0\\ -5xy&D(\mathbf{x}_0 )-5y^2&0\\ 0&0&3D(\mathbf{x}_0 )-5h^2 \end{array} \right). \] By \eqref{hessian in particular} and \eqref{basis regular branch} we obtain that the stability form $\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1^0 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W} $ can be written as: \begin{equation} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \left( \begin {array}{cccccccc} M \xi_1^0 \dfrac{4 h^2-r^2}{r^2+h^2} &0 &0 &0 &0 &0 &0 &0\\ 0 &M {\xi_1^0}^2 r^2 &0 &0 &0 &0 &0 &0\\ 0 &0 &{\dfrac {1}{I_1}} &0 &0 &0 &\xi_1^0 &0\\ 0 &0 &0 &\dfrac{1}{I_1} &0 &0 &0 &- \xi_1^0 \\ 0 &0 &0 &0 &\dfrac{1}{M} &0 &0 &\xi_1^0 r\\ 0 &0 &0 &0 &0 &M{\xi_1^0}^2 \dfrac{3 r ^2 - 2 h ^2 }{r^2+h^2} &M { \xi_1 ^0 }^2 r &0\\ 0 &0 & \xi_1^0 &0 &0 &M{\xi_1^0}^2 r &\dfrac{1}{3}M{\xi_1 ^0 } ^2(r^2+h^2)+\xi _1^0 \Pi _0 &0\\ 0 &0 &0 &- \xi_1^0 & \xi_1^0 r &0 &0 &\dfrac{1}{3}M{\xi_1 ^0 } ^2 (4 r^2+h^2)+\xi _1^0 \Pi _0 \end {array} \right), \end{equation} where $\Pi _0 = I_3 \left( \xi _1 ^0 - \xi _2 \right)$. Notice that this matrix is block diagonal and exhibits two blocks of size two and six. The positivity of the block of size two requires that $M {\xi_1^0}^2 r^2 > 0 $ and $4h ^2 - r ^2 >0$. The first inequality is always satisfied when $\mu q< 0 $ and the second one amounts to \begin{equation} \dfrac{r^2}{h^2} < 4, \end{equation} which yields the right hand side inequality in~(\ref{kozoriez}). We now study the positivity of the lower six dimensional block of the stability form. As we already pointed out in Remark~\ref{why gaussian elimination}, given that by Sylvester's Law of Inertia the signature of a diagonalizable matrix is invariant with respect to conjugation by invertible matrices, it can hence be read out of the pivots of the matrix obtained by performing Gaussian elimination on this block. Indeed, these pivots are \begin{equation} \label{} \dfrac{1}{I _1 }, \enspace \dfrac{1}{I _1 }, \enspace \dfrac{1}{M}, \enspace p _1, \enspace p _2, \enspace p _3, \end{equation} where \begin{eqnarray*} p _1&:=&M{\xi_1^0} ^2 \dfrac{3 r ^2 - 2 h ^2 }{r^2+h^2},\\ p _2&:=&-I_3 \xi_1^0 \xi_2 - {\xi_1^0}^2 \left( \dfrac{2}{3} M \dfrac{ (r^2+h^2) h^2}{3 r^2-2 h^2}+(I_1-I_3) \right), \\ p _3&:=&-I_3 \xi_1^0 \xi_2 - {\xi_1^0}^2 \left( -\dfrac{1}{3} M (r^2+h^2)+(I_1-I_3) \right). \end{eqnarray*} The first three are automatically positive. The positivity of $p _1$ is equivalent to \begin{equation*} \label{radiuscond} {\dfrac{2}{3}}<\dfrac{r^2}{h^2}, \end{equation*} which yields the left hand side inequality in~(\ref{kozoriez}) Finally, we study the positivity of the last two pivots $p _2$ and $p _3$. The simultaneous positivity of $p_2$ and $p_3$ is equivalent to $\min\left\{ p_2,p_3\right\} >0$. It is easy to check that $\min\left\{ p_2,p_3\right\} = p_2$, since the condition \begin{equation} p_3-p_2=M{\xi_1^0}^2 r^2 \dfrac{r^2+h^2}{3r^2-2h^2}>0 \end{equation} is always satisfied due to \eqref{radiuscond} and the condition $\mu q<0$. Regarding the positivity of $p _2$ there are two possible cases:\\ 1) $\xi_1^0 >0$, then \begin{equation*} I_3 (\xi _1^0- \xi _2) >\xi_1^0 \left( I_1+\dfrac{2}{3} M \dfrac{(r^2+h^2)h^2}{3 r^2-2 h^2}\right), \end{equation*} 2) $ \xi_1^0<0$, then \begin{equation*} I_3 (\xi _1^0- \xi _2) <\xi_1^0 \left( I_1+\dfrac{2}{3} M \dfrac{(r^2+h^2)h^2}{3 r^2-2 h^2}\right), \end{equation*} and hence the positivity of $p_2$ can be summarized as \begin{equation*} {\rm sign}(\xi _1^0 ) I_3 \xi _2<-| \xi_1^0 | \left( I_1-I_3+\dfrac{2}{3} M \dfrac{(r^2+h^2)h^2}{3 r^2-2 h^2}\right), \end{equation*} which coincides with~(\ref{signxi2}), as required. \medskip \noindent {\bf Stability study for the singular branch.} We first notice that the matrix expression associated to $\mathbf{d}^2 \left(h- \mathbf{J}^\xi\right) (\mathbf{z}) $ is given by \eqref{Hessian_singular}, where \begin{equation} \mathbf{B}(\mathbf{x}_0) = -\dfrac{\mu_0 q}{2 \pi h ^2} \mathbf{e}_3 \enspace {\rm and} \enspace {\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 ) = -\dfrac{3\mu_0q\mu}{2 \pi h ^4} \left( \begin{array}{ccc} 1&0&0\\ 0&1&0\\ 0&0&-2 \end{array} \right). \end{equation} Consequently, the pivots obtained by Gaussian elimination in the matrix expression of the stability form~$\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W}$ are \begin{equation} \label{singular pivots} \frac{1}{M},\frac{3 \mu _0q \mu}{\pi h ^4}, \frac{1}{M}, \frac{1}{M}, \frac{1}{I _1}, \frac{1}{I _1}, p _1, p _1, p _2, p _2, \end{equation} where \begin{align*} p _1 =- \frac{3}{2}\frac{\mu _0q \mu}{\pi h ^4}-M \xi_1^2,\quad p _2= -\frac{ \mu _0q \mu}{2\pi h ^2}- \xi_1(\xi_1 I _1- \Pi_0), \quad \mbox{and} \quad \Pi_0=I _3(\xi_1- \xi_2). \end{align*} The formal instability of the singular branch is caused by the fact that the pivots in~(\ref{singular pivots}) cannot simultaneously have all the same sign. Indeed, $1/M $ and $1/I _1 $ are always positive which forces $3 \mu _0q \mu/\pi h ^4>0 $. This is in turn incompatible with $p _ 1 > 0$ because that would require $M \xi _1^2<0 $, which is not possible. \medskip \noindent {\bf Proof of part (ii) of the theorem.} \medskip \noindent {\bf Stability study for the regular branch.} In order to prove the second part of Theorem~\ref{kozorez relations}, we follow the same pattern that we used above. Let $f\in C^{\infty}(\mathbb{R} ^2)$ be the function such that $B _z (x,y,z) = f(x ^2 + y ^2,z )$ and $f_0 := f(x ^2+ y ^2, 0)$. Additionally, \begin{equation*} f'_1 := \left. \dfrac{\partial f(v,z)}{\partial v}\right|_{v=x ^2+ y ^2, z= 0}, \enspace f_{1}'' := \left. \dfrac{\partial ^2 f(v,z)}{\partial v ^2}\right|_{v=x ^2+ y ^2, z= 0}, \quad f''_2 := \left. \dfrac{\partial ^2 f(v,z)}{\partial z ^2}\right|_{v=x ^2+ y ^2, z= 0}, \end{equation*} and we recall that $\xi _1 ^0 = \pm{ \left( -\dfrac{2}{M} \mu f'_1\right)}^{1/2}$. We now compute the components of the matrix $D\mathbf{B} (\mathbf{x} _0) $. Using the equations \eqref{Bxy_zero}, \eqref{deriv_Bz_zero} and \eqref{Bz_func_sqr_xy} we obtain \begin{align*} \left. \dfrac{\partial B _x}{\partial x }\right|_{\mathbf{x} _0 } = 0, \ \left. \dfrac{\partial B _x}{\partial y }\right|_{\mathbf{x} _0 } = 0, \ \left. \dfrac{\partial B _y}{\partial x }\right|_{\mathbf{x} _0 }=0, \ \left. \dfrac{\partial B _y}{\partial y }\right|_{\mathbf{x} _0 } = 0, \ \left. \dfrac{\partial B _z}{\partial z }\right|_{\mathbf{x} _0 } = 0, \ \left. \dfrac{\partial B _z}{\partial x }\right|_{\mathbf{x} _0 } = 2x f'_1, \ \left. \dfrac{\partial B _z}{\partial y }\right|_{\mathbf{x} _0 } = 2y f'_1. \end{align*} In order to determine the remaining two components in $D\mathbf{B} (\mathbf{x} _0) $, we use the Amp\`ere-Maxwell equation $\nabla \times \mathbf{B} = 0$ in the absence of additional currents and time-varying electric fields in the region where the body motion takes place. Indeed, $\nabla \times \mathbf{B} = 0$ implies that \begin{equation*} \left. \dfrac{\partial B _x}{\partial z }\right|_{\mathbf{x} _0 }= \left. \dfrac{\partial B _z}{\partial x }\right|_{\mathbf{x} _0 } = 2x f'_1, \enspace \left. \dfrac{\partial B _y}{\partial z }\right|_{\mathbf{x} _0 }= \left. \dfrac{\partial B _z}{\partial y }\right|_{\mathbf{x} _0 } = 2y f'_1, \end{equation*} and hence, \begin{equation*} D\mathbf{B}(\mathbf{x} _0 ) = \left( \begin{array}{ccc} 0 & 0 & 2x f'_1 \\ 0&0&2y f'_1 \\ 2x f'_1 &2y f'_1 &0 \end{array} \right). \end{equation*} By expression \eqref{TxF_is_Hess} \begin{equation} {T} _{\mathbf{x}_0 } \mathbf{F} (\cdot)(\mathbf{e}_3 ) = {\rm Hess} \left( {B_z} \right)(\mathbf{x} _0 )= \left( \begin{array}{ccc} 2f'_1+ 4x ^2 f''_1 & 4xy f''_1 & 0\\ 4xy f_{1}'' &2f'_1+ 4y ^2 f''_1 &0\\ 0&0&f_{2}'' \end{array} \right). \end{equation} \noindent Using the same argument as in the proof of part {\bf (i)} we use, without loss of generality, the relative equilibrium point ${\bf z} _0 $ of the form ${\bf z} _0 = ( (\mathbb{I}_{id}, r \mathbf{e}_1)$, $(I _3 \left( \xi _1 ^0 - \xi _2 \right) \mathbf{e}_3 , M r \xi _1 ^0 \mathbf{e_2} ) )$, where $ r>0$. The matrix expression of $\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1^0 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W} $ is: \begin{equation} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! \left( \begin {array}{cccccccc} -2\mu(f'_1+2r^2f_{1}'' )+3{\xi_1^0}^2M &0 &0 &0 &0 &0 &0 &0\\ 0 &M r ^2 {\xi_1^0}^2 &0 &0 &0 &0 &0 &0\\ 0 &0 &{\dfrac {1}{I_1}} &0 &0 &0 &\xi_1^0 &0\\ 0 &0 &0 &\dfrac{1}{I_1} &0 &0 &0 &- \xi_1^0 \\ 0 &0 &0 &0 &\dfrac{1}{M} &0 &0 &\xi_1^0 r\\ 0 &0 &0 &0 &0 &-\mu f_{2}'' &-2\mu r f_1' &0\\ 0 &0 & \xi_1^0 &0 &0 &-2\mu r f_1' &\mu f_0+\xi_1^0 \Pi_0 &0\\ 0 &0 &0 &- \xi_1^0 & \xi_1^0 r &0 &0 &\mu f_0+\xi_1^0 (Mr^2\xi_1^0+\Pi_0) \end {array} \right), \end{equation} \medskip \noindent where $\Pi _0 = I_3 \left( \xi _1 ^0 - \xi _2 \right)$. Notice that this matrix is block diagonal and exhibits two blocks of size two and six. The positivity of the block of size two requires that $\mu f'_1<0 $ and $\mu(2 f_{1}' +r^2f_{1}'' )<0$ which coincide with~(\ref{f11}) and~(\ref{kozorez generalized1}). We now study the positivity of the lower six dimensional block of the stability form. As we did in the proof of part {\bf (i)}, we will read the signature of this block out of its pivots, which are $ \dfrac{1}{I _1 } $, $ \dfrac{1}{I _1 } $, $ \dfrac{1}{M} $, $ -\mu f_{2}'' $, $ \xi_1^0\left(\Pi_0 - \xi_1^0 I _1\right)+\mu \left(f_0+4 r ^2 \dfrac{{f'_{1}} ^2}{f_{2}'' } \right) $, and $\mu f_0+\xi_1^0 \Pi_0-{\xi_1^0}^2 I_1 $. The first three are automatically positive. The positivity of the fourth requires \begin{equation} \label{pivot4} \mu f_{2}''<0, \end{equation} which corresponds to the inequality~(\ref{f22}) in the statement. Finally, we study the positivity of the last two pivots. Let \begin{equation} \label{b1} p_1:=\xi_1^0\left(\Pi_0 - \xi_1^0 I _1\right)+\mu \left(f_0+4 r ^2 \dfrac{{f'_{1}} ^2}{f_{2}'' } \right) \end{equation} and \begin{equation} \label{b2} p_2:=\mu f_0 +\xi_1^0 \Pi_0-{\xi_1^0}^2 I_1. \end{equation} The simultaneous positivity of $p_1$ and $p_2$ is equivalent to $\min\left\{ p_1,p_2\right\} >0$. It is easy to check that $\min\left\{ p_1,p_2\right\} = p_1$, since the condition \begin{equation} p_2-p_1=-4\mu r^2 \dfrac{{f'_{1}} ^2}{f_{2}'' }>0 \end{equation} is satisfied due to \eqref{pivot4}. The positivity of $p_1$ can be summarized as \begin{equation*} {\rm sign}(\xi _1^0 ) I_3 \xi _2<- |\xi_1^0| \left( (I_1-I_3)+\dfrac{1}{2}M\left(\dfrac{f _0}{f'_1}+4 r ^2\dfrac{{f_{1}' }}{f_{2}'' } \right) \right), \end{equation*} which coincides with \eqref{cond3_xi2}, as required. \medskip \noindent {\bf Stability study for the singular branch.} The matrix expression associated to $\mathbf{d}^2 \left(h- \mathbf{J}^\xi\right) (\mathbf{z}) $ is given by \eqref{Hessian_singular}, where in this case \begin{equation} \mathbf{B}(\mathbf{x}_0) =B _z(\mathbf{x}_0) \mathbf{e}_3 = f_0 \mathbf{e} _3\enspace {\rm and} \enspace {\rm Hess} \left( {B_z} \right)(\mathbf{x}_0 ) =\left( \begin{array}{ccc} 2 f' _1&0&0\\ 0&2 f' _1&0\\ 0&0&f''_2 \end{array} \right). \end{equation} The pivots obtained by Gaussian elimination in the matrix expression of the stability form~$\left.\mathbf{d}^2 \left( h - \boldsymbol{J} ^{(\xi _1 , \xi _2 )}\right)\left( \mathbf{z}_0 \right)\right| _{W \times W}$ are \begin{equation*} p _1 ,p _2 , p _1 , p _1 , p _3 , p _3 , p _4, p _4, p _5, p _5, \end{equation*} where \begin{align*} p _1 = \frac{1}{M}, \enspace p _2 = - \mu f'' _2, \enspace p _3 = \frac{1}{I _1}, \enspace p _4 =- 2\mu f'_1-M \xi_1^2,\enspace p _5= \mu f _0- \xi_1(\xi_1 I _1- \Pi_0), \enspace \mbox{and} \enspace \Pi_0=I _3(\xi_1- \xi_2). \end{align*} The pivots $p _1 $ and $p _3$ are automatically positive. The positivity of the pivots $p _2 $, $p _4 $, and $p _5 $ requires that: \begin{align} &\mu f'' _2 < 0,\label{f22_gen_singular_cond_proof}\\ &\mu f'_1<0,\label{f11_gen_singular_cond_proof} \\ &\xi_1^2 < - \dfrac{2}{M}\mu f'_1,\label{xi1_gen_singular_cond_proof}\\ &{\rm sign}(\xi_1)\Pi_0> \dfrac{I _1\xi_1^2- \mu f_0}{|\xi_1|},\label{Pi0_gen_singular_cond_proof} \end{align} which yields the conditions \eqref{f22_gen_singular_cond}--\eqref{Pi0_gen_singular_cond}. We now derive the optimal stability condition \eqref{Pi0_gen_singular_optimal_cond}. Let $g(\xi_1):={\left(I _1 \xi_1 ^2-\mu f _0\right)}/{\xi_1}$ in \eqref{Pi0_gen_singular_cond}. It is easy to verify that the function $g(\xi_1)$ has a minimum at $\widehat{\xi_1^{+}}=\sqrt{- \mu f _0/I _1}$ and a maximum at $\widehat{\xi_1^{-}}=-\sqrt{- \mu f _0/I _1}$ provided that $\mu f_0<0$. Since the condition \eqref{xi1_gen_singular_cond} has to be satisfied, then $f _0/ f '_1<2I _1/M$ also needs to hold. In that case, the choices $\widehat{\xi_1^{\pm}}= \pm\sqrt{- \mu f _0/I _1}$ and the inequalities \begin{equation*} \Pi_0>\mathop{\rm min}_{\xi_1\in \mathbb{R}^{+}}\left\{g (\xi_1)\right\}=g\left(\widehat{\xi_1^{+}}\right)=2 \sqrt{- \mu f _0 I _1}, \qquad \Pi_0<\mathop{\rm max}_{\xi_1\in \mathbb{R}^{-}}\left\{g (\xi_1)\right\}=g\left(\widehat{\xi_1^{-}}\right)=-2 \sqrt{- \mu f _0 I _1} \end{equation*} determine the largest possible stability region in the $\Pi_0$ (and consequently the $\xi_2$) variable, as required in~(\ref{Pi0_gen_singular_optimal_cond}). \quad $\blacksquare$ \subsection{Proofs of Propositions~\ref{linear tools for instability} and~\ref{linearization for t stars g}} \noindent {\bf Proof of Proposition~\ref{linear tools for instability}} \medskip \noindent {\bf (i)} It is a consequence of the Witt-Artin decomposition (see for example~\cite[Theorem 7.1.1]{Ortega2004}). \medskip \noindent {\bf (ii)} It is a consequence of the fact that the symplectic slice introduced by Marle~\cite{nfm1, nfm}, Guillemin, and Sternberg~\cite{nfgs} can be constructed by Riemannian exponentiation of a symplectic tube. Since we need this construction in the proof of the following parts of the proposition, we briefly recall it using the notation in Chapter 7 of~\cite{Ortega2004}. The first step is the splitting of the Lie algebra $\mathfrak{g} $ of $G$ into three parts. The first summand is $\mathfrak{g}_\mu:= {\rm Lie} \left(G _\mu\right) $. The equivariance of the momentum map $\mathbf{J} $ implies that $G _m \subset G _\mu $ and hence $\mathfrak{g}_{m} \subset \mathfrak{g}_{\mu} $. Hence we can fix an $\mbox{\rm Ad} _{G_{m} } $-invariant inner product $\langle \cdot , \cdot \rangle $ on $\mathfrak{g}$ (always available by the compactness of $G_{m} $) and write \begin{equation} \label{splitting of lie algebras for slice} \mathfrak{g}_\mu= \mathfrak{g}_{m}\oplus \mathfrak{m} \quad\text{and}\quad \mathfrak{g}= \mathfrak{g}_{m}\oplus \mathfrak{m}\oplus \mathfrak{q}, \end{equation} where $\mathfrak{m}$ is the $\langle \cdot , \cdot \rangle $-orthogonal complement of $\mathfrak{g}_{m}$ in $\mathfrak{g}_\mu$ and $\mathfrak{q}$ is the $\langle \cdot , \cdot \rangle $-orthogonal complement of $\mathfrak{g}_\mu$ in $\mathfrak{g}$. The splittings in~(\ref{splitting of lie algebras for slice}) induce similar ones on the duals \begin{equation} \label{splitting of lie algebras for slice dual} \mathfrak{g}_\mu^\ast = \mathfrak{g}_{m}^\ast \oplus \mathfrak{m}^\ast \quad\text{and}\quad \mathfrak{g}^\ast = \mathfrak{g}_{m}^\ast \oplus \mathfrak{m}^\ast \oplus \mathfrak{q}^\ast. \end{equation} Each of the spaces in this decomposition should be understood as the set of covectors in $\mathfrak{g}^\ast $ that can be written as $\langle \xi, \cdot \rangle $, with $\xi $ in the corresponding subspace. For example, $ \mathfrak{q}^\ast =\{ \langle \xi, \cdot \rangle\mid \xi \in \mathfrak{q}\} $. The second ingredient in the construction of the symplectic tube comes from noting that the compact (by the properness of the action) isotropy subgroup $G _m $ acts linearly and canonically on $(W, \omega _W)$ with momentum map $\mathbf{J}_W:W \rightarrow \mathfrak{g}_m^\ast $ given by \begin{equation*} \langle\mathbf{J}_W(w), \eta\rangle= \frac{1}{2}\omega_W \left(\eta_W(w), w\right), \qquad \eta\in \mathfrak{g}_m. \end{equation*} It can be shown~\cite[Proposition 7.2.2]{Ortega2004} that there exist $G _m $--invariant neighborhoods $\mathfrak{m}_r ^\ast $ and $W _r $ of the origin in $\mathfrak{m}^\ast $ and $W$, respectively, such that the twisted product $Y _r:=G \times _{G _m } \left(\mathfrak{m}_r ^\ast \times W _r\right) $ is endowed with a natural symplectic form $\omega_{Y _r} $ whose expression can be found in (7.2.2) of~\cite{Ortega2004}. The Lie group $G $ acts canonically on $\left(Y _r, \omega_{Y _r}\right) $ by $g \cdot \left[h, \eta, w\right]= \left[gh, \eta,w\right] $, for any $g \in G $ and $[h, \eta, w ]\in Y _r $, and has a momentum map $\mathbf{J}_{Y _r}: Y _r \rightarrow \mathfrak{g}^\ast $ associated given by the so called Marle--Guillemin--Sternberg normal form: \begin{equation*} \mathbf{J}_{Y _r} \left(\left[g, \eta, w\right]\right)= {\rm Ad} ^\ast _{g ^{-1}} \left( \mu+ \eta + \mathbf{J}_W (w)\right), \qquad \left[g, \eta, w\right] \in Y _r. \end{equation*} The $G$--symplectic manifold $\left(Y _r, \omega_{Y _r}\right) $ is called a symplectic tube of $(M, \omega) $ at the point $m$. This denomination is justified by the Symplectic Slice Theorem~\cite{nfm1, nfm, nfgs} that proves the existence of a $G$--equivariant symplectomorphism $\phi : U \rightarrow Y _r$ between a $G$--invariant neighborhood $U$ of $m$ in $M$ and $Y _r $ satisfying $\phi(m)=[e,0,0] $. The symplectic slice $S$ in the statement of the proposition is obtained~\cite[Theorem 7.4.1]{Ortega2004} as $S= \phi^{-1} \left(S_{Y _r}\right)$, where $S_{Y _r}:=\left\{[e,0,w]\mid w \in W _r\right\}$ and, more explicitly, as $S= \left\{{\rm Exp} _m (w)\mid w \in W _r\right\} $, with ${\rm Exp} _m $ the Riemannian exponential associated to a $G _m$--invariant metric. The identity $T _mS=W $ is a consequence of the fact that $T _0 {\rm Exp} _m= {\rm Id} $. \medskip \noindent {\bf (iii)} Since $m\in M$ is a relative equilibrium, we have $\mathbf{d} \left(h- \mathbf{J}^\xi\right)(m)=0 $. This implies that $\mathbf{d}h ^\xi_S (m)= \mathbf{d}\left.\left(h- \mathbf{J}^\xi\right)(m)\right|_{T _mS} =0 $ and hence $X_{h _S^\xi}(m) =0 $. \medskip \noindent {\bf (iv)} This statement is a consequence of the combination of {\bf (ii)} and {\bf (iii)} with the following lemma. \begin{lemma} \label{Hessian_Q} Let $(M, \omega )$ be a symplectic manifold, $h \in C^{\infty}(M)$, and $X_h$ the corresponding Hamiltonian vector field. Suppose that $m_0 \in M$ is an equilibrium point of $X_h$, that is $X_h(m_0) = 0$ and consequently $\mathbf{d}h(m_0) = 0 $. Then, the linearization $X'$ of $X_h$ at $m_0$ is a Hamiltonian vector field on the symplectic vector space $(T_{m_0} M, \omega (m_0))$ with Hamiltonian function $Q \in C^{\infty}(T_{m_0}M)$ given by \begin{equation} \label{HessianQ} Q(v) = \dfrac{1}{2} \mathbf{d} ^2 h(m_0) (v,v). \end{equation} \end{lemma} \noindent\textbf{Proof of the Lemma.\ \ } Note $V=T_{m_0} M$ and $\omega _V = \omega (m_0)$. Let $v,w \in V$ arbitrary and let $\left\{ c(s) \vert s \in \mathbb{R}\right\} $ be a curve such that $v = \left. \frac{d}{ds } \right|_{s=0} c(s) $. Then if $F_t$ is the flow of $X_h$, we write \begin{equation} \label{omegaX} \omega _V \left(X'(v), w\right) = \left.\frac{d}{dt}\right|_{t=0} \omega (m_0) \left( T_{m_0} F_t \cdot v, w\right) = \left.\frac{d}{dt}\right|_{t=0} \omega (m_0) \left(\left.\frac{d}{ds}\right|_{s=0} F_t (c(s)), w\right), \quad v,w \in V. \end{equation} We now take a Darboux chart $(U, \phi )$~\cite[page 75]{Abraham1978} around the point $m_0$. Recall that in Darboux coordinates, the symplectic form $\omega _U$ is constant. Additionally if $\phi : U \longrightarrow \phi (U) \subset \mathbb{R}^n$, let $u \in \mathbb{R}^n$ be such that $T_{m_0} \phi \cdot w = \left( \phi (m_0), u \right) \in \phi (U) \times \mathbb{R}^n = T(\phi (U))$. Now, since $\phi ^\ast \omega _U = \omega \vert_U$, then \eqref{omegaX} can be written as \begin{align*} \left.\frac{d}{dt}\right|_{t=0} \omega (m_0) \left( \left.\frac{d}{ds}\right|_{s=0}F _t (c (s)), w\right) &= \left.\frac{d}{dt}\right|_{t=0} \omega _U \left( \left.\frac{d}{ds}\right|_{s=0} \phi \cdot F _t (c(s)), T_{m_0} \phi \cdot w\right) \\ &=\omega _U \left( \left.\frac{d}{ds}\right|_{s=0} T_{c(s)} \phi \cdot X _h (c(s)), \left( \phi (m_0), u\right) \right) \\ &= \left.\frac{d}{ds}\right|_{s=0} \omega _U \left( T_{c(s)} \phi \cdot X _{h \circ \phi ^{-1} \circ \phi } ( c(s)) , ( \phi (c(s)), u) \right) \\ & =\left.\frac{d}{ds}\right|_{s=0} \omega _U \left( X _{h \circ \phi ^{-1}} ( \phi (c(s))) , ( \phi (c(s)), u) \right) \\ &= \left.\frac{d}{ds}\right|_{s=0}\mathbf{d} \left( h\circ \phi ^{-1} \right) \left( \phi (c(s))\right) \cdot \left( \phi (c(s)), T_{m_0} \phi \cdot w \right) \\ & =\mathbf{d^2} \left( h\circ \phi ^{-1} \right) \left( \phi (m_0)\right) \left( (\phi (m_0), T_{m_0} \phi \cdot v), (\phi (m_0), T_{m_0} \phi \cdot w) \right) \\ &= \mathbf{d^2} h(m_0) (v,w) = \mathbf{d} Q(v) \cdot w. \end{align*} Consequently, $\mathbf{i}_{X'} \omega _V = \mathbf{d} Q$, as required. \quad $\blacktriangledown$ \medskip \noindent {\bf (v)} The hypothesis $T_m \left( G_\mu \cdot m \right) = T_m \left( G \cdot m\right) $ implies that $ \mathfrak{q} \cdot m:=\{ \xi_M(m)\mid \xi\in \mathfrak{q}\} $; this fact and the construction of the Witt--Artin decomposition (see for example the expression (7.1.11) in~\cite{Ortega2004}) ensure that \eqref{TmM} holds. In order to prove \eqref{XQ}, notice that for any $w _1 , w _2 \in W$ \begin{align} \label{XQ_proof} \omega _W (X_Q(w _1 ), w _2 ) &= \mathbf{d} Q (w _1 )\cdot w _2 = \mathbf{d} ^2 h ^ \xi (m) (w _1 , w _2 )=\omega (m) (X_{h^ \xi }'(w _1 ), w _2 )\notag \\ &= \omega (m) \left( \mathbb{P}_W X_{h^ \xi }'(w _1 ), w _2 \right) + \omega (m) \left( (\mathbb{I} - \mathbb{P} _W) X_{h^ \xi }' (w _1 ), w _2 \right)\notag \\ &= \omega (m) \left( \mathbb{P}_W X_{h^ \xi }'(w _1 ), w _2 \right) = \omega_W \left( \mathbb{P}_W X_{h^ \xi }'(w _1 ), w _2 \right),\notag \end{align} where we used that $ (\mathbb{I} - \mathbb{P} _W) X_{h^ \xi }' (w _1 ) \in W ^\omega $ and hence $\omega (m) \left( (\mathbb{I} - \mathbb{P} _W) X_{h^ \xi }' (w _1 ), w _2 \right)=0 $. Since $w _1, w _2 \in W$ are arbitrary, the equality $\omega _W (X_Q(w _1 ), w _2 )=\omega_W \left( \mathbb{P}_W X_{h^ \xi }'(w _1 ), w _2 \right) $ implies that \begin{equation*} X_Q(w) = \mathbb{P}_W X_{h^ \xi }' (w), \enspace \forall w \in W, \end{equation*} which is equivalent to \eqref{XQ}. \medskip \noindent {\bf (vi)} Given the local and group invariant character of this statement, we will prove this statement using the so called reconstruction differential equations~\cite{thesis, Roberts2002, Ortega2004} that determine the Hamiltonian vector field associated to a $G$--invariant Hamiltonian $h \in C^{\infty}(Y _r) $ in the symplectic tube $Y _r $. Consider first $ \pi: G \times \mathfrak{m} ^\ast _r \times W _r \longrightarrow G \times _{G _m}\left(\mathfrak{m}^\ast _r \times W _r\right)= Y _r $ the orbit projection; the $G$--invariance of $h$ implies that the composition $h \circ \pi$ can be understood as a $G _m$--invariant function on $G \times \mathfrak{m} ^\ast _r \times W _r $ that does not depend of the first factor, that is, $h \circ \pi \in C^{\infty} \left(\mathfrak{m}_r ^\ast \times W _r\right)^{G _m} $. The reconstruction equations show that for any $[g, \rho, w] \in Y _r $, \[ X _h([g, \rho, w])=T_{(g, \rho, w)} \pi(X_{\mathfrak{m}}( g, \rho,w ), X_{\mathfrak{m}^\ast_r}( g, \rho,w ), X_{W}( g, \rho,w)), \] where $X_{\mathfrak{m}}( g, \rho,w )$, $X_{\mathfrak{m}^\ast_r}( g, \rho,w )$, and $ X_{W_r}( g, \rho,w) $ are determined by the expressions \begin{eqnarray} X_{\mathfrak{m}}( g, \rho,w)&=&T _eL _g(D_{\mathfrak{m}^\ast _{r}}(h\circ\pi)(\rho,w)), \label{field 1 *}\\ X_{W_r}( g, \rho,w) &=&\omega^\sharp_{W}(D_{W_{r}}(h\circ\pi)(\rho,w)),\label{field 2 *}\\ X_{\mathfrak{m}^\ast _{r}}( g, \rho,w) &=&\mathbb{P}_{\mathfrak{m}^\ast }\Bigl({\rm ad}^\ast _{D_{\mathfrak{m}^\ast _{r}}(h\circ\pi)} \rho\Bigr)+{\rm ad}^\ast _{D_{\mathfrak{m}^\ast _{r}}(h\circ\pi)}\mathbf{J}_{W}(w),\label{field 3 *} \end{eqnarray} where $\mathbb{P}_{\mathfrak{m}^\ast }: \mathfrak{g}^\ast\rightarrow \mathfrak{m}^\ast $ is the projection according to the splitting~(\ref{splitting of lie algebras for slice dual}) and $\omega^\sharp_{W}: W ^\ast \rightarrow W $ is the isomorphism associated to the symplectic form $\omega_W$ in $W $. We now assume that $X _Q $ is spectrally unstable, which implies by part {\bf (iv) } that the Hamiltonian vector field $X_{h ^\xi_S } $ on the symplectic slice $S_{Y _r}=\{ [e,0,w]\mid w \in W _r\} $ exhibits an unstable equilibrium at $[e,0,0] $. Notice now that the Hamiltonian vector field $X_{h ^\xi_S } $ is given by the projection onto $Y _r $ via $T \pi $ of the vector field $(0,0, X_{W _r}^{h _S ^\xi}) $ in $G \times \mathfrak{m}^\ast _r \times W _r $ determined by \begin{equation*} X_{W _r}^{h _S ^\xi}= \omega^\sharp _W(D_{W _r}\left(h \circ \pi\right)(0,w))- \left(\mathbb{P}_{\mathfrak{g}_m}\xi\right)_W(w)=X_{W}^h(e,0,w)-\left(\mathbb{P}_{\mathfrak{h}}\xi\right)_W(w), \end{equation*} where $\mathbb{P}_{\mathfrak{g}_m}: \mathfrak{g} \rightarrow \mathfrak{g}_m $ is the projection according to the splitting~(\ref{splitting of lie algebras for slice}), $\left(\mathbb{P}_{\mathfrak{g}_m}\xi\right)_W\in \mathfrak{X}(W _r)$ is the infinitesimal generator associated to $\mathbb{P}_{\mathfrak{g}_m}\xi \in \mathfrak{g}_m $ using the $G _m$--action on $W _r $, and $ X_{W}^h $ is the vector field in~(\ref{field 2 *}) that determines the dynamics induced by $h$ on the space $W _r $. The instability of $X_{h _S ^\xi} $ at $[e,0,0] $ implies the same feature for $X_W ^h $ at $(e,0,0) $ and hence the $K $--instability of the relative equilibrium $[e,0,0] $ of $X _h $, for any subgroup $K \subset G $. \quad $\blacksquare$ \medskip \noindent {\bf Proof of Proposition~\ref{linearization for t stars g}} \medskip \noindent\textbf{(i)} It is a straightforward consequence of the chain rule as $\mathbf{d}h (g, \mu)=0 $ implies that $\mathbf{d}h ^g(e, \mu)=0 $. \medskip \noindent {\bf (ii)} Relation~(\ref{relation quadratic forms}) is a consequence of the following general fact about Hessians: let $m\in M$ and $n\in N$, with $M$ and $N$ smooth manifolds and let $\psi:M\rightarrow N$ be a smooth map such that $\psi(m)=n$. Let $f\in C^{\infty}(N)$ with $\mathbf{d} f(n)=0$. Then $\mathbf{d}^2(\psi^\ast f)(m)=T_m^\ast\psi(\mathbf{d}^2 f(n))$, that is, for any $v,\,w\in T_mM$: \[ \mathbf{d}^2(\psi^\ast f)(m)(v,\,w)= \mathbf{d}^2 f(n)(T_m\psi\cdot v,\,T_m\psi\cdot w). \] In order to establish relation~(\ref{relation linearizations}) note that the map $\Phi _g: \mathfrak{g}\times \mathfrak{g}^\ast \longrightarrow T_{(g, \mu)}\left(G \times \mathfrak{g}^\ast\right) $ is a symplectomorphism and hence a Poisson map. Expression~(\ref{relation linearizations}) follows from~\cite[Proposition 4.1.19]{Ortega2004}. \medskip \noindent {\bf (iii)} By Lemma~\ref{Hessian_Q}, the Hamiltonian vector field $X_{Q ^g} $ is determined by the relation $\mathbf{i}_{X_{Q ^g}} \omega(e, \mu)= \mathbf{d} Q ^g $ or, more explicitly, by \begin{equation*} \omega(e, \mu)\left(X_{Q^g}(\xi, \tau), (\eta, \rho)\right)= \mathbf{d} Q ^g(\xi, \tau)\cdot (\eta, \rho), \quad \mbox{for any} \quad (\xi, \tau), (\eta, \rho) \in \mathfrak{g}\times \mathfrak{g}^\ast. \end{equation*} Using the expression of the canonical symplectic form of $T ^\ast G $ in body coordinates (see for instance~\cite[Expression (6.2.5)]{Ortega2004}), this equality can be rewritten as \begin{equation} \label{intermediate 1} \langle \rho, \pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)\rangle- \langle\pi_{\mathfrak{g}^\ast} \left(X_{Q ^g}(\xi, \tau)\right), \eta\rangle+ \langle\mu, {\rm ad}_{\pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)}\eta\rangle=\langle{\rm Hess}(\xi, \tau), (\eta, \rho)\rangle. \end{equation} Let now ${\rm {\bf pr}}_{\mathfrak{g}}: \mathfrak{g}\times \mathfrak{g}^\ast \rightarrow \mathfrak{g}\times \mathfrak{g}^\ast $ and ${\rm {\bf pr}}_{\mathfrak{g}^\ast }: \mathfrak{g}\times \mathfrak{g}^\ast \rightarrow \mathfrak{g}\times \mathfrak{g}^\ast $ be the maps defined by ${\rm {\bf pr}}_{\mathfrak{g}}(\eta, \rho):=(\eta, 0)$ and ${\rm {\bf pr}}_{\mathfrak{g}^\ast }(\eta, \rho):=(0, \rho)$, for any $(\eta, \rho)\in \mathfrak{g}\times \mathfrak{g}^\ast$, and $\mathbf{i}_{\mathfrak{g}}: \mathfrak{g}\rightarrow \mathfrak{g}\times \mathfrak{g}^\ast $ and $\mathbf{i}_{\mathfrak{g}^\ast }: \mathfrak{g}^\ast \rightarrow \mathfrak{g}\times \mathfrak{g}^\ast $, the canonical injections. Since~(\ref{intermediate 1}) holds for $(\eta, \rho)\in \mathfrak{g}\times \mathfrak{g}^\ast$ arbitrary, we apply it to vectors of the form ${\rm {\bf pr}}_{\mathfrak{g}}(\eta, \rho)=(\eta, 0)$ and ${\rm {\bf pr}}_{\mathfrak{g}^\ast }(\eta, \rho)=(0, \rho)$ and we obtain the following two equalities \begin{eqnarray*} \langle{\rm Hess}(\xi, \tau), {\rm {\bf pr}}_{ \mathfrak{g}^\ast }(\eta, \rho)\rangle&=&\langle \pi_{\mathfrak{g}^\ast }(\eta, \rho), \pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)\rangle,\\ \langle{\rm Hess}(\xi, \tau), {\rm {\bf pr}}_{ \mathfrak{g}}(\eta, \rho)\rangle&=&- \langle\pi_{\mathfrak{g}^\ast} \left(X_{Q ^g}(\xi, \tau)\right), \pi_{\mathfrak{g}}(\eta, \rho)\rangle+ \langle\mu, {\rm ad}_{\pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)}\pi_{\mathfrak{g}}(\eta, \rho)\rangle. \end{eqnarray*} Since $(\eta, \rho)\in \mathfrak{g}\times \mathfrak{g}^\ast$ in these expressions are arbitrary, they can be rewritten as: \begin{eqnarray} {\rm {\bf pr}}_{ \mathfrak{g}^\ast }^\ast {\rm Hess}(\xi, \tau)&=&\pi_{\mathfrak{g}^\ast }^\ast \left( \pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)\right),\label{iw1}\\ {\rm {\bf pr}}_{ \mathfrak{g}}^\ast {\rm Hess}(\xi, \tau)&=&- \pi_{\mathfrak{g}}^\ast \left(\pi_{\mathfrak{g}^\ast} \left(X_{Q ^g}(\xi, \tau)\right)\right)+ \pi_{\mathfrak{g}}^\ast \left({\rm ad}_{\pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right)}^\ast (\mu)\right).\label{iw2} \end{eqnarray} We now apply $\pi_{\mathfrak{g}^\ast} $ and $\pi_{\mathfrak{g}} $ to both sides of~(\ref{iw1}) and~(\ref{iw2}), respectively, and we notice that $\pi_{\mathfrak{g}^\ast} \circ {\rm {\bf pr}}_{ \mathfrak{g}^\ast}^\ast = \pi_{\mathfrak{g}^\ast} $, $\pi_{\mathfrak{g} } \circ {\rm {\bf pr}}_{ \mathfrak{g}} ^\ast = \pi_{\mathfrak{g} } $, $\pi_{\mathfrak{g}} ^\ast = \mathbf{i}_{\mathfrak{g}}$, $\pi_{\mathfrak{g}^\ast } ^\ast = \mathbf{i}_{\mathfrak{g}^\ast }$, $\pi_{\mathfrak{g}} \circ \mathbf{i}_{\mathfrak{g}}= {\rm id}_{\mathfrak{g}} $, and $\pi_{\mathfrak{g}^\ast } \circ \mathbf{i}_{\mathfrak{g}^\ast }= {\rm id}_{\mathfrak{g}^\ast } $. We obtain \begin{eqnarray} \pi_{\mathfrak{g}} \left(X_{Q ^g}(\xi, \tau)\right) &= &\pi_{ \mathfrak{g}^\ast } {\rm Hess}(\xi, \tau), \\ \pi_{\mathfrak{g}^\ast} \left(X_{Q ^g}(\xi, \tau)\right)&=&-\pi_{\mathfrak{g}} \left({\rm Hess}(\xi, \tau)\right)+ {\rm ad}_{\pi_{ \mathfrak{g}^\ast } {\rm Hess}(\xi, \tau)}^\ast \mu, \end{eqnarray} which is equivalent to~(\ref{expression linearization at e}). \quad $\blacksquare$ \subsection{Linear stability and instability of the standard and generalized orbitron relative equilibria} \label{Linear stability and instability of the orbitron relative equilibria} \noindent {\bf The linearization for the regular branches.} The goal in this paragraph is determining the linear Hamiltonian vector fields $X _Q $ in the stability space $W$ used in the proof of Theorem~\ref{kozorez relations} by using the expression~(\ref{XQ}) in Proposition~\ref{linear tools for instability}. Notice that this is indeed possible due to the Abelian character of our symmetry group that ensures that in this situation $G _\mu=G $ and hence the coincidence of the tangent spaces $T _m \left( G_{\mu } \cdot m\right) $ and $T _m \left( G \cdot m\right) $ that is necessary as a hypothesis in this statement. We start by writing down the decomposition~(\ref{TmM}) that in this case can be achieved by noting that \begin{equation} W^{\omega } = {\rm span} \left\{ \boldsymbol{\tau} _1, \boldsymbol{\tau} _2 , \boldsymbol{\tau} _3 , \boldsymbol{\tau} _4 \right\} \end{equation} with $ \boldsymbol{\tau} _1 = (\mathbf{e}_3 , r \mathbf{e}_2, \mathbf{0}, \mathbf{0} )\, \boldsymbol{\tau} _2 = (-\mathbf{e}_3 , \mathbf{0} , \mathbf{0} , - Mr \xi _1 ^0 \mathbf{e}_1),\, \boldsymbol{\tau} _3 = (\mathbf{0} , \mathbf{0}, \mathbf{0} , \mathbf{e}_1 )\, \boldsymbol{\tau} _4 = (\mathbf{0} , \mathbf{e}_1, -Mr \xi _1 ^0 \mathbf{e}_3 , \mathbf{0} )$. If we use as a basis for $W$ the vectors introduced in~(\ref{basis regular branch}) and for $W ^\omega $ the ones those that we just described, it is easy to see that the matrix expressions of the inclusion $\boldsymbol{i}_W: W \hookrightarrow T_{\mathbf{z}_0} \left( SE(3) \times \mathfrak{se}(3)^*\right) \simeq \mathbb{R}^{12}$ and the projection $\mathbb{P}_W: T_{\mathbf{z}_0} \left( SE(3) \times \mathfrak{se}(3)^*\right) \simeq \mathbb{R}^{12} \longrightarrow W$, where $\mathbb{R}^{12}$ is endowed with the canonical basis, are given by \begin{eqnarray} \boldsymbol{i}_W &=& \left(\mathbf{u}_1'\vert \mathbf{u}_2'\vert \mathbf{u}_3'\vert \mathbf{u}_4'\vert \mathbf{u}_5'\vert \mathbf{u}_6'\vert \mathbf{u}_7'\vert \mathbf{u}_8'\right),\label{injection 1}\\ \mathbb{P}_W &=& \left(\mathbf{e}_8'\vert \mathbf{e}_7'\vert \mathbf{e}_2'\vert \mathbf{e}_1'\vert -\frac{1}{r}\mathbf{e}_2'\vert \mathbf{e}_6'\vert \mathbf{e}_3'\vert \mathbf{e}_4' \vert -\frac{1}{Mr \xi _1 ^0 } \mathbf{e}_2' \vert \mathbf{0} \vert \mathbf{e}_5' \right), \label{projection 1} \end{eqnarray} where the apostrophes stand for the transposition operation and the vertical indicate matrix concatenation. The linearization of the Hamiltonian vector field associated to the augmented Hamiltonian at the relative equilibria of the standard orbitron can be immediately obtained by using the expression~(\ref{expression of linearization}) together with the Hessian already computed in~(\ref{hessian in general}) in the context of the proof of Theorem~\ref{kozorez relations}. The resulting expression is inserted in~(\ref{XQ}) using the injection~(\ref{injection 1}) and the projection~(\ref{projection 1}) and yields the following matrix for $X _Q $: \medskip \begin{equation} \label{linearization regular} {\small \!\!\!\!\! \left( \begin {array}{cccccccc} 0&-\xi_1^0 r&0&0&0&0&0&0\\ \xi_1^0 r \dfrac{4 h^2-r^2}{r^2+h^2}&0&0&0&0&0&0&0\\ 0&0&0&\xi_1^0-\dfrac{\Pi_0}{I_1}&0&0&0&-\frac{1}{3} {\xi_1^0}^2 M (r^2+h^2)\\ 0&0&-\xi_1^0+\dfrac{\Pi_0}{I_1}&0&0&-{\xi_1^0}^2 M r&-\frac{1}{3} {\xi_1^0}^2 M (r^2+h^2)&0\\ 0&0&-\dfrac{M \xi_1^0 r}{I_1}&0&0&{\xi_1^0}^2 M\dfrac{2 h^2-3 r^2}{r^2+h^2}&-2 {\xi_1^0}^2 M r&0\\ 0&0&0&0&\dfrac{1}{M}&0&0&\xi_1^0 r\\ 0&0&0&\dfrac{1}{I_1}&0&0&0&-\xi_1^0\\ 0&0&\dfrac{1}{I_1}&0&0&0&\xi_1^0&0 \end {array} \right),} \end{equation} \medskip where $\Pi_0=I_3 (\xi_1^0-\xi_2)$. An analog expression can be obtained for the linearization $X _Q$ at the regular relative equilibria of the generalized orbitron: \begin{equation*} {\small \!\!\!\!\! \left( \begin {array}{cccccccc} 0& -\xi_1 ^0 r& 0& 0& 0& 0& 0& 0\\ -\dfrac{4 \mu }{Mr \xi _1 ^0 }\left( 2 f' _1 + r ^2 f '' _1 \right) & 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& \xi_1^0-\dfrac{\Pi_0}{I_1}& 0& 0& 0& - \mu f _0 \\ 0& 0& -\xi_1^0+\dfrac{\Pi_0}{I_1}& 0& 0& 2 \mu r f _1' & - \mu f _0 & 0\\ 0& 0& -\dfrac{M \xi_1^0 r}{I_1}& 0& 0& \mu f '' _2 & 4 \mu r f' _1 &0\\ 0&0&0&0&\dfrac{1}{M}&0&0&\xi_1^0 r\\ 0&0&0&\dfrac{1}{I_1}&0&0&0&-\xi_1^0\\ 0&0&\dfrac{1}{I_1}&0&0&0&\xi_1^0&0 \end {array} \right).} \end{equation*} \medskip \noindent {\bf The linearization for the singular branches.} The same scheme can be reproduced for the singular branches by using the stability space introduced in~(\ref{stability space singular}). In this case, it can be shown that $W ^ \omega ={\rm span} \left\{ \boldsymbol{\tau} _1 , \boldsymbol{\tau} _2 \right\} $, with $\boldsymbol{\tau} _1 = (\mathbf{e}_3 , \mathbf{0}, \mathbf{0} , \mathbf{0} )$ and $\boldsymbol{\tau} _2 = (\mathbf{0} , \mathbf{0} , \mathbf{e}_3 , \mathbf{0})$, which yields the following matrix expressions for the inclusion and the projection: \begin{eqnarray*} \boldsymbol{i}_W &=& \left(\mathbf{u}_1'\vert \mathbf{u}_2'\vert \mathbf{u}_3'\vert \mathbf{u}_4'\vert \mathbf{u}_5'\vert \mathbf{u}_6'\vert \mathbf{u}_7'\vert \mathbf{u}_8'\vert \mathbf{u}_9'\vert \mathbf{u}_{10}'\right),\\ \mathbb{P}_W &=& \left(\mathbf{e}_{10}'\vert \mathbf{e}_9'\vert \mathbf{0}\vert \mathbf{e}_8'\vert \mathbf{e}_7'\vert \mathbf{e}_2'\vert \mathbf{e}_6'\vert \mathbf{e}_5' \vert \mathbf{0}' \vert \mathbf{e}_4 ' \vert \mathbf{e}_3' \vert \mathbf{e}_1' \right). \end{eqnarray*} Finally, the matrix expression for $X _Q $ at the singular relative equilibria of the standard orbitron is: \begin{equation} \label{xq singular} {\small \left( \begin {array}{cccccccccc} 0&-3\dfrac{\mu _0 \mu q}{\pi h ^4 }&0&0&0&0&0&0&0&0\\ \dfrac{1}{M}&0&0&0&0&0&0&0&0&0\\ 0&0&0&- \xi _1 &0&0&\dfrac{3}{2}\dfrac{\mu _0 \mu q}{\pi h ^4 }&0&0&0\\ 0&0&\xi _1 &0&0&0&0&\dfrac{3}{2}\dfrac{\mu _0 \mu q}{\pi h ^4 }&0&0\\ 0&0&0&0&0&- \xi _1+ \dfrac{\Pi_0}{I _1} &0&0&\dfrac{1}{2}\dfrac{\mu _0 \mu q}{\pi h ^2 }&0\\ 0&0&0&0&\xi _1-\dfrac{\Pi_0}{I _1} &0&0&0&0&\dfrac{1}{2}\dfrac{\mu _0 \mu q}{\pi h ^2 }\\ 0&0&\dfrac{1}{M}&0&0&0&0&- \xi _1 &0&0\\ 0&0&0&\dfrac{1}{M}&0&0&\xi _1 &0&0&0\\ 0&0&0&0&\dfrac{1}{I _1 } &0&0&0&0&-\xi _1 \\ 0&0&0&0&0&\dfrac{1}{I _1 } &0&0&\xi _1 &0\\ \end {array} \right), } \end{equation} \medskip where $\Pi_0=I_3 (\xi_1-\xi_2)$. An analog expression can be obtained for the linearization $X _Q$ at the singular relative equilibria of the generalized orbitron: \begin{equation} \label{xq singular generalized} {\small \left( \begin {array}{cccccccccc} 0& \mu f'' _2 & 0& 0& 0& 0& 0& 0& 0& 0\\ \dfrac{1}{M}& 0& 0& 0& 0& 0& 0& 0& 0& 0\\ 0& 0& 0& - \xi _1 & 0& 0& 2 \mu f' _1 & 0& 0& 0\\ 0& 0& \xi _1 & 0& 0& 0& 0& 2 \mu f' _1 & 0& 0\\ 0& 0& 0& 0& 0& - \xi _1+ \dfrac{\Pi_0}{I _1} & 0& 0& - \mu f _0 & 0\\ 0& 0& 0& 0& \xi _1-\dfrac{\Pi_0}{I _1} & 0& 0& 0& 0& - \mu f _0 \\ 0&0&\dfrac{1}{M}&0&0&0&0&- \xi _1 &0&0\\ 0&0&0&\dfrac{1}{M}&0&0&\xi _1 &0&0&0\\ 0&0&0&0&\dfrac{1}{I _1 } &0&0&0&0&-\xi _1 \\ 0&0&0&0&0&\dfrac{1}{I _1 } &0&0&\xi _1 &0\\ \end {array} \right). } \end{equation} \medskip \addcontentsline{toc}{section}{Bibliography} \bibliographystyle{alpha}
\section{Introduction} \subsection*{Main result} Let $\left\{ \left(Y_{\epsilon}^{x}:x\in[0,1]^{d}\right)\right\} _{\epsilon>0}$ be a family of centered Gaussian fields indexed by the $d$-dimensional unit box $[0,1]^{d}$, where $d$ is any positive integer. Suppose that the family satisfies, for some constant $0<C_{Y}<\infty$ and all $x,y\in[0,1]^{d}$, $\epsilon>0$, \begin{equation} \left|Cov\left(Y_{\epsilon}^{x},Y_{\epsilon}^{y}\right)+\log\left(\max\{\epsilon,\left\Vert x-y\right\Vert \}\right)\right|\leq C_{Y}\label{logcorrY} \end{equation} and \begin{equation} \mathbb{E}\left[\left(Y_{\epsilon}^{x}-Y_{\epsilon}^{y}\right)^{2}\right]\leq C_{Y}\epsilon^{-1}\left\Vert x-y\right\Vert \text{ if }\left\Vert x-y\right\Vert \leq\epsilon,\label{l2distY} \end{equation} where $\left\Vert \cdot\right\Vert $ is Euclidean distance. Display \eqref{logcorrY} implies that the covariance is logarithmic for distant points and that the variance is nearly constant. The second condition is imposed so that the field does not vary too much for close points. Display \eqref{l2distY}, basic relations between the moments of Gaussian random variables and Kolmogorov's continuity criterion (see \cite[Theorem 1.4.17]{k}) imply that the fields have continuous modifications. When $d=2$, an example of a field satisfying \eqref{logcorrY} and \eqref{l2distY} is the bulk of the mollified continuous Gaussian free field (MGFF), which will be defined in Section 3.1, and will be the object of our attention in Section 3. Set $m_{\epsilon}=m_{\epsilon,d}=\sqrt{2d}\log(1/\epsilon)-\frac{3/2}{\sqrt{2d}}\log\log(1/\epsilon)$. The main result of this paper is: \begin{thm} \label{main} There exist constants $0<c,C<\infty$ (depending on $C_{Y}$ and $d$) such that, for all $\epsilon>0$ small enough, \begin{equation} \mathbb{P}\left(\left|\max_{x\in[0,1]^{d}}Y_{\epsilon}^{x}-m_{\epsilon}\right|\geq\lambda\right)\leq Ce^{-c\lambda}\label{rtubY} \end{equation} for all $\lambda\geq0$. \end{thm} \noindent Theorem \ref{main} implies, in particular, that $\left\{ \max_{x\in[0,1]^{d}}Y_{\epsilon}^{x}-m_{\epsilon}:\epsilon>0\right\} $ is tight and that \[ \left|\mathbb{E}\left[\max_{x\in[0,1]^{d}}Y_{\epsilon}^{x}\right]-m_{\epsilon}\right|\leq C \] for some constant $C$ depending on $C_{Y}$ and $d$. The main idea of the proof of Theorem \ref{main} is to use Slepian's Lemma (see \cite[Theorem 2.2.1]{rf}) to compare the maximum of the field $Y_{\epsilon}$ with the maximum of the modified branching random walk (MBRW), a field introduced by Bramson and Zeitouni in \cite{bz}. Since Slepian's Lemma only allows comparison of fields with the same index set, we will add an appropriately chosen independent continuous field to the MBRW. Adding an independent continuous field to the MBRW does not change the maximum much, provided the continuous field is small and smooth enough. These fields are defined in detail in Section 2.1. After defining the fields, we compare the right and left tails in Sections 2.2 and 2.3, respectively. We then show, in Section 3, that Theorem \ref{main} implies tightness of the recentered maximum of the MGFF. A comment on constants: $c$ will always denote a small positive constant and $C$ will always denote a large positive constant. Both constants are allowed to change from line to line. The dependence of the constants will be explicit or will be clear from the context. The phrase ``absolute constant'' will refer to fixed numbers that are independent of everything. \subsection*{Related work} Our approach is motivated by recent advances in the study of the two dimensional discrete Gaussian free field (DGFF). In \cite{bz}, Bramson and Zeitouni computed the expected maximum of the DGFF up to an order 1 error and concluded tightness of the recentered maximum. In \cite{d}, Ding obtained bounds on the right and left tail of the recentered maximum of the DGFF. Later on, in \cite{bdz}, Bramson, Ding and Zeitouni proved convergence in distribution of the recentered maximum. The approach of this line of research is to use first and second moment methods, together with decomposition properties of the DGFF, to obtain good estimates on tail events. Previous work on the DGFF includes \cite{bdg}, where Bolthausen, Deuschel and Giacomin obtained asympotics for the maximum of the DGFF, and \cite{da}, where Daviaud studied the extreme points of the DGFF. On the other hand, previous work on the continuous Gaussian free field (CGFF) includes \cite{hmp}, where Hu, Miller, and Peres studied the Hausdorff dimension of the ``thick points'' of the MGFF, which are closely related to the work of Daviaud. We also mention \cite{dy} for a nice discussion of Gaussian fields induced by Markov processes, and \cite{s} for a survey on the CGFF. Our main result implies, in particular, an analog of \cite[Theorem 1.1]{bz} for the MGFF. Our approach consists on extending the MBRW by Brownian sheet, so that it is possible to compare the extended field with scaled log-correlated continuous fields. Log-correlated Gaussian fields are subject of current interest (see \cite{drsv}, \cite{m}, \cite{mrv}). In particular, in \cite{m}, Madaule proved convergence for stationary centered Gaussian fields $\left(Z_{\epsilon}(x):x\in[0,1]^{d}\right)$ whose covariance satisfies \[ Cov(Z_{\epsilon}(0),Z_{\epsilon}(x))=\int_{0}^{\log(1/\epsilon)}k(e^{r}x)dr, \] where the fixed kernel $k:\mathbb{R}^{d}\to\mathbb{R}$ is of class $C^{1}$, vanishes outside $[-1,1]^{d}$, and satisfies $k(0)=1$. Theorem \ref{main} has weaker conditions on the covariance structure, and consequently, only tightness is achieved. In \cite{mrv}, the authors proved the so called ``Freezing Theorem for GFF in planar domains'' for a sequence of Gaussian fields approximating the continuous GFF by cutting-off white noise, so that the covariance kernel is proportional to the function $G_{t}:[0,1]^{2}\times[0,1]^{2}\to\mathbb{R}$ given by \[ G_{t}(x,y)=\int_{e^{-t}}^{\infty}p_{\partial[0,1]^{2}}(r,x,y)dr, \] where $p_{\partial[0,1]^{2}}(r,x,y)$ is the transition probability density of a Brownian motion killed at $\partial[0,1]^{2}$. In the present paper, we consider a sequence of fields approximating the GFF by mollifying the Green function (see \eqref{green}), and we prove tightness. Convergence for the MGFF is expected to follow by adapting of the arguments given in \cite{bdz}. \section{Comparison to the MBRW} \subsection{Auxiliary fields} In this subsection, we rigorously introduce the fields we mentioned in Section 1. A few properties of these fields will be stated; the proofs of these properties will be given in the Appendix. In order to define these fields, it will be notationally more convenient to use $[0,1)^{d}$ instead of $[0,1]^{d}$ as the index set. This will not affect the main result because the supremum of $Y_{\epsilon}$ over $[0,1)^{d}$ is the same, due to continuity, as the maximum over $[0,1]^{d}$. \subsubsection*{Modified branching random walk} We first divide $[0,1)^{d}$ into boxes of side length $\epsilon>0$. Let $V_{\epsilon}=\left(\epsilon\mathbb{Z}^{d}\right)\cap[0,1)^{d}$ and, for $v\in V_{\epsilon}$, let $\square_{\epsilon}^{v}=[v,v+\epsilon)^{d}\cap[0,1)^{d}$. Moreover, if $x\in\square_{\epsilon}^{v}$, let $[x]:=v$. The set $V_{\epsilon}$ is, of course, a discretized version of $[0,1)^{d}$. We now define the \emph{modified branching random walk} (MBRW) as the centered Gaussian field $\left\{ \xi_{\epsilon}^{v}(t):v\in V_{\epsilon},0\leq t\leq\log(1/\epsilon)\right\} $ with covariance structure \begin{equation} Cov(\xi_{\epsilon}^{v}(t),\xi_{\epsilon}^{u}(s))=\int_{0}^{\min\left\{ t,s\right\} }\prod_{1\leq i\leq d}(1-e^{r}\left|v_{i}-u_{i}\right|)_{+}dr\label{defMBRW} \end{equation} for all $0\leq t,s\leq\log\left(1/\epsilon\right)$ and $v,u\in V_{\epsilon}$, where $v_{i}$ is the $i$-th coordinate of $v$, and $\left(\cdot\right)_{+}=\max\left\{ \cdot,0\right\} $. For simplicity, write $\xi_{\epsilon}^{v}=\xi_{\epsilon}^{v}(\log(1/\epsilon))$. Note that, for each point $v\in V_{\epsilon}$, the process $\left(\xi_{\epsilon}^{v}(t)\right)_{t}$ is a standard Brownian motion. Moreover, for each pair $v,u\in V_{\epsilon}$, the Brownian motions are correlated until $t=-\log\left\Vert v-u\right\Vert _{\infty}$, at which time their increments become independent. The end time is $t=\log\left(1/\epsilon\right)$, because, for the ``usual'' $d$-ary branching random walk, it takes $\log(1/\epsilon)$ units of time to generate $\left|V_{\epsilon}\right|$ particles (see the proof of Proposition \ref{proofrtubMBRW} for a definition of the usual $d$-ary branching random walk). It will be proved in the Appendix (see Proposition \ref{proofdefMBRW}) that the MBRW exists and that it satisfies \begin{equation} Var(\xi_{\epsilon}^{v})=\log(1/\epsilon)\label{varMBRW} \end{equation} and, for $v\neq u$ (so that $\left\Vert v-u\right\Vert _{\infty}\geq\epsilon$), \begin{equation} -\log\left\Vert v-u\right\Vert _{\infty}-C\leq Cov(\xi_{\epsilon}^{v},\xi_{\epsilon}^{u})\leq-\log\left\Vert v-u\right\Vert _{\infty}\label{covMBRW} \end{equation} for some constant $C$ depending on $d$. The MBRW also satisfies (see Proposition \ref{proofrtlbMBRW}) \begin{equation} \mathbb{P}\left(\max_{v\in V_{\epsilon}}\xi_{\epsilon}^{v}\geq m_{\epsilon}\right)\geq c>0,\label{rtlbMBRW} \end{equation} where $c$ is a constant depending only on $d$. It will also be proved in the Appendix (see Proposition \ref{proofrtubMBRW}) that there exist constants $0<c,C<\infty$ (depending on $d$) such that \begin{equation} \mathbb{P}\left(\max_{v\in A}\xi_{\epsilon}^{v}\geq m_{\epsilon}+z\right)\leq C\left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{-cz}\label{rtubMBRW} \end{equation} for all $A\subset V_{\epsilon}$, $z\in\mathbb{R}$ and $\epsilon>0$ small enough, where $\left|A\right|$ is the cardinality of $A$. \subsubsection*{Brownian sheet} As mentioned before, we will need an additional continuous Gaussian field. For $x=(x_{i})_{i\leq d}\in\mathbb{R}_{+}^{d}$, let $\psi^{x}$ denote the centered standard Brownian sheet. Recall that it satisfies \[ \mathbb{E}\left[\psi^{x}\psi^{y}\right]=\prod_{i\leq d}\min\left\{ x_{i},y_{i}\right\} . \] Define a new field $\left(\psi_{\epsilon}^{x}:x\in[0,1)^{d}\right)$, depending on a parameter $p\geq1$, as follows: for $v\in V_{\epsilon}$, let $l$ be the linear map from $\square_{\epsilon}^{v}$ onto $[p,2p)^{d}$ sending $v$ to $(p)_{i\leq d}=(p,p,\ldots,p)$. Set \begin{equation} \left(\psi_{\epsilon}^{x}:x\in\square_{\epsilon}^{v}\right):\overset{d}{=}\left(\psi^{l(x)}:x\in\square_{\epsilon}^{v}\right)=\left(\psi^{x}:x\in[p,2p)^{d}\right),\label{defBS} \end{equation} for each $v\in V_{\epsilon}$, and choose $\psi_{\epsilon}^{x}$ and $\psi_{\epsilon}^{y}$ to be independent if $[x]\neq[y]$. Note that the collection of fields $\left\{ \left(\psi_{\epsilon}^{x}:x\in\square_{\epsilon}^{v}\right)\right\} _{v\in V_{\epsilon}}$ consist of i.i.d copies of Brownian sheet on $[p,2p)^{d}$. Using the covariance structure of the Brownian sheet, it is not hard to see that \begin{equation} p^{d}\leq Var\left(\psi_{\epsilon}^{x}\right)\leq(2p)^{d},\label{varBS} \end{equation} for all $x\in[0,1)^{d}$, and that \begin{equation} p^{d}\epsilon^{-1}\left\Vert x-y\right\Vert _{1}\leq\mathbb{E}\left[\left(\psi_{\epsilon}^{x}-\psi_{\epsilon}^{y}\right)^{2}\right]\leq(2p)^{d}\epsilon^{-1}\left\Vert x-y\right\Vert _{1},\label{l2distBS} \end{equation} for all $[x]=[y]$. Note that $p$ can be chosen as large as desired. To understand the motivation behind the previous definitions, we invite the reader to compare the bounds \eqref{logcorrY} and \eqref{l2distY} with \eqref{covMBRW} and \eqref{l2distBS}, respectively. These bounds will be used in the next sections. We now proceed to the comparison of the right and left tail of the maximum of the field $Y_{\epsilon}$ (which was defined in Section 1 and satisfies \eqref{logcorrY} and \eqref{l2distY}) and the maximum of an appropriate combination of the fields $\xi_{\epsilon}$ and $\psi_{\epsilon}$ (which will be specified in the next section). Note that we will only use Brownian sheet when comparing the right tail; for the left tail, we will compare directly the MBRW with the field $Y_{\epsilon}$ on a discrete index set. \subsection{The right tail} Recall from Section 1 that the field $Y_{\epsilon}$ satisfies \eqref{logcorrY} and \eqref{l2distY}, by definition. \begin{prop} \label{rtcomparison} For $\epsilon>0$, let $\left(\xi_{\epsilon}^{v}:v\in V_{\epsilon}\right)$ and $\left(\psi_{\epsilon}^{x}:x\in[0,1)^{d}\right)$ be independent fields, defined as in \eqref{defMBRW} and \eqref{defBS}, respectively. Then, there exist $\delta>0$ small enough and $p$ large enough (depending on $C_{Y}$ and $d$) such that \[ \mathbb{P}\left(\sup_{x\in[0,1)^{d}}Y_{\delta\epsilon}^{\delta x}\geq\lambda\right)\leq\mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\geq\lambda\right) \] for all $\epsilon>0$ and all $\lambda\in\mathbb{R}$, where $a(x):=\sqrt{\left(Var(Y_{\delta\epsilon}^{\delta x})-Var(\psi_{\epsilon}^{x})\right)/Var(\xi_{\epsilon}^{[x]})}$.\end{prop} \begin{proof} We check the hypotheses of Slepian's Lemma (see \cite[Theorem 2.2.1]{rf}). The variance of the fields $Y_{\delta\epsilon}^{\delta x}$ and $a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}$ are equal by the definition of $a(x)$. We first choose $p$ so that $a(x)\leq1$. Note that \eqref{logcorrY} and \eqref{varBS} imply \[ a(x)^{2}=\frac{Var(Y_{\delta\epsilon}^{\delta x})-Var(\psi_{\epsilon}^{x})}{Var(\xi_{\epsilon}^{[x]})}\leq\frac{\log(1/\epsilon)+\log(1/\delta)+C_{Y}-p^{d}}{\log(1/\epsilon)}, \] so, by choosing $p$ large enough (depending on $C_{Y}$, $d$ and $\delta$), we obtain $a(x)\leq1$, for all $x$. We now compare the covariance for points $x\neq y$, for which we distinguish two cases: 1. $[x]=[y]$ (that is, $\square_{\epsilon}^{[x]}=\square_{\epsilon}^{[y]}$). In this case, \eqref{l2distY} and \eqref{l2distBS} imply \[ \mathbb{E}\left[\left(Y_{\delta\epsilon}^{\delta x}-Y_{\delta\epsilon}^{\delta y}\right)^{2}\right]\leq C_{Y}(\delta\epsilon)^{-1}\left\Vert \delta x-\delta y\right\Vert \leq p^{d}\epsilon^{-1}\left\Vert x-y\right\Vert _{1}\leq\mathbb{E}\left[\left(\psi_{\epsilon}^{x}-\psi_{\epsilon}^{y}\right)^{2}\right] \] \[ \leq\mathbb{E}\left[\left(a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}-a(y)\xi_{\epsilon}^{[y]}-\psi_{\epsilon}^{y}\right)^{2}\right] \] for $p$ large enough (depending on $C_{Y}$). The last inequality is due the independence between $\xi_{\epsilon}$ and $\psi_{\epsilon}$. 2. $[x]\neq[y]$. In this case, we can apply \eqref{covMBRW} and the independence between $\xi_{\epsilon}$, $\psi_{\epsilon}^{[x]}$ and $\psi_{\epsilon}^{[y]}$ to obtain \[ Cov(a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x},a(y)\xi_{\epsilon}^{[y]}+\psi_{\epsilon}^{y})\leq a(x)a(y)Cov(\xi_{\epsilon}^{[x]},\xi_{\epsilon}^{[y]})\leq a(x)a(y)\left(-\log\left\Vert [x]-[y]\right\Vert +C\right). \] But $a(x)a(y)\leq1$, so \[ Cov(a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x},a(y)\xi_{\epsilon}^{[y]}+\psi_{\epsilon}^{y})\leq-\log\left\Vert [x]-[y]\right\Vert +C. \] Note that $-\log\left\Vert [x]-[y]\right\Vert \leq-\log\max\left\{ \epsilon,\left\Vert x-y\right\Vert \right\} +C$. Applying \eqref{logcorrY}, we obtain \[ -\log\max\left\{ \epsilon,\left\Vert x-y\right\Vert \right\} +C\leq-\log\max\left\{ \delta\epsilon,\left\Vert \delta x-\delta y\right\Vert \right\} -C_{Y}\leq Cov(Y_{\delta\epsilon}^{\delta x},Y_{\delta\epsilon}^{\delta y}) \] for some $\delta>0$ small enough (depending on $C_{Y}$ and $d$). Proposition \ref{rtcomparison} follows now from Slepian's Lemma. \end{proof} Proposition \ref{rtcomparison} provides an upper bound for the right tail of the supremum of $Y_{\delta\epsilon}$ taken over the $\delta$-box $\delta[0,1)^{d}$. The same proof works for any $\delta$-box. Therefore, a union bound implies \begin{equation} \mathbb{P}\left(\sup_{x\in[0,1)^{d}}Y_{\delta\epsilon}^{x}\geq\lambda\right)\leq\left(\frac{1}{\delta}\right)^{d}\mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\geq\lambda\right)\label{union} \end{equation} for all $\lambda\in\mathbb{R}$. We now provide an upper bound for the probability on the right hand side of the previous display. We first prove an upper bound on the supremum of the Brownian sheet. \begin{lem} \label{rtubBS} There exist constants $0<c,C<\infty$ (depending on $p$ and $d$) such that \[ \sup_{v\in V_{\epsilon}}\mathbb{P}\left(\sup_{x\in\square_{\epsilon}^{v}}\psi_{\epsilon}^{x}\geq\lambda\right)\leq Ce^{-c\lambda^{2}} \] for all $\lambda\geq0,\epsilon>0$.\end{lem} \begin{proof} Let $v\in V_{\epsilon}$. Fernique's Majorizing Criterion (see \cite[Theorem 4.1]{ce}) implies that \[ \mathbb{E}\left[\sup_{x\in\square_{\epsilon}^{v}}\psi_{\epsilon}^{x}\right]\leq C\sup_{x\in\square_{\epsilon}^{v}}\int_{0}^{\infty}\sqrt{-\log\left(\mu(B(x,r))\right)}dr \] for some absolute constant $C$, where $\mu$ is the normalized $d$-dimensional Lebesgue measure on $\square_{\epsilon}^{v}$ and $B(x,r)=\left\{ y\in\square_{\epsilon}^{v}:\mathbb{E}\left[\left(\psi_{\epsilon}^{x}-\psi_{\epsilon}^{y}\right)^{2}\right]\leq r^{2}\right\} $. But \eqref{l2distBS} implies \[ B(x,r)\supset\left\{ y\in\square_{\epsilon}^{v}:(2p)^{d}\epsilon^{-1}\left\Vert y-x\right\Vert _{1}\leq r^{2}\right\} . \] Therefore, $\mu\left(B(x,r)\right)\geq cr^{2d}$ for some constant $c>0$ depending on $p$ and $d$. Applying the previous display and Fernique's Majorizing Criterion, we obtain \[ \mathbb{E}\left[\sup_{x\in\square_{\epsilon}^{v}}\psi_{\epsilon}^{x}\right]\leq C\int_{0}^{\infty}\sqrt{-\log\left(cr^{2d}\right)}dr\leq C<\infty, \] where $C$ depends on $p$ and $d$. Borell's Inequality (see \cite[Theorem 2.1.1]{rf}) and \eqref{varBS} imply \[ \mathbb{P}\left(\sup_{x\in\square_{\epsilon}^{v}}\psi_{\epsilon}^{x}\geq C+\lambda\right)\leq e^{-\lambda^{2}/(2(2p)^{d})}, \] where $C$ is the constant obtained in the previous display. Lemma \ref{rtubBS} now follows from a change of variables.\end{proof} \begin{prop} \label{rtcomparison2} Let $p$ be as in Proposition \ref{rtcomparison}. There exist constants $0<c,C<\infty$ (depending on $p$ and $d$) such that \[ \mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\geq\lambda+m_{\epsilon}\right)\leq Ce^{-c\lambda} \] for all $\lambda\geq0$ and all $\epsilon>0$ small enough.\end{prop} \begin{proof} By letting $\psi_{\epsilon}^{\ast,[x]}=\sup_{y\in\square_{\epsilon}^{[x]}}\psi_{\epsilon}^{y}$, we have \[ \sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\leq\max_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{\ast,[x]}. \] The previous display implies \[ \sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\geq m_{\epsilon}+\lambda\implies\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{\ast,[x]}\geq m_{\epsilon}+\lambda. \] We now compute an upper bound for the right hand side of the previous display. Define the random sets $\Gamma_{y}=\left\{ v\in V_{\epsilon}:\psi_{\epsilon}^{\ast,v}\in[y-1,y)\right\} $ for $y\geq1$, and $\Gamma_{0}=\left\{ v\in V_{\epsilon}:\psi_{\epsilon}^{\ast,v}\leq0\right\} $. Note that \[ \mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{\ast,[x]}\geq m_{\epsilon}+\lambda\right)\leq\sum_{y\geq0}\mathbb{P}\left(\sup_{x:[x]\in\Gamma_{y}}a(x)\xi_{\epsilon}^{[x]}\geq m_{\epsilon}+\lambda-y\right). \] The definition of $a(x)$ easily implies that $1/2\leq a(x)\leq1$, for $\epsilon>0$ small enough. Therefore, the last display implies \begin{equation} \mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{\ast,[x]}\geq m_{\epsilon}+\lambda\right)\leq\sum_{y\geq0}\mathbb{P}\left(\max_{v\in\Gamma_{y}}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda-2y\right).\label{randomsets} \end{equation} But $\mathbb{P}\left(\max_{v\in\Gamma_{y}}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda-2y\right)=\mathbb{E}\left[\mathbb{P}\left(\max_{v\in\Gamma_{y}}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda-2y\mid\Gamma_{y}\right)\right]$. Since $\psi_{\epsilon}$ and $\xi_{\epsilon}$ are independent, from \eqref{rtubMBRW} we obtain, \[ \mathbb{P}\left(\max_{v\in\Gamma_{y}}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda-2y\mid\Gamma_{y}\right)\leq C\left(\epsilon^{d}\left|\Gamma_{y}\right|\right)^{1/2}e^{-c(\lambda-2y)}. \] Then, \begin{equation} \mathbb{P}\left(\max_{v\in\Gamma_{y}}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda-2y\right)\leq Ce^{-c(\lambda-2y)}\left(\mathbb{E}\left[\epsilon^{d}\left|\Gamma_{y}\right|\right]\right)^{1/2}.\label{randomsets2} \end{equation} But, by Lemma \ref{rtubBS}, $\mathbb{E}\left[\left|\Gamma_{y}\right|\right]=\sum_{v\in V_{\epsilon}}\mathbb{P}\left(\psi_{\epsilon}^{\ast,v}\in[y-1,y)\right)\leq C\epsilon^{-d}e^{-cy^{2}}$. For $y=0$, we simply use $\left|\Gamma_{0}\right|\leq\epsilon^{-d}$. Therefore, from displays \eqref{randomsets} and \eqref{randomsets2}, we obtain \[ \mathbb{P}\left(\sup_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{\ast,[x]}\geq m_{\epsilon}+\lambda\right)\leq Ce^{-c\lambda} \] for some constants $0<c,C<\infty$ (depending on $p$ and $d$). \end{proof} \begin{proof} [Proof of Theorem \ref{main}, \eqref{rtubY}, the right tail] Display \eqref{union} and Proposition \ref{rtcomparison2} imply \[ \mathbb{P}\left(\max_{x\in[0,1)^{d}}Y_{\delta\epsilon}^{x}\geq m_{\epsilon}+\lambda\right)\leq\left(\frac{1}{\delta}\right)^{2}\mathbb{P}\left(\max_{x\in[0,1)^{d}}a(x)\xi_{\epsilon}^{[x]}+\psi_{\epsilon}^{x}\geq\lambda+m_{\epsilon}\right)\leq Ce^{-c\lambda}. \] It is easy to see from the definition that $m_{\delta\epsilon}\leq m_{\epsilon}+C'$ for some $C'$ depending on $\delta$ and $d$. Therefore, \[ \mathbb{P}\left(\max_{x\in[0,1)^{d}}Y_{\delta\epsilon}^{x}\geq m_{\delta\epsilon}+\lambda-C'\right)\leq Ce^{-c\lambda}. \] The upper bound \eqref{rtubY} for the right tail follows by adjusting the constants. \end{proof} \subsection{The left tail} In this subsection we prove the upper bound \eqref{rtubY} for the left tail. As previously mentioned, we can reduce the set under maximization to a discrete set. More precisely, if $\left\{ D_{\epsilon}:\epsilon>0\right\} $ is any collection of subsets of $[0,1)^{d}$, then \begin{equation} \mathbb{P}\left(\sup_{x\in[0,1)^{d}}Y_{\epsilon}^{x}\leq m_{\epsilon}-\lambda\right)\leq\mathbb{P}\left(\sup_{x\in D_{\epsilon}}Y_{\epsilon}^{x}\leq m_{\epsilon}-\lambda\right).\label{discrete} \end{equation} If we select $D_{\epsilon}$ appropriately, we can perform a comparison with the MBRW using Slepian's Lemma. \begin{prop} \label{ltcomparison} There exist $\delta,\rho>0$ small enough (depending on $C_{Y}$ and $d$) such that \[ \mathbb{P}\left(\max_{u\in V_{\epsilon/\rho}}Y_{\delta\epsilon}^{u}\leq\lambda\right)\leq\mathbb{P}\left(\max_{u\in V_{\epsilon}\cap\rho[0,1)^{d}}b(u)\xi_{\epsilon}^{u}\leq\lambda\right) \] for all $\epsilon>0$ and all $\lambda\in\mathbb{R}$, where $b(u):=\sqrt{Var(Y_{\delta\epsilon}^{u})/Var(\xi_{\epsilon}^{u})}$, for $u\in V_{\epsilon/\rho}$.\end{prop} \begin{proof} Note that \eqref{logcorrY} and \eqref{covMBRW} imply that $b(u)\geq\frac{\log(1/\epsilon)+\log(1/\delta)-C_{Y}}{\log(1/\epsilon)}$, which is greater than 1 for $\delta>0$ small enough (depending on $C_{Y}$). Let $u,v\in V_{\epsilon/\rho}$, with $u\neq v$. Then, $\left\Vert u-v\right\Vert \geq\epsilon/\rho\geq\delta\epsilon$. Display \eqref{logcorrY} therefore implies \[ Cov(Y_{\delta\epsilon}^{u},Y_{\delta\epsilon}^{v})\leq-\log\left\Vert u-v\right\Vert +C_{Y}. \] Choose $\rho>0$ small enough so that \[ -\log\left\Vert u-v\right\Vert +C_{Y}\leq-\log\left\Vert \rho u-\rho v\right\Vert -C\leq Cov(\xi_{\epsilon}^{\rho u},\xi_{\epsilon}^{\rho v}), \] where the last bound follows from \eqref{covMBRW}. All the hypotheses of Slepian's Lemma are satisfied, so \[ \mathbb{P}\left(\max_{u\in V_{\epsilon/\rho}}Y_{\delta\epsilon}^{u}\leq\lambda\right)\leq\mathbb{P}\left(\max_{u\in V_{\epsilon/\rho}}b(u)\xi_{\epsilon}^{\rho u}\leq\lambda\right) \] for all $\lambda\in\mathbb{R}$. Proposition \ref{ltcomparison} follows by observing that $\rho V_{\epsilon/\rho}=V_{\epsilon}\cap\rho[0,1]^{2}$.\end{proof} \begin{prop} \label{ltcomparison2} Let $\rho>0$ and $\left\{ b(u):u\in V_{\epsilon}\cap\rho[0,1]^{2}\right\} $ be as in Proposition \ref{ltcomparison}. Then, \[ \mathbb{P}\left(\max_{u\in V_{\epsilon}\cap\rho[0,1]^{2}}b(u)\xi_{\epsilon}^{u}\leq m_{\epsilon}-\lambda\right)\leq\mathbb{P}\left(\max_{u\in V_{\epsilon}\cap\rho[0,1]^{2}}\xi_{\epsilon}^{u}\leq m_{\epsilon}-\lambda/2\right) \] for all $\lambda\geq0$ and all $\epsilon>0$ small enough.\end{prop} \begin{proof} It follows from the definition of $b(u)$ that, for small enough $\epsilon>0$, \[ 1\leq b(u)\leq2 \] for all $u$. Let $\nu$ be the (a.s. well-defined) point that maximizes $\xi_{\epsilon}^{u}$, for $u\in V_{\epsilon}\cap\rho[0,1]^{2}$. Then, \[ b(\nu)\xi_{\epsilon}^{\nu}\leq m_{\epsilon}-\lambda\implies\xi_{\epsilon}^{\nu}\leq m_{\epsilon}/b(\nu)-\lambda/b(\nu)\leq m_{\epsilon}-\lambda/2. \] \end{proof} Our task is now to find an upper bound for the probability on the right hand side of Proposition \ref{ltcomparison2}. \begin{prop} \label{ltcompare3} There exist constants $0<c,C<\infty$ (depending on $\rho$ and $d$) such that \[ \mathbb{P}\left(\max_{v\in V_{\epsilon}\cap\rho[0,1]^{2}}\xi_{\epsilon}^{v}\leq m_{\epsilon}-\lambda\right)\leq Ce^{-c\lambda} \] for all $\lambda\geq0$ and all $\epsilon>0$ small enough.\end{prop} \begin{proof} Assume $0\leq k\leq\log\left(1/\epsilon\right)/2$, where $k$ is a large number, that will be chosen later. Let $\left\{ B^{i}:i=1,\ldots,ce^{k}\right\} $ (where $c>0$ is a small constant, depending on $\rho$) be a collection of boxes of side length $e^{-k}$ inside $\rho[0,1)^{d}$, such that the distance between any pair of boxes is at least $e^{-k}$. Set $B_{\epsilon}^{i}=B^{i}\cap V_{\epsilon}$. We claim that the field \[ \left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in B_{\epsilon}^{i}\right) \] is a copy of $\left(\xi_{\epsilon e^{k}}^{v}:v\in V_{\epsilon e^{k}}\right)$, and that the fields $\left\{ \left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in B_{\epsilon}^{i}\right)\right\} _{i\leq ce^{k}}$ are independent. Indeed, if $v,u\in B_{\epsilon}^{i}$, then \eqref{defMBRW} implies \begin{equation} Cov(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k),\xi_{\epsilon}^{u}-\xi_{\epsilon}^{u}(k))=\int_{k}^{\log(1/\epsilon)}\prod_{j\leq d}\left(1-e^{r}\left|v_{j}-u_{j}\right|\right)_{+}dr\label{copies} \end{equation} \[ =\int_{0}^{-\log(\epsilon e^{k})}\prod_{j\leq d}\left(1-e^{r}\left|e^{k}v_{j}-e^{k}u_{j}\right|\right)_{+}dr, \] and the set $e^{k}B_{\epsilon}^{i}=\left\{ e^{k}v:v\in B_{\epsilon}^{i}\right\} $ coincides with $V_{\epsilon e^{k}}$ after a translation. This shows that $\left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in B_{\epsilon}^{i}\right)\overset{d}{=}\left(\xi_{\epsilon e^{k}}^{v}:v\in V_{\epsilon e^{k}}\right)$. Moreover, from \eqref{copies}, it is easy to see that $\left\Vert v-u\right\Vert \geq e^{-k}$ (which is true for points $v,u$ in different boxes $B_{\epsilon}^{i}$, by construction) implies \[ Cov(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k),\xi_{\epsilon}^{u}-\xi_{\epsilon}^{u}(k))=0, \] as desired. Therefore, independence of the fields $\left\{ \left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in B_{\epsilon}^{i}\right)\right\} _{i\leq ce^{k}}$ and \eqref{rtlbMBRW} imply \[ \mathbb{P}\left(\max_{v\in\bigcup_{i}B_{\epsilon}^{i}}\left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k)\right)\leq m_{\epsilon e^{k}}\right)\leq e^{-ce^{k}} \] for some constant $c>0$ depending on $d$ and $\rho$. By letting $\nu=\arg\max\left\{ \xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in\bigcup_{i}B_{\epsilon}^{i}\right\} $, the previous display implies \[ \mathbb{P}\left(\max_{v\in V_{\epsilon}\cap\rho[0,1)^{d}}\xi_{\epsilon}^{v}\leq m_{\epsilon}-\lambda\right)\leq\mathbb{P}\left(\xi_{\epsilon}^{\nu}\leq m_{\epsilon}-\lambda\right)\leq\mathbb{P}\left(\xi_{\epsilon}^{\nu}(k)\leq m_{\epsilon}-m_{\epsilon e^{k}}-\lambda\right)+\mathbb{P}\left(\xi_{\epsilon}^{\nu}-\xi_{\epsilon}^{\nu}(k)\leq m_{\epsilon e^{k}}\right) \] \[ \leq\mathbb{P}\left(\xi_{\epsilon}^{\nu}(k)\leq m_{\epsilon}-m_{\epsilon e^{k}}-\lambda\right)+e^{-ce^{k}}. \] Moreover, it is clear from \eqref{defMBRW} that the fields $\left(\xi_{\epsilon}^{v}-\xi_{\epsilon}^{v}(k):v\in V_{\epsilon}\right)$ and $\left(\xi_{\epsilon}^{v}(k):v\in V_{\epsilon}\right)$ are independent. Hence, $\nu$ is independent from $\xi_{\epsilon}^{(\cdot)}(k)$, and $\xi_{\epsilon}^{\nu}(k)$ is therefore a Gaussian random variable with mean zero and variance $k$. But \[ m_{\epsilon}-m_{\epsilon e^{k}}\leq\sqrt{2d}k. \] Therefore, by choosing $k=\log\lambda$, the last two displays imply \[ \mathbb{P}\left(\max_{v\in V_{\epsilon}\cap\rho[0,1)^{d}}\xi_{\epsilon}^{v}\leq m_{\epsilon}-\lambda\right)\leq Ce^{-c\frac{\left(\lambda-\sqrt{2d}\log\lambda\right)^{2}}{\log\lambda}}+e^{-c\lambda}\leq Ce^{-c\lambda}, \] proving Proposition \ref{ltcompare3} in the case $k=\log\lambda\leq\log(1/\epsilon)/2$. On the other hand, for $\lambda\geq\sqrt{1/\epsilon}$, \[ \mathbb{P}\left(\max_{v\in V_{\epsilon}\cap\rho[0,1)^{d}}\xi_{\epsilon}^{v}\leq m_{\epsilon}-\lambda\right)\leq\mathbb{P}\left(\xi_{\epsilon}^{v}\leq m_{\epsilon}-\lambda\right)\leq Ce^{-c\frac{\left(\lambda-m_{\epsilon}\right)^{2}}{\log\left(1/\epsilon\right)}}\leq Ce^{-c\lambda} \] (where $v$ is any point), which implies Proposition \ref{ltcompare3} in this case. \end{proof} Using Propositions \ref{ltcomparison}, \ref{ltcomparison2} and \ref{ltcompare3}, we are now ready to finish the proof of Theorem \ref{main}. \begin{proof} [Proof of \ref{main}, \eqref{rtubY}, the left tail] Propositions \ref{ltcomparison}, \ref{ltcomparison2} and \ref{ltcompare3} imply the existence of constants $0<\delta,\rho,c,C<\infty$, depending on $C_{Y}$ and $d$, such that \[ \mathbb{P}\left(\max_{u\in V_{\epsilon/\rho}}Y_{\delta\epsilon}^{u}\leq m_{\epsilon}-\lambda\right)\leq Ce^{-c\lambda} \] for all $\lambda\geq0$. But $m_{\delta\epsilon}\leq m_{\epsilon}+C'$, where $C'$ depends on $\delta$ and $d$. Therefore, \[ \mathbb{P}\left(\max_{u\in V_{\epsilon/\rho}}Y_{\delta\epsilon}^{u}\leq m_{\delta\epsilon}-\lambda-C'\right)\leq Ce^{-c\lambda}. \] The bound \eqref{rtubY} for the left tail follows by adjusting the constants. \end{proof} \section{Example: a mollified Gaussian free field in $d=2$} The Gaussian free field in two dimensions provides an important example of a log-correlated field. Intuitively speaking, the reason for the log-correlation is simply that, in $d=2$, the Green function for the Laplacian is logarithmic. We begin by recalling in Section 3.1 the definitions of the Dirichlet product and the Hilbert space induced by it. We then use this Hilbert space to define the continuous Gaussian free field and the mollified Gaussian free field. After that, we prove some useful properties of these fields, which will be used to check the hypotheses of Theorem \ref{main}. Finally, in section 3.2, we use Theorem \ref{main} to prove tightness of the recentered maximum of the family of mollified Gaussian free fields. \subsection{Continuous and mollified Gaussian free fields} \subsubsection*{Dirichlet product} We begin by recalling the definition of the Dirichlet product. Let $C_{c}^{\infty}\left((0,1)^{2}\right)$ denote the set of real valued $C^{\infty}$ functions with compact support in $(0,1)^{2}$. For $\phi,\psi\in C_{c}^{\infty}\left((0,1)^{2}\right)$, let \[ \langle\phi,\psi\rangle_{\nabla}=\int\nabla\phi(x)\nabla\psi(x)dx \] denote the Dirichlet product, where $\nabla$ is the gradient and $dx$ is two-dimensional Lebesgue measure. Note that the Dirichlet product satisfies \begin{equation} \langle\phi,\psi\rangle_{\nabla}=\int\phi(x)(-\Delta\psi)(x)dx,\label{parts} \end{equation} where $\Delta$ is the standard Laplacian. The Dirichlet product induces a norm on $C_{c}^{\infty}\left((0,1)^{2}\right)$ by \[ \left\Vert \phi\right\Vert _{\nabla}=\sqrt{\langle\phi,\phi\rangle_{\nabla}}, \] called the Dirichlet norm. Denote by $W=W\left((0,1)^{2}\right)$ the completion of $C_{c}^{\infty}\left((0,1)^{2}\right)$ with respect to the Dirichlet norm. The set $W$, together with the Dirichlet product on $W$, defines a Hilbert space. The Dirichlet norm satisfies Poincare's Inequality: there exists a constant $C$ (which depends only on the domain $(0,1)^{2}$) such that \[ \left\Vert \phi\right\Vert _{L^{2}}\leq C\left\Vert \nabla\phi\right\Vert _{L^{2}} \] for all $\phi\in C_{c}^{\infty}$. Poincare's Inequality implies that the Dirichlet norm is equivalent to the norm \[ \left\Vert \phi\right\Vert _{L^{2}}+\left\Vert \frac{\partial}{\partial x_{1}}\phi\right\Vert _{L^{2}}+\left\Vert \frac{\partial}{\partial x_{2}}\phi\right\Vert _{L^{2}}. \] Recall that the completion of $C_{c}^{\infty}\left((0,1)^{2}\right)$ with respect to the latter norm is called a $(1,2)$-Sobolev space (i.e., measurable functions such that their weak derivatives up to order 1 exist and belong to $L^{2}\left((0,1)^{2}\right))$. Since the norms are equivalent, the space $\left(W,\left\Vert \cdot\right\Vert _{\nabla}\right)$ is also a Sobolev space. Therefore, for any $g\in W$ and any measurable set $E\subset[0,1]^{2}$, the integral $\int_{E}g(x)dx$ is well-defined. For a given open set $U\subset(0,1)^{2}$, Poincare's Inequality implies that the linear mapping $W\to\mathbb{R}$ given by \[ g\mapsto\int_{U}g(x)dx \] is $\left\Vert \cdot\right\Vert _{\nabla}$-continuous. Note that, since $W$ is a Hilbert space, the Riesz representation theorem implies the existence of a function $f=f_{U}\in W$ such that \begin{equation} \langle g,f_{U}\rangle_{\nabla}=\int_{U}g(x)dx\label{riesz} \end{equation} for all $g\in W$. \subsubsection*{Gaussian free fields} The continuous Gaussian free field is defined as follows: since $\langle\cdot,\cdot\rangle_{\nabla}$ is positive definite, there exists a family $\left\{ X^{f}:f\in W\right\} $ of centered Gaussian variables, defined on some probability space $\left(\Omega,\mathbb{P}\right)$, such that \[ Cov(X^{f},X^{g})=\langle f,g\rangle_{\nabla}. \] The family $\left\{ X^{f}:f\in W\right\} $ is called the \emph{continuous Gaussian free field}. We next define a field indexed by the set $[0,1]^{2}$. Fix $\epsilon>0$, and let $x\in[0,1]^{2}$. By \eqref{riesz}, there exists a function $f_{x,\epsilon}\in W$ such that \begin{equation} \langle f_{x,\epsilon},g\rangle_{\nabla}=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap(0,1)^{2}}g(u)du\label{def} \end{equation} for all $g\in W$, where $D(x,\epsilon)$ is the disk of radius $\epsilon$ centered at $x$. Using \eqref{parts} and \eqref{def}, it is not hard to show that \begin{equation} f_{x,\epsilon}(u)=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap(0,1)^{2}}G(u,v)dv,\label{GRRREN} \end{equation} where $G=G_{(0,1)^{2}}$ is the Green function of $\left(0,1\right)^{2}$ for the operator $-\Delta$, with Dirichlet boundary conditions on $\partial\left(0,1\right)^{2}$. For the domain $\left(0,1\right)^{2}$, the Green function can be explicitly stated as: \[ G(u,v)=\frac{4}{\pi^{2}}\sum_{n,m\geq1}\frac{1}{n^{2}+m^{2}}\sin\left(n\pi u_{1}\right)\sin\left(m\pi u_{2}\right)\sin\left(n\pi v_{1}\right)\sin\left(m\pi v_{2}\right), \] where $u=(u_{1},u_{2})\in[0,1]^{2}$. The field $\left(X^{f_{x,\epsilon}}:x\in[0,1]^{2}\right)$ will be called \emph{$\epsilon$-mollified Gaussian free field} (MGFF). To simplify notation, set $X_{\epsilon}^{x}=X^{f_{x,\epsilon}}$. Note that, by definition, \[ Cov(X_{\epsilon}^{x},X_{\epsilon}^{y})=\langle f_{x,\epsilon},f_{y,\epsilon}\rangle_{\nabla}=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap(0,1)^{2}}f_{y,\epsilon}(u)du \] and, from \eqref{GRRREN}, we obtain \begin{equation} Cov(X_{\epsilon}^{x},X_{\epsilon}^{y})=\frac{1}{\left(\pi\epsilon^{2}\right)^{2}}\int_{D(x,\epsilon)\cap(0,1)^{2}\times D(y,\epsilon)\cap(0,1)^{2}}G(u,v)\, dudv,\label{green} \end{equation} for all $x,y\in[0,1]^{2}$. \subsubsection*{Orthogonal decomposition} The next proposition shows that the MGFF satisfies a tree-like decomposition property. \begin{prop} \label{prop3.1} Let $Q=\frac{1}{2}(0,1)^{2}\subset(0,1)^{2}$ be a sub-square of side length $1/2$. Then, $X_{\epsilon}^{x}$ can be decomposed as \[ X_{\epsilon}^{x}=\hat{X}_{\epsilon}^{x}+\phi^{x}, \] where $\left(\hat{X}_{\epsilon}^{x}:x\in\overline{Q}\right)$ is a copy of $\left(X_{2\epsilon}^{x}:x\in[0,1]^{2}\right)$, and $\left(\hat{X}_{\epsilon}^{x}:x\in\overline{Q}\right)$ is independent of $\left(\phi^{x}:x\in[0,1]^{2}\right)$.\end{prop} \begin{proof} Denote by $C_{c}^{\infty}\left(Q\right)$ the set of real valued $C^{\infty}$ functions with compact support in $Q$, and let $W(Q)$ be the corresponding Hilbert space induced by the Dirichlet product in $C_{c}^{\infty}\left(Q\right)$. Note that $C_{c}^{\infty}\left(Q\right)\subset C_{c}^{\infty}\left((0,1)^{2}\right)$ and \begin{equation} \langle f,g\rangle_{\nabla,Q}:=\int_{Q}\nabla f(u)\cdot\nabla g(u)\, du=\int_{(0,1)^{2}}\nabla f(u)\cdot\nabla g(u)\, du\label{isom} \end{equation} for all $f,g\in C_{c}^{\infty}(Q)$. By taking the completion of $C_{c}^{\infty}\left(Q\right)$ with respect to the Dirichlet product, we see that $W(Q)$ is a Hilbert subspace of $W\left((0,1)^{2}\right)$ and that \eqref{isom} holds for all $f,g\in W(Q)$. Let $f_{x,\epsilon}$ be as in \eqref{def} and decompose it as \[ f_{x,\epsilon}=g_{x,\epsilon}+h_{x,\epsilon}, \] where $g_{x,\epsilon}\in W(Q)$ and $h_{x,\epsilon}\in W(Q)^{\perp}$ (the orthogonal space). Set \[ \hat{X}_{\epsilon}^{x}=X^{g_{x,\epsilon}} \] and \[ \phi^{x}=X^{h_{x,\epsilon}}. \] Since $g_{x,\epsilon}\perp h_{y,\epsilon}$ for all $x,y\in[0,1]^{2}$, the families $\left(\hat{X}_{\epsilon}^{x}:x\in[0,1]^{2}\right)$ and $\left(\phi^{x}:x\in[0,1]^{2}\right)$ are independent. Also, since $f\mapsto X^{f}$ is a linear embedding of $W$ into $L^{2}\left(\Omega,\mathbb{P}\right)$, \[ X_{\epsilon}^{x}=\hat{X}_{\epsilon}^{x}+\phi^{x}\text{ a.s.} \] for every $x\in[0,1]^{2}$. We show now that $\left(\hat{X}_{\epsilon}^{x}:x\in\overline{Q}\right)$ is a copy of $\left(X_{2\epsilon}^{x}:x\in[0,1]^{2}\right)$. \begin{claim} For every $k\in W(Q)$, \[ \langle g_{x,\epsilon},k\rangle_{\nabla,Q}=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap Q}k(u)\, du. \] \end{claim} \begin{proof} [Proof of Claim 3.2] By \eqref{isom}, \[ \langle g_{x,\epsilon},k\rangle_{\nabla,Q}=\langle g_{x,\epsilon},k\rangle_{\nabla} \] and, since $g_{x,\epsilon}=f_{x,\epsilon}-h_{x,\epsilon}$, \[ \langle g_{x,\epsilon},k\rangle_{\nabla}=\langle f_{x,\epsilon},k\rangle_{\nabla}-\langle h_{x,\epsilon},k\rangle_{\nabla}. \] But $h_{x,\epsilon}\perp k$, so the second term on the right hand side of the previous display vanishes. Using \eqref{def} and the two previous displays, we obtain \[ \langle g_{x,\epsilon},k\rangle_{\nabla,Q}=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap(0,1)^{2}}k(u)\, du. \] Since $k\in W(Q)$, the function $k$ vanishes outside of $Q$. Therefore, \[ \langle g_{x,\epsilon},k\rangle_{\nabla,Q}=\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)\cap Q}k(u)\, du, \] as desired. \end{proof} Claim 3.2 implies, in analogy with \eqref{green}, that the following is true for all $x,y\in\overline{Q}$: \[ Cov\left(\hat{X}_{\epsilon}^{x},\hat{X}_{\epsilon}^{y}\right)=\langle g_{x,\epsilon},g_{y,\epsilon}\rangle_{\nabla}=\langle g_{x,\epsilon},g_{y,\epsilon}\rangle_{\nabla,Q}=\frac{1}{\left(\pi\epsilon^{2}\right)^{2}}\int_{D(x,\epsilon)\cap Q\times D(y,\epsilon)\cap Q}G_{Q}(u,v)\, dudv, \] where $G_{Q}$ is the Green function of $Q$ for the operator $-\Delta$, with Dirichlet boundary conditions on $\partial Q$. \begin{claim} For every $u,v\in[0,1]^{2}$, \[ G_{Q}\left(u/2,v/2\right)=G(u,v). \] \end{claim} \begin{proof} [Proof of Claim 3.3]Let $\phi\in C_{c}^{\infty}\left(\left(0,1\right)^{2}\right)$ and note that $\left(\Delta\phi\right)(2u)=\frac{1}{4}\Delta\left(\phi(2u)\right)$. By the change of variables $u'=u/2$, \[ \int_{(0,1)^{2}}G_{Q}\left(u/2,v/2\right)\left(\Delta\phi\right)(u)\, du=\int_{Q}G_{Q}\left(u',v/2\right)\Delta\left(\phi\left(2u'\right)\right)\, du'=-\phi(2v/2), \] where the last equality holds by definition of $G_{Q}$. On the other hand, \[ \int_{(0,1)^{2}}G(u,v)\left(\Delta\phi\right)(u)\, du=-\phi(v), \] by definition of $G$. Since \[ \int_{(0,1)^{2}}G_{Q}\left(u/2,v/2\right)\left(\Delta\phi\right)(u)\, du=\int_{(0,1)^{2}}G(u,v)\left(\Delta\phi\right)(u)\, du \] for every $\phi\in C_{c}^{\infty}\left((0,1)^{2}\right)$, the functions $G_{Q}(u/2,v/2)$ and $G(u,v)$ are identical (Lebesgue-a.e.). \end{proof} The change of variables $u'=2u,v'=2v$ implies \[ Cov\left(\hat{X}_{\epsilon}^{x},\hat{X}_{\epsilon}^{y}\right)=\frac{1}{\left(\pi\epsilon^{2}\right)^{2}}\int_{D(x,\epsilon)\cap Q\times D(y,\epsilon)\cap Q}G_{Q}(u,v)\, dudv \] \[ =\frac{1}{\left(\pi(2\epsilon)^{2}\right)^{2}}\int_{D(2x,2\epsilon)\cap(0,1)^{2}\times D(2y,2\epsilon)\cap(0,1)^{2}}G_{Q}(u'/2,v'/2)\, du'dv', \] and Claim 3.3 implies that the previous display is \[ =\frac{1}{\left(\pi(2\epsilon)^{2}\right)^{2}}\int_{D(2x,2\epsilon)\cap(0,1)^{2}\times D(2y,2\epsilon)\cap(0,1)^{2}}G(u',v')\, du'dv'=Cov\left(X_{2\epsilon}^{2x},X_{2\epsilon}^{2y}\right). \] For Gaussian fields, equality of the covariance structure implies that the fields have the same distribution. Therefore, \[ \left(\hat{X}_{\epsilon}^{x}:x\in\overline{Q}\right)\overset{d}{=}\left(X_{2\epsilon}^{2x}:x\in\overline{Q}\right), \] and the right hand side is clearly equal to $\left(X_{2\epsilon}^{x}:x\in[0,1]^{2}\right)$, which finishes the proof of Proposition 3.1. \end{proof} Proposition 3.1 is true for any sub-square $Q\subset(0,1)^{2}$ of side length $1/2$, because Green functions are translation invariant (i.e., $G_{Q+z}\left(u+z,v+z\right)=G_{Q}(u,v)$ for any $z\in\mathbb{R}^{2},u,v\in Q$, where $G_{Q+z}$ is the Green function of $Q+z$ for the operator $-\Delta$, with Dirichlet boundary conditions on $\partial Q+z$). \subsubsection*{Estimates on the covariance} In this subsection, we prove that the ``bulk'' of the field $\left\{ \sqrt{\frac{\pi}{2}}X_{\epsilon}^{x}:x\in[0,1]^{2}\right\} $ satisfies both \eqref{logcorrY} and \eqref{l2distY}. Recall that $\Gamma(\cdot,\cdot)=\Gamma\left(\left\Vert \cdot-\cdot\right\Vert \right)=\frac{2}{\pi}\log(1/\left\Vert \cdot-\cdot\right\Vert )$ is the Green function of $\mathbb{R}^{2}$ for the operator $-\Delta$. \begin{prop} \label{prop3.4} Let $K\subset(0,1)^{2}$ be such that $k=dist(\partial(0,1)^{2},K)>0$, and let $0<\epsilon<k/2$. Then, there exists a constant $C<\infty$, depending on $k$ only, such that, for all $x\in K,y\in[0,1]^{2}$, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}G(u,y)du-\frac{2}{\pi}\log(1/\epsilon)\right|\leq C \] if $\left\Vert y-x\right\Vert <\epsilon$, and \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}G(u,y)du-\frac{2}{\pi}\log(1/\left\Vert x-y\right\Vert )\right|\leq C \] if $\left\Vert y-x\right\Vert \geq\epsilon$. \end{prop} \begin{proof} The function $(G-\Gamma)(x,y)$ is symmetric, harmonic in each variable, and continuous. Hence, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}(G-\Gamma)(u,y)\, du\right|\leq\sup_{\{u:dist(u,\partial(0,1)^{2})\geq k/2\}}\sup_{\{y\in[0,1]^{2}\}}\left|(G-\Gamma)(u,y)\right| \] \[ \leq\sup_{u:dist(u,\partial(0,1)^{2})\geq k/2}|\Gamma(dist(u,\partial(0,1)^{2}))|=\Gamma(k/2), \] where the second bound is obtained by applying the maximum principle to $(G-\Gamma)(u,\cdot)$, noting that $G(u,\cdot)$ vanishes at the boundary of $(0,1)^{2}$, and using that $\Gamma$ is decreasing. Therefore, it is enough to prove Proposition \ref{prop3.4} with $G$ replaced by $\Gamma$. Suppose that $\left\Vert x-y\right\Vert <\epsilon$. Then, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}\Gamma(u,y)-\Gamma(\epsilon)du\right|=\left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}\Gamma\left(\frac{\left\Vert u-y\right\Vert }{\epsilon}\right)du\right|. \] The change of variables $u'=(u-y)/\epsilon$ implies that the previous display is \[ =\left|\frac{1}{\pi}\int_{D((x-y)/\epsilon,1)}\Gamma(u')du'\right|\leq\sup_{z\in\overline{D(0,1)}}\left|\frac{1}{\pi}\int_{D(z,1)}\Gamma(u')du'\right|\leq C, \] by continuity in $z$ and compactness of $\overline{D(0,1)}$, where $C$ is an absolute constant. Suppose now that $\left\Vert x-y\right\Vert \geq\epsilon$. Then, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}\Gamma(u,y)du-\Gamma(\left\Vert x-y\right\Vert )\, du\right|=\left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}\Gamma\left(\left\Vert \frac{u-y}{\left\Vert x-y\right\Vert }\right\Vert \right)du\right|. \] The change of variables $u'=(u-y)/\left\Vert x-y\right\Vert $ implies that the previous line is \[ =\left|\frac{1}{\pi(\epsilon/\left\Vert x-y\right\Vert )^{2}}\int_{D\left(\frac{x-y}{\left\Vert x-y\right\Vert },\frac{\epsilon}{\left\Vert x-y\right\Vert }\right)}\Gamma(u')\, du'\right|\leq\sup_{0\leq r\leq1}\sup_{\left\Vert z\right\Vert =1}\left|\frac{1}{\pi r^{2}}\int_{D(z,r)}\Gamma(u')\, du'\right|<C, \] by continuity in $r,z$ and compactness of $\left\{ 0\leq r\leq1\right\} \times\left\{ \left\Vert z\right\Vert =1\right\} $, where $C$ is an absolute constant. \end{proof} Note that the fact that we are integrating over disks is not essential. We could define similar MGFF for other mollifiers. A trivial corollary (which follows from elementary properties of $\log$) of the previous proposition is \begin{cor} \label{cor3.5} Let $K,k,\epsilon$ be as in Proposition \ref{prop3.4} and let $c_{0}>0$. Then, there exists a constant $C$ (depending on $k$ and $c_{0}$) such that, for all $x\in K,y\in[0,1]^{2}$, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}G(u,y)du-\frac{2}{\pi}\log(1/\epsilon)\right|\leq C \] whenever $\left\Vert x-y\right\Vert <c_{0}\epsilon$, and \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(x,\epsilon)}G(u,y)du-\frac{2}{\pi}\log(1/\left\Vert x-y\right\Vert )\right|\leq C \] whenever$\left\Vert x-y\right\Vert \geq c_{0}\epsilon$. \end{cor} Now we prove an important corollary of Proposition \ref{prop3.4}. \begin{cor} \label{cor3.6} Let $K,k$ be as in Proposition \ref{prop3.4}. Then, there exists a constant $C$ (depending only on $k$) such that, for all $x,y\in K,\epsilon>0$, \begin{equation} |Cov(X_{\epsilon}^{x},X_{\epsilon}^{y})+\frac{2}{\pi}\log(\max\{\epsilon,\left\Vert x-y\right\Vert \})|\leq C.\label{cov} \end{equation} Moreover, if $\left\Vert x-y\right\Vert \leq\epsilon$, then \begin{equation} \mathbb{E}\left(X_{\epsilon}^{x}-X_{\epsilon}^{y}\right)^{2}\leq C\epsilon^{-1}\left\Vert x-y\right\Vert .\label{l2_1} \end{equation} \end{cor} \begin{proof} Let us prove \eqref{cov}. If $\left\Vert x-y\right\Vert \leq2\epsilon$, by Corollary \ref{cor3.5}, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(y,\epsilon)}G(u,v)\, dv-\Gamma(\epsilon)\right|\leq C \] for every $u\in D(x,\epsilon)$. Integrating the last inequality over $u\in D(x,\epsilon)$ and using \eqref{green}, we obtain that \[ \left|Cov(X_{\epsilon}^{x},X_{\epsilon}^{y})-\Gamma(\epsilon)\right|\leq C \] for all $\left\Vert x-y\right\Vert \leq2\epsilon$ (and in particular, for $\left\Vert x-y\right\Vert \leq\epsilon$). If $\left\Vert x-y\right\Vert \geq2\epsilon$, Corollary \ref{cor3.5} implies \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(y,\epsilon)}G(u,v)dv-\Gamma(\left\Vert y-u\right\Vert )\right|\leq C \] for every $u\in D(x,\epsilon)$. But $\Gamma(3/2)\leq\Gamma(1+\tfrac{\epsilon}{\left\Vert x-y\right\Vert })\leq\Gamma(\left\Vert y-u\right\Vert )-\Gamma(\left\Vert x-y\right\Vert )\leq\Gamma(1-\tfrac{\epsilon}{\left\Vert x-y\right\Vert })\leq\Gamma(1/2)$ for all $u\in D(x,\epsilon)$. Therefore, \[ \left|\frac{1}{\pi\epsilon^{2}}\int_{D(y,\epsilon)}G(u,v)dv-\Gamma(\left\Vert x-y\right\Vert )\right|\leq C. \] The same (with a different constant) holds for $\left\Vert x-y\right\Vert \geq\epsilon$, because $\Gamma$ is logarithmic. Integrating over $u\in D(x,\epsilon)$ finishes the proof of \eqref{cov}. We now prove \eqref{l2_1}. Display \eqref{green} implies \[ Cov(X_{\epsilon}^{x},X_{\epsilon}^{x}-X_{\epsilon}^{y})=\frac{1}{\pi^{2}\epsilon^{4}}\int_{D(x,\epsilon)}\int_{D(x,\epsilon)}G-\frac{1}{\pi^{2}\epsilon^{4}}\int_{D(x,\epsilon)}\int_{D(y,\epsilon)}G \] \[ =\frac{1}{\pi^{2}\epsilon^{4}}\left(\int_{D(x,\epsilon)\backslash D(y,\epsilon)}\int_{D(x,\epsilon)}G-\int_{D(y,\epsilon)\backslash D(x,\epsilon)}\int_{D(x,\epsilon)}G\right). \] We can use Corollary \ref{cor3.5} to obtain an upper bound of the first term and a lower bound of the second term of the previous display. Then, the previous display is \[ \leq\frac{1}{\pi\epsilon^{2}}\left(\int_{D(x,\epsilon)\backslash D(y,\epsilon)}(\Gamma(\epsilon)+C)-\int_{D(y,\epsilon)\backslash D(x,\epsilon)}(\Gamma(\epsilon)-C)\right)=\frac{C}{\pi\epsilon^{2}}\left|D(x,\epsilon)\backslash D(y,\epsilon)\right|, \] where $\left|D(x,\epsilon)\backslash D(y,\epsilon)\right|$ is the Lebesgue measure of the set $D(x,\epsilon)\backslash D(y,\epsilon)$. Elementary geometry implies $|D(x,\epsilon)\backslash D(y,\epsilon)|\leq C\epsilon\left\Vert x-y\right\Vert $. Repeating the previous argument for $Cov(X_{\epsilon}^{y},X_{\epsilon}^{y}-X_{\epsilon}^{x})$ finishes the proof. \end{proof} \subsection{Tightness for the MGFF} In the next theorem we provide upper bounds on the left and right tail of the MGFF, and we compute the expected maximum up to an order 1 term. \begin{thm} For $\epsilon>0$, let $X_{\epsilon}^{x},x\in[0,1]^{2}$ be the MGFF. Then, there exist absolute constants $0<c,C<\infty$ such that \begin{equation} \mathbb{P}\left(\left|\max_{x\in[0,1]^{2}}X_{\epsilon}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\right|\geq+\lambda\right)\leq Ce^{-c\lambda}\label{rtail} \end{equation} for all $\lambda\geq0$. Moreover, \[ \mathbb{E}\left[\max_{x\in[0,1]^{2}}X_{\epsilon}^{x}\right]=\sqrt{\frac{2}{\pi}}m_{\epsilon}+O(1). \] \end{thm} \begin{proof} Let $Q$ be the open square of side length $1/2$, which is concentric with $(0,1)^{2}$, and let $q:[0,1]^{2}\to\overline{Q}$ be the natural concentric contraction. Consider the field $Y_{\epsilon}^{x}:=X_{\epsilon/2}^{q(x)};x\in[0,1]^{2}$. By Corollary \ref{cor3.6}, \[ Cov(Y_{\epsilon}^{x},Y_{\epsilon}^{y})=Cov(X_{\epsilon/2}^{q(x)},X_{\epsilon/2}^{q(y)})=\frac{2}{\pi}\log\left(\max\left\{ \epsilon/2,\left\Vert q(x)-q(y)\right\Vert \right\} \right)+O(1) \] \[ =\frac{2}{\pi}\log\left(\max\left\{ \epsilon,\left\Vert x-y\right\Vert \right\} \right)+O(1) \] for all $x,y\in[0,1]^{2}$, and \[ \mathbb{E}\left(Y_{\epsilon}^{x}-Y_{\epsilon}^{y}\right)^{2}=\mathbb{E}\left(X_{\epsilon/2}^{q(x)}-X_{\epsilon/2}^{q(y)}\right)^{2}\leq C\epsilon^{-1}2\left\Vert q(x)-q(y)\right\Vert =C\epsilon^{-1}\left\Vert x-y\right\Vert \] for all $x,y\in[0,1]^{2}$ such that $\left\Vert x-y\right\Vert \leq\epsilon$. An application of Theorem \ref{main} yields the existence of absolute constants $0<c,C<\infty$ such that \begin{equation} \mathbb{P}\left(\max_{x\in[0,1]^{2}}Y_{\epsilon}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right)=\mathbb{P}\left(\max_{x\in\overline{Q}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right)\leq Ce^{-c\lambda}\label{asdtail} \end{equation} and \begin{equation} \mathbb{P}\left(\max_{x\in\overline{Q}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\leq\lambda\right)\leq Ce^{-c\lambda}\label{asdasdtail} \end{equation} for all $\lambda\geq0$. Bound \eqref{asdasdtail} easily implies that \[ \mathbb{P}\left(\max_{x\in[0,1]^{2}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\leq\lambda\right)\leq\mathbb{P}\left(\max_{x\in\overline{Q}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\leq\lambda\right)\leq Ce^{-c\lambda} \] for all $\lambda\geq0$, proving \eqref{rtail} for the left tail (after using $m_{\epsilon/2}=m_{\epsilon}+O(1)$, and adjusting the constants). In order to prove the bound \eqref{rtail} for the right tail, we use Proposition \ref{prop3.1} and the comment that follows it to decompose \[ X_{\epsilon/2}^{x}=\hat{X}_{\epsilon/2}^{x}+\phi^{x}, \] where $\left(\hat{X}_{\epsilon/2}^{x}:x\in\overline{Q}\right)\overset{d}{=}\left(X_{\epsilon}^{x}:x\in[0,1]^{2}\right)$ and the fields $\left(\phi^{x}:x\in\overline{Q}\right),\left(\hat{X}_{\epsilon/2}^{x}:x\in\overline{Q}\right)$ are independent. If $\chi=\arg\max\left\{ \hat{X}_{\epsilon/2}^{x}:x\in\overline{Q}\right\} $, then \[ \left\{ \phi^{\chi}\geq0,\hat{X}_{\epsilon/2}^{\chi}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right\} \subset\left\{ \max_{x\in\overline{Q}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right\} . \] But independence of $\phi$ and $\chi$ implies \[ \mathbb{P}\left(\phi^{\chi}\geq0,\hat{X}_{\epsilon/2}^{\chi}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right)=\frac{1}{2}\mathbb{P}\left(\hat{X}_{\epsilon/2}^{\chi}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right) \] because $\phi$ is a centered field. By using the last display and \eqref{asdtail}, we obtain \[ \mathbb{P}\left(\hat{X}_{\epsilon/2}^{\chi}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right)\leq2\mathbb{P}\left(\max_{x\in\overline{Q}}X_{\epsilon/2}^{x}-\sqrt{\frac{2}{\pi}}m_{\epsilon}\geq\lambda\right)\leq Ce^{-c\lambda} \] for some absolute constants $0<c,C<\infty$. The bound \eqref{rtail} and $m_{\epsilon/2}=m_{\epsilon}+O(1)$ implies tightness of the family \[ \left\{ \max_{x\in[0,1]^{2}}X_{\epsilon}^{x}-m_{\epsilon}:\epsilon>0\right\} , \] and the same bound also implies \[ \mathbb{E}\left[\max_{x\in[0,1]^{2}}X_{\epsilon}^{x}\right]=\sqrt{\frac{2}{\pi}}m_{\epsilon}+O(1), \] finishing the proof. \end{proof} \section{Appendix} We prove here the claims made in Section 2.1. \begin{prop} \label{proofdefMBRW}The MBRW, defined by display \eqref{defMBRW}, exists and satisfies \[ Var(\xi_{\epsilon}^{v}(t))=t \] for all $0\leq t\leq\log(1/\epsilon)$ and all $v\in V_{\epsilon}$, and \[ t-C\leq Cov(\xi_{\epsilon}^{v}(t),\xi_{\epsilon}^{w}(t))\leq t \] for all $0\leq t\leq-\log\left\Vert v-w\right\Vert _{\infty}$ and all $v,w\in V_{\epsilon}$, where $C$ is a constant depending on the dimension. \end{prop} \begin{proof} We show that the mapping $\left(V_{\epsilon}\times[0,\log(1/\epsilon)]\right)^{2}\to\mathbb{R}$ given by \[ \left((v,t),(u,s)\right)\mapsto\int_{0}^{\min\left\{ t,s\right\} }\prod_{i\leq d}(1-e^{r}\left|v_{i}-u_{i}\right|)_{+}dr \] is positive definite. Note first that \[ \prod_{i\leq d}(1-e^{r}\left|v_{i}-u_{i}\right|)_{+}=\int_{\mathbb{R}^{d}}1_{A(v,r)}(z)1_{A(u,r)}(z)dz, \] where $dz$ is $d$-dimensional Lebesgue measure and $A(v,r)$ is the $d$-dimensional box of side length 1, centered at $e^{r}v$. Let $\left\{ (v^{\alpha},t^{\alpha})\right\} _{\alpha}$ be any finite subset of $V_{\epsilon}\times[0,\log(1/\epsilon)]$, and let $\left\{ c_{\alpha}\right\} _{\alpha}$ be arbitrary real numbers. Then, applying the previous display, we obtain \[ \sum_{\alpha,\beta}c_{\alpha}c_{\beta}\int_{0}^{\min\left\{ t^{\alpha},t^{\beta}\right\} }\prod_{i\leq d}(1-e^{r}\left|v_{i}^{\alpha}-v_{i}^{\beta}\right|)_{+}dr \] \[ =\int_{0}^{\infty}\int_{\mathbb{R}^{d}}\sum_{\alpha,\beta}c_{\alpha}c_{\beta}1_{[0,t^{\alpha}]}(r)1_{[0,t^{\beta}]}(r)1_{A(v^{\alpha},r)}(z)1_{A(v^{\beta},r)}(z)\, dz\, dr \] \[ =\int_{0}^{\infty}\int_{\mathbb{R}^{d}}\left(\sum_{\alpha}c_{\alpha}1_{[0,t^{\alpha}]}(r)1_{A(v^{\alpha},r)}(z)\right)^{2}\, dz\, dr\geq0, \] as desired. This shows that the MBRW exists. For any $v\in V_{\epsilon}$ and $t\leq\log(1/\epsilon)$, \[ Var(\xi_{\epsilon}^{v}(t))=\int_{0}^{t}\prod_{i\leq d}\left(1\right)dr=t. \] Moreover, if $v\neq w$, \[ \prod_{i\leq d}\left(1-e^{r}\left|v_{i}-w_{i}\right|\right)_{+}\begin{cases} >0 & \text{ if }r<-\log\left\Vert v-w\right\Vert _{\infty}\\ =0 & \text{ if }r\geq-\log\left\Vert v-w\right\Vert _{\infty} \end{cases}. \] Therefore, if $t<-\log\left\Vert v-w\right\Vert _{\infty}$, \[ t\geq Cov(\xi_{\epsilon}^{v}(t),\xi_{\epsilon}^{w}(t))\geq\int_{0}^{t}\prod_{i\leq d}\left(1-e^{r}\left|v_{i}-w_{i}\right|\right)dr\geq\int_{0}^{t}\left(1-e^{r}\left\Vert v-w\right\Vert _{\infty}\right)^{d}dr \] \[ \geq t+\sum_{k=1}^{d}\left(\begin{array}{c} d\\ k \end{array}\right)(-1)^{k}\left\Vert v-w\right\Vert _{\infty}^{k}\left(\frac{e^{kt}-1}{k}\right)\geq t-C \] for some constant $C<\infty$ depending on $d$ only. Similarly, if $t\geq-\log\left\Vert v-w\right\Vert _{\infty}$, \[ -\log\left\Vert v-w\right\Vert _{\infty}\geq Cov(\xi_{\epsilon}^{v}(t),\xi_{\epsilon}^{w}(t))\geq-\log\left\Vert v-w\right\Vert _{\infty}-C. \] \end{proof} \begin{prop} \label{proofrtlbMBRW} Let $\left(\xi_{\epsilon}^{v}:v\in V_{\epsilon}\right)$ be the MBRW and let $m_{\epsilon}$ be the number defined in the line preceding Theorem \ref{main}. Then, there exists a constant $c>0$ (depending on the dimension) such that \[ \mathbb{P}\left(\max_{v\in V_{\epsilon}}\xi_{\epsilon}^{v}\geq m_{\epsilon}\right)\geq c. \] \end{prop} \begin{proof} We use a second moment method. Let $T=T_{\epsilon}=\log(1/\epsilon)$ and let \[ A_{v}=\left\{ \xi_{\epsilon}^{v}\geq m_{\epsilon},\xi_{\epsilon}^{v}(t)\leq\frac{m_{\epsilon}}{T}t+1\text{ for all }0\leq t\leq T\right\} , \] \[ Z=\sum_{v\in V_{\epsilon}}1_{A_{v}}. \] Note that \begin{equation} \mathbb{P}\left(\max_{v\in V_{\epsilon}}\xi_{\epsilon}^{v}\geq m_{\epsilon}\right)\geq\mathbb{P}\left(Z>0\right)\geq\frac{\left(\mathbb{E}\left[Z\right]\right)^{2}}{\mathbb{\mathbb{E}}\left[Z^{2}\right]},\label{secondmomentmethod} \end{equation} where the second inequality follows by Cauchy-Schwarz. We first compute a lower bound for $\mathbb{E}\left[Z\right]$. Note that \[ \mathbb{E}\left[Z\right]=\epsilon^{-d}\mathbb{P}\left(A_{v}\right). \] Let $\bar{\xi}_{\epsilon}^{v}(t)=\xi_{\epsilon}^{v}(t)-\frac{m_{\epsilon}}{T}t$. Define a probability measure $\mathbb{Q}$ by \[ \frac{d\mathbb{P}}{d\mathbb{Q}}=\exp\left(-\frac{m_{\epsilon}}{T}\bar{\xi}_{\epsilon}^{v}(T)-\frac{m_{\epsilon}^{2}}{2T}\right). \] Girsanov's Theorem (see \cite[Theorem 5.1]{ks}) implies that $\bar{\xi}_{\epsilon}^{v}(t)$ is Brownian motion under $\mathbb{Q}$. Note that \[ \mathbb{P}\left(A_{v}\right)=\int_{A_{v}}\exp\left(-\frac{m_{\epsilon}}{T}\bar{\xi}_{\epsilon}^{v}(T)-\frac{m_{\epsilon}^{2}}{2T}\right)d\mathbb{Q}\geq\exp\left(-\frac{m_{\epsilon}}{T}-\frac{m_{\epsilon}^{2}}{2T}\right)\mathbb{Q}(A_{v}) \] \[ \geq ce^{-\sqrt{2d}}\epsilon^{d}T^{3/2}\mathbb{Q}(A_{v}) \] for some absolute constant $c>0$. It follows easily from the Reflection Principle (see \cite[Proposition 6.19]{ks}) that $\mathbb{Q}(A_{v})=\mathbb{Q}(\bar{\xi}_{\epsilon}^{v}\geq0,\bar{\xi}_{\epsilon}^{v}(t)\leq1\text{ for all }0\leq t\leq T)\geq cT^{-3/2}$ for some absolute constant $c>0$. Combining the three previous displays, we obtain \begin{equation} \mathbb{E}\left[Z\right]\geq c\label{lbfirstmoment} \end{equation} for some constant $c>0$, depending on the dimension $d$. We now compute an upper bound for $\mathbb{E}\left[Z^{2}\right]$. Note that \begin{equation} \mathbb{E}\left[Z^{2}\right]=\sum_{v,w\in V_{\epsilon}}\mathbb{P}\left(A_{v}\cap A_{w}\right)=\sum_{v,w\in V_{\epsilon}}\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v},\bar{\xi}_{\epsilon}^{w}\geq0,\bar{\xi}_{\epsilon}^{v}(t),\bar{\xi}_{\epsilon}^{w}(t)\leq1\text{ for all }0\leq t\leq T\right).\label{ZZ} \end{equation} Both $\xi_{\epsilon}^{v}(\cdot),\xi_{\epsilon}^{w}(\cdot)$ are Brownian motions, which have independent increments starting at time $s=s_{v,w}=-\log\left(\max\left\{ \epsilon,\left\Vert v-w\right\Vert _{\infty}\right\} \right)$. Therefore, \[ \mathbb{P}\left(A_{v}\cap A_{w}\right)\leq\sum_{-\infty<x,y\leq1}p(x)p(y)\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t),\bar{\xi}_{\epsilon}^{w}(t)\leq1\text{ for all }t\in[0,s],\bar{\xi}_{\epsilon}^{v}(s)\in[x-1,x],\bar{\xi}_{\epsilon}^{w}(s)\in[y-1,y]\right) \] \begin{equation} \leq\sum_{-\infty<y\leq x\leq1}2p(x)p(y)\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t),\bar{\xi}_{\epsilon}^{w}(t)\leq1\text{ for all }t\in[0,s],\bar{\xi}_{\epsilon}^{v}(s)\in[x-1,x],\bar{\xi}_{\epsilon}^{w}(s)\in[y-1,y]\right),\label{conditioning} \end{equation} where \[ p(x)=\sup_{z\in[x-1,x]}\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t)\leq1-z\text{ for all }t\in[0,T-s],\bar{\xi}_{\epsilon}^{v}(T-s)\geq-z\right). \] Assume $0<s<T$. Applying Girsanov's Theorem and the Reflection Principle, we obtain \[ p(x)\leq C\exp\left(\frac{m_{\epsilon}}{T}x-\frac{m_{\epsilon}^{2}}{2T^{2}}(T-s)\right)\frac{(1-x)}{\left(T-s\right)^{3/2}} \] for some constant $C$. Therefore, from \eqref{conditioning} and the last display, \[ \mathbb{P}\left(A_{v}\cap A_{w}\right)\leq\sum_{-\infty<y\leq x\leq1}Cp(x)^{2}\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t),\bar{\xi}_{\epsilon}^{w}(t)\leq1\text{ for all }t\in[0,s],\bar{\xi}_{\epsilon}^{v}(s)\in[x-1,x],\bar{\xi}_{\epsilon}^{w}(s)\in[y-1,y]\right) \] \[ \leq\sum_{-\infty<x\leq1}Cp(x)^{2}\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t)\leq1\text{ for all }t\in[0,s],\bar{\xi}_{\epsilon}^{v}(s)\in[x-1,x]\right). \] Applying Girsanov's Theorem and the Reflection Principle again, \[ \mathbb{P}\left(A_{v}\cap A_{w}\right)\leq C\sum_{-\infty<x\leq1}p(x)^{2}\exp\left(-\frac{m_{\epsilon}}{T}x-\frac{m_{\epsilon}^{2}}{2T^{2}}s\right)\frac{(1-x)}{s^{3/2}} \] \begin{equation} \leq C\frac{1}{\left(T-s\right)^{3}s^{3/2}}\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right)\label{intersection1} \end{equation} for some constant $C$. Consider now the case $s=0$. Then, the independence of $\xi_{\epsilon}^{v}(\cdot)$ and $\xi_{\epsilon}^{v}(\cdot)$ implies \[ \mathbb{P}\left(A_{v}\cap A_{w}\right)=\mathbb{P}\left(A_{v}\right)^{2}=\mathbb{P}\left(\bar{\xi}_{\epsilon}^{v}(t)\leq1\text{ for all }t\in[0,T],\bar{\xi}_{\epsilon}^{v}(T)\geq0\right)^{2} \] \begin{equation} \leq C\frac{1}{T^{3}}\exp\left(-\frac{m_{\epsilon}^{2}}{T}\right),\label{intersection2} \end{equation} where the last bound follows from Girsanov's Theorem and the Reflection Principle. In the case $s=T$, \begin{equation} \mathbb{P}\left(A_{v}\cap A_{w}\right)\leq\mathbb{P}\left(A_{v}\right)\leq C\frac{1}{T^{3/2}}\exp\left(-\frac{m_{\epsilon}^{2}}{2T}\right).\label{intersection3} \end{equation} In consequence, for any pair $v,w\in V_{\epsilon}$, displays \eqref{intersection1}, \eqref{intersection2} and \eqref{intersection3} imply \[ \mathbb{P}\left(A_{v}\cap A_{w}\right)\leq C\frac{1}{\left(\left(T-s\right)\vee1\right)^{3}\left(s\vee1\right)^{3/2}}\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right), \] where $\cdot\vee\cdot=\max\left\{ \cdot,\cdot\right\} $. For any fixed $v\in V_{\epsilon}$, there are $O(e^{(d-1)(T-s)})$ points $w$ such that $-\log\left\Vert v-w\right\Vert _{\infty}=s$. Therefore, from \eqref{ZZ}, the last display, we obtain \[ \mathbb{E}\left[Z^{2}\right]\leq C\sum_{0\leq s\leq T}\left|V_{\epsilon}\right|e^{(d-1)(T-s)}\frac{1}{\left(\left(T-s\right)\vee1\right)^{3}\left(s\vee1\right)^{3/2}}\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right) \] \[ \leq C+C\sum_{0<s<T}\left|V_{\epsilon}\right|e^{(d-1)(T-s)}\frac{\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right)}{\left(T-s\right)^{3}s^{3/2}} \] \[ =C+C\sum_{0<s<T}e^{dT}e^{(d-1)(T-s)}\frac{\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right)}{\left(T-s\right)^{3}s^{3/2}}. \] But, \[ \sum_{0<s<T}e^{dT}e^{(d-1)(T-s)}\frac{\exp\left(-\frac{m_{\epsilon}^{2}}{2T^{2}}(2T-s)\right)}{\left(T-s\right)^{3}s^{3/2}}\leq\sum_{0<s<T}e^{d(2T-s)}\frac{\exp\left(\left(-d+\frac{3\log T}{2T}\right)(2T-s)\right)}{\left(T-s\right)^{3}s^{3/2}} \] \[ =\sum_{0<s<T}\frac{\exp\left(\frac{3}{2}\frac{\log T}{T}(2T-s)\right)}{(T-s)^{3}s^{3/2}}\leq C\sum_{0<s<T/2}\frac{1}{s^{3/2}}+\sum_{T/2\leq s<T}\frac{\exp\left(\frac{3}{2}\frac{\log T}{T}(T-s)\right)}{(T-s)^{3}}\frac{T^{3/2}}{s^{3/2}} \] \[ \leq C+C\sum_{0<s\leq T/2}\frac{T^{3s/2T}}{s^{3}}\leq C<\infty, \] because the last expression is (eventually) decreasing in $T$. Proposition \ref{proofrtlbMBRW} follows from the last display, \eqref{secondmomentmethod} and \eqref{lbfirstmoment}. \end{proof} \begin{prop} \label{proofrtubMBRW} Let $\left(\xi_{\epsilon}^{v}:v\in V_{\epsilon}\right)$ be the MBRW and let $m_{\epsilon}$ be the number defined in the line preceding Theorem \ref{main}. Then, there exist constants $0<c,C<\infty$ (depending on the dimension $d$) such that \[ \mathbb{P}\left(\max_{v\in A}\xi_{\epsilon}^{v}\geq m_{\epsilon}+z\right)\leq C\left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{-cz} \] for all $A\subset V_{\epsilon}$, $z\in\mathbb{R}$ and $\epsilon>0$ small enough.\end{prop} \begin{proof} We introduce the $d$-ary branching random walk (BRW) as follows: let $\epsilon=2^{-n}$ for some $n\in\mathbb{N}$. At each time $T_{k}=k\log2;k=0,1,\ldots,n$, we partition $[0,1)^{d}$ into $2^{kd}$ disjoint boxes of side length $2^{-k}$. For a pair $v,w\in V_{\epsilon}$, denote by $l(v,w)$ the first time that $v,w$ lie in different boxes of the partition. With this notation, define the BRW as the Gaussian field $\left(\eta_{\epsilon}^{v}(t):v\in V_{\epsilon},t\in[0,T_{n}]\right)$ with \[ Cov(\eta_{\epsilon}^{v}(t),\eta_{\epsilon}^{w}(s))=\min\left\{ t,s,l(v,w)\right\} . \] For simplicity, let $T=T_{n}$ and $\eta_{\epsilon}^{v}=\eta_{\epsilon}^{v}(T)$. It is not hard to show that such a field exists. Note that our BRW can be interpreted as a branching Brownian motion that splits every $\log2$ units of time into $2^{d}$ independent Brownian motions. Following the argument given in \cite[Lemma 3.7]{dz}, one can show that there exists $C$ (depending on the dimension) such that \[ \mathbb{P}\left(\max_{v\in A}\xi_{\epsilon}^{v}\geq m_{\epsilon}+\lambda\right)\leq C\mathbb{P}\left(\max_{v\in A}\eta_{\epsilon/C}^{v}\geq m_{\epsilon}+\lambda\right) \] for all $A\subset V_{\epsilon}\subset V_{\epsilon/C}$ and all $\lambda\in\mathbb{R}$. Therefore, it is enough to prove Proposition \ref{proofrtubMBRW} for the BRW. We do so by following very closely the proof in \cite[Lemma 3.8]{bdz}. We will use the following estimate, which is proved in \cite[Lemma 3.6]{bdz}: let $W_{s}$ be standard Brownian motion under $\mathbb{P}$ and fix a large constant $C_{1}$. Then, if \[ \mu_{q,r}^{\ast}(x)=\mathbb{P}\left(W_{q}\in dx,W_{s}\leq r+C_{1}(\min\left\{ s,q-s\right\} )^{1/20}\text{ for all }0\leq s\leq q\right)/dx, \] we have \begin{equation} \mu_{q,r}^{\ast}(x)\leq C_{2}r(r-x)/q^{3/2}\label{quasibridge} \end{equation} for all $x\leq r$, where $C_{2}$ depends on $C_{1}$. We next define the event \[ G(\lambda)=\left\{ \exists t\leq T,v\in V_{\epsilon}:\eta_{\epsilon}^{v}(t)-\frac{m_{\epsilon}}{T}t-10\log\left(\min\left\{ t,T-t\right\} \right)_{+}\geq\lambda\right\} \] and we prove the following claim: \begin{claim} There exists a constant $C>0$ (depending on $d$) such that \[ \mathbb{P}\left(G(\lambda)\right)\leq C\lambda e^{-\sqrt{2d}\lambda} \] for all $\lambda\geq1$. \end{claim} \begin{proof} Following the proof of \cite[Lemma 3.7]{bdz}, we define $\psi_{t}=\lambda+10\log\left(\min\left\{ t,T-t\right\} \right)_{+}$ and $\chi_{T_{k}}(x)=\mathbb{P}\left(\eta_{\epsilon}^{v}(t)-\frac{m_{\epsilon}}{T}t\leq\psi_{t}\text{ for all }t\leq T_{k},\eta_{\epsilon}^{v}(T_{k})-\frac{m_{\epsilon}}{T}T_{k}\in dx\right)/dx$. Then, by decomposing based on the first time such that $\eta_{\epsilon}^{v}(t)-\frac{m_{\epsilon}}{T}t\geq\psi_{t}$, we obtain that \[ \mathbb{P}\left(G(\lambda)\right)\leq\sum_{k=1}^{n}2^{dk}\int_{-\infty}^{\psi_{T_{k}}}\chi_{T_{k}}(x)\mathbb{P}\left(\max_{s\leq\log2}\eta_{\epsilon}^{v}(t)\geq\psi_{T_{k}}-x-C\right)dx, \] where $C$ is an absolute constant. Display \eqref{quasibridge} and Girsanov's Theorem imply that \[ \chi_{T_{k}}(x)\leq C2^{-dk}e^{-x\left(\sqrt{2d}-O(\log T/T)\right)}\psi_{T_{k}}(\psi_{T_{k}}-x), \] where $C$ depends on $d$. On the other hand, \[ \mathbb{P}\left(\max_{s\leq\log2}\eta_{\epsilon}^{v}(t)\geq\psi_{T_{k}}-x-C\right)\leq Ce^{-\left(\psi_{T_{k}}-x-C\right)^{2}/2\log2} \] for some absolute constant $C$. Therefore, by the three previous displays, we obtain \[ \mathbb{P}\left(G(\lambda)\right)\leq C\sum_{k=1}^{n}\psi_{T_{k}}\int_{-\infty}^{\psi_{T_{k}}}e^{-x\left(\sqrt{2d}-O(\log T/T)\right)}(\psi_{T_{k}}-x)e^{-\left(\psi_{T_{k}}-x-C\right)^{2}/2\log2}dx. \] A change of variables $u=\psi_{T_{k}}-x$ yields \[ \mathbb{P}\left(G(\lambda)\right)\leq C\sum_{k=1}^{n}\psi_{T_{k}}e^{-\sqrt{2d}\psi_{T_{k}}} \] \[ =C\sum_{k=1}^{n}\left(\lambda+10\log\left(\min\left\{ T_{k},T-T_{k}\right\} \vee1\right)\right)e^{-\sqrt{2d}\left(\lambda+10\log\left(\min\left\{ T_{k},T-T_{k}\right\} \vee1\right)\right)} \] \[ =C\sum_{k=1}^{n}\frac{\left(\lambda+10\log\left(\min\left\{ T_{k},T-T_{k}\right\} \vee1\right)\right)}{\left(\min\left\{ T_{k},T-T_{k}\right\} \vee1\right)^{10}}e^{-\sqrt{2d}\lambda}\leq C\lambda e^{-\sqrt{2d}\lambda}, \] where $\cdot\vee\cdot=\max\left\{ \cdot,\cdot\right\} $, and the convergence of the last sum is due the exponent $10$ in the denominator (with room to spare). \end{proof} We now finish the proof of Proposition \ref{proofrtubMBRW}. Fix $A\subset V_{\epsilon}$ and $z\in\mathbb{R}$. For $z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}\geq1$, let $\lambda=z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}$, and continuing with the notation of Claim 4.4, we let \[ F_{v}=\left\{ \eta_{\epsilon}^{v}(t)\leq\frac{m_{\epsilon}}{T}t+\psi_{t}\text{ for all }0\leq t\le T,\eta_{\epsilon}^{v}\geq m_{\epsilon}+z\right\} , \] where $v\in V_{\epsilon}$. We now compute \[ \mathbb{P}\left(F_{v}(\lambda)\right)=\int_{z}^{\psi_{T}}\frac{d\mathbb{P}}{d\mathbb{Q}}(x+m_{\epsilon})\chi_{T}(x)dx \] \[ \leq C\int_{z}^{\psi_{T}}2^{-dn}e^{-x\left(\sqrt{2d}-O(\log T/T)\right)}\psi_{T}\left(\psi_{T}-x\right)dx \] \[ \leq C2^{-dn}\psi_{T}e^{-\sqrt{2d}\psi_{T}}\int_{0}^{\psi_{T}-z}e^{u}u\, du\leq C2^{-dn}\psi_{T}e^{-\sqrt{2d}z}\left(\psi_{T}-z\right). \] Recalling that $\psi_{T}=\lambda=z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}$, we obtain \[ \mathbb{P}\left(F_{v}(\lambda)\right)\leq C2^{-dn}\left(z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}\right)\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}e^{-\sqrt{2d}z} \] \[ \leq C2^{-dn}\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/2}e^{-cz}. \] Adding the last display for $v\in A$ and using Claim 4.4, we obtain \[ \mathbb{P}\left(\max_{v\in A}\eta_{\epsilon}^{v}\geq m_{\epsilon}+z\right)\leq C\left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{-cz}+C\left(z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}\right)e^{-\sqrt{2d}\left(z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}\right)} \] \[ \leq C\left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{-cz} \] for some $0<c,C<\infty$ (depending on $d$ only), as desired. The previous computation was made under the assumption $z+\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}\geq1$. Assume now $\left(\left|V_{\epsilon}\right|/\left|A\right|\right)^{1/4}-1\leq-z$. In this case, \[ \left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{-cz}\geq c\left(\epsilon^{d}\left|A\right|\right)^{1/2}e^{c\left(\epsilon^{d}\left|A\right|\right)^{-1/4}}. \] But $\inf_{0<x<1}x^{1/2}e^{cx^{-1/4}}\geq c>0$, where $c$ depends only $d$. Therefore, in this case, Proposition \ref{proofrtubMBRW} holds trivially by adjusting the constant $C$.\end{proof}
\section{Introduction} A model of the non-degenerate parametric amplifier was introduced and studied in detail in classical papers \cite{LoisellYarivSiegman61}, \cite{GordonLouisellWalker63}, \cite{MollowGlauberI}, \cite{MollowGlauberII} (see also reviews \cite{Hillery09} and \cite{Yariv11} for a historical perspective and textbooks/reviews \cite{BachorRalph04}, \cite{Klyshko88}, \cite{Klyshko94}, \cite{Klyshko96}, \cite{Klyshko11}, \cite{MandelWolf}, \cite{Scully:Zubairy97} and \cite{Walls:Milburn} for a standard paradigm in quantum optics). In a nonlinear dielectric medium, one adds to the linear susceptibility tensor the second and third terms in the power expansion of the polarization in the electric field. This nonlinear polarization couples back to the electric field and a subsequent field quantization results in interaction Hamiltonians that are cubic in the field. Among typical nonlinear effects of this kind are the optical parametric amplification, the second-harmonic generation (both third-order), and the degenerate four-wave mixing (fourth-order) (see \cite{MollowGlauberI}, \cite{BachorRalph04}, \cite{Hanamuraetal07}, \cite{Hillery09} and the references therein for more details). Nonlinear media are essential for the generation of squeezed and entangled states of light \cite{AdessoIlluminati07}, \cite{BraunsteinvanLoock05}, \cite{Breit:Schill:Mlyn97}, \cite{Giedketal01}, \cite{LvRay09}, \cite{Mairetal01}, \cite{MolinaTerrizaetal07}. The squeezed states and the quantum non-demolition experiment are expected to be utilized in the detection of gravitational waves \cite{Caves81}, \cite{Abadetal11}, \cite{Demkowiczetal13} and also to enhance the performance of optical communication systems \cite{BachorRalph04}, \cite{Hanamuraetal07}. The degenerate parametric amplifier was investigated in \cite{Taka65}, \cite{Mollow67}, \cite{Raiford70}, \cite{Keating71}, \cite{PrakashChandraVach74}, \cite{Raiford74}, \cite{Stoler74}, \cite{Yuen76}, \cite{Rowe77}, \cite{HilleryZubairy82}, \cite{WodkiewichZubairy83}, \cite{Angelow:Trifonov95}, \cite{Angelow98} in the so-called parametric approximation, when the pump mode is treated classically. Quantum description beyond this approximation can be found in \cite{Raiford70}, \cite{HilleryZubairy84}, \cite{CohenBraunst95}, \cite{Hillery09} (see also the references therein). Various aspects of the corresponding photon statistics and photon-counting were studied in \cite{Glauber63}, \cite{Stoler70}, \cite{Stoler71}, \cite{Mollow73}, \cite{Caves81}, \cite{Caves82}, \cite{CollettGardiner84}, \cite{VenkSatya85}, \cite{Agarwal87}, \cite{CollettLoundon87}, \cite{VourdasWeiner87}, \cite{AgarwalAdam88}, \cite{FernCollett88}, \cite{FearnLoudon89}, \cite{KimOlivKnight89}, \cite{Marian91}, \cite{Marian92}, \cite{MarianMarianI93}, \cite{MarianMarianII93}, \cite{Dod:Man:Man94} in detail. Connections with the experimentally observed dynamical Casimir effect \cite{Man'koCasimir}, \cite{Dodonov10}, \cite{Wilsonetal11}, \cite{Lahetal11} are discussed in \cite{Dod09}, \cite{DezaelLambrecht10}, \cite{Johanetal10}, \cite{FujiietalZeilinger11}, \cite{Nationetal12}. All these results are also of interest to the theory of quantum noise and measurement (see, for example, \cite{Clerketal10}, \cite{Cavesetal12}, \cite{ChiXie13} and the references therein). Nowadays, advanced experimental techniques allow one to measure photon correlation functions of input microwave signals \cite{Bozyigitetal10}, to do quantum tomography on itinerant microwave photons \cite{Eichleretal11a}, and to study squeezing of microwave fields \cite{Eichleretal11b}, \cite{Malletetal11} (see also \cite{Braggioetal13}, \cite{Galeazzietal13} for experimental study of a single-mode thermal field using a microwave parametric amplifier). In spite of the considerable literature on the degenerate parametric amplifiers, the general case of multi-parameter squeezed input photons (corresponding, say, to a cascade of nonlinear crystals), to the best of our knowledge, has never been discussed. Traditionally, the interaction picture is commonly used \cite{Mollow67}, \cite{Scully:Zubairy97}, \cite{Walls:Milburn} even though the statistics is postulated in the Schr\"{o}dinger picture \cite{Klyshko94}, \cite{Klyshko98}. In this article, we study the statistical properties of output squeezed quanta in terms of explicit solutions of certain Ermakov-type system introduced in \cite{Lan:Lop:Sus}. In particular, explicit formulas for the mean number and variance of generated photons are found together with the corresponding time-dependent photon statistics. We elaborate on the dynamical aspects of this problem related to evolution of the corresponding photon states in Fock's space. In order to achieve this goal, we utilize a unified approach to generalized harmonic oscillators discussed in detail in several recent publications \cite{Cor-Sot:Lop:Sua:Sus}, \cite{Me:Co:Su}, \cite{Cor-Sot:Sus}, \cite{CorSus11}, \cite{Lan:Lop:Sus}, \cite{SanSusVin}, \cite{Lop:Sus:VegaGroup}, \cite{KretalSus13} (see also \cite{Malk:Man:Trif73}, \cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cite{Malkin:Man'ko79} for the classical accounts). A similar treatment may be useful for the Josephson metamaterial dynamical Casimir effect \cite{Johanetal10}, \cite{Nationetal12}, \cite{Wilsonetal11}, \cite{Lahetal11 , \cite{FujiietalZeilinger11} (see also \cite{Dod95}, \cite{Dod09}, \cite{Law94}, \cite{Braggioetal05}), for experimental recognition of squeezed microwave photons \cite{Eichleretal11b}, \cite{Malletetal11}, and for study of a single-mode thermal microwave field \cite{Braggioetal13}, \cite{Galeazzietal13}. The paper is organized as follows. In sections~2 and 3, we describe two exactly solvable models of optical degenerate parametric amplifiers. The generalized Fock states are constructed in section~4. The mean and variance of the number operator are evaluated in Schr\"{o}dinger's picture in section~5. The eigenfunction expansions of the generalized harmonic states of light in terms of the standard Fock ones are derived in section~6 in coordinate representation. In section~7, the Wigner and Moyal functions of the multi-parameter squeezed states are evaluated directly from the corresponding wavefunctions and their time evolution is verified with the help of a computer algebra system. A brief summary is provided in the end. Our explicit solutions of the corresponding Ermakov-type systems are given in Appendix~A together with the means and variances of position and momentum operators. A convenient expansion for the single photon mode Hamiltonian is presented in Appendix~B, solutions in interaction picture are briefly discussed in Appendix~C, and a canonical transformation of creation and annihilation operators is derived in Appendix~D. An attempt to collect relevant references is made. \section{Degenerate Parametric Amplifiers} In this article, we consider the quantization of radiation field in a variable dielectric medium in the Schr\"{o}dinger picture, as outlined in \cite{Heitler57}, \cite{Ber:Lif:Pit} (in vacuum), by using the method of dynamical invariants originally developed in \cite{Dod:Mal:Man75}, \cite{Dodonov:Man'koFIAN87}, \cite{Malk:Man:Trif73}, \cite{Malkin:Man'ko79} and recently revisited in \cite{Cor-Sot:Sua:SusInv}, \cite{Suslov10}, \cite{SanSusVin}. To that end, we follow the mathematical technique of the field quantization for a variable quadratic system in an abstract (Fock-)Hilbert space discussed in \cite{KretalSus13} and concentrate on a single mode of the radiation field. In this picture, the time evolution of degenerate parametric amplifier is governed by the time-dependent Schr\"{o}dinger equation for the state vector $\left\vert \psi\left( t\right) \right\rangle :$ \begin{equation} i\frac{d}{dt}\left\vert \psi\left( t\right) \right\rangle =\widehat{H \left( t\right) \left\vert \psi\left( t\right) \right\rangle \label{Schroedinger \end{equation} with a certain variable quadratic Hamiltonian considered in original publications \cite{Mollow67}, \cite{Raiford70}, \cite{Stoler74}. In a more general setting, the degenerate parametric amplification with time-dependent amplitude and phase was discussed by Raiford \cite{Raiford74}. The corresponding Hamiltonian, without damping and neglecting high-frequency terms, has the for \begin{equation} \widehat{H}\left( t\right) =\frac{\omega}{2}\left( \widehat{a \ \widehat{a}^{\dagger}+\widehat{a}^{\dagger}\ \widehat{a}\right) -\frac{\lambda\left( t\right) }{2}\left( e^{i\left( 2\omega t+\phi\left( t\right) \right) }\ \left. \widehat{a}\right. ^{2}+e^{-i\left( 2\omega t+\phi\left( t\right) \right) }\left( \widehat{a}^{\dagger}\right) ^{2}\right) . \label{RaiHam \end{equation} In this model, the phenomenological coupling parameter $\lambda\left( t\right) ,$ which describes the strength of the interaction between the quantized signal of frequency $\omega$ and the classical pump of frequency $2\omega,$ and the pump phase $\phi\left( t\right) $ are in general functions of time. (It includes the special case of the pump and signal being off-resonance by a given amount $\epsilon,$ i.~e., the pump frequency being $2\omega+\epsilon,$ by letting $\phi\left( t\right) =\epsilon t$ and $\lambda\left( t\right) =\lambda,$ a constant \cite{Raiford74}.) One can use the standard annihilation and creation operators for a given mode $\omega, \begin{equation} \widehat{a}=\frac{1}{\sqrt{2\omega}}\left( \omega\widehat{q}+i\left. \widehat{p}\right. \right) ,\qquad\widehat{a}^{\dagger}=\frac{1 {\sqrt{2\omega}}\left( \omega\widehat{q}-i\left. \widehat{p}\right. \right) ,\qquad\left[ \widehat{a},\ \widehat{a}^{\dagger}\right] =1, \label{aaspq \end{equation} where $\widehat{q}$ and $\widehat{p}$ are time-independent operators in an abstract Hilbert space with the canonical commutation relation $\left[ \left. \widehat{q}\right. ,\left. \widehat{p}\right. \right] =i$ (in the units of $\hbar).$ The \begin{align} & \widehat{H}\left( t\right) =\frac{1}{2}\left( 1+\frac{\lambda\left( t\right) }{\omega}\cos\left( 2\omega t+\phi\left( t\right) \right) \right) \left. \widehat{p}\right. ^{2}\label{RaiHamPX}\\ & \qquad+\frac{\omega^{2}}{2}\left( 1-\frac{\lambda\left( t\right) }{\omega}\cos\left( 2\omega t+\phi\left( t\right) \right) \right) \left. \widehat{q}\right. ^{2}\nonumber\\ & \qquad\quad+\frac{\lambda\left( t\right) }{2}\sin\left( 2\omega t+\phi\left( t\right) \right) \left( \widehat{p}\ \widehat{q +\widehat{q}\ \widehat{p}\right) \nonumber \end{align} and the corresponding characteristic equation (classical equation of motion \cite{Cor-Sot:Sua:SusInv}, \cite{Lan:Lop:Sus}) takes the form \cite{CorSus11} \begin{align} & \mu^{\prime\prime}+\frac{\lambda\sin\left( 2\omega t+\phi\right) \left( 2\omega+\phi^{\prime}\right) -\lambda^{\prime}\cos\left( 2\omega t+\phi\right) }{\omega+\lambda\cos\left( 2\omega t+\phi\right) }\mu ^{\prime}\label{RaiCharMu}\\ & \quad\ +\frac{\omega\left( \omega^{2}-3\lambda^{2}\right) -\lambda \phi^{\prime}-\lambda\left( \omega^{2}+\lambda^{2}+\omega\phi^{\prime }\right) \cos\left( 2\omega t+\phi\right) -\lambda^{\prime}\omega \sin\left( 2\omega t+\phi\right) }{\omega+\lambda\cos\left( 2\omega t+\phi\right) }\mu=0,\nonumber \end{align} which can be thought of as an extension of Ince's equation \cite{Mag:Win}, \cite{Menn68}. Here, we present two explicit solutions of this model when $\lambda^{\prime}=0$ and $\phi=0,$ $\pi/2$ as usually accepted in the literature. A general case can be discussed in a similar fashion. \section{Two Integrable Cases} For the Hamiltonian (\ref{RaiHamPX}) with $\lambda=$constant and $\phi=0: \begin{align} \widehat{H}\left( t\right) & =\frac{1}{2}\left( 1+\frac{\lambda}{\omega }\cos2\omega t\right) \left. \widehat{p}\right. ^{2}+\frac{\omega^{2} {2}\left( 1-\frac{\lambda}{\omega}\cos2\omega t\right) \left. \widehat{q}\right. ^{2}\label{HamI}\\ & +\frac{\lambda}{2}\sin2\omega t\ \left( \widehat{p}\ \widehat{q +\widehat{q}\ \widehat{p}\right) ,\nonumber \end{align} the corresponding Ince's equation \cite{CorSus11 \begin{align} & \left( \omega+\lambda\cos2\omega t\right) \mu^{\prime\prime +2\lambda\omega\sin2\omega t\ \mu^{\prime}\label{CharI}\\ & \quad+\left( \omega\left( \omega^{2}-3\lambda^{2}\right) -\lambda\left( \omega^{2}+\lambda^{2}\right) \cos2\omega t\right) \mu=0\nonumber \end{align} has the following standard solutions \begin{align} \mu_{0}\left( t\right) & =\left( \sinh\lambda t\ \cos\omega t+\cosh\lambda t\ \sin\omega t\right) /\omega,\label{Stand1}\\ \mu_{1}\left( t\right) & =\cosh\lambda t\ \cos\omega t+\sinh\lambda t\ \sin\omega t,\nonumber \end{align} which have been recently found with the aid of differential Galois theory \cite{AcostaSuazo13}, in particular, by using techniques for solving the one-dimensional stationary Schr{\"{o}}dinger equation such as algebrization procedure and Kovacic algorithm \cite{ac}, \cite{acmowe}. The Wronskian is given by $W\left( \mu_{0},\mu_{1}\right) =-1-\left( \lambda/\omega\right) \cos2\omega t.$ (Traditionally, Ince's equation was studied for the sake of periodic solutions, which do not exist for the degenerate parametric oscillators under consideration \cite{CorSus11}, \cite{Mag:Win}.) In the second case, when $\phi=\pi/2$ an \begin{align} \widehat{H}\left( t\right) & =\frac{1}{2}\left( 1-\frac{\lambda}{\omega }\sin2\omega t\right) \left. \widehat{p}\right. ^{2}+\frac{\omega^{2} {2}\left( 1+\frac{\lambda}{\omega}\sin2\omega t\right) \left. \widehat{q}\right. ^{2}\label{HamII}\\ & +\frac{\lambda}{2}\cos2\omega t\ \left( \widehat{p}\ \widehat{q +\widehat{q}\ \widehat{p}\right) ,\nonumber \end{align} we find standard solutions of the corresponding characteristic equation \begin{align} & \left( \omega-\lambda\sin2\omega t\right) \mu^{\prime\prime +2\lambda\omega\cos2\omega t\ \mu^{\prime}\label{CharII}\\ & \quad+\left( \omega\left( \omega^{2}-3\lambda^{2}\right) +\lambda\left( \omega^{2}+\lambda^{2}\right) \sin2\omega t\right) \mu=0\nonumber \end{align} in a similar fashion \begin{align} \mu_{0}\left( t\right) & =\frac{1}{\omega}e^{-\lambda t}\sin\omega t,\label{Stand2}\\ \mu_{1}\left( t\right) & =e^{\lambda t}\cos\omega t-\frac{\lambda}{\omega }e^{-\lambda t}\sin\omega t\nonumber \end{align} with the Wronskian $W\left( \mu_{0},\mu_{1}\right) =1-\left( \lambda /\omega\right) \sin2\omega t.$ \section{Generalized Fock States for Multi-Parameter Squeezed Photons} The linear dynamical invariants have the form \cite{SanSusVin} (see also Theorem~1 of Ref.~\cite{KretalSus13} for an abstract setting, which is adapted here)\footnote{It represents a general time-dependent Bogoliubov's transformation of the creation and annihilation operators.} \begin{align} \widehat{b}\left( t\right) & =\frac{e^{-2i\gamma\left( t\right) } {\sqrt{2}}\left( \beta\left( t\right) \widehat{q}+\varepsilon\left( t\right) +i\frac{\widehat{p}-2\alpha\left( t\right) \widehat{q -\delta\left( t\right) }{\beta\left( t\right) }\right) , \label{LinDynInv}\\ \widehat{b}^{\dagger}\left( t\right) & =\frac{e^{2i\gamma\left( t\right) }}{\sqrt{2}}\left( \beta\left( t\right) \widehat{q}+\varepsilon\left( t\right) -i\frac{\widehat{p}-2\alpha\left( t\right) \widehat{q -\delta\left( t\right) }{\beta\left( t\right) }\right) .\nonumber \end{align} The real-valued solution of the corresponding Ermakov-type system \cite{Lan:Lop:Sus}, \cite{Lop:Sus:VegaGroup} (subject to arbitrary real-valued initial data $\alpha\left( 0\right) ,$ $\beta\left( 0\right) \neq0,$ $\gamma\left( 0\right) ,$ $\delta\left( 0\right) ,$ $\varepsilon\left( 0\right) ,$ $\kappa\left( 0\right) ,$ which provides a natural multi-parameter description of the squeezing at $t=0$ \cite{KrySusVega13}) is given b \begin{align} & \alpha\left( t\right) =\alpha_{0}\left( t\right) -\beta_{0}^{2}\left( t\right) \frac{\alpha\left( 0\right) +\gamma_{0}\left( t\right) {\beta^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma _{0}\left( t\right) \right) ^{2}},\label{SolErmakovI}\\ & \beta\left( t\right) =-\frac{\beta\left( 0\right) \beta_{0}\left( t\right) }{\sqrt{\beta^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) ^{2}}},\label{SolIBeta}\\ & \gamma\left( t\right) =\gamma\left( 0\right) -\frac{1}{2}\arctan \frac{\beta^{2}\left( 0\right) }{2\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) }, \label{SolIGamma \end{align} an \begin{align} \delta\left( t\right) & =-\beta_{0}\left( t\right) \frac{\varepsilon \left( 0\right) \beta^{3}\left( 0\right) +2\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) \delta\left( 0\right) }{\beta ^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) ^{2}},\label{SolErmakovII}\\ \varepsilon\left( t\right) & =\frac{2\varepsilon\left( 0\right) \left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) -\beta\left( 0\right) \delta\left( 0\right) }{\sqrt{\beta^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) ^{2} },\label{SolIIEpsilon}\\ \kappa\left( t\right) & =\kappa\left( 0\right) -\varepsilon\left( 0\right) \beta^{3}\left( 0\right) \frac{\delta\left( 0\right) }{\beta ^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) ^{2}}\label{SolIIKappa}\\ & +\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) \frac{\varepsilon^{2}\left( 0\right) \beta^{2}\left( 0\right) -\delta ^{2}\left( 0\right) }{\beta^{4}\left( 0\right) +4\left( \alpha\left( 0\right) +\gamma_{0}\left( t\right) \right) ^{2}}\nonumber \end{align} in terms of the fundamental solutions: \begin{align} \alpha_{0}(t) & =\frac{1}{4a(t)}\frac{\mu_{0}^{\prime}(t)}{\mu_{0}(t) -\frac{d(t)}{2a(t)},\quad\beta_{0}(t)=-\frac{1}{\mu_{0}(t)}, \label{SolFundam \\ \gamma_{0}(t) & =\frac{\mu_{1}(t)}{2\mu_{0}(t)}+\frac{d(0)}{2a(0)}.\nonumber \end{align} Here \begin{equation} a\left( t\right) =\left\{ \begin{array} [c]{c \left( 1+\left( \lambda/\omega\right) \cos2\omega t\right) /2,\medskip\\ \left( 1-\left( \lambda/\omega\right) \sin2\omega t\right) /2 \end{array} \right. \qquad d\left( t\right) =\left\{ \begin{array} [c]{c \left( \lambda\sin2\omega t\right) /2\medskip,\\ \left( \lambda\cos2\omega t\right) /2 \end{array} \right. \label{CoeffsI-II \end{equation} an \begin{equation} \alpha_{0}(t)=\left\{ \begin{array} [c]{c \dfrac{\omega}{2}\medskip\dfrac{\cosh\lambda t\ \cos\omega t-\sinh\lambda t\ \sin\omega t}{\cosh\lambda t\ \sin\omega t+\sinh\lambda t\ \cos\omega t},\\ \dfrac{\omega}{2}\cot\omega t \end{array} \right. \label{AlphaI-II \end{equation \begin{equation} \beta_{0}(t)=\left\{ \begin{array} [c]{c -\dfrac{\omega}{\cosh\lambda t\ \sin\omega t+\sinh\lambda t\ \cos\omega t}\medskip,\\ -e^{\lambda t}\dfrac{\omega}{\sin\omega t \end{array} \right. \label{BetaI-II \end{equation \begin{equation} \gamma_{0}(t)=\left\{ \begin{array} [c]{c \dfrac{\omega}{2}\medskip\dfrac{\cosh\lambda t\ \cos\omega t+\sinh\lambda t\ \sin\omega t}{\cosh\lambda t\ \sin\omega t+\sinh\lambda t\ \cos\omega t},\\ \dfrac{\omega}{2}e^{2\lambda t}\cot\omega t \end{array} \right. \label{GammaI-II \end{equation} for the Hamiltonians (\ref{HamI}) and (\ref{HamII}), respectively. Equations (\ref{AlphaI-II})--(\ref{GammaI-II}) define the corresponding Green's functions (see \cite{AcostaSuazo13}, \cite{Cor-Sot:Lop:Sua:Sus}, \cite{CorSus11} and \cite{Suslov11} for more details; in the limit $\omega\rightarrow0$ for the second Hamiltonian, we obtain an interesting model of a strong coupling; the propagator in the interaction picture is traditionally used in the physical literature \cite{Mollow67}, \cite{Scully:Zubairy97}, \cite{Walls:Milburn}\footnote{See also Appendix~C for a brief summary.}). Explicit forms of solutions (\ref{SolErmakovI )--(\ref{SolIIKappa}) for both Hamiltonians are presented in Appendix~A; see (\ref{AlphaI})--(\ref{M(t)}) and (\ref{AlphaII})--(\ref{KappaII}), respectively. The corresponding dynamical Fock states $\left\vert \psi_{n}(t)\right\rangle ,$ where the phase $\kappa\left( t\right) $ finally shows up, can be obtained from now on in a standard fashion \cite{Fock32-2}, \cite{Fock34-3}, \cite{Ber:Lif:Pit} with the aid of our variable creation and annihilation operators (\ref{LinDynInv}) (see also \cite{KobManin89}, in particular, dialogues 8 and 9 and section 3.4, and (\ref{TimeSqueezeOper})). Under a certain condition, they do satisfy the time-dependent Schr\"{o}dinger equation (\ref{Schroedinger}) (see \cite{Fock28-2}, Lemma~2 of Ref.~\cite{KretalSus13} and (\ref{TimeSqueezeOper}) for more details). Moreover, the wave functions of degenerate parametric oscillators in coordinate representation are given by equation (18) of \cite{Lan:Lop:Sus} in terms of our explicit solutions of the Ermakov-type system; see also (\ref{WaveFunctionN}) below. \section{Mean Photon Number, Variances and Identities} The time-dependent variances \cite{KretalSus13} \begin{align} & \sigma_{p}=\langle\left( \left. \Delta\widehat{p}\right. \right) ^{2}\rangle=\left( n+\frac{1}{2}\right) \frac{4\alpha^{2}+\beta^{4} {\beta^{2}},\quad\sigma_{q}=\langle\left( \left. \Delta\widehat{q}\right. \right) ^{2}\rangle=\left( n+\frac{1}{2}\right) \frac{1}{\beta^{2 },\label{Variences}\\ & \quad\quad\qquad\sigma_{pq}=\frac{1}{2}\langle\Delta\widehat{p \ \Delta\widehat{q}+\Delta\widehat{q}\ \Delta\widehat{p}\rangle=\left( n+\frac{1}{2}\right) \frac{2\alpha}{\beta^{2}}\nonumber \end{align} with an invariant \cite{Dod:Man:Man94} \begin{equation} \left\vert \begin{array} [c]{cc \sigma_{p} & \sigma_{pq}\\ \sigma_{pq} & \sigma_{q \end{array} \right\vert =\sigma_{p}\sigma_{q}-\sigma_{pq}^{2}=\left( n+\frac{1 {2}\right) ^{2} \label{Dinvariant \end{equation} can be evaluated in terms of solutions of the Ermakov-type system for the generalized Fock states (described in general by Lemma~2 of \cite{KretalSus13 ; see also (\ref{TimeSqueezeOper})). Expressions (4.9)--(4.12) of \cite{KretalSus13} provide a convenient generic form for all quadratic operators under consideration. As a result, directly in the Schr\"{o}dinger picture, the average number of photons for these states $\left\vert \psi_{n}\left( t\right) \right\rangle : \begin{equation} \left\langle \widehat{N}\right\rangle (t)=\left\langle \widehat{a}^{\dagger }\ \widehat{a}\right\rangle =\left\langle \psi_{n}(t)\left\vert \frac {1}{2\omega}\left( \left. \widehat{p}\right. ^{2}+\omega^{2}\left. \widehat{q}\right. ^{2}\right) -\frac{1}{2}\right\vert \psi_{n (t)\right\rangle \label{AverageN \end{equation} is given b \begin{align} \left\langle \widehat{N}\right\rangle & =\left( n+\frac{1}{2}\right) \frac{4\alpha^{2}+\beta^{4}+\omega^{2}}{2\omega\beta^{2}}-\frac{1 {2}\label{AverageNFinal}\\ & +\frac{1}{2\omega}\left[ \left( \delta-\frac{2\alpha\varepsilon {\beta^{2}}\right) ^{2}+\frac{\omega^{2}\varepsilon^{2}}{\beta^{2}}\right] .\nonumber \end{align} For the reader's convenience, a useful expansion of the single photon mode Hamiltonian, $\widehat{H}=\left( \left. \widehat{p}\right. ^{2}+\omega ^{2}\left. \widehat{q}\right. ^{2}\right) /2,$ in terms of our variable creation and annihilation operators and solutions of the corresponding Ermakov-type system is given in Appendix~B. The general expression for the mean photon number can be significantly simplified for the Hamiltonians (\ref{HamI}) and (\ref{HamII}) of the degenerate parametric oscillators under consideration. Indeed, one can get in terms of \textquotedblleft slow\textquotedblright\ variables only \begin{align} & A\left( t\right) =\frac{4\alpha^{2}+\beta^{4}+\omega^{2}}{\beta^{2 }\label{Ainvariant}\\ & =\left\{ \begin{array} [c]{c \medskip\dfrac{\left( 2\alpha\left( 0\right) +\omega\right) ^{2}+\beta ^{4}\left( 0\right) \ }{2\beta^{2}\left( 0\right) }e^{2\lambda t \ \quad\qquad\\ +\dfrac{\left( 2\alpha\left( 0\right) -\omega\right) ^{2}+\beta^{4}\left( 0\right) \ }{2\beta^{2}\left( 0\right) }e^{-2\lambda t},\bigskip\\ \dfrac{\left( 4\alpha^{2}\left( 0\right) +\beta^{4}\left( 0\right) \right) e^{-2\lambda t}\ +\omega^{2}\ e^{2\lambda t}}{\beta^{2}\left( 0\right) \end{array} \right. \nonumber \end{align} an \begin{align} & B\left( t\right) =\left( \delta-\frac{2\alpha\varepsilon}{\beta}\right) ^{2}+\frac{\omega^{2}\varepsilon^{2}}{\beta^{2}}\ \label{Binvariant}\\ & =\left\{ \begin{array} [c]{c \dfrac{1}{2}\left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }-\dfrac{\omega \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) ^{2 e^{2\lambda t}\medskip\quad\quad\\ +\dfrac{1}{2}\left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }+\dfrac{\omega \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) ^{2 e^{-2\lambda t},\bigskip\\ \left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) ^{2 e^{-2\lambda t}+\dfrac{\varepsilon^{2}\left( 0\right) \omega^{2}}{\beta ^{2}\left( 0\right) }\ e^{2\lambda t \end{array} \right. \nonumber \end{align} for (\ref{HamI}) and (\ref{HamII}), respectively, with the common invariant \begin{align} \varepsilon^{2}+\frac{\delta^{2}}{\beta^{2}}\ & =\varepsilon^{2}\left( 0\right) +\frac{\delta^{2}\left( 0\right) }{\beta^{2}\left( 0\right) }=C,\label{Cinvariant}\\ \kappa-\frac{\delta\varepsilon}{2\beta} & =\kappa\left( 0\right) -\frac{\delta\left( 0\right) \varepsilon\left( 0\right) }{2\beta\left( 0\right) }=D. \label{Kinvariant \end{align} In compact form \begin{equation} \left\langle \widehat{N}\right\rangle =\left( n+\frac{1}{2}\right) \frac{A\left( t\right) }{2\omega}+\frac{B\left( t\right) }{2\omega -\frac{1}{2} \label{NCompact \end{equation} (see also \cite{Breit:Schill:Mlyn97}, \cite{Dod:Man:Man94}, \cite{KrySusVega13} and the references therein). For the initial coherent state, when $n=\alpha\left( 0\right) =0$ and $\beta^{2}\left( 0\right) =\omega,$ one get \begin{align} & \left\langle \widehat{N}\right\rangle =\sinh^{2}\lambda t\nonumber\\ & +\left\{ \begin{array} [c]{cc \medskip\dfrac{1}{2}\left( \varepsilon^{2}\left( 0\right) +\dfrac {\delta^{2}\left( 0\right) }{\omega}\right) \cosh2\lambda t-\dfrac {\delta\left( 0\right) \varepsilon\left( 0\right) }{\sqrt{\omega} \sinh2\lambda t & \text{for (\ref{HamI}),}\\ \dfrac{1}{2}\left( \varepsilon^{2}\left( 0\right) e^{2\lambda t +\dfrac{\delta^{2}\left( 0\right) }{\omega}e^{-2\lambda t}\right) \qquad\qquad & \text{for (\ref{HamII}). \end{array} \right. \label{Ncoherent \end{align} For the vacuum state, when $\delta\left( 0\right) =\varepsilon\left( 0\right) =0,$ we obtain a familiar expression from the theory of dynamical Casimir effect and spontaneous parametric fluorence \begin{equation} \left\langle \widehat{N}\right\rangle =\sinh^{2}\lambda t \label{Nvacuum \end{equation} (see, for example, \cite{Dod95}, \cite{Dod09}, \cite{Dodonov10}, \cite{Klyshko88}, \cite{Yariv11} and the references therein). The corresponding time-dependent variance \begin{equation} \text{Var\ }\widehat{N}=\left\langle \left( \widehat{N}-\langle \widehat{N}\rangle\right) ^{2}\right\rangle =\left\langle \left( \widehat{H}/\omega\right) ^{2}\right\rangle -\left\langle \left( \widehat{H}/\omega\right) \right\rangle ^{2} \label{VarN \end{equation} can also be evaluated in terms of \textquotedblleft slow\textquotedblrigh \ variables only \begin{align} \text{Var\ }\widehat{N} & =\frac{A^{2}\left( t\right) -4\omega^{2 }{8\omega^{2}}\left[ \left( n+\frac{1}{2}\right) ^{2}+\frac{3}{4}\right] \label{VarNFinal}\\ & +\frac{A\left( t\right) B\left( t\right) -C}{\omega^{2}}\left( n+\frac{1}{2}\right) \nonumber \end{align} with the help of expansion (\ref{HAB}). For the initial vacuum state, in particular, \begin{equation} \text{Var\ }\widehat{N}=\frac{1}{2}\sinh^{2}2\lambda t. \label{VarNvacuum \end{equation} Moreover, the second-order intensity correlation function \cite{Glauber63} \begin{equation} g^{\left( 2\right) }=\frac{\left\langle \left( \widehat{a}^{\dagger }\right) ^{2}\ \left. \widehat{a}\right. ^{2}\right\rangle }{\left\langle \widehat{a}^{\dagger}\ \widehat{a}\right\rangle ^{2}}=1+\frac{\text{Var\ \widehat{N}-\langle\widehat{N}\rangle}{\left. \langle\widehat{N \rangle\right. ^{2}}, \label{SecondCorr \end{equation} is explicitly given in terms of (\ref{NCompact}) and (\ref{VarNFinal}) for the multi-parameter squeezed number states $\left\vert \psi_{n}\left( t\right) \right\rangle $ (see also \cite{Agarwal87}, \cite{VourdasWeiner87}, \cite{KimOlivKnight89} and \cite{Marian91} for special cases). Explicit expressions for averages of operators $\widehat{q}$ and $\widehat{p}$ and for their time-dependent variances are given in Appendix~A. For a complete quantum mechanical description of the nonclassical states of light generated in the process of degenerate parametric amplification, the corresponding photon statistics are required. In vacuum, an explicit connection between the multi-parameter squeezed states and photon distributions is found in our recent publication \cite{KrySusVega13} (in coordinate representation, when $\widehat{q}=x$ and $\widehat{p =-i\partial/\partial x$). It can be readily extended to the models of optical degenerate parametric oscillators under consideration. One should replace $\alpha\rightarrow\alpha/\omega,$ $\beta\rightarrow\beta/\sqrt{\omega},$ $\delta\rightarrow\delta/\sqrt{\omega}$ in (6.9)--(6.11) of \cite{KrySusVega13} and a required modification of identities (6.11)--(6.20) is as follows. A joint complex identity \begin{equation} \frac{\delta}{\beta}+i\varepsilon=\left( \frac{\delta\left( 0\right) }{\beta\left( 0\right) }+i\varepsilon\left( 0\right) \right) e^{2i\gamma }, \label{ComplexIdentityBoth \end{equation} holds for both Hamiltonians (\ref{HamI}) and (\ref{HamII}), which implies (\ref{Cinvariant}). Moreover, by separating the \textquotedblleft fast\textquotedblright,\ $\omega,$\ and \textquotedblleft slow\textquotedblright,\ $\lambda,$ variables \begin{equation} \delta-\frac{2\alpha\varepsilon}{\beta}+i\frac{\omega\varepsilon}{\beta }=e^{-i\omega t}\ \xi\left( t\right) , \label{ComplexIdentitiesII \end{equation} wher \begin{equation} \xi\left( t\right) =\left\{ \begin{array} [c]{c \left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) \cosh\lambda t-\dfrac{\omega\varepsilon\left( 0\right) }{\beta\left( 0\right) \sinh\lambda t\medskip\quad\qquad\quad\\ +i\left( \dfrac{\omega\varepsilon\left( 0\right) }{\beta\left( 0\right) }\cosh\lambda t-\left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) \sinh\lambda t\right) ,\bigskip\\ \left( \delta\left( 0\right) -\dfrac{2\alpha\left( 0\right) \varepsilon\left( 0\right) }{\beta\left( 0\right) }\right) e^{-\lambda t}+i\dfrac{\omega\varepsilon\left( 0\right) }{\beta\left( 0\right) }e^{\lambda t \end{array} \right. \label{KSI \end{equation} for the Hamiltonian (\ref{HamI}) and (\ref{HamII}), respectively. As a by-product, we derive identities (\ref{Binvariant}). In a similar fashion \begin{equation} \frac{\omega+\beta^{2}}{2}-i\alpha=\frac{1}{2}e^{i\omega t}\ \frac{\eta\left( t\right) }{z\left( t\right) },\qquad\frac{\omega-\beta^{2}}{2 +i\alpha=\frac{1}{2}e^{-i\omega t}\ \frac{\zeta\left( t\right) }{z\left( t\right) }, \label{ComplexIdentitiesIandII \end{equation} wher \begin{equation} z\left( t\right) =\left\{ \begin{array} [c]{c \dfrac{\omega\cos\omega t+\left( 2\alpha\left( 0\right) +i\beta^{2}\left( 0\right) \right) \sin\omega t}{\omega}\cosh2\lambda t\medskip\quad\\ +\dfrac{\left( 2\alpha\left( 0\right) +i\beta^{2}\left( 0\right) \right) \cos\omega t+\omega\sin\omega t}{\omega}\sinh2\lambda t,\bigskip\\ e^{2\lambda t}\omega\cos\omega t+2\alpha\left( 0\right) \sin\omega t+i\beta^{2}\left( 0\right) \sin\omega t \end{array} \right. \label{Zeta \end{equation} an \begin{equation} \eta\left( t\right) =\left\{ \begin{array} [c]{c \left( \omega+\beta^{2}\left( 0\right) \right) \cosh2\lambda t\ +2\alpha\left( 0\right) \sinh\lambda t\medskip\qquad\quad\\ -i\left( 2\alpha\left( 0\right) \cosh2\lambda t\ +\medskip\left( \omega-\beta^{2}\left( 0\right) \right) \sinh\lambda t\right) ,\bigskip\\ -2i\alpha\left( 0\right) +\beta^{2}\left( 0\right) +\omega e^{2\lambda t \end{array} \right. \label{Uta \end{equation} $\medskip$ \begin{equation} \zeta\left( t\right) =\left\{ \begin{array} [c]{c \medskip\left( \omega-\beta^{2}\left( 0\right) \right) \cosh2\lambda t\ +2\alpha\left( 0\right) \sinh\lambda t\qquad\quad\\ +i\left( 2\alpha\left( 0\right) \cosh2\lambda t\ +\medskip\left( \omega+\beta^{2}\left( 0\right) \right) \sinh\lambda t\right) ,\bigskip\\ 2i\alpha\left( 0\right) -\beta^{2}\left( 0\right) +\omega e^{2\lambda t \end{array} \right. \label{Vta \end{equation} for the Hamiltonians (\ref{HamI}) and (\ref{HamII}), respectively. (Computational details are left to the reader; we use Mathematica and Maple to verify our calculations.) Vector $z\left( t\right) $ is related to the complex parametrization found in \cite{KretalSus13}, \cite{KrySusVega13} (see also \cite{Dod:Man79} and \cite{Har:Ben-Ar:Mann11} ). Having these modifications in mind, in the next section, we will be able to derive variable probability amplitudes and photon statistics in terms of hypergeometric series which are similar to those in \cite{KrySusVega13} (see also \cite{Taka65}, \cite{VenkSatya85}, \cite{Marian91}, \cite{Marian92}). \section{Time-Dependent Probability Amplitudes and Photon Statistics} In coordinate representation, when $\widehat{q}=x$ and $\widehat{p =-i\partial/\partial x,$ the wave functions of the optical degenerate parametric oscillators under consideration take the form\footnote{A direct Mathematica verification of these solutions is given by Christoph Koutschan. \begin{equation} \psi_{n}\left( x,t\right) =e^{i\left( \alpha x^{2}+\delta x+\kappa\right) +i\left( 2n+1\right) \gamma}\sqrt{\frac{\beta}{2^{n}n!\sqrt{\pi} }\ e^{-\left( \beta x+\varepsilon\right) ^{2}/2}\ H_{n}\left( \beta x+\varepsilon\right) , \label{WaveFunctionN \end{equation} where $H_{n}\left( x\right) $ are the Hermite polynomials \cite{Ni:Su:Uv} and explicit solutions of the corresponding Ermakov-type system are given by (\ref{AlphaI})--(\ref{KappaI}) and (\ref{AlphaII})--(\ref{KappaII}) for the Hamiltonians (\ref{HamI}) and (\ref{HamII}), respectively. (An important special case $\lambda=0$ was originally investigated in \cite{Marhic78}; see also \cite{Dod:Man79}, \cite{LopSusVegaHarm}, \cite{Lop:Sus:VegaGroup}, \cite{KrySusVega13} and the references therein. In paraxial optics, this solution may be associated with a new model of periodic lens-like medium; see, for example, \cite{Mah:Sua:Sus13}.) In terms of the stationary harmonic oscillator wavefunctions \begin{equation} \Psi_{n}\left( x\right) =\left( \frac{\omega}{\pi}\right) ^{1/4 \frac{e^{-\omega x^{2}/2}}{\sqrt{2^{n}n!}}\ H_{n}\left( x\sqrt{\omega }\right) , \label{HarmonicWaveFunctions \end{equation} the eigenfunction expansion has the form \cite{KrySusVega13} \begin{equation} \psi_{n}\left( x,t\right) =e^{i\left( 2n+1\right) \gamma}\sqrt{\frac {\beta}{\omega^{1/2}}}\sum_{m=0}^{\infty}C_{mn}\left( t\right) \ \Psi _{m}\left( x\right) . \label{ExpansionGeneral \end{equation} The time-dependent coefficients can be found in terms of our solutions of the Ermakov-type system as follows \begin{align} & C_{mn}\left( t\right) =\sum_{k=0}^{\infty}M_{mk}\left( \alpha ,\beta\right) \ T_{kn}\left( \varepsilon,\frac{\delta}{\beta},\kappa\right) \label{ExpansionCoeffs}\\ & =\sum_{k=0}^{\infty}T_{mk}\left( \frac{\omega^{1/2}\varepsilon}{\beta },\omega^{-1/2}\left( \delta-\frac{2\alpha\varepsilon}{\beta}\right) ,\kappa-\frac{\alpha\varepsilon^{2}}{\beta^{2}}\right) \ M_{kn}\left( \alpha,\beta\right) .\nonumber \end{align} Here \begin{align} & T_{mn}\left( A,B,\Gamma\right) =i^{m-n}\frac{e^{i\left( \Gamma -AB/2\right) }\ e^{-\nu/2}\ }{\sqrt{m!n!}}\ \left( \frac{iA+B}{\sqrt{2 }\right) ^{m}\left( \frac{iA-B}{\sqrt{2}}\right) ^{n \label{HeisenbergGroupMatrix}\\ & \qquad\times\ _{2}F_{0}\left( -n,\ -m;\ -\frac{1}{\nu}\right) ,\qquad \nu=\left( A^{2}+B^{2}\right) /2\nonumber \end{align} an \begin{align} & M_{mn}\left( \alpha,\beta\right) =i^{n}\sqrt{\frac{2^{m+n}\omega {m!n!\pi}}\ \Gamma\left( \frac{m+n+1}{2}\right) \label{MartixSU(1,1)Hyper}\\ & \times\ \frac{\left( \dfrac{\omega-\beta^{2}}{2}+i\alpha\right) ^{m/2}\left( \dfrac{\omega-\beta^{2}}{2}-i\alpha\right) ^{n/2}}{\left( \dfrac{\omega+\beta^{2}}{2}-i\alpha\right) ^{\left( m+n+1\right) /2 }\nonumber\\ & \times~_{2}F_{1}\left( \begin{array} [c]{c -m,\quad-n\\ \dfrac{1}{2}\left( 1-m-n\right) \end{array} ;\dfrac{1}{2}\left( 1\pm\frac{2i\beta\sqrt{\omega}}{\sqrt{4\alpha^{2}+\left( \beta^{2}-\omega\right) ^{2}}}\right) \right) .\nonumber \end{align} (We use the standard definition of generalized hypergeometric series and an integral evaluated by Bailey \cite{Bailey48}; see \cite{KrySusVega13} for more details.) These general expressions can be significantly simplified, once again, with the help of identities found in the previous section. By separating the \textquotedblleft fast\textquotedblright\ and \textquotedblleft slow\textquotedblright\ variables, one get \begin{align} & T_{mn}\left( \varepsilon,\frac{\delta}{\beta},\kappa\right) =e^{2i\left( m-n\right) \gamma}\ S_{mn},\qquad S_{mn}=\left. T_{mn}\left( \varepsilon ,\frac{\delta}{\beta},\kappa\right) \right\vert _{t=0}\ ,\label{Mat1}\\ & T_{mn}\left( \frac{\omega^{1/2}\varepsilon}{\beta},\omega^{-1/2}\left( \delta-\frac{2\alpha\varepsilon}{\beta}\right) ,\kappa-\frac{\alpha \varepsilon^{2}}{\beta^{2}}\right) =e^{i\omega\left( n-m\right) t \ R_{mn}\left( t\right) \label{Mat2 \end{align} an \begin{equation} M_{mn}\left( \alpha,\beta\right) =e^{-i\omega\left( m+1/2\right) t}e^{-i\left( 2n+1\right) \gamma}\sqrt{\frac{\beta\left( 0\right) }{\beta }}\ N_{mn}\left( t\right) . \label{Mat3 \end{equation} Here, by definitio \begin{align} & R_{mn}\left( t\right) =\frac{i^{m+n}}{\sqrt{m!n!\left( 2\omega\right) ^{m+n}}}\ \ \xi^{m}\left( \xi^{\ast}\right) ^{n}\label{RT}\\ & \quad\times\ e^{iD-B\left( t\right) /\left( 4\omega\right) }\ _{2 F_{0}\left( -n,\ -m;\ -\frac{2\omega}{B\left( t\right) }\right) \nonumber \end{align} an \begin{align} & N_{mn}\left( t\right) =i^{n}\sqrt{\frac{2^{m+n+1}\omega}{m!n!\pi }\ \Gamma\left( \frac{m+n+1}{2}\right) \frac{\zeta^{m/2}\left( \zeta^{\ast }\right) ^{n/2}}{\eta^{\left( m+n+1\right) /2}}\label{NT}\\ & \quad\times~_{2}F_{1}\left( \begin{array} [c]{c -m,\quad-n\\ \dfrac{1}{2}\left( 1-m-n\right) \end{array} ;\dfrac{1}{2}\left( 1+2i\sqrt{\frac{\omega}{A\left( t\right) -2\omega }\right) \right) .\nonumber \end{align} The asterisk denotes complex conjugation. (Quadratic transformation (6.8) of \cite{KrySusVega13} can be used in order to complete our evaluation; see also \cite{Taka65}, \cite{VenkSatya85}, \cite{Marian91}, \cite{Marian92}, \cite{MarianMarianI93} and Appendix~D. The special case $\lambda=0,$ $\omega=1$ corrects a typo in \cite{KrySusVega13}.) As a result, we finally obtai \begin{equation} \psi_{n}\left( x,t\right) =\sqrt{\frac{\beta\left( 0\right) }{\omega ^{1/2}}}\sum_{m=0}^{\infty}c_{mn}\left( t\right) \ e^{-i\omega\left( m+1/2\right) t}\ \Psi_{m}\left( x\right) , \label{ExpansionFinalIndependent \end{equation} where the time-dependent probability amplitudes are given b \begin{equation} c_{mn}\left( t\right) =\sum_{k=0}^{\infty}N_{mk}\left( t\right) \ S_{kn}=\sum_{k=0}^{\infty}R_{mk}\left( t\right) \ N_{kn}\left( t\right) \label{CoeffFinalDependent \end{equation} in terms of our \textquotedblleft slow\textquotedblright\ variables $\xi,$ $\eta,$ and $\zeta$ (\textquotedblleft adiabatic invariants\textquotedblright) only for all real-valued initial data/constants of motion (of the corresponding Ermakov-type system). Thus, the total probability amplitudes are associated with the product of two infinite matrices related to the Poisson and Pascal distributions; see \cite{KrySusVega13} for more details. The quantities $\left\vert c_{mn}\left( t\right) \right\vert ^{2}$ explicitly determine the corresponding variable photon statistics. In mathematical terms, our expansion (\ref{ExpansionFinalIndependent}) gives a mapping between two complete sets of vectors in $\mathcal{L}^{2}\left( \mathbb{R} \right) ;$ namely, the transition matrix (\ref{CoeffFinalDependent}) between the generalized harmonic, or \textquotedblleft squeezed\textquotedblrigh \ states, $\left\{ \psi_{n}\left( x,t\right) \right\} _{n=0}^{\infty},$ and the standard Fock ones, $\left\{ e^{-i\omega\left( m+1/2\right) t}\ \Psi_{m}\left( x\right) \right\} _{n=0}^{\infty},$ in coordinate representation; the latter are usually being recorded by a photodetector in clever quantum nonlinear optics experiments (see, for example, \cite{Klyshko94}, \cite{Klyshko96}, \cite{Breit:Schill:Mlyn97}, \cite{LvRay09 , \cite{Scully:Zubairy97}, \cite{Klyshko98}, \cite{BachorRalph04}, \cite{Klyshko11}, \cite{Haroche13} and the references therein). \section{The Wigner and Moyal Functions} In coordinate representation, Wigner's functions can be derived by following our analysis in the simplest case, when $\lambda=0$ \cite{KrySusVega13} (see also \cite{Dod:Man:Man94}, \cite{HilletyetalWigner84}, \cite{Schleich01} and \cite{Schradeetal95} for a general approach). For the multi-parameter \textquotedblleft dynamical vacuum state\textquotedblright, when $n=0,$ for example, the final result is given b \begin{equation} W\left( x,p,t\right) =\frac{1}{\pi\sqrt{\omega}}\exp\left( -Q\left( U,V\right) \right) , \label{Wigner \end{equation} wher \begin{equation} Q\left( U,V\right) =\frac{\beta^{2}\left( 0\right) \left[ \beta\left( 0\right) U+\omega\varepsilon\left( 0\right) \right] ^{2}+\left[ 2\alpha\left( 0\right) U-\omega\left( V-\delta\left( 0\right) \right) \right] ^{2}}{\beta^{2}\left( 0\right) \omega^{2}} \label{QuadraticWigner \end{equation} in the rotating $X=\omega x\cos\omega t-p\sin\omega t,$ $P=\omega x\sin\omega t+p\sin\omega t$ and \textquotedblleft squeezing\textquotedblrigh \ coordinate \begin{equation} U=\left\{ \begin{array} [c]{c \left[ \left( X-P\right) e^{\lambda t}+\left( X+P\right) e^{-\lambda t}\right] \medskip/2\ ,\\ Xe^{-\lambda t \end{array} \right. \label{SqueezingU \end{equation \begin{equation} V=\left\{ \begin{array} [c]{c \medskip\left[ \left( P-X\right) e^{\lambda t}+\left( X+P\right) e^{-\lambda t}\right] /2\ ,\\ Pe^{\lambda t \end{array} \right. \label{SqueezingV \end{equation} in the quantum phase space for the Hamiltonians (\ref{HamI}) and (\ref{HamII}), respectively. Our result is consistent with \cite{Mollow67}, \cite{Agarwal87}, \cite{FernCollett88}, \cite{Dod:Man:Man94} and \cite{Walls:Milburn} in the case of the initial vacuum state for the second Hamiltonian. A similar consideration is valid for Moyal's functions \cite{KrySusVega13}, \cite{Schleich01}. An example of generation of the squeezed vacuum state is presented in Figure~1. \begin{figure}[htbp] \centerin \scalebox{.75 {\includegraphics{DPAfigure1.eps}} \caption{Phase space animation showing rotation and squeezing of contours $Q\left(U, V\right)=2$ for the first Hamiltonian in (\ref{QuadraticWigner})--(\ref{SqueezingV}) with $\omega=1$ and $\lambda=0.25$ for the initial vacuum state. The minimum-uncertainty squeezed states occur at $t_{\text{min}} = 0,\ \pi/4,\ 3\pi/4,$ etc. (The color version of this figure is available only in the electronic edition.)} \end{figure} In summary, we have investigated, as explicitly as possible in coordinate representation, all quantum statistical properties of the single mode multi-parameter squeezed photon states in two classical models of optical degenerate parametric oscillators. In principle, one can reproduce our explicit time-dependent photon statistics in a more general operator approach with the help of methods of representation theory (see, for example, \cite{Marian91}, \cite{MarianMarianI93}, \cite{Schumaker86}, \cite{SchumakerCaves85} and Appendix~D). It will be of interest to everyone who studies quantum optics and cavity QED. From the mathematical standpoint, our results motivate further investigations of the Hamiltonian (\ref{RaiHam}) and the corresponding generalized Ince equation (\ref{RaiCharMu}). Last but not least, it is worth noting that variable linear terms in creation and annihilation operators, which correspond to a classical current \cite{Glauber63}, \cite{Hanamuraetal07}, \cite{Scully:Zubairy97}, \cite{Walls:Milburn}, can be easily incorporated into this Hamiltonian in our approach. Lossy medium models may also be taken into consideration. \medskip \noindent\textbf{Acknowledgments.\/} We would like to thank Albert Boggess for help and encouragement. This research was partially supported by AFOSR grant FA9550-11-1-0220. The first named author was partially supported by MICIIN/FEDER MTM2009--06973, gencat 2009SGR859 and DIDI -- Universidad del Norte. One of the authors (ES) was also supported by the AMS-Simons Travel Grants, with support provided by the Simons Foundation. Sergei Suslov thanks Marlan O. Scully, Wolfgang Schleich and M. Suhail Zubairy for valuable discussions during his visit to Institute for Quantum Science and Engineering, Texas A\&M University. We are grateful to Victor V. Dodonov, Vladimir I. Man'ko, Geza Giedke, Boris A. Malomed, Paulina Marian, Giuseppe Ruoso, Christoph Koutschan, Jos\'{e} M. Vega-Guzm\'{a}n and Andreas Ruffing for valuable comments and important references and to Kamal Barley for graphics enhancement.
\section{Introduction} In the study of knots and links in $3$-dimensional spaces (such as handlebodies, knot complements, c.c.o. $3$-manifolds) it can prove very useful to take an approach via braids, as the use of braids provides more structure and more control on the braid equivalence moves. \smallbreak In \cite{LR1} geometric braid equivalence has been given for isotopic knots and links in knot complements and in c.c.o. $3$-manifolds with integral surgery description. This was done by fixing (pointwise) a description for the knot complement or the $3$-manifold via a closed braid $\widehat{B}$ in $S^3$. Then, knots and links in such a $3$-manifold are represented unambiguously by {\it mixed links} in $S^3$ (Figure~\ref{A mixed link}) and braids in the manifold are represented unambiguously by geometric {\it mixed braids} in $S^3$. A geometric mixed braid is a braid in $S^3$ that contains $B$ as a fixed subbraid (Figure~\ref{mixedbraidsLmoves}). Isotopy for links in a c.c.o. 3-manifold $M$ is then translated into appropriate moves for mixed links in $S^3$, which comprise isotopy in the complement $S^3\backslash \widehat{B}$ together with the {\it band moves} coming from the handle sliding moves in $M$, and are related to the surgery description of $M$ (Figure~\ref{Band Moves}). For the braid equivalence, the authors sharpened first the classic Markov theorem giving only one type of moves, the {\it $L$-moves} (Figure~\ref{mixedbraidsLmoves}), which are geometric as well as algebraic. Then, with the use of the $L$-moves and the {\it braid band moves} (Figure~\ref{pbbm}) they formulated geometric mixed braid equivalence for knots and links in knot complements and in c.c.o. $3$-manifolds obtained from $S^3$ by integral surgery. Further, in \cite{LR2} the same authors provided algebraic formulations for the geometric mixed braid equivalences in knot complements and in c.c.o. $3$-manifolds, using the {\it mixed braid groups} $B_{m,n}$ (Eq.~\ref{B} and Figure~\ref{Loops and Crossings}), introduced and studied in \cite{La2}, and the techniques of parting and combing mixed braids (Figure~\ref{partingandcombing}). {\it Parting} a geometric mixed braid means to separate its strands into two sets: the strands of the fixed subbraid $B$ and the moving strands of the braid representing a link in the $3$-manifold. {\it Combing} a parted mixed braid means to separate the braiding of the fixed subbraid $B$ from the braiding of the moving strands (Figure~\ref{combing}). The above techniques have been also applied in \cite{HL} for links in a handlebody. \smallbreak Integral surgery is a special case of rational surgery and, by a classic result of topology, every c.c.o. $3$-manifold can be obtained by surgery (integral or rational) along a framed link in $S^3$. Moreover, in the case of integral surgery the components of the framed link may be all assumed to be simple closed curves, giving rise to a closed pure braid. There are $3$-manifolds which have simpler description when obtained from $S^3$ by rational surgery. Representative examples are the lens spaces $L(p,q)$ which with rational surgery description $p/q$ are obtained from the trivial knot, while with integral surgery description a non-trivial link is needed, see for example \cite{Ro} p.322. A simpler surgery description of a $3$-manifold $M$ is expected to induce simpler algebraic expressions for the braid equivalence in $M$. As an example, compare Section~4 in \cite{LR2} with Section~5.1 in this paper for the case of lens spaces. In this paper the braid equivalence is in the mixed braid groups $B_{1,n}$ and there is only one expression for the braid band moves. In \cite{LR2} there are many, according to the surgery coefficient of each strand of the surgery pure braid and, on top of that, combing is also needed. \smallbreak The purpose of this paper is to provide mixed braid equivalence, geometric as well as algebraic, for isotopic oriented links in c.c.o. $3$-manifolds obtained by rational surgery along framed links in $S^3$. We follow the setting and the techniques of \cite{LR1,LR2} and we use the results therein. More precisely, let $M$ be a c.c.o. $3$-manifold obtained by rational surgery along a framed link $\widehat{B}$ in $S^3$. Note that the surgery braid $B$ may not be assumed to be a pure braid. Let $s$ be a surgery component of $B$ with surgery description $p/q$ consisting of $k$ strands, $s_1, \ldots, s_k$. When a {\it geometric braid band move} on $s$ occurs, $k$ sets of $q$ new strands appear, each one running in parallel to a strand of $s$, and also a $(p,q)$-torus braid wraps around the last strand, $s_k$, $p$ times (see Figure~\ref{homtref} for an example). These moves together with the $L$-moves lead to the {\it geometric mixed braid equivalence} in $M$ (Theorem~\ref{geommarkovrat}). The geometric braid band moves are much more complicated than in the case of integral surgery \cite{LR1}. However, a sharpened version of the Reidemeister theorem for links in $M$ (Theorem~\ref{reidemrational}; see also \cite{LR1, Sk, Su}), whereby only one type of band moves is used in the isotopy equivalence (Figure~\ref{bmoves}), makes the proof of Theorem~\ref{geommarkovrat} lighter. In order to move toward algebraic statements we use the notion of a $q$-strand cable and we apply the techniques and results from \cite{LR2}. A {\it $q$-strand cable} encloses a set of $q$ new strands arising from the performance of a geometric braid band move. We show first that parting a $q$-strand cable is equivalent to parting each strand of the cable one by one; that is, parting and cabling commute (Figure~\ref{stndpartcable}). Treating now each one of the $k$ $q$-strand cables as one thickened strand leads to the {\it parted mixed braid equivalence} (Theorem~\ref{qparted}), assuming the corresponding result in \cite{LR2}. We continue by providing algebraic expressions for parted cables (Lemma~\ref{cableloop}) and also for the loopings between the strands of $B$ and the remaining strands of the mixed braid after a braid band move is performed (Figures~\ref{cablesa}, \ref{cablesa2}). Then, we part locally the $(p,q)$-torus braid and the crossing of a parted mixed braid band move (Figures~\ref{cables2}, \ref{cablesc}), obtaining algebraic expressions for these parts. From the above we obtain the algebraic expression of an {\it algebraic braid band move}. Finally, we do combing through $B$ and we show that combing a $q$-strand cable is equivalent to combing each strand of the cable one by one; that is, combing and cabling commute (Figures~\ref{combcable1} and \ref{combcableproof2}). So, assuming the corresponding result in \cite{LR2}, we obtain the {\it algebraic mixed braid equivalence} for links in $M$ in terms of the mixed braid groups $B_{m,n}$, which is our main result (Theorem~\ref{algmarkov}). The paper is organized as follows. In Section~1 we recall the setting and the essential techniques and results from \cite{LR1,LR2}. In Section~2 we prove the sharpened version of the Reidemeister theorem for knots and links in c.c.o. $3$-manifolds with rational surgery description. In Section~3 we give the geometric braid equivalence for links in such $3$-manifolds (Theorem~\ref{geommarkovrat}) and in Section~4 we give the corresponding algebraic version (Theorem~\ref{algmarkov}). Finally, in Section~5 we give the concrete algebraic expressions for the mixed braid equivalences in lens spaces, in Seifert manifolds and in homology spheres obtained from $S^3$ by rational surgery along the trefoil knot. Also, we provide some figures (Figures~\ref{seifcomb1}--\ref{seifcomb6}) illustrating in a generic example all the techniques developed and used in this paper step-by-step. \smallbreak The algebraic mixed braid equivalence gives good control over the band moves, so our results can be applied for the study of skein modules of c.c.o. $3$-manifolds obtained by rational surgery along a framed link in $S^3$ and for constructing invariants of links in such $3$-manifolds, using algebraic means. For a construction of the analogues of the HOMFLYPT polynomial for knots and links in the solid torus via this approach, the reader is referred to \cite{La3}. In a sequel paper we shall use Theorem~\ref{algmarkov} and the results of \cite{La3} in order to compute the HOMFLYPT skein module of the lens spaces $L(p,q)$ with this method. \section{Background results in the case of integral surgery description} In this section we recall the topological settings and results from [LR1] and [LR2] for expressing link isotopy in a c.c.o. $3$-manifold $M$ in terms of geometric and algebraic mixed braid equivalence in $S^3$. For the rest of this section we fix a c.c.o. $3$-manifold $M$, which is obtained by integral surgery on a framed link in $S^3$ given in the form of a closed braid $\widehat{B}$. We shall denote $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$. \subsection{Mixed links and isotopy} Let $L$ be an oriented link in $M$. Fixing $\widehat{B}$ pointwise, $L$ can be represented unambiguously by a \textit{mixed link} in $S^3$ denoted $\widehat{B}\cup L$, that is, a link in $S^{3}$ consisting of the \textit{fixed part} $\widehat{B}$ and the \textit{moving part} $L$ that links with $\widehat{B}$. A \textit{mixed link diagram }is a diagram $\widehat{B}\cup \widetilde{L}$ of $\widehat{B}\cup L$ on the plane of $\widehat{B}$, where this plane is equipped with the top-to-bottom direction of the braid $B$, see Figure \ref{A mixed link} for an example. \begin{figure} \begin{center} \includegraphics[width=1.3in]{mixedlink} \end{center} \caption{ A mixed link in $S^{3}$. } \label{A mixed link} \end{figure} An isotopy of $L$ in $M$ can be translated into a finite sequence of moves of the mixed link $\widehat{B} \bigcup L$ in $S^3$ as follows. As we know, surgery along $\widehat{B}$ is realized by taking first the complement $S^3\backslash \widehat{B}$ and then attaching to it solid tori according to the surgery description. Thus, isotopy in $M$ can be viewed as certain moves in $S^3$, namely, isotopy in $S^3\backslash \widehat{B}$ together with the band moves in $S^3$, which are similar to the second Kirby move. Isotopy in $S^3\backslash \widehat{B}$ is realized by the classical Reidemeister moves and planar moves for the moving part together with the {\it extended Reidemeister moves}. These are the Reidemeister II and III moves involving the fixed and the moving part of the mixed link (cf. Definition~5.1 \cite{LR1}). A \textit{band move} is a non-isotopy move in $S^3 \backslash \widehat{B}$ that reflects isotopy in $M$ and is the band connected sum of a component, say $s$, of $L$ with the specified (from the framing) parallel curve $l$ of a surgery component, say $c$, of $\widehat{B}$. Note that $l$ bounds a disc in $M$. There are two types of band moves according to the orientations of the component $s$ of $L$ and of the surgery curve $c$, as illustrated and exemplified in Figure~\ref{Band Moves}. In the $\alpha$-type the orientation of $s$ is opposite to the orientation of $c$ (and of its parallel curve $l$), but after the performance of the move their orientations agree. In the $\beta$-type the orientation of $s$ agrees initially with the orientation of $c$, but disagrees after the performance of the move. Note that the two types of band moves are related by a twist of $s$ (Reidemeister I move in $S^3\backslash \widehat{B}$). \begin{figure} \begin{center} \includegraphics[width=5.5in]{integralbandmoves} \end{center} \caption{ The two types of $\mathbb{Z}$-band moves. } \label{Band Moves} \end{figure} The above are summarized in the following analogue of the Reidemeister theorem for oriented links in $M$. \begin{thm}[Reidemeister for $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$, Thm. 5.8 \cite{LR1}] \label{reidemintegral} Two oriented links $L_1$, $L_2$ in $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding mixed link diagrams of theirs, $\widehat{B} \cup \widetilde{L_1}$ and $\widehat{B} \cup \widetilde{L_2}$, differ by isotopy in $S^3 \backslash \widehat{B}$ together with a finite sequence of the two types $\alpha$ and $\beta$ of band moves. \end{thm} \subsection{Geometric mixed braids and the $L$-moves} In order to translate isotopy of links in the $3$-manifold $M$ to braid equivalence, we need to introduce the notion of a geometric mixed braid. A \textit{geometric mixed braid} related to $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$ and to a link $K$ in $M$, is an element of the group $B_{m+n}$, where $m$ strands form the fixed \textit{surgery braid} $B$ and $n$ strands form the {\it moving subbraid\/} $\beta$ representing the link $K$ in $M$. For an illustration see the middle picture of Figure~\ref{mixedbraidsLmoves}. We further need the notions of the $L$-moves and the braid band moves. \begin{defn} [$L$-moves and $\mathbb{Z}$-braid band moves, Definitions 2.1 and 5.6 \cite{LR1}] \rm \noindent {\it(i)} Let $B \bigcup \beta$ be a geometric mixed braid in $S^3$ and $P$ a point of an arc of the moving subbraid $\beta$, such that $P$ is not vertically aligned with any crossing or endpoint of a braid strand. Doing an {\it $L$-move\/} at $P$ means to cut the arc at $P$, to bend the two resulting smaller arcs slightly apart by a small isotopy and to stretch them vertically, the upper downward and the lower upward, and both {\it over\/} or {\it under\/} all other arcs of the diagram, so as to introduce two new corresponding moving strands with endpoints on the vertical line of the point $P$. Stretching the new strands over will give rise to an {\it $L_o$-move\/} and under to an {\it $L_u$-move\/}. For an illustration see Figure~\ref{mixedbraidsLmoves}. Two geometric mixed braids shall be called {\it $L$-equivalent} if and only if they differ by a sequence of $L$-moves and braid isotopy. \noindent {\it(ii)} A {\it geometric $\mathbb{Z}$-braid band move\/} is a move between geometric mixed braids which is a band move between their closures. It starts with a little band oriented downward, which, before sliding along a surgery strand, gets one twist {\it positive\/} or {\it negative\/} (see Figure~\ref{pbbm} (a) and (b)). \end{defn} \begin{figure} \begin{center} \includegraphics[width=5in]{Lmoves} \end{center} \caption{A geometric mixed braid and the two types of $L$-moves. } \label{mixedbraidsLmoves} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.4in]{pbbm} \end{center} \caption{ A geometric, a parted and an algebraic $\mathbb{Z}$-braid band move (top part of (d)). } \label{pbbm} \end{figure} \begin{remark} \label{Lmove} \rm {\it (i)} In \cite{LR1} it is shown that classical braid equivalence in $S^3$ is generated only by the $L$-moves. This implies that braid conjugation and in particular change of the order of the endpoints of a braid can be realized by $L$-moves. A demonstration can be found in \cite{LR2} Figure~14. \\ \noindent {\it (ii)} A geometric $\mathbb{Z}$-braid band move may be always assumed, up to $L$-equivalence, to take place at the top part of a mixed braid and on the right of the specific surgery strand (\cite{LR2} Lemma~5). \end{remark} In [LR1] the following theorem was proved for isotopic links in $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$. \begin{thm}[Geometric braid equivalence for $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$, Theorem 5.10 \cite{LR1}] \label{geommarkov} Two oriented links in $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding geometric mixed braids in $S^3$ differ by mixed braid isotopy, by $L$-moves that do not touch the fixed subbraid $B$ and by the geometric $\mathbb{Z}$-braid band moves. \end{thm} \subsection{Algebraic mixed braids and their equivalence} Let $M=\chi_{_{\mathbb{Z}}}(S^3, \widehat{B})$. We will pass from the geometric braid equivalence to an algebraic statement for links in $M$. An \textit{algebraic mixed braid} is a mixed braid on $m+n$ strands such that the first $m$ strands are fixed and form the identity braid on $m$ strands and the next $n$ strands are moving strands and represent a link in the manifold $M$. The set of all algebraic mixed braids on $m+n$ strands forms a subgroup of $B_{m+n}$, denoted $B_{m,n}$, and called {\it mixed braid group}. In [La2] the mixed braid groups $B_{m,n}$ have been introduced and studied and it is shown that $B_{m,n}$ has the presentation: \begin{equation} \label{B} B_{m,n} = \left< \begin{array}{ll} \begin{array}{l} a_1, \ldots, a_m, \\ \sigma_1, \ldots ,\sigma_{n-1} \\ \end{array} & \left| \begin{array}{l} \sigma_k \sigma_j=\sigma_j \sigma_k, \ \ |k-j|>1 \\ \sigma_k \sigma_{k+1} \sigma_k = \sigma_{k+1} \sigma_k \sigma_{k+1}, \ \ 1 \leq k \leq n-1 \\ {a_i} \sigma_k = \sigma_k {a_i}, \ \ k \geq 2, \ 1 \leq i \leq m \\ {a_i} \sigma_1 {a_i} \sigma_1 = \sigma_1 {a_i} \sigma_1 {a_i}, \ \ 1 \leq i \leq m \\ {a_i} (\sigma_1 {a_r} {\sigma^{-1}_1}) = (\sigma_1 {a_r} {\sigma^{-1}_1}) {a_i}, \ \ r < i \end{array} \right. \end{array} \right>, \end{equation} \bigbreak \noindent where the {\it loop generators} $a_i$ and the {\it braiding generators} $\sigma_j$ are as illustrated in Figure~\ref{Loops and Crossings}. \begin{figure} \begin{center} \includegraphics[width=5.5in]{loops} \end{center} \caption{ The loop generators $a_i, \ {a^{-1}_i}$ and the braiding generators $\sigma_j$ of $B_{m,n}$. } \label{Loops and Crossings} \end{figure} \begin{figure} \begin{center} \includegraphics[width=6.2in]{mixedbraids} \end{center} \caption{ Parting and combing a geometric mixed braid. } \label{partingandcombing} \end{figure} In order to give an algebraic statement for braid equivalence in $M$, we first part the mixed braids and we translate the geometric $L$-equivalence of Theorem~\ref{geommarkov} to an equivalence of parted mixed braids. {\it Parting} a geometric mixed braid $B \bigcup\beta$ on $m+n$ strands means to separate its endpoints into two different sets, the first $m$ belonging to the subbraid $B$ and the last $n$ to $\beta$, and so that the resulting braids have isotopic closures. This is realized by pulling each pair of corresponding moving strands to the right and {\it over\/} or {\it under\/} each strand of $B$ that lies on their right. We start from the rightmost pair respecting the position of the endpoints. This process is called {\it parting} of a geometric mixed braid and the result is a {\it parted mixed braid}. If the strands are pulled always over the strands of $B$, then this parting is called {\it standard parting}. See the middle illustration of Figure~\ref{partingandcombing} for the standard parting of an abstract mixed braid. For more details the reader is referred to \cite{LR2}. Then, in order to restrict Theorem~\ref{geommarkov} to the set of all parted mixed braids related to the manifold $M$, we need the following moves between parted mixed braids. {\it Loop conjugation} of a parted mixed braid $\beta$ is its concatenation by a loop $a_i$ (or by ${a^{-1}_i}$) from above and from ${a^{-1}_i}$ (corr. $a_i$) from below, that is $\beta \sim a^{\pm1} \beta a^{\mp1}$. As it turns out, two partings of a geometric mixed braid differ by loop conjugations (cf. Lemma~2 \cite{LR2}). A {\it parted $L$-move} is an $L$-move between parted mixed braids. Further, a mixed braid with an $L$-move performed can be parted to a parted mixed braid with a parted $L$-move performed. Namely we make the parting consistent with the label of the $L$-move: an $L_o$ move will be parted by pulling over all other strands, while an $L_u$ move will be parted by pulling under all other strands (cf. Lemma~3 \cite{LR2}). A {\it parted $\mathbb{Z}$-braid band move} is a geometric $\mathbb{Z}$-braid band move between parted mixed braids, such that it takes place at the top part of the braid and the little band {\it starts from the last strand} of the moving subbraid and it {\it moves over\/} each moving strand and each component of the surgery braid, until it reaches from the right the specific component, and then is followed by parting (see Figure~\ref{pbbm}(c)). Moreover, performing a $\mathbb{Z}$-braid band move on a mixed braid and then parting, the result is equivalent, up to $L$-moves and loop conjugation, to performing a parted $\mathbb{Z}$-braid band move (cf. Lemma~5 \cite{LR2}). \begin{thm}[Parted mixed braid equivalence for $M=\chi_{_{\mathbb{Z}}} (S^3, \widehat B)$, Theorem~3 \cite{LR2}]\label{partedcco} Two oriented links in $M=\chi_{_{\mathbb{Z}}} (S^3, \widehat B)$ are isotopic if and only if any two corresponding parted mixed braids differ by a finite sequence of parted mixed braid isotopies, parted $L$-moves, loop conjugations and parted $\mathbb{Z}$-braid band moves. \end{thm} We now comb the parted mixed braids in order to translate the parted mixed braid equivalence to an equivalence between algebraic mixed braids. {\it Combing} a parted mixed braid means to separate the knotting and linking of the moving part away from the fixed subbraid using mixed braid isotopy. More precisely, let $\Sigma_k$ denote the crossing between the $k^{th}$ and the $(k+1)^{st}$ strand of the fixed subbraid. Then, for all $j=1,\ldots,n-1$ and $k=1,\ldots,m-1$ we have: $\Sigma_k \sigma_j = \sigma_j \Sigma_k$. Thus, the only generating elements of the moving part that are affected by the combing are the loops $a_i$. This is illustrated in Figure~\ref{combing}. In Lemma~6 \cite{LR2} formuli are given for the effect of combing on the $a_i$'s (see Lemma~\ref{combinglem} below). \begin{figure} \begin{center} \includegraphics[width=5.2in]{combing} \end{center} \caption{ Combing a parted mixed braid. } \label{combing} \end{figure} The effect of combing a parted mixed braid is to separate it into two distinct parts: the {\it algebraic part} at the top, which has all fixed strands forming the identity braid, so is an element of some mixed braid group $B_{m,n}$, and which contains all the knotting and linking information of the link $L$ in $M$; the {\it coset part} at the bottom, which contains only the fixed subbraid $B$ and an identity braid for the moving part (see right hand most illustration in Figure~\ref{partingandcombing}). Let now $C_{m,n}$ denote the set of parted mixed braids on $n$ moving strands with fixed subbraid $B$. Concatenating two elements of $C_{m,n}$ is not a closed operation since it alters the braid description of the manifold. However, as a result of the combing, for the fixed subbraid $B$ the set $C_{m,n}$ is a coset of $B_{m,n}$ in $B_{m+n}$. Fore details on the above the reader is referred to \cite{La2}. Translating the parted braid equivalence into an equivalence between algebraic mixed braids, we will obtain an algebraic statement of Theorem~\ref{partedcco}. Since loop conjugation does not take into account the combing of the loop through the fixed subbraid, we need the notion of {\it combed loop conjugation}. A combed loop conjugation is a move between algebraic mixed braids and is the result of a loop conjugation on a combed mixed braid followed by combing, so it can be described algebraically as: $\beta \sim \alpha_i^{\mp1} \beta \rho_i^{\pm1}$ for $\beta, a_i, \rho_i \in B_{m,n}$, where $\rho_i$ is the combing of the loop $a_i$ through the fixed subbraid $B$. We also define {\it algebraic $M$-conjugation} of an algebraic mixed braid to be its conjugation by a crossing $\sigma_j$ (or by ${\sigma^{-1}_j}$). An {\it algebraic $M$-move} is defined to be the insertion of a crossing ${\sigma^{\pm 1}_n}$ on the right hand side of an algebraic mixed braid. Finally, an {\it algebraic $L$-move} is defined to be a $L$-move between algebraic mixed braids. An algebraic $L$-move has the following algebraic expression for an $L_o$-move and an $L_u$-move respectively: \begin{equation} \begin{array}{lll} \alpha &=& \alpha_1\alpha_2 \stackrel{L_o}{\sim} \sigma_i^{-1}\ldots \sigma_n^{-1} \alpha_1' \sigma_{i-1}^{-1}\ldots \sigma_{n-1}^{-1}\sigma_n^{\pm 1}\sigma_{n-1} \ldots \sigma_i \alpha_2' \sigma_n \ldots \sigma_i \\ \alpha&=&\alpha_1\alpha_2 \stackrel{L_u}{\sim} \sigma_i\ldots \sigma_n \alpha_1' \sigma_{i-1}\ldots \sigma_{n-1}\sigma_n^{\pm 1}\sigma_{n-1}^{-1}\ldots\sigma_i^{-1} \alpha_2' \sigma_n^{-1}\ldots\sigma_i^{-1} \end{array} \end{equation} \noindent where $\alpha_1$, $\alpha_2$ are elements of $B_{m,n}$ and $\alpha_1'$, $\alpha_2' \in B_{m,n+1}$ are obtained from $\alpha_1$, $\alpha_2$ by replacing each $\sigma_j$ by $\sigma_{j+1}$ for $j=i,\ldots,n-1$. Note that algebraic $M$-conjugation, the algebraic $M$-moves and the algebraic $L$-moves commute with combing. Note also that Remark~\ref{Lmove}(i) applies equally to the case of algebraic mixed braids (cf. Lemma~4 \cite{LR2}). We finally need to understand how a parted $\mathbb{Z}$-braid band move is combed through the surgery braid $B$. \begin{defn}[Definition~7 \cite{LR2}] \label{twistband} \rm An {\it algebraic $\mathbb{Z}$-braid band move\/} is defined to be a parted band move between algebraic mixed braids (see top part of Figure~\ref{pbbm}(d)). Setting: \[ \lambda_{n-1,1} := \sigma_{n-1} \ldots \sigma_1 \mbox{ \ \ and \ \ } t_{k,n} := \sigma_n \ldots \sigma_1 a_k {\sigma^{-1}_1} \ldots {\sigma^{-1}_n}, \] an algebraic band move has the following algebraic expression: \[ \beta_1 \beta_2 \ \sim \ \beta^{\prime}_1 \, {t^{p_k}_{k,n}} \, {\sigma^{\pm 1}_n} \, \beta^{\prime}_2, \] \noindent where $\beta_1, \beta_2 \in B_{m,n}$ and $\beta^{\prime}_1, \beta^{\prime}_2 \in B_{m,n+1}$ are the words $\beta_1, \beta_2$ respectively with the substitutions: \[ \begin{array}{lcl} {a^{\pm 1}_k} & \longleftrightarrow & {[({\lambda^{-1}_{n-1,1}} {\sigma^{2}_n} \lambda_{n-1,1}) \, a_k]}^{\pm 1} \\ {a^{\pm 1}_i} & \longleftrightarrow & ({\lambda^{-1}_{n-1,1}} {\sigma^{2}_n} \lambda_{n-1,1}) \, {a^{\pm 1}_i} \, ({\lambda^{-1}_{n-1,1}} {\sigma^{2}_n} \lambda^{-1}_{n-1,1}), \mbox{ \ \ if \ } i < k \\ {a^{\pm 1}_i} & \longleftrightarrow & {a^{\pm 1}_i}, \mbox{ \ \ if \ } i > k. \\ \end{array} \] Further, a {\it combed algebraic $\mathbb{Z}$-braid band move\/} is a move between algebraic mixed braids and is defined to be a parted $\mathbb{Z}$-braid band move that has been combed through $B$. So it is the composition of an algebraic $\mathbb{Z}$-braid band move with the combing of the parallel strand and it has the following algebraic expression: \[ \beta_1 \beta_2 \ \sim \ \beta^{\prime}_1 \, {t^{p_k}_{k,n}} \, {\sigma^{\pm 1}_n} \, \beta^{\prime}_2 \, r_k, \] \noindent where $r_k$ is the combing of the parted parallel strand to the $k^{th}$ surgery strand through the surgery braid. \end{defn} The group $B_{m,n}$ embeds naturally into the group $B_{m,n+1}$. We shall denote $B_{m,\infty}=\bigcup_{n=1}^{\infty }B_{m,n}$ and similarly $C_{m,\infty}=\bigcup_{n=1}^{\infty }C_{m,n}$. We are now in position to give the algebraic Markov theorem for $M=\chi_{_{\mathbb{Z}}} (S^3, \widehat B)$. \begin{thm}[Algebraic Markov Theorem for $M=\chi_{_{\mathbb{Z}}} (S^3, \widehat B)$, Theorem~5 \cite{LR2}] \label{algcco} \ Two oriented links in $M=\chi_{_{\mathbb{Z}}} (S^3, \widehat B)$ are isotopic if and only if any two corresponding algebraic mixed braid representatives in $B_{m,\infty}$ differ by a finite sequence of the following moves: \vspace{.03in} \noindent (1) \ Algebraic $M$-moves: \ ${\beta}_1 {\beta}_2 \sim {\beta}_1{\sigma^{\pm 1}_n}{\beta}_2, \ \ for \ {\beta}_1, {\beta}_2 \in B_{m,n}$, \noindent (2) \ Algebraic $M$-conjugation: \ $\beta \sim {\sigma^{\pm 1}_j} \beta {\sigma^{\mp 1}_j}$, \ \ for \ $\beta, \sigma_j \in B_{m,n}$, \vspace{.03in} \noindent (3) \ Combed loop conjugation: \ $\beta \sim {a^{\mp 1}_i} \beta {\rho^{\pm 1}_i}$, \ \ for \ $\beta, a_i, \rho_i \in B_{m,n}$, where $\rho_i$ is the combing of the loop $a_i$ through $B$, \vspace{.03in} \noindent (4) \ Combed algebraic braid band moves: \ For for every $k=1,\ldots,m$ we have: \[ \beta_1 \beta_2 \ \sim \ \beta^{\prime}_1 \, {t^{p_k}_{k,n}} \, {\sigma^{\pm 1}_n} \, \beta^{\prime}_2 \, r_k, \] \noindent where $\beta_1, \beta_2 \in B_{m,n}$ and $\beta^{\prime}_1, \beta^{\prime}_2 \in B_{m,n+1}$ are as in Definition~\ref{twistband} and where $r_k$ is the combing of the parted parallel strand to the $k$th surgery strand through $B$. \smallbreak Equivalently, by a finite sequence of algebraic mixed braid relations and the following moves: \vspace{.03in} \noindent (1$'$) \, algebraic $L$-moves, \noindent (2$'$) \, combed loop conjugations, \noindent (3$'$) \ combed algebraic $\mathbb{Z}$-braid band moves. \end{thm} \section{The Reidemeister Theorem for links in $3$-manifolds with rational surgery description} Let $M$ be a c.c.o. $3$-manifold obtained from $S^3$ by rational surgery, that is surgery along a framed link $\widehat{B}$ with rational coefficients. We shall denote $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$. Let also $L$ be an oriented link in $M$. By the discussion in Section~1.1, isotopy in $M$ is translated into isotopy in $S^3\backslash \widehat{B}$ together with the two types, $\alpha$ and $\beta$, of band moves for mixed links in $S^3$. The band moves in this case are described as follows. Let $c$ be a component of $\widehat{B}$ with framing $p/q$. The specified parallel curve $l$ of $c$ is a $(p,q)$-torus knot on the boundary of a tubular neighborhood of $c$ which, by construction, bounds a disc in $M$. Then, a {\it $\mathbb{Q}$-band move} along $c$ is the connected sum of a component of $L$ with the $(p,q)$-torus knot $l$ and there are two types, $\alpha$ and $\beta$, according to the orientations. The two types of band moves are illustrated in Figure~\ref{bmoves}, where $c$ is a trefoil knot with $2/3$ surgery coefficient and where ``band move" is shortened to ``b.m.". \begin{figure} \begin{center} \includegraphics[width=3.6in]{bandmovestrefoil} \end{center} \caption{ The two types of $\mathbb{Q}$-band moves. } \label{bmoves} \end{figure} Clearly, Theorem~\ref{reidemintegral} applies also to $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$. Namely we have: \begin{thm}[Reidemeister for $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ with two types of band moves] \label{reidemrationalboth} Two oriented links $L_1$, $L_2$ in $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding mixed link diagrams of theirs, $\widehat{B} \cup \widetilde{L_1}$ and $\widehat{B} \cup \widetilde{L_2}$, differ by isotopy in $S^3 \backslash \widehat{B}$ together with a finite sequence of the two types $\alpha$ and $\beta$ of band moves. \end{thm} In this section we sharpen Theorem~\ref{reidemrationalboth}. More precisely, we show that only one of the two types of band moves is necessary in order to describe isotopy for links in $M$. The proof is based on a known contrivance, which was used in the proof of Theorem~5.10 \cite{LR1} (recall Theorem~\ref{geommarkov}) for establishing the sufficiency of the geometric braid band moves in the mixed braid equivalence where the case of integral surgery only is considered (see Figure~\ref{Unknotted surgery}). Theorem~\ref{reidemrational} simplifies the proof of Theorem~\ref{geommarkovrat}. \begin{thm} [Reidemeister for $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ with one type of band moves] \label{reidemrational} Two oriented links $L_1$, $L_2$ in $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding mixed link diagrams of theirs, $\widehat{B}\bigcup L_1$ and $\widehat{B}\bigcup L_2$, differ by a finite sequence of the band moves of type $\alpha$ (or equivalently of type $\beta$) and isotopy in $S^3\backslash \widehat{B}$. \end{thm} \begin{proof} Let $L$ be an oriented link in $M$. By Theorem~\ref{reidemrationalboth}, it suffices to show that a band move of type $\beta$ can be obtained from a band move of type $\alpha$ and isotopy in the knot complement. We will first demonstrate the proof for an unknotted surgery component $c$ with integral coefficient $p$. (Note that integral surgery description can be considered as a special case of rational surgery description.) We shall follow the steps of the proof in Figure~\ref{Unknotted surgery} where $p=2$. We start with performing a band move of type $\beta$ using a component $s$ of the link $L$. In Figure~\ref{Unknotted surgery}(b) we see the two twists of the band move wrapping around the surgery curve $c$ in the righthand sense. Then, using an arc of the same link component $s$, we perform a second band move of type $\alpha$. This will take place within a thinner tubular neighborhood than the first band move. So, the two twists of the second band move, which also wrap around $c$ in the righthand sense, commute with the two twists of the first band move (see Figure~\ref{Unknotted surgery}(c)). We arrange all $2p$ twists in pairs as follows. We pass one twist from the second band move (the closest) through all twists of the first band move, see Figure~\ref{Unknotted surgery}(f). Since all twists follow the righthand sense, the two innermost twists coming from the second and the first band move, create a little band (Figure~\ref{Unknotted surgery}(f)) which can be eliminated using isotopy in the knot complement of $c$ (Figure~\ref{Unknotted surgery}(g) and Figure~\ref{Twist Cancelation}). This is the cancelation of the first pair of the $2p$ twists. Repeating the same procedure we cancel all $p$ pairs and we end up with the component $s$ of the link $L$ as it was in the initial position before the band moves (see Figures~\ref{Unknotted surgery}(i), (a)). \begin{figure} \begin{center} \includegraphics[width=5.9in]{unknotsurgery} \end{center} \caption{A type-$\beta$ band move follows from a type-$\alpha$ band move in the case of integral surgery coefficient. } \label{Unknotted surgery} \end{figure} \begin{figure} \begin{center} \includegraphics[width=3.8in]{reidcancel} \end{center} \caption{ Twist cancelation.} \label{Twist Cancelation} \end{figure} For the more general case of rational surgery along any knot $c$ we follow the same idea. More precisely, we perform a $\mathbb{Q}$-band move of type $\beta$ along $c$ and we obtain an outer $(p,q)$-torus knot. Then, we perform a $\mathbb{Q}$-band move of type $\alpha$ along $c$ and we obtain an inner $(p,q)$-torus knot. In Figure~\ref{Reidemeister Theorem1} we illustrate this for the case where $p=2$, $q=3$ and $c$ a trefoil knot. \begin{figure} \begin{center} \includegraphics[width=6in]{ReidTref1} \end{center} \caption{ A band move of type $\beta$ followed by a band move of type $\alpha$. } \label{Reidemeister Theorem1} \end{figure} Without loss of generality (by isotopy in the complement of $c$), the second band move is performed on the innermost arc of the $q$ arcs parallel to $c$, creating $q$ new parallel arcs even closer to $c$. After the second $\mathbb{Q}$-band move is performed, the outer arc of the $q$ new arcs and the inner arc of the $q$ arcs coming from the first band move of type $\alpha$ form a band (see shaded area in Figure~\ref{Reidemeister Theorem2}). Then, using isotopy in the complement of $c$, we eliminate this band by pulling it along $c$. This will result in the elimination of $p-q$ pairs of parallel arcs to $c$. In our example, this is done in two steps (see Figures~\ref{Reidemeister Theorem2} and \ref{Reidemeister Theorem3}). \begin{figure} \begin{center} \includegraphics[width=5.1in]{ReidTref2} \end{center} \caption{ Band with boundary two parallel arcs of opposite orientations. } \label{Reidemeister Theorem2} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.1in]{ReidTref3} \end{center} \caption{ Retracting the band along the surgery component. } \label{Reidemeister Theorem3} \end{figure} As in the case of integral surgery the twists coming from the two band moves commute (see Figure~\ref{Reidemeister Theorem4}). Arranging these $2p$ twists pairwise (Figure~\ref{Reidemeister Theorem5}), they cancel out by the fact that all twists have the same handiness, but opposite orientation. In the end, $s$ is left as in its initial position. \begin{figure} \begin{center} \includegraphics[width=5.1in]{ReidTref4} \end{center} \caption{ Arranging twists in pairs to form bands. } \label{Reidemeister Theorem4} \end{figure} \begin{figure} \begin{center} \includegraphics[width=4.7in]{ReidTref5} \end{center} \caption{ Twist cancelation. } \label{Reidemeister Theorem5} \end{figure} So, a $\mathbb{Q}$-band move of type $\beta$ can be performed using a $\mathbb{Q}$-band move of type $\alpha$ and isotopy in the complement of the surgery component $c$. The proof of Theorem~\ref{reidemrational} is now concluded. \end{proof} \section{Geometric braid equivalence for links in a $3$-manifold with rational surgery description} In this section we extend Theorem~\ref{geommarkov} to manifolds with rational surgery description, that is $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$, taking into account the sharpened Reidemeister theorem for $M$ (Theorem~\ref{reidemrational}). We first need the following. \begin{defn} \rm A {\it geometric $\mathbb{Q}$-braid band move\/} is a move between geometric mixed braids which is a $\mathbb{Q}$-band move of type $\alpha$ between their closures. It starts with a little band (an arc of the moving subbraid) close to a surgery strand with surgery coefficient $p/q$. The little band gets first one twist {\it positive\/} or {\it negative\/}, which shall be denoted as $c^{\prime}_{\pm}$ and then is replaced by $q$ strands that run in parallel to all strands of the same surgery component and link only with that surgery strand, wrapping around it $p$ times and, thus, forming a $(p,q)$-torus knot, Figure~\ref{lemma}. This braided $(p,q)$-torus knot is denoted as $d^{\prime}$. A geometric $\mathbb{Q}$-braid band move with a positive (resp. negative) twist shall be called a positive geometric $\mathbb{Q}$-braid band move (resp. negative geometric $\mathbb{Q}$-braid band move). \end{defn} \begin{figure} \begin{center} \includegraphics[width=3in]{bbm} \end{center} \caption{ A $\mathbb{Q}$-braid band move.} \label{lemma} \end{figure} By Remark~1(ii) a $\mathbb{Q}$-braid band move may be assumed to take place at the top part of a mixed braid and all strands from a $\mathbb{Q}$-braid band move may be assumed to lie on the righthand side of the surgery strands. We shall now prove the following. \begin{thm} [Geometric braid equivalence for $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$] \label{geommarkovrat} Two oriented links in $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding geometric mixed braids in $S^3$ differ by mixed braid isotopy, by $L$-moves that do not touch the fixed subbraid $B$ and by the geometric $\mathbb{Q}$-braid band moves. \end{thm} \begin{proof} The proof is completely analogous to and is based on the proof of Theorem 5.10 \cite{LR1} (Theorem~\ref{geommarkov}). Let $K_1$ and $K_2$ be two isotopic oriented links in $M$. By Theorem~\ref{reidemrational}, the corresponding mixed links $\widehat{B} \bigcup K_1$ and $\widehat{B} \bigcup K_2$ differ by isotopy in the complement of $\widehat{B}$ and $\mathbb{Q}$-band moves of type $\alpha$. Note that, by Theorem~\ref{reidemrational} we do not need to consider band moves of type $\beta$. By Theorem 5.10 \cite{LR1}, isotopy in the complement of $\widehat{B}$ translates into geometric braid isotopy and the $L$-moves. Let now $\widehat{B} \bigcup K_1$ and $\widehat{B} \bigcup K_2$ differ by a $\mathbb{Q}$-band move of type $\alpha$ (recall Figure~\ref{bmoves}). Let $\widehat{B} \bigcup \widetilde{K_1}$ and $\widehat{B} \bigcup \widetilde{K_2}$ be two mixed link diagrams of the mixed links $\widehat{B} \bigcup K_1$ and $\widehat{B} \bigcup K_2$ which differ only by the places illustrated in Figure~\ref{markovtheorem}. As in \cite{LR1}, by the braiding algorithm given therein, the diagrams $\widehat{B} \bigcup \widetilde{K_1}$ and $\widehat{B} \bigcup \widetilde{K_2}$ may be assumed braided everywhere except for the places where the $\mathbb{Q}$-band move is performed. We now braid the up-arc in Figure~\ref{markovtheorem}(b) and obtain a geometric mixed braid $\widehat{B} \bigcup b_1$ corresponding to the diagram $\widehat{B} \bigcup \widetilde{K_1}$ (see Figure~\ref{markovtheorem}(a)). Note that Figure~\ref{markovtheorem}(c) is already in braided form and let $B \bigcup b_2$ denote the geometric mixed braid corresponding to the diagram $\widehat{B} \bigcup \widetilde{K_2}$. \begin{figure} \begin{center} \includegraphics[width=4.8in]{markovtheorem} \end{center} \caption{ A type $\alpha$ band move (locally) and its braiding. } \label{markovtheorem} \end{figure} We would like to show that the two mixed braids $B \bigcup b_1$ and $B \bigcup b_2$ differ by the moves given in the statement of the Theorem. We perform a Reidemeister I move on $\widehat{B} \bigcup \widetilde{K_1}$ with a {\it negative} crossing and obtain the diagram $\widehat{B} \bigcup \widetilde{K'_1}$. Then, the corresponding mixed braids, $B \bigcup b_1$ and $B \bigcup b'_1$, differ by mixed braid isotopy and $L$-moves (see Figure~\ref{markovproof}(a) and (b)). We then perform a {\it positive} $\mathbb{Q}$-braid band move on $B \bigcup b'_1$ and obtain the mixed braid $B \bigcup b'_2$. In the closure of $B \bigcup b'_2$ we unbraid and re-introduce the two up-arcs illustrated in Figure~\ref{markovproof}(b), obtaining a diagram $\widehat{B} \bigcup \widetilde{K'_2}$ with the formation of a Reidemeister II move. Performing this move on $\widehat{B} \bigcup \widetilde{K'_2}$ we obtain the diagram $\widehat{B} \bigcup \widetilde{K_2}$, which is already in braided form and its corresponding mixed braid is $B \bigcup b_2$ (see Figure~\ref{markovproof}(c) and (d)). So, the mixed braids $B \bigcup b'_2$ and $B \bigcup b_2$ differ by mixed braid isotopy and $L$-moves. Therefore, we showed that the braids $B \bigcup b_1$ and $B \bigcup b_2$ in Figure~\ref{markovtheorem}(a) and (c) differ by mixed braid isotopy, $L$-moves and a braid band move. This concludes the proof. \end{proof} \begin{figure} \begin{center} \includegraphics[width=6.3in]{markovproof} \end{center} \caption{ The steps of the proof of Theorem~\ref{geommarkovrat}. } \label{markovproof} \end{figure} \section{Algebraic braid equivalence for links in a $3$-manifold with rational surgery description} In order to translate the geometric mixed braid equivalence to an equivalence of algebraic mixed braids we follow the strategy in \cite{LR2} for integral surgery description. Namely, we apply first parting and then combing to the geometric mixed braids. What makes things more complicated in the case of rational surgery description is that the surgery braid $B$ is in general not a pure braid and when we apply a $\mathbb{Q}$-braid band move on a mixed braid, the little band that approaches the surgery strand is replaced by $q$ strands that run in parallel to all strands of the same surgery component. In order to adopt and apply results from \cite{LR2} we need the notion of a {\it $\mathbb{Q}$-strand cable}. \begin{defn} \label{cable} \rm We define a {\it $q$-strand cable} to be a set of $q$ parallel strands coming from a $\mathbb{Q}$-braid band move and following one strand of the specified surgery component. \end{defn} Let now $B\bigcup \beta$ be a geometric mixed braid and suppose that a $\mathbb{Q}$-braid band move is performed on $B\bigcup \beta$. Treating the new strands coming from the braid band move as cables running in parallel to the strands of a surgery component, that is, treating each cable as one thickened strand, we may part the geometric mixed braid following the exact procedure as in \cite{LR2}. More precisely, we have the following. \begin{lemma} \label{partcable} Cabling and standard parting commute. That is, standard parting of a mixed braid with a $\mathbb{Q}$-braid band move performed and then cabling, is the same as cabling first the set of new strands and then standard parting. \end{lemma} \begin{proof} Let $B\bigcup \beta$ be a geometric mixed braid on $m+n$ strands and let a $\mathbb{Q}$-braid band move be performed on a surgery component $s$ of $B$. Let also $s_1, \ldots , s_k \in \{ 1, \ldots, m \}$ be the numbers of the strands of the surgery component $s$ and let $c_1, \ldots , c_k$ denote the $q$-strand cables corresponding to $s_1, \ldots , s_k$. On the one hand, after the $\mathbb{Q}$-braid band move is performed and before any cablings occur, we part the geometric mixed braid following the procedure of the standard parting as described in Section~1.3 (see Figure~\ref{stndpartcable}(a)). On the other hand we cable first each set of $q$-strands resulting from the $\mathbb{Q}$-braid band move and then we part the geometric mixed braid with the standard parting, treating each cable as one (thickened) strand, see Figure~\ref{stndpartcable}(b). Since both cabling and parting a geometric mixed braid respect the position of the endpoints of each pair of corresponding moving strands, it follows that cabling and parting commutes. \end{proof} \begin{figure} \begin{center} \includegraphics[width=6.2in]{partcable} \end{center} \caption{ Standard parting and cabling.} \label{stndpartcable} \end{figure} Recall from Section~1.3 that a geometric $L$-move can be turned to a parted $L$-move. In order to give the analogue of Theorem~\ref{partedcco} in the case of rational surgery we also need to introduce the following adaptation of a parted $\mathbb{Z}$-braid band move. \begin{defn} \label{partqbbm} \rm A {\it parted $\mathbb{Q}$-braid band move} is defined to be a geometric $\mathbb{Q}$-braid band move between parted mixed braids, such that it takes place at the top part of the braid and on the right of the rightmost strand, $s_k$, of the specific surgery component, $s$, consisting of the strands $s_1, \ldots , s_k$. Moreover, the little band starts from the last strand of the moving subbraid and it moves over each moving strand and each component of the surgery braid, until it reaches the last strand of $s$, and then is followed by parting of the resulting mixed braid. See Figures~\ref{partedqbbm} and \ref{partecables}. \end{defn} \begin{figure} \begin{center} \includegraphics[width=6in]{partqbbm} \end{center} \caption{ A parted $\mathbb{Q}$-braid band move.} \label{partedqbbm} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.9in]{bbmcabled} \end{center} \caption{ A parted $\mathbb{Q}$-braid band move using cables.} \label{partecables} \end{figure} Then Theorem~\ref{geommarkovrat} restricts to the following. \begin{thm}[Parted version of braid equivalence for $M=\chi_{_{\mathbb{Q}}} (S^3, \widehat B)$] \label{qparted} \ Two oriented links in $M=\chi_{_{\mathbb{Q}}} (S^3, \widehat B)$ are isotopic if and only if any two corresponding parted mixed braids in $C_{m,\infty}$ differ by a finite sequence of parted $L$-moves, loop conjugations and parted $\mathbb{Q}$-braid band moves. \end{thm} \begin{proof} By Lemma~\ref{partcable} the cables resulting from a geometric $\mathbb{Q}$-braid band move are treated as one strand, so we can apply Theorem~\ref{partedcco}. Moreover, by Lemma~9 in \cite{LR2} a geometric $\mathbb{Q}$-braid band move may be always assumed, up to $L$-equivalence, to take place on the right of the rightmost strand of the specific surgery component. \end{proof} In order to translate Theorem~\ref{qparted} into an algebraic equivalence between elements of $B_{m,\infty}$ we need the following lemmas. \begin{lemma}[Combing Lemma, Lemma 6 \cite{LR2}] \label{combinglem} { \ The crossings $\Sigma_k$, $k=1,\ldots,m-1$ of the fixed subbraid $B$, and the loops $a_i$, for $i=1,\ldots,m$, satisfy the following {\rm `combing' relations}: \[ \begin{array}{llcll} \ \ & \Sigma_k {a^{\pm 1}_k} & = & {a^{\pm 1}_{k+1}} \Sigma_k & \\ \ \ & \Sigma_k {a^{\pm 1}_{k+1}} & = & {a^{-1}_{k+1}} {a^{\pm 1}_k} a_{k+1} \Sigma_k & \\ \ \ & \Sigma_k {a^{\pm 1}_i} & = & {a^{\pm 1}_i} \Sigma_k & \mbox{if \ } i \neq k, k+1 \\ \ \ & {\Sigma^{-1}_k} {a^{\pm 1}_k} & = & a_k {a^{\pm 1}_{k+1}} {a^{-1}_k} {\Sigma^{-1}_k} & \\ \ \ & {\Sigma^{-1}_k} {a^{\pm 1}_{k+1}} & = & {a^{\pm 1}_k} {\Sigma^{-1}_k} & \\ & {\Sigma^{-1}_k} {a^{\pm 1}_i} & = & {a^{\pm 1}_i} {\Sigma^{-1}_k} & \mbox{if \ } i \neq k, k+1. \\ \end{array} \] } \end{lemma} \noindent \textbf{Notation:} We set $\lambda_{k,r}:=\sigma_k \sigma_{k+1} \ldots \sigma_{r-1} \sigma_r$, for $k<r$ and $\lambda_{k,r}:=\sigma_k \sigma_{k-1} \ldots \sigma_{r+1} \sigma_r$, for $r<k$. We note that $\lambda_{i,i}:=\sigma_i$. Also, by convention we set $\lambda_{0,i}=\lambda_{i,0}:=1$. Then we have the following: \begin{lemma} \label{cableloop} A positive looping between a $q$-strand cable and the $j^{th}$ fixed strand of the fixed subbraid $B$ has the algebraic expression: $$\prod_{i=0}^{q-1} \lambda_{i,1} a_j\lambda_{i,1}^{-1}=\prod_{i=0}^{q-1} \lambda_{1,(q-1)-i}^{-1} a_j\lambda_{1,(q-1)-i} \ ,$$ \noindent while a negative looping has the algebraic expression: $$\prod_{i=0}^{q-1} \lambda_{1,i}^{-1} a_j^{-1} \lambda_{1,i}=\prod_{i=0}^{q-1} \lambda_{(q-1)-i,1} a_j^{-1} \lambda_{(q-1)-i,1}^{-1} \ .$$ \end{lemma} \begin{proof} We start with Figure~\ref{lemma2}(a) where a positive looping between a $q$-strand cable and a fixed stand of the mixed braid is shown. In Figure~\ref{lemma2}(b) the cable is replaced by the $q$ strands according to Definition~\ref{cable}. Then, using mixed braid isotopy, we end up with Figure~\ref{lemma2}(c), top, whereby we can read directly the algebraic expression $\prod_{i=0}^{q-1} \lambda_{i,1} a_j \lambda_{i,1}^{-1}$. The second algebraic expression comes from the bottom illustration of Figure~\ref{lemma2}. Similarly, in Figure~\ref{lemma2b} we illustrate the case where a negative looping between a $q$-strand cable and a fixed strand of the mixed braid occurs. \end{proof} \begin{figure} \begin{center} \includegraphics[width=5.8in]{lemma1} \end{center} \caption{ A positive looping between a cable and a fixed strand.} \label{lemma2} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.8in]{lemma1b} \end{center} \caption{ A negative looping between a cable and a fixed strand.} \label{lemma2b} \end{figure} \begin{lemma} \label{cablecomb} Cabling and combing commute. That is, treating a $q$-strand cable as a thickened moving strand and combing it through the fixed subbraid $B$, the result is equivalent to combing one by one each strand of the cable. \end{lemma} \begin{proof} According to the Combing Lemma we have to consider all cases between looping and crossings of the subbraid $B$. We will only examine the four cases illustrated in Figure~\ref{combing} as representative cases. All others are completely analogous. The first case is illustrated in Figure~\ref{combcable1}, where a positive looping between the cable and the $k^{th}$ fixed strand of $B$ is being considered and the crossing of the fixed strands is positive. For a negative looping the proof is similar. \begin{figure} \begin{center} \includegraphics[width=5.8in]{combcable1} \end{center} \caption{ Combing and cabling commute: Proof of Case 1.} \label{combcable1} \end{figure} We now consider the case illustrated in Figure~\ref{combcable2}, where a positive looping between the cable and the $(k+1)^{th}$ fixed strand of $B$ is being considered, and the crossing in $B$ is positive. We shall prove this case by induction on the number of strands that belong to the cable, since, as we can see from Figure~\ref{combcable2}, the resulting algebraic expressions are not directly comparable. \begin{figure} \begin{center} \includegraphics[width=5.8in]{combcable2} \end{center} \caption{ Combing and cabling commute: Case 2. } \label{combcable2} \end{figure} The case where the cable consists of one strand is trivial. For a two-strand cable, combing the cable first and then uncabling (see top part part of Figure~\ref{2sccomb}) results in the algebraic expression: $$ \alpha_2^{-1}\ (\sigma_1^{-1} \alpha_2^{-1} \sigma_1)\ \alpha_1\ (\sigma_1^{-1} \alpha_1 \sigma_1)\ \alpha_2\ (\sigma_1^{-1} \alpha_2 \sigma_1), $$ while uncabling first and then combing (bottom part of Figure~\ref{2sccomb}) results in the algebraic expression: $$(\alpha_2^{-1}\alpha_1\alpha_2)\ (\sigma_1^{-1} \alpha_2^{-1}\alpha_1\alpha_2 \sigma_1).$$ \begin{figure} \begin{center} \includegraphics[width=5.8in]{2sccomb} \end{center} \caption{ Combing a $2$-strand cable: Case 2. } \label{2sccomb} \end{figure} We show below that these algebraic expressions are equal, whereby we have underlined expressions which are crucial for the next step. Indeed: $$ \begin{array}{rclr} \underline{\alpha_2^{-1}} (\sigma_1^{-1} \alpha_2^{-1} \sigma_1) \alpha_1 (\sigma_1 \alpha_1 \sigma_1^{-1}) \alpha_2 (\sigma_1 \underline{\alpha_2 \sigma_1^{-1}}) & = & (\underline{\alpha_2^{-1}}\alpha_1\alpha_2)(\sigma_1 \alpha_2^{-1}\alpha_1\underline{\alpha_2 \sigma_1^{-1}}) & \Leftrightarrow \\ (\sigma_1^{-1} \alpha_2^{-1} \sigma_1) \alpha_1 (\underline{\sigma_1 \alpha_1 \sigma_1^{-1}) \alpha_2} (\sigma_1)& = & (\alpha_1\alpha_2)(\sigma_1 \alpha_2^{-1}\alpha_1) & \Leftrightarrow \\ \sigma_1^{-1} \alpha_2^{-1} \sigma_1 \alpha_1 \alpha_2 \sigma_1 \underline{\alpha_1\sigma_1^{-1}\sigma_1}& = & \alpha_1\alpha_2\sigma_1 \alpha_2^{-1}\underline{\alpha_1}&\Leftrightarrow \\ \sigma_1^{-1} \underline{\alpha_2^{-1} \sigma_1 \alpha_1 (\sigma_1^{-1}} \sigma_1) \alpha_2 \sigma_1 & = & \alpha_1\alpha_2\sigma_1 \alpha_2^{-1}&\Leftrightarrow \\ \underline{\sigma_1^{-1} \sigma_1 \alpha_1} \sigma_1^{-1} \alpha_2^{-1} \sigma_1 \alpha_2 \sigma_1 & = & \underline{\alpha_1} \alpha_2\sigma_1 \alpha_2^{-1}&\Leftrightarrow \\ \underline{\sigma_1^{-1} \alpha_2^{-1}} \sigma_1 \alpha_2 \sigma_1 & = & \alpha_2\sigma_1 \underline{\alpha_2^{-1}}&\Leftrightarrow \\ \sigma_1 \alpha_2 \sigma_1 \alpha_2 & = & \alpha_2\sigma_1 \alpha_2 \sigma_1 \\ \end{array} $$ \noindent We ended up with one of the defining relations of the mixed braid group $B_{m,n}$, recall (\ref{B}). We now consider a $(q+1)$-strand cable and we let the first $q$ strands form a $q$-strand subcable. We first comb the $q$-strand cable and then the $(q+1)^{st}$ strand and the result follows by applying the case of a 2-strand cable and the induction hypothesis for the $q$-strand cable (see Figure~\ref{combcableproof2}). \end{proof} \begin{figure} \begin{center} \includegraphics[width=6in]{combcableproof2} \end{center} \caption{ Combing and cabling commute: Proof of Case 2.} \label{combcableproof2} \end{figure} Let now $B\bigcup \beta$ be a parted mixed braid and let a parted $\mathbb{Q}$-braid band move be performed on the last strand, $s_k$, of a surgery component consisting of the strands $s_1, \ldots , s_k$. Recall Figures~\ref{partedqbbm} and \ref{partecables}. In order to give an algebraic expression for the parted $\mathbb{Q}$-braid band move, we part locally the subbraids $d^{\prime}$ and $c_{\pm}^{\prime}$ and the loop generators $a _i$, $i=1, \ldots, m$, and we use mixed braid isotopy in order to transform $d^{\prime}$ into $d$ and $c_{\pm}^{\prime}$ into $c_{\pm}$. See Figures~\ref{cables2}, \ref{cablesc}, \ref{cablesa} and \ref{cablesa2}. Then, $d$ has the algebraic expression: \begin{equation} d\ =\ [\ \lambda_{n+kq-1,n+(k-1)q+1}\ \lambda_{n+1,n+(k-1)q}^{-1}\ \lambda_{n,1}\ a_{s_k}\ \lambda_{n,1}^{-1}\ \lambda_{n+1,n+(k-1)q}^{-1}\ ]^p \end{equation} and $c_{\pm}$ has the algebraic expression: \begin{equation} c_{\pm}\ =\ \lambda_{n,n+kq-2}\ \sigma_{n+kq-1}^{\pm1}\ \lambda_{n,n+kq-2}^{-1}. \end{equation} \begin{figure} \begin{center} \includegraphics[width=4.7in]{dd} \end{center} \caption{The parting of a $\mathbb{Q}$-braid band move is an algebraic $\mathbb{Q}$-braid band move. } \label{cables2} \end{figure} \begin{figure} \begin{center} \includegraphics[width=3.7in]{partingc} \end{center} \caption{Algebraization of the crossing part $c^{\prime}_{\pm}$ of the mixed braid to $c_{\pm}$ after a $\mathbb{Q}$-braid band move is performed. } \label{cablesc} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.8in]{localparttwist} \end{center} \caption{Algebraization of the loop generators $a_j$ for $j \in \{s_1, \ldots, s_k \}$ after a $\mathbb{Q}$-braid band move is performed. } \label{cablesa} \end{figure} \begin{figure} \begin{center} \includegraphics[width=6in]{localparttwist2} \end{center} \caption{Algebraization of the loop generators $a_j$ for $j \notin \{s_1, \ldots, s_k \}$ after a $\mathbb{Q}$-braid band move is performed. } \label{cablesa2} \end{figure} We are now in the position to give the definition of an algebraic $\mathbb{Q}$-braid band move. \begin{defn} \label{algbm} \rm (i) An {\it algebraic $\mathbb{Q}$-braid band move} is defined to be a parted $\mathbb{Q}$-braid band move between elements of $B_{n,\infty}$ and it has the following algebraic expression: $$ \beta \sim d \ c_{\pm} \ \beta^{\prime}, $$ \noindent where $\beta^{\prime}$ is the algebraic mixed braid $\beta$ with the substitutions: $$ \begin{array}{ccl} {a_i}^{\pm 1} & \longleftrightarrow & {a_i}^{\pm 1}, \mbox{ \ for} \ i > s_k, \\ && \\ {a_i}^{\pm 1} & \longleftrightarrow & \lambda^{-1}_{n-1,1} \lambda_{n,n+kq-1} \lambda_{n+kq-1,1} \ {a_i}^{\pm 1} \\ & & \lambda^{-1}_{n-1,1} \lambda^{-1}_{n+kq-1,n} \lambda^{-1}_{n,n+kq-1} \lambda_{n-1,1}, \mbox{\ for} \ i < s_1,\\ \end{array} $$ $$ \begin{array}{ccl} a_{s_{j}} & \longleftrightarrow & \lambda^{-1}_{n-1,1} \lambda_{n,n+kq-1} \lambda_{n+kq-1,n+(j-1)q} \lambda^{-1}_{n,n+(j-1)q-1} \lambda_{n-1,1} \ a_{s_{j}}\\ & & \lambda^{-1}_{n-1,1} \lambda_{n,n+jq-1} \lambda^{-1}_{n+kq-1,n+jq} \lambda^{-1}_{n,n+kq-1} \lambda_{n-1,1} \\ && \\ & & and \\ && \\ a^{-1}_{s_{j}} & \longleftrightarrow & \lambda^{-1}_{n-1,1} \lambda_{n,n+kq-1} \lambda_{n+kq-1,n+jq} \lambda^{-1}_{n+(j-1)q,n+jq-1} \lambda^{-1}_{n,n+(j-1)q-1} \lambda_{n-1,1} \ a^{-1}_{s_{j}}\\ & & \lambda^{-1}_{n-1,1} \lambda_{n,n+(j-1)q-1} \lambda^{-1}_{n+jq-1,n+(j-1)q} \lambda^{-1}_{n+kq-1,n+jq} \lambda^{-1}_{n+kq-1,n} \lambda_{n-1,1}, \\ && \mbox{ \ for} \ s_j \in \{s_1, \ldots, s_k \},\\ && \\ a^{\pm 1}_{j} & \longleftrightarrow & \lambda^{-1}_{n-1,1} \lambda_{n,n+kq-1} \lambda_{n+kq-1,n+(r-1)q} \lambda^{-1}_{n,n+(r-1)q-1} \lambda_{n-1,1} a^{\pm 1}_{j} \\ && \lambda^{-1}_{n-1,1} \lambda_{n,n+(r-1)q-1} \lambda^{-1}_{n+kq-1,n+(r-1)q} \lambda^{-1}_{n,n+kq-1} \lambda_{n-1,1}, \mbox{ \ for} \ s_{r-1}<j<s_r.\\ \end{array} $$ (ii) A {\it combed algebraic $\mathbb{Q}$-braid band move\/} is a move between algebraic mixed braids and is defined to be a parted $\mathbb{Q}$-braid band move that has been combed through $B$. Moreover, it has the following algebraic expression: $$ \beta \sim d \ c_{\pm} \ \beta^{\prime} \ comb_B(c_1, \ldots , c_k), $$ \noindent where $comb_B(c_1, \ldots , c_k)$ is the combing of the parted $q$-strand cables $c_1, \ldots , c_k$ through the surgery braid $B$ (see Figure~\ref{cables}). \end{defn} \begin{figure} \begin{center} \includegraphics[width=6.1in]{bbmcabledd} \end{center} \caption{ A combed algebraic $\mathbb{Q}$-braid band move. } \label{cables} \end{figure} We are, finally, in the position to state the following main result of the paper. \begin{thm}[Algebraic Markov Theorem for $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$] \label{algmarkov} Let $s_1, \ldots , s_k$ be the numbers of the strands of a surgery component $s$ and let $c_1, \ldots , c_k$ be the corresponding $q$-strand cables arising from a $\mathbb{Q}$-braid band move performed on $s$. Then, two oriented links in $M=\chi_{_{\mathbb{Q}}}(S^3, \widehat{B})$ are isotopic if and only if any two corresponding algebraic mixed braid representatives in $B_{m,\infty}$ differ by a finite sequence of the following moves: \\ (i) \ Algebraic $M$-moves: \ $\beta_1 \beta_2\sim \beta_1 \sigma_n^{\pm1} \beta_2$, for $\beta_1, \beta_2 \in B_{m,n}$, \\ (ii) \ Algebraic $M$-conjugation: \ $\beta \sim \sigma_j^{\mp1} \beta \sigma_j^{\pm1}$, for $\beta, \sigma_j \in B_{m,n}$, \\ (iii) \ Combed loop conjugation: $\beta \sim \alpha_i^{\mp1} \beta \ \rho_i^{\pm1}$, for $\beta \in B_{m,n}$, where $\rho_i$ is the combing of the loop $\alpha_i$ through $B$, \\ (iv) \ Combed algebraic braid band moves: $\beta \ \sim \ d \ c_{\pm} \ \beta^{\prime} \ comb_{B}(c_1, \ldots, c_k)$, where the algebraic expressions of $d$ and $c_{\pm}$ are as in Eqs.~(3) and (4) respectively, $\beta^{\prime}$ is $\beta$ with the substitutions of the loop generators as in Definition~\ref{algbm} and $comb_{B}(c_1, \ldots, c_k)$ is the combing of the resulting $q$-strand cables $c_1, \ldots, c_k$ through the fixed subbraid $B$, \smallbreak \noindent or equivalently, by a finite sequence of the following moves: \\ $\bullet$ algebraic $L$-moves (see algebraic expressions in Eqs~2), \\ $\bullet$ combed loop conjugation, \\ $\bullet$ combed algebraic braid band moves (Definition~\ref{algbm}). \end{thm} \begin{proof} The arguments for passing from parted braid equivalence (Theorem~\ref{qparted}) to algebraic braid equivalence are the same as in those in the proof of the transition from Theorem~\ref{partedcco} to Theorem~\ref{algcco} in the case of integral surgery. The only part we need to analyze in detail is the algebraization of a parted $\mathbb{Q}$-braid band move. Namely, we will show that the following diagram commutes. \begin{center} $$ \begin{CD} C_{m,n} \ni B \bigcup \beta @>{\text{Parted Q-b.b.m.}}>> B\bigcup \beta^{\prime} \in C_{m,n+kq} \\ @| @| \\ comb_B{\beta} @. comb_B{\beta^{\prime}} \\ @VVV @VVV \\ B_{m,n} \ni alg_B(\beta) @>{\text{\text{Algebraic Q-b.b.m.}}}>> alg_B(\beta^{\prime}) \in B_{m+n+kq} \end{CD} $$ \end{center} In words, we start with a parted mixed braid $B\bigcup \beta \in C_{m,n}$ and we perform on it a parted $\mathbb{Q}$-braid band move (Definition~\ref{partqbbm}) obtaining a parted mixed braid $B\bigcup \beta^{\prime} \in C_{m,n+kq}$, where $k$ is the number of strands forming the surgery component. We then comb both parted mixed braids obtaining $comb_B(\beta)$ and $comb_B(\beta^{\prime})$ respectively. We will show that the corresponding algebraic parts, $alg_B(\beta) \in B_{m,n}$ and $alg_B(\beta^{\prime}) \in B_{m,n+kq}$ differ by the algebraic braid equivalence given in the statement of the theorem. We apply Lemma~8 in \cite{LR2}, where the $q$ strands of a braid band move are placed in the cable and the cable is treated as one strand. More precisely, we note that the parted $\mathbb{Q}$-braid band move takes place at the top of the braid, so it forms an algebraic $\mathbb{Q}$-braid band move. We now comb away $\beta$ to the top of $B$ and on the other side we comb away $\beta^{\prime}$. Since the $q$-strands cable of the parted $\mathbb{Q}$-braid band move lie very close to the surgery strands, this ensures that the loops $\alpha_j^{\pm1}$ around any strand of the $k$ strands of the specific surgery components get combed in the same way before and after the $\mathbb{Q}$-braid band move. So, having combed away $\beta$ we are left at the bottom with the identity moving braid on the one hand, and with the combing of all cables of the braid band move on the other hand, which is precisely what we denote $comb_B()$. Finally, by Lemma~\ref{cablecomb}, combing and cable commute. Thus, the Theorem is proved. \end{proof} \section{Examples} In this section we give the braid equivalences for knots in specific $3$-manifolds that play a very important role to $3$-dimensional topology such as the lens spaces $L(p,q)$, homology spheres and Seifert manifolds. \subsection{Lens spaces $L(p,q)$} It is known that the lens spaces $L(p,q)$ can be obtained by surgery on the unknot with surgery coefficient $p/q$. So, the fixed braid $\widehat{B}$ that represents $L(p,q)$ is the identity braid of one single strand and thus, no combing is needed. We have the following: \smallbreak \textit{Two oriented links in $L(p,q)$ are isotopic if and only if any two corresponding algebraic mixed braids in $B_{1, \infty}$ differ by a finite sequence of the following moves (compare with \cite{LR2}, Section~4) :} \smallbreak {\it (1)} \ {\it Algebraic $M$-moves: \ $\beta \sim \beta {\sigma^{\pm 1}_n}, \ \ \ \ \beta \in B_{1,n}$} \vspace{.03in} {\it (2)} \ {\it Algebraic $M$-conjugation: \ $\beta \sim {\sigma^{\mp 1}_i} \beta {\sigma^{\pm 1}_i}, \ \ \ \ \beta, \sigma_i \in B_{1,n}$} \vspace{.03in} {\it (3)} \ {\it Loop conjugation: \ $\beta \sim {t}^{\mp 1}\ \beta \ {t}^{\pm 1}, \ \ \ \ \beta \in B_{1,n}$ } \vspace{.03in} {\it (4)} \ {\it Algebraic braid band moves: } \ For $\beta \in B_{1,n}$ we have: \[ \beta \ \sim \ {d} \, {c_{\pm}} \, \beta^{\prime}, \] \noindent where: $$ \begin{array}{ccl} d &= & [\lambda_{n+q-1,1}\ t\ \lambda^{-1}_{1,n+q-1}]^p,\\ &&\\ c_{\pm} & = & \lambda_{n,n+q-1}\ \sigma_{n+q-1}^{\pm1}\ \lambda^{-1}_{n,n+q-1},\\ \end{array} $$ \noindent and where $\beta^{\prime} \in B_{1,n+q}$ is the word $\beta$ with the substitutions: $$ \begin{array}{ccl} t & \longleftrightarrow & ({\lambda^{-1}_{n-1,1}}\ {\lambda_{n,n+q-1}}\ {\lambda_{n+q-1,1}})\ t,\\ &&\\ {t}^{-1} & \longleftrightarrow & t^{-1}\ ({\lambda^{-1}_{n+q-1,1}}\ {\lambda^{-1}_{n,n+q-1}}\ {\lambda_{n-1,1}}).\\ \end{array} $$ \begin{figure} \begin{center} \includegraphics[width=6in]{Leneg} \end{center} \caption{ A $\mathbb{Q}$-braid band move in $L(p,q)$ and its algebraic expression. } \label{figure20} \end{figure} In Figure~\ref{lpqeg} the case where $p=2$ and $q=3$ is illustrated. \begin{figure} \begin{center} \includegraphics[width=3.3in]{lenspacesexample23} \end{center} \caption{ An algebraic $\mathbb{Q}$-braid band move in $L(2,3)$. } \label{lpqeg} \end{figure} \subsection{Homology spheres} It is known that a Dehn surgery on a knot yields a homology sphere exactly when the surgery coefficient is the reciprocal of an integer (see [Ro] p.262). For example, surgery on the right-handed trefoil, with surgery coefficient $-1$ yields the Poincare Manifold also known as ``dodecahedral space" (for the algebraic braid equivalence in this case see \cite{LR2}, Section~4). In this subsection we give the algebraic braid equivalence for knots in a homology sphere $M$ obtained from $S^3$ by surgery on the trefoil knot with rational surgery coefficient $1/q$, where $q \in \mathbb{Z}$. \smallbreak \textit{Two oriented links in $M$ are isotopic if and only if any two corresponding algebraic mixed braids in $B_{2, \infty}$ differ by a finite sequence of the following moves:} \smallbreak {\it (1)} \ {\it Algebraic $M$-moves: \ $\beta \sim \beta {\sigma^{\pm 1}_n}, \ \ \ \ \beta \in B_{2,n}$} \vspace{.03in} {\it (2)} \ {\it Algebraic $M$-conjugation: \ $\beta \sim {\sigma^{\mp 1}_i} \beta {\sigma^{\pm 1}_i}, \ \ \ \ \beta, \sigma_i \in B_{2,n}$} \vspace{.03in} {\it (3)} \ {\it Combed Loop conjugation: \ $\beta \sim {a_i}^{\mp 1} \beta {\rho_i}^{\pm 1}, \ \ \ \ \beta \in B_{2,n}$, \textit{where} $\rho_i$ \textit{is the combing of the loop} $a_i$ \textit{through} $\widehat{B}$,} \vspace{.03in} {\it (4)} \ {\it Combed algebraic braid band moves: } \ $\beta \sim d \ c_{\pm} \ \beta^{\prime} \ comb_B(c_1, c_2)$, \textit{where:} $\beta \in B_{2,n}$, $$ \begin{array}{ccl} d & = & (\lambda_{n+2q-1,n+q+1}\ \lambda_{n+1,n+q}^{-1}\ \lambda_{n,1})\ a_2\ (\lambda_{n,1}^{-1}\ \lambda_{n+1,n+q}),\\ &&\\ c_{\pm} & = & \lambda_{n,n+2q-1}\ \sigma_{n+2q-1}^{\pm1}\ \lambda_{n,n+2q-1}^{-1},\\ \end{array} $$ \noindent $\beta^{\prime}$ \textit{is the word} $\beta$ \textit{with the substitutions:} $$ \begin{array}{ccl} a_1 & \longleftrightarrow & (\lambda_{n-1,1}^{-1} \ \lambda_{n,n+2q-1}\ \lambda_{n+2q-1,n+q}\ \lambda^{-1}_{n,n+q-1}\ \lambda_{n-1,1}) a_1,\\ & & \\ a^{-1}_1 & \longleftrightarrow & a^{-1}_1 \ (\lambda_{n-1,1}^{-1}\ \lambda_{n,n+q-1}\ \lambda^{-1}_{n+2q-1,n+q}\ \lambda^{-1}_{n,n+2q-1}\ \lambda_{n-1,1}),\\ & & \\ a_2 & \longleftrightarrow & (\lambda_{n-1,1}^{-1}\ \lambda_{n,n+2q-1}\ \lambda_{n+2q-1,1})\ a_2\\ & & (\lambda_{n-1,1}^{-1}\ \lambda_{n,n+q-1}\ \lambda^{-1}_{n+2q-1,n+q}\ \lambda^{-1}_{n,n+2q-1}\ \lambda_{n-1,1}),\\ & & \\ a^{-1}_2 & \longleftrightarrow & (\lambda_{n-1,1}^{-1}\ \lambda_{n,n+2q-1}\ \lambda_{n+2q-1,n+q}\ \lambda_{n,n+q}^{-1}\ \lambda_{n-1,1})\ a^{-1}_2 \\ & & (\lambda^{-1}_{n+2q-1,1}\ \lambda^{-1}_{n,n+2q-1}\ \lambda_{n-1,1}),\\ \end{array} $$ \noindent \textit{and} $comb_B(c_1, c_2)$ \textit{is the combing of the $q$-strand cables ($c_1$ and $c_2$) through the fixed braid:} $$ \begin{array}{ccl} comb_B(c_1, c_2) & = & \prod_{i=0}^{q-1}{\lambda_{n+i,1} \ a_2\ \lambda^{-1}_{n+i,1}} \ \prod_{i=0}^{q-1}{\lambda_{n+2q-1-i,1}\ a^{-1}_2\ \lambda^{-1}_{n+2q-1-i,1}} \\ &&\\ && \prod_{i=0}^{q-1}{\lambda_{n+q+i,1}\ a_1\ \lambda^{-1}_{n+q+i,1}} \ \lambda_{n+q,1} \ a_2\ \lambda^{-1}_{n,1}\ \lambda_{n+1,n+q}\\ &&\\ && \prod_{i=1}^{q-1}{\lambda_{n+q+i,1}\ a_2\ \lambda^{-1}_{n,1}\ \lambda_{n+1,n+q}\ \lambda^{-1}_{n+q+i,n+q+1}}\\ &&\\ && \prod_{i=0}^{q-1}{\lambda_{n+q-1-i,1}\ a_2\ \lambda^{-1}_{n+q-1-i,1}} \ \prod_{i=0}^{q-1}{\lambda_{n+i,1}\ a_1\ \lambda^{-1}_{n+i,1}}\\ &&\\ && \prod_{i=0}^{q-1}{\lambda_{n+i,1}\ a_2\ \lambda^{-1}_{n+i,1}} \ \prod_{i=0}^{q-1}{\lambda_{n+i,1}\ a_1\ \lambda^{-1}_{n+i,1}}\\ &&\\ && \prod_{i=0}^{q-1}{\lambda_{n+i,1}\ a_2\ \lambda^{-1}_{n+i,1}} \ \prod_{i=0}^{q-1}{\lambda_{n+q+i,n+1+i}}.\\ \end{array} $$ \subsection{Seifert Manifolds} It is known that a Seifert manifold $M((p_1,q_1), \ldots, (p_{m-1},q_{m-1}))$ has a rational surgery description as shown in Figure~\ref{seifsurg} (see \cite{Sa}, p.33)). \begin{figure} \begin{center} \includegraphics[width=3.2in]{SeifertSaveliev} \end{center} \caption{ Surgery description of a Seifert manifold. } \label{seifsurg} \end{figure} \textit{Two oriented links in a Seifert manifold $M((p_1,q_1), \ldots, (p_{m-1},q_{m-1}))$ are isotopic if and only if any two corresponding algebraic mixed braids differ by a finite sequence of the following moves:} \smallbreak {\it (1)} \ {\it Algebraic $M$-moves: \ $\beta \sim \beta {\sigma^{\pm 1}_n}, \ \ \ \ \beta \in B_{m,n}$} \vspace{.03in} {\it (2)} \ {\it Algebraic $M$-conjugation: \ $\beta \sim {\sigma^{\mp 1}_i} \beta {\sigma^{\pm 1}_i}, \ \ \ \ \beta, \sigma_i \in B_{m,n}$} \vspace{.03in} {\it (3)} \ {\it Combed loop conjugation: \ $\beta \sim {a_j}^{\mp 1} \beta {a_j}^{\pm 1}, \ \ \ \ \beta \in B_{m,n}$. } \vspace{.03in} {\it (4)} \ {\it Combed algebraic braid band moves: } \ For $\beta \in B_{m,n}$ we distinguish the cases: \smallbreak $\bullet$ If a $\mathbb{Q}$-braid band move is performed on the $j^{th}$ strand of the fixed braid with rational coefficient $p/q$ (see Figure~\ref{seifcomb}) then: $\beta \sim d \ c_{\pm} \ \beta^{\prime} \ comb_B(c_j)$, where $comb_B(c_j)$ is the combing of the $c_j$ cable through $B$, $$d\ =\ [\lambda_{n+q-1,1}\ \alpha_i\ \lambda_{n-1,1}^{-1}]^p \ \ {\rm and} \ \ c_{\pm}\ =\ \lambda_{n,n+q-1}\ \sigma_{n+q-1}^{-1}\ \lambda_{n,n+q-1}^{-1},$$ \smallbreak and where $\beta^{\prime}$ is $\beta$ with the substitutions: $$ \begin{array}{ccll} a^{\pm 1}_i & \longleftrightarrow & a^{\pm 1}_i,& \quad i\ > \ j,\\ &&& \\ a_i^{\pm1} & \longleftrightarrow & \lambda_{n-1,1}^{-1} \ \lambda_{n,n+q-1}\ \lambda_{n+q-1,1}\ a_i^{\pm1}& \\ && \lambda_{n+q-1,1}^{-1}\ \lambda_{n,n+q-1}^{-1}\ \lambda_{n-1,1},& \quad i\ < \ j,\\ &&& \\ a_j &\longleftrightarrow & \lambda_{n-1,1}^{-1} \ \lambda_{n,n+q-1} \ \lambda_{n+q-1,1} \ a_j & \\ &&& \\ a^{-1}_j & \longleftrightarrow & a^{-1}_j \lambda_{n+q-1,1}^{-1} \ \lambda^{-1}_{n,n+q-1} \ \lambda_{n-1,1}.& \\ \end{array} $$ \noindent $\bullet$ If a $\mathbb{Q}$-braid band move is performed on the last strand of the fixed braid with surgery coefficient $0$, then: $$\beta \sim \sigma_{n}^{\pm 1}\ \beta^{\prime},$$ \ \noindent where $\beta^{\prime}$ is $\beta$ with the substitutions: $$ \begin{array}{ccl} a^{\pm 1}_{j} & \longleftrightarrow & \lambda_{n-1,1}^{-1} \ \sigma^{2}_n\ \lambda_{n-1,1} a^{\pm 1}_{j}\ \lambda_{n,1}^{-1}\ \sigma_n^{-1} \ \lambda_{n-1,1},\ \textrm{for}\ j=1, \ldots, m-1,\\ && \\ a_{m} & \longleftrightarrow & \lambda_{n-1,1}^{-1}\ \sigma^{2}_n\ \lambda_{n-1,1}\ a_{m},\\ && \\ a^{-1}_{m} & \longleftrightarrow & a^{-1}_{m}\ \lambda_{n-1,1}^{-1}\ \sigma^{-2}_n\ \lambda_{n-1,1}.\\ \end{array} $$ \begin{figure} \begin{center} \includegraphics[width=6in]{seifert} \end{center} \caption{ A $\mathbb{Q}$-braid band move in a Seifert manifold and its algebraic expression. } \label{seifcomb} \end{figure} \subsection{Rational surgery along a torus knot} It is well-known that a manifold $M$ obtained by rational surgery from $S^3$ along an $(m,r)$-torus knot with rational coefficient $p/q$ is either the lens space $L(|q|,pr^2)$, or the connected sum of two lens spaces $L(m,r)\sharp L(r,m)$, or a Seifert manifold (for more details the reader is referred to \cite{LM}). \textit{Two oriented links in $M$ are isotopic if and only if any two corresponding algebraic mixed braids differ by a finite sequence of the following moves: } \smallbreak {\it (1)} \ {\it Algebraic $M$-moves: \ $\beta \sim \beta {\sigma^{\pm 1}_n}, \ \ \ \ \alpha \in B_{m,n}$} \vspace{.03in} {\it (2)} \ {\it Algebraic $M$-conjugation: \ $\beta \sim {\sigma^{\mp 1}_i} \beta {\sigma^{\pm 1}_i}, \ \ \ \ \beta, \sigma_i \in B_{m,n}$} \vspace{.03in} {\it (3)} \ {\it Combed loop conjugation: \ $\beta \sim {a_i}^{\mp 1} \beta {\rho_i}^{\pm 1}, \ \ \ \ \beta \in B_{m,n}$, \textit{where} $\rho_i$ \textit{is the combing of the loop} $a_i$ \textit{through} $\widehat{B}$,} \vspace{.03in} {\it (4)} \ {\it Combed algebraic braid band moves:} For $\beta \in B_{m,n}$ we have: $$\beta \sim d \ c_{\pm} \ \beta^{\prime} \ comb_B(c_1, \ldots, c_m),$$ \noindent where $$ \begin{array}{lll} d & = & [\ \lambda_{n+mq-1,n+(m-1)q+1}\ \lambda_{n,n+(m-1)q}^{-1}\ \lambda_{n-1,1}\ \alpha_j\ \lambda_{n-1,1}^{-1}\ \lambda_{n,n+(m-1)q}\ ]^p,\\ c_{\pm} & = & \lambda_{n,n+mq-2}\ \sigma_{n+mq-1}^{\pm1}\ \lambda_{n,n+mq-2}^{-1}, \end{array} $$ \noindent $comb_B(c_1, \ldots, c_m)$ \textit{is the combing through the fixed braid braid of the parted moving cables parallel to the surgery strands and $\beta^{\prime}$ is the word $\beta$ with the substitutions:} $$ \begin{array}{ccl} a_j & \longleftrightarrow & (\lambda_{n-1,1}^{-1} \ \lambda_{n,n+mq-1} \ \lambda_{n+mq-1,n+(j-1)q} \ \lambda^{-1}_{n,n+(j-1)q-1} \ \lambda_{n-1,1}) \ a_j\\ && (\lambda^{-1}_{n-1,1} \ \lambda_{n,n+jq-1} \ \lambda^{-1}_{n+mq-1,n+jq} \ \lambda^{-1}_{n,n+mq-1} \ \lambda_{n-1,1}),\\ && \\ a^{-1}_j & \longleftrightarrow & (\lambda_{n-1,1}^{-1} \ \lambda_{n,n+mq-1} \ \lambda_{n+mq-1,n+jq} \ \lambda^{-1}_{n,n+jq-1} \ \lambda_{n-1,1}) \ a^{-1}_j\\ && (\lambda^{-1}_{n-1,1} \ \lambda_{n,n+(j-1)q-1} \ \lambda^{-1}_{n+mq-1,n+(j-1)q} \ \lambda^{-1}_{n,n+mq-1} \ \lambda_{n-1,1}), \ for\ j \in \{ 1, \ldots , m \}.\\ \end{array} $$ In Figures~\ref{homtref} and \ref{hompart} we illustrate an example where the $(m,r)$-torus knot is the $(2,3)$-torus knot, $p=2$ and $q=3$ (see Proposition~3.1 in \cite{LM} for details about the manifold obtained). \begin{figure} \begin{center} \includegraphics[width=3.5in]{trefoileg} \end{center} \caption{ A geometric $(2,3)$-braid band move along a trefoil. } \label{homtref} \end{figure} \begin{figure} \begin{center} \includegraphics[width=5.2in]{trefoilegparting} \end{center} \caption{ Turning the geometric $(2,3)$-braid band move into a combed algebraic $(2,3)$-braid band move. } \label{hompart} \end{figure} \subsection{ Illustrations for an abstract generic example} Let $M$ be the manifold obtained by rational surgery along a framed link $\widehat{B}$ in $S^3$. Let also $B \bigcup \beta$ be a parted mixed braid representing a link in $M$. In Figures~\ref{seifcomb1} to \ref{seifcomb6} we illustrate step-by-step the algebraization of a geometric $\mathbb{Q}$-braid band move. More precisely, in Figure~\ref{seifcomb1} a geometric $\mathbb{Q}$-braid band move takes place on the last strand of a surgery component $(s_1,\ldots,s_k)$ of $B$. In Figure~\ref{seifcomb2} we part all cables $c_1,\ldots,c_k$ arising from the geometric $\mathbb{Q}$-braid band move, turning the initial geometric $\mathbb{Q}$-braid band move to a parted $\mathbb{Q}$-braid band move. In Figure~\ref{seifcomb2} we also part locally the $(p,q)$-torus subbraid $d^{\prime}$. This leads to the algebraic expression $d$ of $d^{\prime}$, illustrated in Figure~\ref{seifcomb3}, where the local parting of the crossing subbraid $c^{\prime}_{\pm}$ is also initiated. In Figure~\ref{seifcomb4} the algebraic expression $c_{\pm}$ of $c^{\prime}_{\pm}$ is illustrated and the local parting of all loop generators is also initiated. This leads to the algebraic expressions of the loop generators, in Figure~\ref{seifcomb5}, where also the preparation for combing of the cables $c_1,\ldots,c_k$ through $B$ is illustrated. Note that the top part of Figure~\ref{seifcomb5} (above the dotted line) illustrates an algebraic $\mathbb{Q}$-braid band move. Finally, in Figure~\ref{seifcomb6} the combing of the cables $c_1,\ldots,c_k$ through $B$ is performed and the final result is a combed algebraic $\mathbb{Q}$-braid band move. \begin{figure} \begin{center} \includegraphics[width=3.2in]{ap1} \end{center} \caption{ A geometric $\mathbb{Q}$-braid band move. } \label{seifcomb1} \end{figure} \begin{figure} \begin{center} \includegraphics[width=4.1in]{ap2} \end{center} \caption{ Parting locally $d^{\prime}$. } \label{seifcomb2} \end{figure} \clearpage{} \begin{figure} \begin{center} \includegraphics[width=3.7in]{ap3} \end{center} \caption{ Algebraization of $d^{\prime}$ to $d$ and local parting of $c^{\prime}_{\pm}$. } \label{seifcomb3} \end{figure} \begin{figure} \begin{center} \includegraphics[width=3.9in]{ap4} \end{center} \caption{ Algebraization of $c^{\prime}_{\pm}$ to $c$ and local parting of the loop generators $a_{i}$. } \label{seifcomb4} \end{figure} \begin{figure} \begin{center} \includegraphics[width=4in]{ap5} \end{center} \caption{ Algebraization of the loop generators and preparation for combing. } \label{seifcomb5} \end{figure} \begin{figure} \begin{center} \includegraphics[width=3.3in]{ap6} \end{center} \caption{ The combing of the cables through $B$. } \label{seifcomb6} \end{figure}
\section{Introduction} An astronomical image might display multiple superimposed features, such as point sources, compact objects, diffuse emission, or background radiation. The raw photon count images delivered by high energy telescopes are far from perfect; they suffer from shot noise and distortions caused by instrumental effects. The analysis of such astronomical observations demands elaborate denoising, deconvolution, and decomposition strategies. The data obtained by the detection of individual photons is subject to Poissonian shot noise which is more severe for low count rates. This causes difficulty for the discrimination of faint sources against noise, and makes their detection exceptionally challenging. Furthermore, uneven or incomplete survey coverage and complex instrumental response functions leave imprints in the photon data. As a result, the data set might exhibit gaps and artificial distortions rendering the clear recognition of different features a difficult task. Point-like sources are especially afflicted by the instrument's point spread function (PSF) that smooths them out in the observed image, and therefore can cause fainter ones to vanish completely in the background noise. In addition to such noise and convolution effects, it is the superposition of the different objects that makes their separation ambiguous, if possible at all. In astrophysics, photon emitting objects are commonly divided into two morphological classes, diffuse sources and point sources. Diffuse sources span rather smoothly across large fractions of an image, and exhibit apparent internal correlations. Point sources, on the contrary, are local features that, if observed perfectly, would only appear in one pixel of the image. In this work, we will not distinguish between diffuse sources and background, both are diffuse contributions. Intermediate cases, which are sometimes classified as extended or compact sources, are also not considered here. The question arises of how to reconstruct the original source contributions, the individual signals, that caused the observed photon data. This task is an ill-posed inverse problem without a unique solution. There are a number of heuristic and probabilistic approaches that address the problem of denoising, deconvolution, and decomposition in partial or simpler settings. \textsc{SExtractor}~\citep{BA96} is one of the heuristic tools and the most prominent for identifying sources in astronomy. Its popularity is mostly based on its speed and easy operability. However, \textsc{SExtractor} produces a catalog of fitted sources rather than denoised and deconvolved signal estimates. \textsc{CLEAN}~\citep{H74} is commonly used in radio astronomy and attempts a deconvolution assuming there are only contributions from point sources. Therefore, diffuse emission is not optimally reconstructed in the analysis of real observations using \textsc{CLEAN}. \onecolumn \begin{figure*}[!t] \centering \begin{tabular}{c} \begin{overpic} [scale=0.5]{images/120a2.pdf} \put(-2,29){(a)} \end{overpic} \\ \begin{overpic} [scale=0.5]{images/120b2.pdf} \put(-2,29){(b)} \end{overpic} \\ \begin{overpic} [scale=0.5]{images/120c2.pdf} \put(-2,29){(c)} \end{overpic} \\ \begin{overpic} [scale=0.5]{images/120d2.pdf} \put(-2,29){(d)} \end{overpic} \end{tabular} \caption{Illustration of a 1D reconstruction scenario with $1024$ pixels. Panel (a) shows the superimposed diffuse and point-like signal components (green solid line) and its observational response (gray contour). Panel (b) shows again the signal response representing noiseless data (gray contour) and the generated Poissonian data (red markers). Panel (c) shows the reconstruction of the point-like signal component (blue solid line), the diffuse one (orange solid line), its $2\sigma$ reconstruction uncertainty interval (orange dashed line), and again the original signal response (gray contour). The point-like signal comprises $1024$ point-sources of which only five are not too faint. Panel (d) shows the reproduced signal response representing noiseless data (black solid line), its $2\sigma$ shot noise interval (black dashed line), and again the data (gray markers).} \label{fig:motivation} \end{figure*} \twocolumn \noindent Multiscale extensions of CLEAN \citep{C_08,RC11} improve on this, but are also not prefect \citep{JBSE13}. Decomposition techniques for diffuse backgrounds, based on the analysis of angular power spectra have recently been proposed by \citet{HPS12}. Inference methods, in contrast, investigate the probabilistic relation between the data and the signals. Here, the signals of interest are the different source contributions. Probabilistic approaches allow a transparent incorporation of model and \emph{a~priori} assumptions, but often result in computationally heavier algorithms. As an initial attempt, a maximum likelihood analysis was proposed by \citet{V82}. In later work, maximum entropy \citep{S03} and minimum $\chi^2$ methods \citep[e.g.,][]{BABR13} were applied to the INTEGRAL/SPI data reconstructing a single signal component, though. On the basis of sparse regularization a number of techniques exploiting waveforms \citep[based on the work by][]{H10,H11} have proven successful in performing denoising and deconvolution tasks in different settings \citep{G+06,W07,D09,F10,D11}. For example, \citet{S10,S12} analyzed simulated (single and multi-channel) data from the Fermi $\gamma$-ray space telescope focusing on the removal of Poisson noise and deconvolution or background separation. Furthermore, a (generalized) morphological component analysis denoised, deconvolved and decomposed simulated radio data assuming Gaussian noise statistics \citep{B07,C13}. Still in the regime of Gaussian noise, \citet{GC08} derived a deconvolution algorithm for point and extended sources minimizing regularized least squares. They introduce an efficient convex regularization scheme at the price of \emph{a~priori} unmotivated fine tuning parameters. The fast algorithm PowellSnakes~I/II by \citet{CRH09,CRHL11} is capable of analyzing multi-frequency data sets and detecting point-like objects within diffuse emission regions. It relies on matched filters using PSF templates and Bayesian filters exploiting, among others, priors on source position, size, and number. PowellSnakes~II has been successfully applied to the Planck data \citep{P711}. The approach closest to ours is the background-source separation technique used to analyze the ROSAT data \citep{GFD09}. This Bayesian model is based on a two-component mixture model that reconstructs extended sources and (diffuse) background concurrently. The latter is, however, described by a spline model with a small number of spline sampling points. The strategy presented in this work aims at the simultaneous reconstruction of two signals, the diffuse and point-like photon flux. Both fluxes contribute equally to the observed photon counts, but their morphological imprints are very different. The proposed algorithm, derived in the framework of information field theory (IFT) \citep{EFK09,E13}, therefore incorporates prior assumptions in form of a hierarchical parameter model. The fundamentally different morphologies of diffuse and point-like contributions reflected in different prior correlations and statistics. The exploitation of these different prior models is crucial to the signal decomposition. In this work, we exclusively consider Poissonian noise, in particular, but not exclusively, in the low count rate regime, where the signal-to-noise ratio becomes challengingly low. The D$^3$PO algorithm presented here targets the simultaneous denoising, deconvolution, and decomposition of photon observations into two signals, the diffuse and point-like photon flux. This task includes the reconstruction of the harmonic power spectrum of the diffuse component from the data themselves. Moreover, the proposed algorithm provides \emph{a~posteriori} uncertainty information on both inferred signals. The fluxes from diffuse and point-like sources contribute equally to the observed photon counts, but their morphological imprints are very different. The proposed algorithm, derived in the framework of information field theory (IFT) \citep{EFK09,E13,E14}, therefore incorporates prior assumptions in form of a hierarchical parameter model. The fundamentally different morphologies of diffuse and point-like contributions reflected in different prior correlations and statistics. The exploitation of these different prior models is key to the signal decomposition. The diffuse and point-like signal are treated as two separate signal fields. A signal field represents an original signal appearing in nature; e.g., the physical photon flux distribution of one source component as a function of real space or sky position. In theory, a field has infinitely many degrees of freedom being defined on a continuous position space. In computational practice, however, a field needs of course to be defined on a finite grid. It is desirable that the signal field is reconstructed independently from the grid's resolution, except for potentially unresolvable features.\footnote{If the resolution of the reconstruction were increased gradually, the diffuse signal field might exhibit more and more small scale features until the information content of the given data is exhausted. From this point on, any further increase in resolution would not change the signal field reconstruction significantly. In a similar manner, the localization accuracy and number of detections of point sources might increase with the resolution until all relevant information of the data was captured. All higher resolution grids can then be regarded as acceptable representations of the continuous position space.} We note that the point-like signal field hosts one point source in every pixel, however, most of them might be invisibly faint. Hence, a complicated determination of the number of point sources, as many algorithms perform, is not required in our case. The derivation of the algorithm makes use of a wide range of Bayesian methods that are discussed below in detail with regard to their implications and applicability. For now, let us consider an example to emphasize the range and performance of the D$^3$PO algorithm. Figure~\ref{fig:motivation} illustrates a reconstruction scenario in one dimension, where the coordinate could be an angle or position (or time, or energy) in order to represent a 1D sky (or a time series, or an energy spectrum). The numerical implementation uses the \textsc{NIFTy}\footnote{\textsc{NIFTy} homepage \url{http://www.mpa-garching.mpg.de/ift/nifty/}} package \citep{S+13}. \textsc{NIFTy} permits an algorithm to be set up abstractly, independent of the finally chosen topology, dimension, or resolution of the underlying position space. In this way, a 1D prototype code can be used for development, and then just be applied in 2D, 3D, or even on the sphere. The remainder of this paper is structured as follows. Sec.~\ref{sec:problem} discusses the inference on photon observations; i.e., the underlying model and prior assumptions. The D$^3$PO algorithm solving this inference problem by denoising, deconvolution, and decomposition is derived in Sec.~\ref{sec:solution}. In Sec.~\ref{sec:application} the algorithm is demonstrated in a numerical application on simulated high energy photon data. We conclude in Sec.~\ref{sec:conclusion}. \section{Inference on photon observations} \label{sec:problem} \subsection{Signal inference} \label{sec:ift} Here, a signal is defined as an unknown quantity of interest that we want to learn about. The most important information source on a signal is the data obtained in an observation to measure the signal. Inferring a signal from an observational data set poses a fundamental problem because of the presence of noise in the data and the ambiguity that several possible signals could have produced the same data, even in the case of negligible noise. For example, given some image data like photon counts, we want to infer the underlying photon flux distribution. This physical flux is a continuous scalar field that varies with respect to time, energy, and observational position. The measured photon count data, however, is restricted by its spatial and energy binning, as well as its limitations in energy range and observation time. Basically, all data sets are finite for practical reasons, and therefore cannot capture all of the infinitely many degrees of freedom of the underlying continuous signal field. There is no exact solution to such signal inference problems, since there might be (infinitely) many signal field configurations that could lead to the same data. This is why a probabilistic data analysis, which does not pretend to calculate the correct field configuration but provides expectation values and uncertainties of the signal field, is appropriate for signal inference. Given a data set $\bb{d}$, the \emph{a~posteriori} probability distribution $P(\bb{s}|\bb{d})$ judges how likely a potential signal $\bb{s}$ is. This posterior is given by Bayes' theorem, \begin{align} P(\bb{s}|\bb{d}) &= \frac{P(\bb{d}|\bb{s}) \, P(\bb{s})}{P(\bb{d})} , \label{eq:bayes} \end{align} as a combination of the likelihood $P(\bb{d}|\bb{s})$, the signal prior $P(\bb{s})$, and the evidence $P(\bb{d})$, which serves as a normalization. The likelihood characterizes how likely it is to measure data set $\bb{d}$ from a given signal field $\bb{s}$. It covers all processes that are relevant for the measurement of $\bb{d}$. The prior describes the knowledge about $\bb{s}$ without considering the data, and should, in general, be less restrictive than the likelihood. IFT is a Bayesian framework for the inference of signal fields exploiting mathematical methods for theoretical physics. A signal field, $\bb{s} = s(x)$, is a function of a continuous position $x$ in some position space $\Omega$. In order to avoid a dependence of the reconstruction on the partition of $\Omega$, the according calculus regarding fields is geared to preserve the continuum limit \citep{E13,E14,S+13}. In general, we are interested in the \emph{a~posteriori} mean estimate $\bb{m}$ of the signal field given the data, and its (uncertainty) covariance $\bb{D}$, defined as \begin{align} \bb{m} &= \left< \bb{s} \right>_{(\bb{s}|\bb{d})} = \int\mathcal{D}\bb{s} \; \bb{s} \; P(\bb{s}|\bb{d}) , \label{eq:map} \\ \bb{D} &= \left< (\bb{m}-\bb{s})(\bb{m}-\bb{s})^\dagger}%{\intercal \right>_{(\bb{s}|\bb{d})} , \label{eq:error} \end{align} where $\dagger}%{\intercal$ denotes adjunction and $\left< \,\cdot\, \right>_{(\bb{s}|\bb{d})}$ the expectation value with respect to the posterior probability distribution $P(\bb{s}|\bb{d})$.\footnote{This expectation value is computed by a path integral, $\int \mathcal{D}\bb{s}$, over the complete phase space of the signal field $\bb{s}$; i.e., all possible field configurations.} In the following, the posterior of the physical photon flux distribution of two morphologically different source components given a data set of photon counts is build up piece by piece according to Eq.~\eqref{eq:bayes}. \subsection{Poissonian likelihood} \label{sec:likelihood} The images provided by astronomical high energy telescopes typically consist of integer photon counts that are binned spatially into pixels. Let $d_i$ be the number of detected photons, also called events, in pixel $i$, where $i \in \{1,\dots,N_\mathrm{pix}\} \subset \mathds{N}$. The kind of signal field we would like to infer from such data is the causative photon flux distribution. The photon flux, $\bb{\rho} = \rho(x)$, is defined for each position $x$ on the observational space $\Omega$. In astrophysics, this space $\Omega$ is typically the $\mathcal{S}^2$ sphere representing an all-sky view, or a region within $\mathds{R}^2$ representing an approximately plane patch of the sky. The flux $\bb{\rho}$ might express different morphological features, which can be classified into a diffuse and point-like component. The exact definitions of the diffuse and point-like flux should be specified \emph{a~priori}, without knowledge of the data, and are addressed in Sec.~\ref{sec:s_prior} and \ref{sec:u_prior}, respectively. At this point it shall suffices to say that the diffuse flux varies smoothly on large spatial scales, while the flux originating from point sources is fairly local. These two flux components are superimposed, \begin{align} \bb{\rho} &= \bb{\rho}_\mathrm{diffuse} + \bb{\rho}_\mathrm{point-like} = \rho_0 \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) , \label{eq:superposition} \end{align} where we introduced the dimensionless diffuse and point-like signal fields, $\bb{s}$ and $\bb{u}$, and the constant $\rho_0$ which absorbs the physical dimensions of the photon flux; i.e., events per area per energy and time interval. The exponential function in Eq.~\eqref{eq:superposition} is applied componentwise. In this way, we naturally account for the strict positivity of the photon flux at the price of a non-linear change of variables, from the flux to its natural logarithm. A measurement apparatus observing the photon flux $\bb{\rho}$ is expected to detect a certain number of photons $\bb{\lambda}$. This process can be modeled by a linear response operator $\bb{R}_0$ as follows, \begin{align} \bb{\lambda} &= \bb{R}_0 \bb{\rho} = \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) , \label{eq:l} \end{align} where $\bb{R} = \bb{R}_0 \rho_0$. This reads for pixel $i$, \begin{align} \lambda_i &= \int_\Omega \d x \; R_i(x) \left( \mathrm{e}^{s(x)} + \mathrm{e}^{u(x)} \right) . \label{eq:l_i} \end{align} The response operator $\bb{R}_0$ comprises all aspects of the measurement process; i.e., all instrument response functions. This includes the survey coverage, which describes the instrument's overall exposure to the observational area, and the instrument's PSF, which describes how a point source is imaged by the instrument. The superposition of different components and the transition from continuous coordinates to some discrete pixelization, cf. Eq.~\eqref{eq:l_i}, cause a severe loss of information about the original signal fields. In addition to that, measurement noise distorts the signal's imprint in the data. The individual photon counts per pixel can be assumed to follow a Poisson distribution $\P$ each. Therefore, the likelihood of the data $\bb{d}$ given an expected number of events $\bb{\lambda}$ is modeled as a product of statistically independent Poisson processes, \begin{align} P(\bb{d}|\bb{\lambda}) &= \prod_i \P(d_i,\lambda_i) = \prod_i \frac{1}{d_i!} \; \lambda_i^{d_i} \mathrm{e}^{-\lambda_i} . \label{eq:likelihood} \end{align} The Poisson distribution has a signal-to-noise ratio of $\sqrt{\bb{\lambda}}$ which scales with the expected number of photon counts. Therefore, Poissonian shot noise is most severe in regions with low photon fluxes. This makes the detection of faint sources in high energy astronomy a particularly challenging task, as X- and $\gamma$-ray photons are sparse. The likelihood of photon count data given a two component photon flux is hence described by the Eqs.~\eqref{eq:l} and \eqref{eq:likelihood}. Rewriting this likelihood $P(\bb{d}|\bb{s},\bb{u})$ in form of its negative logarithm yields the information Hamiltonian $H(\bb{d}|\bb{s},\bb{u})$,\footnote{Throughout this work we define $H(\,\cdot\,) = - \log P(\,\cdot\,)$, and absorb constant terms into a normalization constant $H_0$ in favor of clarity.} \begin{align} H(\bb{d}|\bb{s},\bb{u}) &= -\log P(\bb{d}|\bb{s},\bb{u}) \\ &= H_0 + \bb{1}^\dagger}%{\intercal \bb{\lambda} - \bb{d}^\dagger}%{\intercal \log(\bb{\lambda}) \\ &= H_0 + \bb{1}^\dagger}%{\intercal \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) - \bb{d}^\dagger}%{\intercal \log \left( \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) \right) , \label{eq:H_likelihood} \end{align} where the ground state energy $H_0$ comprises all terms constant in $\bb{s}$ and $\bb{u}$, and $\bb{1}$ is a constant data vector being $1$ everywhere. \subsection{Prior assumptions} \label{sec:prior} The diffuse and point-like signal fields, $\bb{s}$ and $\bb{u}$, contribute equally to the likelihood defined by Eq.~\eqref{eq:H_likelihood}, and thus leaving it completely degenerate. On the mere basis of the likelihood, the full data set could be explained by the diffuse signal alone, or only by point-sources, or any other conceivable combination. In order to downweight intuitively implausible solutions, we introduce priors. The priors discussed in the following address the morphology of the different photon flux contributions, and define diffuse and point-like in the first place. These priors aid the reconstruction by providing some remedy for the degeneracy of the likelihood. The likelihood describes noise and convolution properties, and the prior describe the individual morphological properties. Therefore, the denoising and deconvolution of the data towards the total photon flux $\bb{\rho}$ is primarily likelihood driven, but for a decomposition of the total photon flux into $\bb{\rho}^{(s)}$ and $\bb{\rho}^{(u)}$, the signal priors are imperative. \subsubsection{Diffuse component} \label{sec:s_prior} The diffuse photon flux, $\bb{\rho}^{(s)} = \rho_0 \mathrm{e}^\bb{s}$, is strictly positive and might vary in intensity over several orders of magnitude. Its morphology shows cloudy patches with smooth fluctuations across spatial scales; i.e., one expects similar values of the diffuse flux in neighboring locations. In other words, the diffuse component exhibits spatial correlations. A log-normal model for $\bb{\rho}^{(s)}$ satisfies those requirements according to the maximum entropy principle \citep{OSBE12,K13}. If the diffuse photon flux follows a multivariate log-normal distribution, the diffuse signal field $\bb{s}$ obeys a multivariate Gaussian distribution $\mathcal{G}$, \begin{align} P(\bb{s}|\bb{S}) &= \mathcal{G}(\bb{s},\bb{S}) = \frac{1}{\sqrt{\mathrm{det}[2\pi\bb{S}]}} \; \exp\left( - \frac{1}{2} \bb{s}^\dagger}%{\intercal \bb{S}^{-1} \bb{s} \right) , \label{eq:s_prior} \end{align} with a given covariance $\bb{S} = \left< \bb{s} \bb{s}^\dagger}%{\intercal \right>_{(\bb{s}|\bb{S})}$. This covariance describes the strength of the spatial correlations, and thus the smoothness of the fluctuations. A convenient parametrization of the covariance $\bb{S}$ can be found, if the signal field $\bb{s}$ is \emph{a~priori} not known to distinguish any position or orientation axis; i.e., its correlations only depend on relative distances. This is equivalent to assume $\bb{s}$ to be statistically homogeneous and isotropic. Under this assumption, $\bb{S}$ is diagonal in the harmonic basis\footnote{The basis in which the Laplace operator is diagonal is denoted harmonic basis. If $\Omega$ is a $n$-dimensional Euclidean space $\mathds{R}^n$ or Torus $\mathcal{T}^n$, the harmonic basis is the Fourier basis; if $\Omega$ is the $\mathcal{S}^2$ sphere, the harmonic basis is the spherical harmonics basis.} of the position space $\Omega$ such that \begin{align} \bb{S} &= \sum_k \mathrm{e}^{\tau_k} \bb{S}_k , \end{align} where $\tau_k$ are spectral parameters and $\bb{S}_k$ are projections onto a set of disjoint harmonic subspaces of $\Omega$. These subspaces are commonly denoted as spectral bands or harmonic modes. The set of spectral parameters, $\bb{\tau} = \{\tau_k\}_k$, is then the logarithmic power spectrum of the diffuse signal field $\bb{s}$ with respect to the chosen harmonic basis denoted by $k$. However, the diffuse signal covariance is in general unknown \emph{a~priori}. This requires the introduction of another prior for the covariance, or for the set of parameters $\bb{\tau}$ describing it adequately. This approach of hyperpriors on prior parameters creates a hierarchical parameter model. \subsubsection{Unknown power spectrum} \label{sec:p_prior} The lack of knowledge of the power spectrum, requires its reconstruction from the same data the signal is inferred from \citep{WLL04,JKWE10,EF11,J13}. Therefore, two \emph{a~priori} constraints for the spectral parameters $\bb{\tau}$, which describe the logarithmic power spectrum, are incorporated in the model. The power spectrum is unknown and might span over several orders of magnitude. This implies a logarithmically uniform prior for each element of the power spectrum, and a uniform prior for each spectral parameter $\tau_k$, respectively. We initially assume independent inverse-Gamma distributions $\mathcal{I}$ for the individual elements, \begin{align} P(\mathrm{e}^{\bb{\tau}}|\bb{\alpha},\bb{q}) &= \prod_k \mathcal{I}(\mathrm{e}^{\tau_k},\alpha_k,q_k) \\ &= \prod_k \frac{q_k^{\alpha_k-1}}{\Gamma(\alpha_k-1)} \; \mathrm{e}^{-\left(\alpha_k \tau_k + q_k \mathrm{e}^{-\tau_k} \right)} , \end{align} and hence \begin{align} P_\mathrm{un}(\bb{\tau}|\bb{\alpha},\bb{q}) &= \prod_k \mathcal{I}(\mathrm{e}^{\tau_k},\alpha_k,q_k) \left| \frac{\d\mathrm{e}^{\tau_k}}{\d\tau_k} \right| \\ &\propto \exp\Big( - (\bb{\alpha} - \bb{1})^\dagger}%{\intercal \bb{\tau} - \bb{q}^\dagger}%{\intercal \mathrm{e}^\bb{-\tau} \Big) , \end{align} where $\bb{\alpha} = \{\alpha_k\}_k$ and $\bb{q} = \{q_k\}_k$ are the shape and scale parameters, and $\Gamma$ denotes the Gamma function. In the limit of $\alpha_k \rightarrow 1$ and $q_k \rightarrow 0 \; \forall k$, the inverse-Gamma distributions become asymptotically flat on a logarithmic scale, and thus $P_\mathrm{un}$ constant.\footnote{If $P(\tau_k = \log z) = \mathrm{const.}$, then a substitution yields $P(z) = P(\log z) \; |\d(\log z)/\d z| \propto z^{-1} \sim \mathcal{I}(z,\alpha \rightarrow 1, q \rightarrow 0)$.} Small non-zero scale parameters, $0 < q_k$, provide lower limits for the power spectrum that, in practice, lead to more stable inference algorithms. So far, the variability of the individual elements of the power spectrum is accounted for, but the question about their correlations has not been addressed. Empirically, power spectra of a diffuse signal field do not exhibit wild fluctuation or change drastically over neighboring modes. They rather show some sort of spectral smoothness. Moreover, for diffuse signal fields that were shaped by local and causal processes, we might expect a finite correlation support in position space. This translates into a smooth power spectrum. In order to incorporate spectral smoothness, we employ a prior introduced by \citet{EF11,OSBE12}. This prior is based on the second logarithmic derivative of the spectral parameters $\bb{\tau}$, and favors power spectra that obey a power law. It reads \begin{align} P_\mathrm{sm}(\bb{\tau}|\bb{\sigma}) &\propto \exp\left( - \frac{1}{2} \bb{\tau}^\dagger}%{\intercal \bb{T} \bb{\tau} \right) , \end{align} with \begin{align} \bb{\tau}^\dagger}%{\intercal \bb{T} \bb{\tau} &= \int\d(\log k) \frac{1}{\sigma_k^2} \left( \frac{\partial^2 \tau_k}{\partial (\log k)^2} \right)^2 , \end{align} where $\bb{\sigma} = \{\sigma_k\}_k$ are Gaussian standard deviations specifying the tolerance against deviation from a power-law behavior of the power spectrum. A choice of $\sigma_k = 1 \; \forall k$ would typically allow for a change in the power law's slope of $1$ per e-fold in $k$. In the limit of $\sigma_k \rightarrow \infty \; \forall k$, no smoothness is enforced upon the power spectrum. The resulting prior for the spectral parameters is given by the product of the priors discussed above, \begin{align} P(\bb{\tau}|\bb{\alpha},\bb{q},\bb{\sigma}) &= P_\mathrm{un}(\bb{\tau}|\bb{\alpha},\bb{q}) \; P_\mathrm{sm}(\bb{\tau}|\bb{\sigma}) . \label{eq:t_prior} \end{align} The parameters $\bb{\alpha},\bb{q}$ and $\bb{\sigma}$ are considered to be given as part of the hierarchical Bayesian model, and provide a flexible handle to model our knowledge on the scaling and smoothness of the power spectrum. \subsubsection{Point-like component} \label{sec:u_prior} The point-like photon flux, $\bb{\rho}^{(u)} = \rho_0 \mathrm{e}^\bb{u}$, is supposed to originate from very distant astrophysical sources. These sources appear morphologically point-like to an observer because their actual extent is negligible given the extreme distances. This renders point sources to be spatially local phenomena. The photon flux contributions of neighboring point sources can (to zeroth order approximation) be assumed to be statistically independent of each other. Even if the two sources are very close on the observational plane, their physical distance might be huge. Even in practice, the spatial cross-correlation of point sources is negligible. Therefore, statistically independent priors for the photon flux contribution of each point-source are assumed in the following. Because of the spatial locality of a point source, the corresponding photon flux signal is supposed to be confined to a single spot, too. If the point-like signal field, defined over a continuous position space $\Omega$, is discretized properly\footnote{The numerical discretization of information fields is described in great detail in \citet{S+13}.}, this spot is sufficiently identified by an image pixel in the reconstruction. A discretization, $\rho(x \in \Omega) \rightarrow (\rho_x)_x$, is an inevitable step since the algorithm is to be implemented in a computer environment anyway. Nevertheless, we have to ensure that the \emph{a~priori} assumptions do not depend on the chosen discretization but satisfy the continuous limit. Therefore, the prior for the point-like signal component factorizes spatially, \begin{align} P(\bb{\rho}^{(u)}) &= \prod_x P(\rho_x^{(u)}) , \label{eq:prod} \end{align} but the functional form of the priors are yet to be determined. This model allows the point-like signal field to host one point source in every pixel. Most of these point sources are expected to be invisibly faint contributing negligibly to the total photon flux. However, the point sources which are just identifiable from the data are pinpointed in the reconstruction. In this approach, there is no necessity for a complicated determination of the number and position of sources. For the construction of a prior, that the photon flux is a strictly positive quantity also needs to be considered. Thus, a simple exponential prior, \begin{align} P(\rho_x^{(u)}) &\propto \exp\left( - \rho_x^{(u)}/\rho_0 \right) , \end{align} has been suggested \citep[e.g.,][]{GFD09}. It has the advantage of being (easily) analytically treatable, but its physical implications are questionable. This distribution strongly suppresses high photon fluxes in favor of lower ones. The maximum entropy prior, which is also often applied, is even worse because it corresponds to a brightness distribution,\footnote{The so-called maximum entropy regularization $\sum_x (\rho_x^{(u)}/\rho_0) \log(\rho_x^{(u)}/\rho_0)$ of the log-likelihood can be regarded as log-prior, cf. Eqs.~\eqref{eq:prod} and \eqref{eq:ME}.} \begin{align} P(\rho_x^{(u)}) &\propto \left( \rho_x^{(u)}/\rho_0 \right)^{\left( - \rho_x^{(u)}/\rho_0 \right)} . \label{eq:ME} \end{align} The following (rather crude) consideration might motivate a more astrophysical prior. Say the universe hosts a homogeneous distribution of point sources. The number of point sources would therefore scale with the observable volume; i.e., with distance cubed. Their apparent brightness, which is reduced because of the spreading of the light rays; i.e., a proportionality to the distance squared. Consequently, a power-law behavior between the number of point sources and their brightness with a slope $\beta = \tfrac{3}{2}$ is to be expected \citep{F68,MH11}. However, such a plain power law diverges at $0$, and is not necessarily normalizable. Furthermore, Galactic and extragalactic sources cannot be found in arbitrary distances owing to the finite size of the Galaxy and the cosmic (past) light cone. Imposing an exponential cut-off above $0$ onto the power law yields an inverse-Gamma distribution, which has been shown to be an appropriate prior for point-like photon fluxes \citep{GFD09,CRH09,CRHL11}. The prior for the point-like signal field is therefore derived from a product of independent inverse-Gamma distributions,\footnote{A possible extension of this prior model that includes spatial correlations would be an inverse-Wishart distribution for $\mathrm{diag}[ \bb{\rho}^{(u)} ]$.} \begin{align} P(\bb{\rho}^{(u)}|\bb{\beta},\bb{\eta}) &= \prod_x \mathcal{I}(\rho_x^{(u)},\beta_x,\rho_0\eta_x) \\ &= \prod_x \frac{(\rho_0\eta_x)^{\beta_x-1}}{\Gamma(\beta_x-1)} \left( \rho_x^{(u)} \right)^{-\beta_x} \exp\left( - \frac{\rho_0\eta_x}{\rho_x^{(u)}} \right) , \end{align} yielding \begin{align} P(\bb{u}|\bb{\beta},\bb{\eta}) &= \prod_x \mathcal{I}(\rho_0\mathrm{e}^{u_x},\beta_k,\rho_0\eta_k) \left| \frac{\d\rho_0\mathrm{e}^{u_x}}{\d u_x} \right| \\ &\propto \exp\Big( - (\bb{\beta} - \bb{1})^\dagger}%{\intercal \bb{u} - \bb{\eta}^\dagger}%{\intercal \mathrm{e}^\bb{-\bb{u}} \Big) , \label{eq:u_prior} \end{align} where $\bb{\beta} = \{\beta_x\}_x$ and $\bb{\eta} = \{\eta_x\}_x$ are the shape and scale parameters. The latter is responsible for the cut-off of vanishing fluxes, and should be chosen adequately small in analogy to the spectral scale parameters $\bb{q}$. The determination of the shape parameters is more difficile. The geometrical argument above suggests a universal shape parameter, $\beta_x = \tfrac{3}{2} \; \forall x$. A second argument for this value results from demanding \emph{a~priori} independence of the discretization. If we choose a coarser resolution that would add up the flux from two point sources at merged pixels, then our prior should still be applicable. The universal value of $\tfrac{3}{2}$ indeed fulfills this requirement as shown in Appendix~\ref{app:stacking}. There it is also shown that $\eta$ has to be chosen resolution dependent, though. \subsection{Parameter model} \label{sec:parameter_model} Figure~\ref{fig:hierarchy} gives an overview of the parameter hierarchy of the suggested Bayesian model. The data $\bb{d}$ is given, and the diffuse signal field $\bb{s}$ and the point-like signal field $\bb{u}$ shall be reconstructed from that data. The logarithmic power spectrum $\bb{\tau}$ is a set of nuisance parameters that also need to be reconstructed from the data in order to accurately model the diffuse flux contributions. The model parameters form the top layer of this hierarchy and are given to the reconstruction algorithm. This set of model parameters can be boiled down to five scalars, namely $\alpha$, $q$, $\sigma$, $\beta$, and $\eta$, if one defines $\bb{\alpha} = \alpha \bb{1}$, etc. The incorporation of the scalars in the inference is possible in theory, but this would increase the computational complexity dramatically. We discussed reasonable values for these scalars to be chosen \emph{a~priori}. If additional information sources, such as theoretical power spectra or object catalogs, are available the model parameters can be adjusted accordingly. In Sec.~\ref{sec:application}, different parameter choices for the analysis of simulated data are investigated. \section{Denoising, deconvolution, and decomposition} \label{sec:solution} \begin{figure}[t!] \centering \begin{tikzpicture} [c/.style={circle,minimum size=2em,text centered,thin}, r/.style={rectangle,minimum size=2em,text centered,thin}, v/.style={->,shorten >=1pt,>=stealth,thick}] \node(a)at(-2,5)[r,text width=3em,draw]{$\alpha,\;q$}; \node(z)at(0,5)[r,draw]{$\sigma$}; \node(b)at(2,5)[r,text width=3em,draw]{$\beta,\;\eta$}; \node(t)at(-1,4)[c,draw]{$\bb{\tau}$}; \node(s)at(-1,3)[c,draw]{$\bb{s}$}; \node(u)at(2,3)[c,draw]{$\bb{u}$}; \node(g)at(0.5,2)[c,draw,dashed]{$\bb{\rho}$}; \node(l)at(0.5,1)[c,draw,dashed]{$\bb{\lambda}$}; \node(d)at(0.5,0)[r,draw]{$\bb{d}$}; \draw[v](a.south)--(t); \draw[v](z.south)--(t); \draw[v](t)--(s); \draw[v](b)--(u); \draw[v](s)--(g); \draw[v](u)--(g); \draw[v](g)--(l); \draw[v](l)--(d); \end{tikzpicture} \flushleft \caption{Graphical model of the model parameters $\alpha$, $q$, $\sigma$, $\beta$, and $\eta$, the logarithmic spectral parameters $\bb{\tau}$, the diffuse signal field $\bb{s}$, the point-like signal field $\bb{u}$, the total photon flux $\bb{\rho}$, the expected number of photons $\bb{\lambda}$, and the observed photon count data $\bb{d}$.} \label{fig:hierarchy} \end{figure} The likelihood model, describing the measurement process, and the prior assumptions for the signal fields and the power spectrum of the diffuse component yield a well-defined inference problem. The corresponding posterior is given by \begin{align} P(\bb{s},\bb{\tau},\bb{u}|\bb{d}) &= \frac{P(\bb{d}|\bb{s},\bb{u})\; P(\bb{s}|\bb{\tau}) \; P(\bb{\tau}|\alpha,q,\sigma) \; P(\bb{u}|\beta,\eta)}{P(\bb{d})} , \label{eq:posterior} \end{align} which is a complex form of Bayes' theorem~\eqref{eq:bayes}. Ideally, we would now calculate the \emph{a~posteriori} expectation values and uncertainties according to Eqs.~\eqref{eq:map} and \eqref{eq:error} for the diffuse and point-like signal fields, $\bb{s}$ and $\bb{u}$, as well as for the logarithmic spectral parameters $\bb{\tau}$. However, an analytical evaluation of these expectation values is not possible because of the complexity of the posterior. The posterior is non-linear in the signal fields and, except for artificially constructed data, non-convex. It, however, is more flexible and therefore allows for a more comprehensive description of the parameters to be inferred \citep{KGV83,GG84}. Numerical approaches involving Markov chain Monte Carlo methods \citep{MU49,M+53} are possible, but hardly feasible because of the huge parameter phase space. Nevertheless, similar problems have been addressed by elaborate sampling techniques \citep{WLL04,JKWE10,JK10,J13}. Here, two approximative algorithms with lower computational costs are derived. The first one uses the maximum \emph{a~posteriori} (MAP) approximation, the second one minimizes the Gibbs free energy of an approximate posterior ansatz in the spirit of variational Bayesian methods. The fidelity and accuracy of these two algorithms are compared in a numerical application in Sec.~\ref{sec:application}. \subsection{Posterior maximum} \label{sec:MAP} The posterior maximum and mean coincide, if the posterior distribution is symmetric and single peaked. In practice, this often holds -- at least in good approximation --, so that the maximum \emph{a~posteriori} approach can provide suitable estimators. This can either be achieved using a $\delta$-distribution at the posterior's mode, \begin{align} \left< \bb{s} \right>_{(\bb{s}|\bb{d})} &\overset{\mathrm{MAP}\text{-}\delta}{\approx} \int\mathcal{D}\bb{s} \; \bb{s} \; \delta(\bb{s} - \bb{s}_\mathrm{mode}) , \label{eq:MAPmap} \end{align} or using a Gaussian approximation around this point, \begin{align} \left< \bb{s} \right>_{(\bb{s}|\bb{d})} &\overset{\mathrm{MAP}\text{-}\mathcal{G}}{\approx} \int\mathcal{D}\bb{s} \; \bb{s} \; \mathcal{G}(\bb{s} - \bb{s}_\mathrm{mode},\bb{D}_\mathrm{mode}) , \label{eq:MAPmap2} \end{align} Both approximations require us to find the mode, which is done by extremizing the posterior. Instead of the complex posterior distribution, it is convenient to consider the information Hamiltonian, defined by its negative logarithm, \begin{align} H(\bb{s},\bb{\tau},\bb{u}|\bb{d}) &= - \log P(\bb{s},\bb{\tau},\bb{u}|\bb{d}) \\ &= H_0 + \bb{1}^\dagger}%{\intercal \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) - \bb{d}^\dagger}%{\intercal \log \left( \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) \right) \notag \\ &\quad + \frac{1}{2} \log\left( \mathrm{det}\left[ \bb{S} \right] \right) + \frac{1}{2} \bb{s}^\dagger}%{\intercal \bb{S}^{-1} \bb{s} \label{eq:H} \\ &\quad + (\bb{\alpha} - \bb{1})^\dagger}%{\intercal \bb{\tau} + \bb{q}^\dagger}%{\intercal \mathrm{e}^\bb{-\tau} + \frac{1}{2} \bb{\tau}^\dagger}%{\intercal \bb{T} \bb{\tau} \notag \\ &\quad + (\bb{\beta} - \bb{1})^\dagger}%{\intercal \bb{u} + \bb{\eta}^\dagger}%{\intercal \mathrm{e}^{-\bb{u}} \notag , \end{align} where all terms constant in $\bb{s}$, $\bb{\tau}$, and $\bb{u}$ have been absorbed into a ground state energy $H_0$, cf. Eqs.~\eqref{eq:likelihood}, \eqref{eq:s_prior}, \eqref{eq:t_prior}, and \eqref{eq:u_prior}, respectively. The MAP solution, which maximizes the posterior, minimizes the Hamiltonian. This minimum can thus be found by taking the first (functional) derivatives of the Hamiltonian with respect to $\bb{s}$, $\bb{\tau}$, and $\bb{u}$ and equating them with zero. Unfortunately, this yields a set of implicit, self-consistent equations rather than an explicit solution. However, these equations can be solved by an iterative minimization of the Hamiltonian using a steepest descent method for example, see Sec.~\ref{sec:algorithm} for details. In order to better understand the structure of the MAP solution, we consider the minimum $(\bb{s},\bb{\tau},\bb{u}) = (\bb{m}^{(s)},\bb{\tau}^\star,\bb{m}^{(u)})$. The resulting filter formulas for the diffuse and point-like signal field read \begin{align} \frac{\partial H}{\partial\bb{s}} \bigg|_\mathrm{min} &= \bb{0} = \left( \bb{1} - \bb{d}/\bb{l} \right)^\dagger}%{\intercal \bb{R} \ast \mathrm{e}^{\bb{m}^{(s)}} + {\bb{S}^\star}^{-1} \bb{m}^{(s)} , \label{eq:s_MAP} \\ \frac{\partial H}{\partial\bb{u}} \bigg|_\mathrm{min} &= \bb{0} = \left( \bb{1} - \bb{d}/\bb{l} \right)^\dagger}%{\intercal \bb{R} \ast \mathrm{e}^{\bb{m}^{(u)}} + \bb{\beta} - \bb{1} - \bb{\eta} \ast \mathrm{e}^{-\bb{m}^{(u)}} , \label{eq:u_MAP} \end{align} with \begin{align} \bb{l} &= \bb{R} \left( \mathrm{e}^{\bb{m}^{(s)}} + \mathrm{e}^{\bb{m}^{(u)}} \right) , \label{eq:l_MAP} \\ \bb{S}^\star &= \sum_k \mathrm{e}^{\tau_k^\star} \bb{S}_k . \end{align} Here, $\ast$ and $/$ denote componentwise multiplication and division, respectively. The first term in Eq.~\eqref{eq:s_MAP} and \eqref{eq:u_MAP}, which comes from the likelihood, vanishes in case $\bb{l} = \bb{d}$. We note that $\bb{l} = \bb{\lambda}|_\mathrm{min}$ describes the most likely number of photon counts, not the expected number of photon counts $\bb{\lambda} = \left< \bb{d} \right>_{(\bb{d}|\bb{s},\bb{u})}$, cf. Eqs.~\eqref{eq:l} and \eqref{eq:likelihood}. Disregarding the regularization by the priors, the solution would overfit; i.e., noise features are partly assigned to the signal fields in order to achieve an unnecessarily close agreement with the data. However, the \emph{a~priori} regularization suppresses this tendency to some extend. The second derivative of the Hamiltonian describes the curvature around the minimum, and therefore approximates the (inverse) uncertainty covariance, \begin{align} \frac{\partial^2 H}{\partial\bb{s}\partial\bb{s}^\dagger}%{\intercal} \bigg|_\mathrm{min} &\approx {\bb{D}^{(s)}}^{-1} , & \frac{\partial^2 H}{\partial\bb{u}\partial\bb{u}^\dagger}%{\intercal} \bigg|_\mathrm{min} &\approx {\bb{D}^{(u)}}^{-1} . \label{eq:D_MAP} \end{align} The closed form of $\bb{D}^{(s)}$ and $\bb{D}^{(u)}$ is given explicitly in Appendix~\ref{app:covariance}. The filter formula for the power spectrum, which is derived from a first derivative of the Hamiltonian with respect to $\bb{\tau}$, yields \begin{align} \mathrm{e}^{\bb{\tau}^\star} &= \frac{\bb{q} + \frac{1}{2} \left( \mathrm{tr}\left[ \bb{m}^{(s)}{\bb{m}^{(s)}}^\dagger}%{\intercal \bb{S}_k^{-1} \right] \right)_k}{\bb{\gamma} + \bb{T} \bb{\tau}^\star} , \label{eq:t_MAP} \end{align} where $\bb{\gamma} = (\bb{\alpha} - \bb{1}) + \frac{1}{2} \left( \mathrm{tr}\left[ \bb{S}_k {\bb{S}_k}^{-1} \right] \right)_k$. This formula is in accordance with the results by \citet{EF11,OSBE12}. It has been shown by the former authors that such a filter exhibits a perception threshold; i.e., on scales where the signal-response-to-noise ratio drops below a certain bound the reconstructed signal power becomes vanishingly low. This threshold can be cured by a better capture of the \emph{a~posteriori} uncertainty structure. \subsection{Posterior approximation} \label{sec:Gibbs} In order to overcome the analytical infeasibility as well as the perception threshold, we seek an approximation to the true posterior. Instead of approximating the expectation values of the posterior, approximate posteriors are investigated in this section. In case the approximation is good, the expectation values of the approximate posterior should then be close to the real ones. The posterior given by Eq.~\eqref{eq:posterior} is inaccessible because of the entanglement of the diffuse signal field $\bb{s}$, its logarithmic power spectrum $\bb{\tau}$, and the point-like signal field $\bb{u}$. The involvement of $\bb{\tau}$ can been simplified by a mean field approximation, \begin{align} P(\bb{s},\bb{\tau},\bb{u}|\bb{d}) &\approx Q = Q_s(\bb{s},\bb{u}|\bb{\mu},\bb{d}) \; Q_\tau(\bb{\tau}|\bb{\mu},\bb{d}) , \label{eq:PQ} \end{align} where $\bb{\mu}$ denotes an abstract mean field mediating some information between the signal field tuple $(\bb{s},\bb{u})$ and $\bb{\tau}$ that are separated by the product ansatz in Eq.~\eqref{eq:PQ}. This mean field is fully determined by the problem, as it represents effective (rather than additional) degrees of freedom. It is only needed implicitly for the derivation, an explicit formula can be found in Appendix~\ref{app:about_t}, though. Since the \emph{a~posteriori} mean estimates for the signal fields and their uncertainty covariances are of primary interest, a Gaussian approximation for $Q_s$ that accounts for correlation between $\bb{s}$ and $\bb{u}$ would be sufficient. Hence, our previous approximation is extended by setting \begin{align} Q_s(\bb{s},\bb{u}|\bb{\mu},\bb{d}) = \mathcal{G}(\bb{\varphi},\bb{D}) , \end{align} with \begin{align} \bb{\varphi} &= \begin{pmatrix} \bb{s} - \bb{m}^{(s)} \\ \bb{u} - \bb{m}^{(u)} \end{pmatrix} , & \bb{D} &= \left( \begin{array}{ll} \bb{D}^{(s)} & \bb{D}^{(su)} \\ {\bb{D}^{(su)}}^\dagger}%{\intercal & \bb{D}^{(u)} \end{array} \right) . \label{eq:gau} \end{align} This Gaussian approximation is also a convenient choice in terms of computational complexity because of its simple analytic structure. The goodness of the approximation $P \approx Q$ can be quantified by an information theoretical measure, see Appendix~\ref{app:KLG}. The Gibbs free energy of the inference problem, \begin{align} G &= \big< H(\bb{s},\bb{\tau},\bb{u}|\bb{d}) \big>_Q - \big< - \log Q(\bb{s},\bb{\tau},\bb{u}|\bb{d}) \big>_Q , \end{align} which is equivalent to the Kullback-Leibler divergence $D_\mathrm{KL}(Q,P)$, is chosen as such a measure \citep{EW10}. In favor of comprehensibility, we suppose the solution for the logarithmic power spectrum $\bb{\tau}^\star$ is known for the moment. The Gibbs free energy is then calculated by plugging in the Hamiltonian, and evaluating the expectation values\footnote{The second likelihood term in Eq.~\eqref{eq:G}, $\bb{d}^\dagger}%{\intercal \log(\bb{\lambda})$, is thereby expanded according to \begin{align} \log(x) &= \log\left<x\right> - \sum_{\nu=2}^\infty \frac{(-1)^\nu}{\nu} \left< \left( \frac{x}{\left<x\right>} - 1 \right)^\nu \right> \notag \\ &\approx \log\left<x\right> + \mathcal{O}\left( \left< x^2 \right> \right) , \notag \end{align} under the assumption $x \approx \left<x\right>$.}, \begin{align} G &= G_0 + \big< H(\bb{s},\bb{u}|\bb{d}) \big>_{Q_s} - \frac{1}{2} \log\left( \mathrm{det}\left[ \bb{D} \right] \right) \\ &= G_1 + \bb{1}^\dagger}%{\intercal \bb{l} - \bb{d}^\dagger}%{\intercal \left\{ \log(\bb{l}) - \sum_{\nu=2}^\infty \frac{(-1)^\nu}{\nu} \big< \left( \bb{\lambda} / \bb{l} - 1 \right)^\nu \big>_{Q_s} \right\} \notag \\ &\quad + \frac{1}{2} {\bb{m}^{(s)}}^\dagger}%{\intercal {\bb{S}^\star}^{-1} \bb{m}^{(s)} + \frac{1}{2} \mathrm{tr}\left[ \bb{D}^{(s)} {\bb{S}^\star}^{-1} \right] \label{eq:G} \\ &\quad + (\bb{\beta} - \bb{1})^\dagger}%{\intercal \bb{m}^{(u)} + \bb{\eta}^\dagger}%{\intercal \mathrm{e}^{-\bb{m}^{(u)} + \tfrac{1}{2} \hh{D}^{(u)}} \notag \\ &\quad - \frac{1}{2} \log\left( \mathrm{det}\left[ \bb{D} \right] \right) , \notag \end{align} with \begin{align} \bb{\lambda} &= \bb{R} \left( \mathrm{e}^\bb{s} + \mathrm{e}^\bb{u} \right) , \\ \bb{l} &= \left< \bb{\lambda}\right>_{Q_s} = \bb{R} \left( \mathrm{e}^{\bb{m}^{(s)} + \tfrac{1}{2} \hh{D}^{(s)}} + \mathrm{e}^{\bb{m}^{(u)} + \tfrac{1}{2} \hh{D}^{(u)}} \right) , \label{eq:l_Gibbs} \\ \bb{S}^\star &= \sum_k \mathrm{e}^{\tau_k^\star} \bb{S}_k , \mathrm{\ and} \\ \hh{D} &= \mathrm{diag}\left[ \bb{D} \right] . \end{align} Here, $G_0$ and $G_1$ carry all terms independent of $\bb{s}$ and $\bb{u}$. In comparison to the Hamiltonian given in Eq.~\eqref{eq:H}, there are a number of correction terms that now also consider the uncertainty covariances of the signal estimates properly. For example, the expectation values of the photon fluxes differ comparing $\bb{l}$ in Eq.~\eqref{eq:l_MAP} and \eqref{eq:l_Gibbs} where it now describes the expectation value of $\bb{\lambda}$ over the approximate posterior. In case $\bb{l} = \bb{\lambda}$ the explicit sum in Eq.~\eqref{eq:G} vanishes. Since this sum includes powers of $\left< \bb{\lambda}^{\nu > 2} \right>_{Q_s}$ its evaluation would require all entries of $\bb{D}$ to be known explicitly. In order to keep the algorithm computationally feasible, this sum shall hereafter be neglected. This is equivalent to truncating the corresponding expansion at second order; i.e., $\nu = 2$. It can be shown that, in consequence of this approximation, the cross-correlation $\bb{D}^{(su)}$ equals zero, and $\bb{D}$ becomes block diagonal. Without these second order terms, the Gibbs free energy reads \begin{align} G &= G_1 + \bb{1}^\dagger}%{\intercal \bb{l} - \bb{d}^\dagger}%{\intercal \log(\bb{l}) \notag \\ &\quad + \frac{1}{2} {\bb{m}^{(s)}}^\dagger}%{\intercal {\bb{S}^\star}^{-1} \bb{m}^{(s)} + \frac{1}{2} \mathrm{tr}\left[ \bb{D}^{(s)} {\bb{S}^\star}^{-1} \right] \\ &\quad + (\bb{\beta} - \bb{1})^\dagger}%{\intercal \bb{m}^{(u)} + \bb{\eta}^\dagger}%{\intercal \mathrm{e}^{-\bb{m}^{(u)} + \tfrac{1}{2} \hh{D}^{(u)}} \notag \\ &\quad - \frac{1}{2} \log\left( \mathrm{det}\left[ \bb{D}^{(s)} \right] \right) - \frac{1}{2} \log\left( \mathrm{det}\left[ \bb{D}^{(u)} \right] \right) . \notag \end{align} Minimizing the Gibbs free energy with respect to $\bb{m}^{(s)}$, $\bb{m}^{(u)}$, $\bb{D}^{(s)}$, and $\bb{D}^{(u)}$ would optimize the fitness of the posterior approximation $P \approx Q$. Filter formulas for the Gibbs solution can be derived by taking the derivative of $G$ with respect to the approximate mean estimates, \begin{align} \frac{\partial G}{\partial\bb{m}^{(s)}} = \bb{0} &= \left( \bb{1} - \bb{d}/\bb{l} \right)^\dagger}%{\intercal \bb{R} \ast \mathrm{e}^{\bb{m}^{(s)} + \tfrac{1}{2} \hh{D}^{(s)}} + {\bb{S}^\star}^{-1} \bb{m}^{(s)} , \label{eq:s_Gibbs} \\ \frac{\partial G}{\partial\bb{m}^{(u)}} = \bb{0} &= \left( \bb{1} - \bb{d}/\bb{l} \right)^\dagger}%{\intercal \bb{R} \ast \mathrm{e}^{\bb{m}^{(u)} + \tfrac{1}{2} \hh{D}^{(u)}} \label{eq:u_Gibbs}\\ &\quad + \bb{\beta} - \bb{1} - \bb{\eta} \ast \mathrm{e}^{-\bb{m}^{(u)} + \tfrac{1}{2} \hh{D}^{(u)}} , \notag \end{align} This filter formulas again account for the uncertainty of the mean estimates in comparison to the MAP filter formulas in Eq.~\eqref{eq:s_MAP} and \eqref{eq:u_MAP}. The uncertainty covariances can be constructed by either taking the second derivatives, \begin{align} \frac{\partial^2 G}{\partial\bb{m}^{(s)}\partial{\bb{m}^{(s)}}^\dagger}%{\intercal} &\approx {\bb{D}^{(s)}}^{-1} , & \frac{\partial^2 G}{\partial\bb{m}^{(u)}\partial{\bb{m}^{(u)}}^\dagger}%{\intercal} &\approx {\bb{D}^{(u)}}^{-1} , \label{eq:D_Gibbs} \end{align} or setting the first derivatives of $G$ with respect to the uncertainty covariances equal to zero matrices, \begin{align} \frac{\partial G}{\partial D_{xy}^{(s)}} &= 0 , & \frac{\partial G}{\partial D_{xy}^{(u)}} &= 0 . \label{eq:D_Gibbs_2} \end{align} The closed form of $\bb{D}^{(s)}$ and $\bb{D}^{(u)}$ is given explicitly in Appendix~\ref{app:covariance}. So far, the logarithmic power spectrum $\bb{\tau}^\star$, and with it $\bb{S}^\star$, have been supposed to be known. The mean field approximation in Eq.~\eqref{eq:PQ} does not specify the approximate posterior $Q_\tau(\bb{\tau}|\bb{\mu},\bb{d})$, but it can be retrieved by variational Bayesian methods \citep{JGJS99,WW13}, according to the procedure detailed in Appendix~\ref{app:variation}. The subsequent Appendix~\ref{app:about_t} discusses the derivation of an solution for $\bb{\tau}$ by extremizing $Q_\tau$. This result, which was also derived in \citet{OSBE12}, applies to the inference problem discussed here, yielding \begin{align} \mathrm{e}^{\bb{\tau}^\star} &= \frac{\bb{q} + \frac{1}{2} \left( \mathrm{tr}\left[ \left( \bb{m}^{(s)}{\bb{m}^{(s)}}^\dagger}%{\intercal + \bb{D}^{(s)} \right) \bb{S}_k^{-1} \right] \right)_k}{\bb{\gamma} + \bb{T} \bb{\tau}^\star} . \label{eq:t_Gibbs} \end{align} Again, this solution includes a correction term in comparison to the MAP solution in Eq.~\eqref{eq:t_MAP}. Since $\bb{D}^{(s)}$ is positive definite, it contributes positive to the (logarithmic) power spectrum, and therefore reduces the possible perception threshold further. We note that this is a minimal Gibbs free energy solution that maximizes $Q_\tau$. A proper calculation of $\left< \bb{\tau} \right>_{Q_\tau}$ might include further correction terms, but their derivation is not possible in closed form. Moreover, the above used diffuse signal covariance ${\bb{S}^\star}^{-1}$ should be replaced by $\left< \bb{S}^{-1} \right>_{Q_\tau}$ adding further correction terms to the filter formulas. In order to keep the computational complexity on a feasible level, all these higher order corrections are not considered here. The detailed characterization of their implications and implementation difficulties is left for future investigation. \subsection{Physical flux solution} To perform calculations on the logarithmic fluxes is convenient for numerical reasons, but it is the physical fluxes that are actually of interest to us. Given the chosen approximation, we can compute the posterior expectation values of the diffuse and point-like photon flux, $\bb{\rho}^{(s)}$ and $\bb{\rho}^{(u)}$, straight forwardly, \begin{align} \left< \bb{\rho}^{(\,\cdot\,)} \right>_P \overset{\mathrm{MAP}\text{-}\delta}{\approx} \left< \bb{\rho}^{(\,\cdot\,)} \right>_\delta &= \rho_0 \mathrm{e}^{\bb{m}_\mathrm{mode}^{(\,\cdot\,)}} , \label{eq:r_MAP} \\ \overset{\mathrm{MAP}\text{-}\mathcal{G}}{\approx} \left< \bb{\rho}^{(\,\cdot\,)} \right>_\mathcal{G} &= \rho_0 \mathrm{e}^{\bb{m}_\mathrm{mode}^{(\,\cdot\,)} + \tfrac{1}{2} \hh{D}_\mathrm{mode}^{(\,\cdot\,)}} , \label{eq:r_MAP2} \\ \overset{\mathrm{Gibbs}}{\approx} \left< \bb{\rho}^{(\,\cdot\,)} \right>_Q &= \rho_0 \mathrm{e}^{\bb{m}_\mathrm{mean}^{(\,\cdot\,)} + \tfrac{1}{2} \hh{D}_\mathrm{mean}^{(\,\cdot\,)}} , \label{eq:r_Gibbs} \end{align} in accordance with Eq.~\eqref{eq:MAPmap}, \eqref{eq:MAPmap2}, or \eqref{eq:PQ}, respectively. Those solutions differ from each other in terms of the involvement of the posterior's mode or mean, and in terms of the inclusion of the uncertainty information, see subscripts. In general, the mode approximation holds for symmetric, single peaked distributions, but can perform poorly in other cases \citep[e.g.,][]{EF11}. The exact form of the posterior considered here is highly complex because of the many degrees of freedom. In a dimensionally reduced frame, however, the posterior appears single peaked and exhibits a negative skewness.\footnote{For example, the posterior $P(s|d)$ for a one-dimensional diffuse signal is proportional to $\exp(-\tfrac{1}{2} s^2 + d s - \exp(s))$, whereby all other parameters are fixed to unity. Analogously, $P(u|d) \propto \exp(d u - 2\,\mathrm{cosh}(u))$.} Although this is not necessarily generalizable, it suggest a superiority of the posterior mean compared to the MAP because of the asymmetry of the distribution. Nevertheless, the MAP approach is computationally cheaper compared to the Gibbs approach that requires permanent knowledge of the uncertainty covariance. The uncertainty of the reconstructed photon flux can be approximated as for an ordinary log-normal distribution, \begin{align} \left< {\bb{\rho}^{(\,\cdot\,)}}^2 \right>_P - \left< \bb{\rho}^{(\,\cdot\,)} \right>_P^2 \overset{\mathrm{MAP}}{\approx} \left< \bb{\rho}^{(\,\cdot\,)} \right>_\mathcal{G}^2 &\left( \mathrm{e}^{\hh{D}_\mathrm{mode}^{(\,\cdot\,)}} - 1 \right) , \label{eq:var_MAP} \\ \overset{\mathrm{Gibbs}}{\approx} \left< \bb{\rho}^{(\,\cdot\,)} \right>_Q^2 &\left( \mathrm{e}^{\hh{D}_\mathrm{mean}^{(\,\cdot\,)}} - 1 \right) , \label{eq:var_Gibbs} \end{align} where the square root of the latter term would describe the relative uncertainty. \subsection{Imaging algorithm} \label{sec:algorithm} \begin{figure*}[t] \centering \begin{tabular}{ccc} \begin{overpic} [scale=0.35]{images/345_d.png} \put(-3,97){(a)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_Rk.png} \put(-3,97){(b)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_Rm.png} \put(-3,97){(c)} \end{overpic} \\ \begin{overpic} [scale=0.35]{images/345_l_MAP.png} \put(-3,97){(d)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_l_MAP2.png} \put(-3,97){(e)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_l_Gibbs.png} \put(-3,97){(f)} \end{overpic} \\ \end{tabular} \flushleft \caption{Illustration of the data and noiseless, but reconvolved, signal responses of the reconstructions. Panel (a) shows the data from a mock observation of a $32 \times 32 \,\mathrm{arcmin}^2$ patch of the sky with a resolution of $0.1 \,\mathrm{arcmin}$ corresponding to a total of $102\,400$ pixels. The data had been convolved with a Gaussian-like PSF (FWHM $\approx 0.2 \,\mathrm{arcmin}$ $= 2$ pixels, finite support of $1.1 \,\mathrm{arcmin}$ $= 11$ pixels) and masked because of an uneven exposure. Panel (b) shows the centered convolution kernel. Panel (c) shows the exposure mask. The bottom panels show the reconvolved signal response $\bb{R}\left<\bb{\rho}\right>$ of a reconstruction using a different approach each, namely (d) MAP-$\delta$, (e) MAP-$\mathcal{G}$, and (f) Gibbs. All reconstructions shown here and in the following figures used the same model parameters: $\alpha=1$, $q=10^{-12}$, $\sigma=10$, $\beta = \tfrac{3}{2}$, and $\eta = 10^{-4}$.} \label{fig:d} \end{figure*} \begin{figure*}[t] \centering \begin{tabular}{ccc} \begin{overpic} [scale=0.35]{images/345_s_exp_MAP.png} \put(-3,97){(a)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_s_exp_MAP2.png} \put(-3,97){(b)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_s_exp_Gibbs.png} \put(-3,97){(c)} \end{overpic} \\ \begin{overpic} [scale=0.35]{images/345_ds_exp_MAP.png} \put(-3,97){(d)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_ds_exp_MAP2.png} \put(-3,97){(e)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_ds_exp_Gibbs.png} \put(-3,97){(f)} \end{overpic} \\ \end{tabular} \flushleft \caption{Illustration of the diffuse reconstruction. The top panels show the denoised and deconvolved diffuse contribution $\langle\bb{\rho}^{(s)}\rangle/\rho_0$ reconstructed using a different approach each, namely (d) MAP-$\delta$, (e) MAP-$\mathcal{G}$, and (f) Gibbs. The bottom panels (d) to (f) show the difference between the originally simulated signal and the respective reconstruction.} \label{fig:rs} \end{figure*} \begin{figure*}[t] \centering \begin{tabular}{ccc} \begin{overpic} [scale=0.35]{images/345_s.png} \put(-3,97){(a)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_s_MAP.png} \put(-3,97){(b)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_s_Gibbs.png} \put(-3,97){(c)} \end{overpic} \\ & \begin{overpic} [scale=0.35]{images/345_ds_MAP.png} \put(-3,97){(d)$\,=\,$(a)$\,-\,$(b)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_ds_Gibbs.png} \put(-3,97){(e)$\;=\,$(a)$\,-\,$(c)} \end{overpic} \\ \begin{overpic} [scale=0.35]{images/345_ds_s_Gibbs.png} \put(-3,97){(f)$\;=|$(e)$|\,/\,$(c)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_D_MAP.png} \put(-3,97){(g)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_D_Gibbs.png} \put(-3,97){(h)} \end{overpic} \\ \end{tabular} \flushleft \caption{Illustration of the reconstruction of the diffuse signal field $\bb{s} = \log\bb{\rho}^{(s)}$ and its uncertainty. The top panels show diffuse signal fields. Panel (a) shows the original simulation $\bb{s}$, panel (b) the reconstruction $\bb{m}_\mathrm{mode}^{(s)}$ using a MAP approach, and panel (c) the reconstruction $\bb{m}_\mathrm{mean}^{(s)}$ using a Gibbs approach. The panels (d) and (e) show the differences between original and reconstruction. Panel (f) shows the relative difference. The panels (g) and (h) show the relative uncertainty of the above reconstructions.} \label{fig:s} \end{figure*} \begin{figure*}[!t] \centering \begin{tabular}{cc} \begin{overpic} [scale=0.5]{images/345_p_10b.pdf} \put(-3,57){(a)} \end{overpic} & \begin{overpic} [scale=0.5]{images/345_p_1000b.pdf} \put(-3,57){(b)} \end{overpic} \\ \end{tabular} \flushleft \caption{Illustration of the reconstruction of the logarithmic power spectrum $\bb{\tau}$. Both panels show the default power spectrum (black dashed line), and the simulated realization (black dotted line), as well as the reconstructed power spectra using a MAP (orange solid line), plus second order corrections (orange dashed line), and a Gibbs approach (blue solid line). Panel (a) shows the reconstruction for a chosen $\sigma$ parameter of $10$, panel (b) for a $\sigma$ of $1000$.} \label{fig:t} \end{figure*} \begin{figure*}[t] \centering \begin{tabular}{ccc} \begin{overpic} [scale=0.35]{images/345_u_expb.pdf} \put(-3,97){(a)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_u_exp_MAPb.pdf} \put(-3,97){(b)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_u_exp_Gibbsb.pdf} \put(-3,97){(c)} \end{overpic} \\ & \begin{overpic} [scale=0.35]{images/345_uuF_MAP.pdf} \put(-3,100){(d)} \end{overpic} & \begin{overpic} [scale=0.35]{images/345_uuF_Gibbs.pdf} \put(-3,100){(e)} \end{overpic} \\ \end{tabular} \flushleft \caption{Illustration of the reconstruction of the point-like signal field $\bb{u} = \log\bb{\rho}^{(u)}$ and its uncertainty. The top panels show the location (markers) and intensity (gray scale) of the point-like photon fluxes, underlaid is the respective diffuse contribution (contours) to guide the eye, cf. Fig~\ref{fig:rs}. Panel (a) shows the original simulation, panel (b) the reconstruction using a MAP approach, and panel (c) the reconstruction using a Gibbs approach. The bottom panels (d) and (e) show the match between original and reconstruction in absolute and relative fluxes, the $2\sigma$ shot noise interval (gray contour), as well as some reconstruction uncertainty estimate (error bars).} \label{fig:u} \end{figure*} The problem of denoising, deconvolving, and decomposing photon observations is a non-trivial task. Therefore, this section discusses the implementation of the D$^3$PO algorithm given the two sets of filter formulas derived in Sec.~\ref{sec:MAP} and \ref{sec:Gibbs}, respectively. The information Hamiltonian, or equivalently the Gibbs free energy, are scalar quantities defined over a huge phase space of possible field and parameter configurations including, among others, the elements of $\bb{m}^{(s)}$ and $\bb{m}^{(u)}$. If we only consider those, and no resolution refinement from data to signal space, two numbers need to be inferred from one data value each. Including $\bb{\tau}$ and the uncertainty covariances $\bb{D}^{(s)}$ and $\bb{D}^{(u)}$ in the inference, the problem of underdetermined degrees of freedom gets worse. This is reflected in the possibility of a decent number of local minima in the non-convex manifold landscape of the codomain of the Hamiltonian, or Gibbs free energy, respectively \citep{KGV83,GG84,GC08}. The complexity of the inference problem goes back to the, in general, non-linear entanglement between the individual parameters. The D$^3$PO algorithm is based on an iterative optimization scheme, where certain subsets of the problem are optimized alternately instead of the full problem at once. Each subset optimization is designed individually, see below. The global optimization cycle is in some degree sensitive to the starting values because of the non-convexity of the considered potential; i.e., the information Hamiltonian or Gibbs free energy, respectively. We can find such appropriate starting values by solving the inference problem in a reduced frame in advance, see below. So far, a step-by-step guide of the algorithm looks like the following. \vspace{0.5em} \begin{enumerate \item \label{i} Initialize the algorithm with primitive starting values; e.g., $m_x^{(s)} = m_x^{(u)} = 0$, $D_{xy}^{(s)} = D_{xy}^{(u)} = \delta_{xy}$, and $\tau_k^\star = \log(k^{-2})$. -- Those values are arbitrary. Although the optimization is rather insensitive to them, inappropriate values can cripple the algorithm for numerical reasons because of the high non-linearity of the inference problem. \item \label{s0} Optimize $\bb{m}^{(s)}$, the diffuse signal field, coarsely. -- The preliminary optimization shall yield a rough estimate of the diffuse only contribution. This can be achieved by reconstructing a coarse screened diffuse signal field that only varies on large scales; i.e., limiting the bandwidth of the diffuse signal in its harmonic basis. Alternatively, obvious point sources in the data could be masked out by introducing an artificial mask into the response, if feasible. \item \label{u0} Optimize $\bb{m}^{(u)}$, the point-like signal field, locally. -- This initial optimization shall approximate the brightest, most obvious, point sources that are visible in the data image by eye. Their current disagreement with the data dominates the considered potential, and introduces some numerical stiffness. The gradient of the potential can be computed according to Eq.~\eqref{eq:u_MAP} or \eqref{eq:u_Gibbs}, and its minima will be at the expected position of the brightest point source which has not been reconstructed, yet. It is therefore very efficient to increase $\bb{m}^{(u)}$ at this location directly until the sign of the gradient flips, and repeat this procedure until the obvious point sources are fit. \item \label{u1} Optimize $\bb{m}^{(u)}$, the point-like signal field. -- This task can be done by a steepest descent minimization of the potential combined with a line search following the Wolfe conditions \citep{NW06}. The potentials can be computed according to Eq.~\eqref{eq:H} or \eqref{eq:G} neglecting terms independent of $\bb{m}^{(u)}$, and the gradient according to Eq.~\eqref{eq:u_MAP} or \eqref{eq:u_Gibbs}. A more sophisticated minimization scheme, such as a non-linear conjugate gradient \citep{S94}, is conceivable but would require the application of the full Hessian, cf. step~\ref{u2}. In the first run, it might be sufficient to restrict the optimization to the locations identified in step~\ref{u0}. \item \label{u2} Update $\hh{D}^{(u)}$, the point-like uncertainty variance, in case of a Gibbs approach. -- It is not feasible to compute the full uncertainty covariance $\bb{D}^{(u)}$ explicitly in order to extract its diagonal. A more elegant way is to apply a probing technique relying on the application of $\bb{D}^{(u)}$ to random fields $\bb{\xi}$ that project out the diagonal \citep{H89,SOE12}. The uncertainty covariance is given as the inverse Hessian by Eq.~\eqref{eq:D_MAP} or \eqref{eq:D_Gibbs}, and should be symmetric and positive definite. For that reason, it can be applied to a field using a conjugate gradient \citep{S94}; i.e., solving $(\bb{D}^{(u)})^{-1}\bb{y} = \bb{\xi}$ for $\bb{y}$. However, if the current phase space position is far away from the minimum, the Hessian is not necessarily positive definite. One way to overcome this temporal instability, would be to introduce a Levenberg damping in the Hessian \citep[inspired by][]{TMS09,TS12}. \item \label{s1} Optimize $\bb{m}^{(s)}$, the diffuse signal field. -- An analog scheme as in step~\ref{u1} using steepest descent and Wolfe conditions is effective. The potentials can be computed according to Eq.~\eqref{eq:H} or \eqref{eq:G} neglecting terms independent of $\bb{m}^{(s)}$, and the gradient according to Eq.~\eqref{eq:s_MAP} or \eqref{eq:s_Gibbs}, respectively. It has proven useful to first ensure a convergence on large scales; i.e., small harmonic modes $k$. This can be done repeating steps~\ref{s1}, \ref{s2}, and \ref{t} for all $k < k_\mathrm{max}$ with growing $k_\mathrm{max}$ using the corresponding projections $\bb{S}_k$. \item \label{s2} Update $\hh{D}^{(s)}$, the diffuse uncertainty variance, in case of a Gibbs approach in analogy to step~\ref{u2}. \item \label{t} Optimize $\bb{\tau}^\star$, the logarithmic power spectrum. -- This is done by solving Eq.~\eqref{eq:t_MAP} or \eqref{eq:t_Gibbs}. The trace term can be computed analog to the diagonal; e.g., by probing. Given this, the equation can be solved efficiently by a Newton-Raphson method. \item Repeat the steps~\ref{u1} to \ref{t} until convergence. -- This scheme will take several cycles until the algorithm reaches the desired convergence level. Therefore, it is not required to achieve a convergence to the final accuracy level in all subsets in all cycles. It is advisable to start with weak convergence criteria in the first loop and increase them gradually. \end{enumerate} \noindent A few remarks are in order. The phase space of possible signal field configurations is tremendously huge. It is therefore impossible to judge if the algorithm has converged to the global or some local minima, but this does not matter if both yield reasonable results that do not differ substantially. In general, the converged solution is also subject to the choice of starting values. Solving a non-convex, non-linear inference problem without proper initialization can easily lead to nonsensical results, such as fitting (all) diffuse features by point sources. Therefore, the D$^3$PO algorithm essentially creates its own starting values executing the initial steps~\ref{i} to \ref{u0}. The primitive starting values are thereby processed to rough estimates that cover coarsely resolved diffuse and prominent point-like features. These estimates serve then as actual starting values for the optimization cycle. Because of the iterative optimization scheme starting with the diffuse component in step~\ref{s0}, the algorithm might be prone to explaining some point-like features by diffuse sources. Starting with the point-like component instead would give rise to the opposite bias. To avoid such biases, it is advisable to restart the algorithm partially. To be more precise, we propose to discard the current reconstruction of $\bb{m}^{(u)}$ after finishing step~\ref{t} for the first time, then start the second iteration again with step~\ref{u0}, and to discard the current $\bb{m}^{(s)}$ before step~\ref{s1}. The above scheme exploits a few numerical techniques, such as probing or Levenberg damping, that are described in great detail in the given references. The code of our implementation of the D$^3$PO algorithm will be made public in the future under \url{http://www.mpa-garching.mpg.de/ift/d3po/}. \section{Numerical application} \label{sec:application} Exceeding the simple 1D scenario illustrated in Fig.~\ref{fig:motivation}, the D$^3$PO algorithm is now applied to a realistic, but simulated, data set. The data set represents a high energy observation with a field of view of $32 \times 32 \,\mathrm{arcmin}^2$ and a resolution of $0.1 \,\mathrm{arcmin}$; i.e., the photon count image comprises $102\,400$ pixels. The instrument response includes the convolution with a Gaussian-like PSF with a FWHM of roughly $0.2 \,\mathrm{arcmin}$, and an uneven survey mask attributable to the inhomogeneous exposure of the virtual instrument. The data image and those characteristics are shown in Fig.~\ref{fig:d}. In addition, the top panels of Fig.~\ref{fig:d} show the reproduced signal responses of the reconstructed (total) photon flux. The reconstructions used the same model parameters, $\alpha=1$, $q=10^{-12}$, $\sigma=10$, $\beta = \tfrac{3}{2}$, and $\eta = 10^{-4}$ in a MAP-$\delta$, MAP-$\mathcal{G}$ and a Gibbs approach, respectively. They all show a very good agreement with the actual data, and differences are barely visible by eye. We note that only the quality of denoising is visible, since the signal response shows the convolved and superimposed signal fields. The diffuse contribution to the deconvolved photon flux is shown Fig.~\ref{fig:rs} for all three estimators, cf. Eqs.~\eqref{eq:r_MAP} to \eqref{eq:r_Gibbs}. There, all point-like contributions as well as noise and instrumental effects have been removed presenting a denoised, deconvolved and decomposed reconstruction result for the diffuse photon flux. Figure~\ref{fig:rs} also shows the absolute difference to the original flux. Although the differences in the MAP estimators are insignificant, the Gibbs solution seems to be slightly better. In order to have a quantitative statement about the goodness of the reconstruction, we define a relative residual error $\epsilon^{(s)}$ for the diffuse contribution as follows, \begin{align} \epsilon^{(s)} &= \left| \bb{\rho}^{(s)} - \left< \bb{\rho}^{(s)} \right> \right|_2 \left| \bb{\rho}^{(s)} \right|_2^{-1} , \label{eq:errs} \end{align} where $|\,\cdot\,|_2$ is the Euclidean L$^2$-norm. For the point-like contribution, however, we have to consider an error in brightness and position. For this purpose we define, \begin{align} \epsilon^{(u)} &= \int_1^N \d n \; \left| \bb{R}_\mathrm{PSF}^n \bb{\rho}^{(u)} - \bb{R}_\mathrm{PSF}^n \left< \bb{\rho}^{(u)} \right> \right|_2 \left| \bb{R}_\mathrm{PSF}^n \bb{\rho}^{(u)} \right|_2^{-1} , \label{eq:erru} \end{align} where $\bb{R}_\mathrm{PSF}$ is a (normalized) convolution operator, such that $\bb{R}_\mathrm{PSF}^N$ becomes the identity for large $N$. These errors are listed in Table~\ref{tab:errs}. When comparing the MAP-$\delta$ and MAP-$\mathcal{G}$ approach, the incorporation of uncertainty corrections seems to improve the results slightly. The full regularization treatment within the Gibbs approach outperforms MAP solutions in terms of the chosen error measure $\epsilon^{(\,\cdot\,)}$. For a discussion of how such measures can change the view on certain Bayesian estimators, we refer to the work by \citet{BL14}. \begin{table}[!b] \caption{Overview of the relative residual errors in the photon flux reconstructions for the respective approaches, all using the same model parameters, cf. text.} \centering \begin{tabular}{|rrr|} \hline \multicolumn{1}{|c}{MAP-$\delta$} & \multicolumn{1}{c}{MAP-$\mathcal{G}$} & \multicolumn{1}{c|}{Gibbs} \\ \hline \hline $\epsilon^{(s)} = 4.442$\% & $\epsilon^{(s)} = 4.441$\% & $\epsilon^{(s)} = 2.078$\% \\% 4.44190179 4.44112794 2.07831735 $\epsilon^{(u)} = 1.540$\% & $\epsilon^{(u)} = 1.540$\% & $\epsilon^{(u)} = 1.089$\% \\% 1.53970515 1.53970516 1.08854039 \hline \end{tabular} \label{tab:errs} \end{table} Figure~\ref{fig:s} illustrates the reconstruction of the diffuse signal field, now in terms of logarithmic flux. The original and the reconstructions agree well, and the strongest deviations are found in the areas with low amplitudes. With regard to the exponential ansatz in Eq.~\eqref{eq:superposition}, it is not surprising that the inference on the signal fields is more sensitive to higher values than to lower ones. For example, a small change in the diffuse signal field, $\bb{s} \rightarrow (1 \pm \epsilon) \bb{s}$, translates into a factor in the photon flux, $\bb{\rho}^{(s)} \rightarrow \bb{\rho}^{(s)} \mathrm{e}^{\pm\epsilon\bb{s}}$, that scales exponentially with the amplitude of the diffuse signal field. The Gibbs solution shows less deviation from the original signal than the MAP solution. Since the latter lacks the regularization by the uncertainty covariance it exhibits a stronger tendency to overfitting compared to the former. This includes overestimates in noisy regions with low flux intensities, as well as underestimates at locations where point-like contributions dominate the total flux The reconstruction of the power spectrum, as shown in Fig.~\ref{fig:t}, gives further indications of the reconstruction quality of the diffuse component. The simulation used a default power spectrum of \begin{align} \exp(\tau_k)&= 42 \, (k+1)^{-7} . \end{align} This power spectrum was on purpose chosen to deviate from a strict power law supposed by the smoothness prior. From Fig.~\ref{fig:t} it is apparent that the reconstructed power spectra track the original well up to a harmonic mode $k$ of roughly $0.4 \,\mathrm{arcmin}^{-1}$. Beyond that point, the reconstructed power spectra fall steeply until they hit a lower boundary set by the model parameter $q = 10^{-12}$. This drop-off point at $0.4 \,\mathrm{arcmin}^{-1}$ corresponds to a physical wavelength of roughly $2.5 \,\mathrm{arcmin}$, and thus (half-phase) fluctuations on a spatial distances below $1.25 \,\mathrm{arcmin}$. The Gaussian-like PSF of the virtual observatory has a finite support of $1.1 \,\mathrm{arcmin}$. The lack of reconstructed power indicates that the algorithm assigns features on spatial scales smaller than the PSF support preferably to the point-like component. This behavior is reasonable because solely the point-like signal can cause PSF-like shaped imprints in the data image. However, there is no strict threshold in the distinction between the components on the mere basis of their spatial extend. We rather observe a continuous transition from assigning flux to the diffuse component to assigning it to the point-like component while reaching smaller spatial scales because strict boundaries are blurred out under the consideration of noise effects. The differences between the reconstruction using a MAP and a Gibbs approach are subtle. The difference in the reconstruction formulas given by Eqs.~\eqref{eq:t_MAP} and \eqref{eq:t_Gibbs} is an additive trace term involving $\bb{D}^{(s)}$, which is positive definite. Therefore, a reconstructed power spectrum regularized by uncertainty corrections is never below the one with out given the same $\bb{m}^{(s)}$. However, the reconstruction of the signal field follows different filter formulas, respectively. Since the Gibbs approach considers the uncertainty covariance $\bb{D}^{(s)}$ properly in each cycle, it can present a more conservative solution. The drop-off point is apparently at higher $k$ for the MAP approach, leading to higher power on scales between roughly $0.3$ and $0.7 \,\mathrm{arcmin}^{-1}$. In turn, the MAP solution tends to overfit by absorbing some noise power into $\bb{m}^{(s)}$ as discussed in Sec.~\ref{sec:solution}. Thus, the higher MAP power spectrum in Fig.~\ref{fig:t} seems to be caused by a higher level of noise remnants in the signal estimate. The influence of the choice of the model parameter $\sigma$ is also shown in Fig.~\ref{fig:t}. Neither a smoothness prior with $\sigma = 10$, nor a weak one with $\sigma = 1000$ influences the reconstruction of the power spectrum substantially in this case.\footnote{For a discussion of further log-normal reconstruction scenarios please refer to the work by \citet{OSBE12}.} The latter choice, however, exhibits some more fluctuations in order to better track the concrete realization. The results for the reconstruction of the point-like component are illustrated in Fig.~\ref{fig:u}. Overall, the reconstructed point-like signal field and the corresponding photon flux are in good agreement with the original ones. The point-sources have been located with an accuracy of $\pm 0.1 \,\mathrm{arcmin}$, which is less than the FWHM of the PSF. The localization tends to be more precise for higher flux values because of the higher signal-to-noise ratio. The reconstructed intensities match the simulated ones well, although the MAP solution shows a spread that exceeds the expected shot noise uncertainty interval. This is again an indication of the overfitting known for MAP solutions. Moreover, neither reconstruction shows a bias towards higher or lower fluxes. The uncertainty estimates for the point-like photon flux $\bb{\rho}^{(u)}$ obtained from $\bb{D}^{(u)}$ according to Eqs.~\eqref{eq:var_MAP} and \eqref{eq:var_Gibbs} are, in general, consistent with the deviations from the original and the shot noise uncertainty, cf. Fig.~\ref{fig:u}. They show a reasonable scaling being higher for lower fluxes and vice versa. However, some uncertainties seem to be underestimated. There are different reasons for this. On the one hand, the Hessian approximation for $\bb{D}^{(u)}$ in Eq.~\eqref{eq:D_MAP} or \eqref{eq:D_Gibbs} is in individual cases in so far poor as that the curvature of the considered potential does not describe the uncertainty of the point-like component adequately. The data admittedly constrains the flux intensity of a point source sufficiently, especially if it is a bright one. However, the rather narrow dip in the manifold landscape of the considered potential can be asymmetric, and thus not always well described by the quadratic approximation of Eq.~\eqref{eq:D_MAP} or \eqref{eq:D_Gibbs}, respectively. On the other hand, the approximation leading to vanishing cross-correlation $\bb{D}^{(su)}$, takes away the possibility of communicating uncertainties between diffuse and point-like components. However, omitting the used simplification or incorporating higher order corrections would render the algorithm too computationally expensive. The fact that the Gibbs solution, which takes $\bb{D}^{(u)}$ into account, shows improvements backs up this argument. The reconstructions shown in Fig.~\ref{fig:s} and \ref{fig:u} used the model parameters $\sigma=10$, $\beta = \tfrac{3}{2}$, and $\eta = 10^{-4}$. In order to reflect the influence of the choice of $\sigma$, $\beta$, and $\eta$, Table~\ref{tab:runs} summarizes the results from several reconstructions carried out with varying model parameters. Accordingly, the best parameters seem to be $\sigma = 10$, $\beta = \tfrac{5}{4}$, and $\eta = 10^{-4}$, although we caution that the total error is difficile to determine as the residual errors, $\epsilon^{(s)}$ and $\epsilon^{(u)}$, are defined differently. Although the errors vary significantly, $2$--$15\%$ for $\epsilon^{(s)}$, we like to stress that the model parameters were changed drastically, partly even by orders of magnitude. The impact of the prior clearly exists, but is moderate. We note that the case of $\sigma \rightarrow \infty$ corresponds to neglecting the smoothness prior completely. The $\beta = 1$ case that corresponds to a logarithmically flat prior on $\bb{u}$ showed a tendency to fit more noise features by point-like contributions. In summary, the D$^3$PO algorithm is capable of denoising, deconvolving and decomposing photon observations by reconstructing the diffuse and point-like signal field, and the logarithmic power spectrum of the former. The reconstruction using MAP and Gibbs approaches perform flawlessly, except for a little underestimation of the uncertainty of the point-like component. The MAP approach shows signs of overfitting, but those are not overwhelming. Considering the simplicity of the MAP approach that goes along with a numerically faster performance, this shortcoming seems acceptable. Because of the iterative scheme of the algorithm, a combination of the MAP approach for the signal fields and a Gibbs approach for the power spectrum is possible. \section{Conclusions \& summary} \label{sec:conclusion} The D$^3$PO algorithm for the denoising, deconvolving and decomposing photon observations has been derived. It allows for the simultaneous but individual reconstruction of the diffuse and point-like photon fluxes, as well as the harmonic power spectrum of the diffuse component, from a single data image that is exposed to Poissonian shot noise and effects of the instrument response functions. Moreover, the D$^3$PO algorithm can provide \emph{a~posteriori} uncertainty information on the reconstructed signal fields. With these capabilities, D$^3$PO surpasses previous approaches that address only subsets of these complications. The theoretical foundation is a hierarchical Bayesian parameter model embedded in the framework of IFT. The model comprises \emph{a~priori} assumptions for the signal fields that account for the different statistics and correlations of the morphologically different components. The diffuse photon flux is assumed to obey multivariate log-normal statistics, where the covariance is described by a power spectrum. The power spectrum is \emph{a~priori} unknown and reconstructed from the data along with the signal. Therefore, hyperpriors on the (logarithmic) power spectra have been introduced, including a spectral smoothness prior \citep{EF11,OSBE12}. The point-like photon flux, in contrast, is assumed to factorize spatially in independent inverse-Gamma distributions implying a (regularized) power-law behavior of the amplitudes of the flux. An adequate description of the noise properties in terms of a likelihood, here a Poisson distribution, and the incorporation of all instrumental effects into the response operator renders the denoising and deconvolution task possible. The strength of the proposed approach is the performance of the additional decomposition task, which especially exploits the \emph{a~priori} description of diffuse and point-like. The model comes down to five scalar parameters, for which all \emph{a~priori} defaults can be motivated, and of which none is driving the inference predominantly. We discussed maximum \emph{a~posteriori} (MAP) and Gibbs free energy approaches to solve the inference problem. The derived solutions provide optimal estimators that, in the considered examples, yielded equivalently excellent results. The Gibbs solution slightly outperforms MAP solutions (in terms of the considered L$^2$-residuals) thanks to the full regularization treatment, however, for the price of a computationally more expensive optimization. Which approach is to be preferred in general might depend on the concrete problem at hand and the trade-off between reconstruction precision against computational effort. The performance of the D$^3$PO algorithm has been demonstrated in realistic simulations carried out in 1D and 2D. The implementation relies on the \textsc{NIFTy} package \citep{S+13}, which allows for the application regardless of the underlying position space. In the 2D application example, a high energy observation of a $32 \times 32 \,\mathrm{arcmin}^2$ patch of a simulated sky with a $0.1 \,\mathrm{arcmin}$ resolution has been analyzed. The D$^3$PO algorithm successfully denoised, deconvolved and decomposed the data image. The analysis yielded a detailed reconstruction of the diffuse photon flux and its logarithmic power spectrum, the precise localization of the point sources and accurate determination of their flux intensities, as well as \emph{a~posteriori} estimates of the reconstructed fields. The D$^3$PO algorithm should be applicable to a wide range of inference problems appearing in astronomical imaging and related fields. Concrete applications in high energy astrophysics, for example, the analysis of data from the Chandra X-ray observatory or the Fermi $\gamma$-ray space telescope, are currently considered by the authors. In this regard, the public release of the D$^3$PO code is planned. \section*{Acknowledgments} We thank Niels Oppermann, Henrik Junklewitz and two anonymous referees for the insightful discussions and productive comments. Furthermore, we thank the DFG Forschergruppe 1254 ``Magnetisation of Interstellar and Intergalactic Media: The Prospects of Low-Frequency Radio Observations'' for travel support in order to present this work at their annual meeting in 2013. Some of the results in this publication have been derived using the \textsc{NIFTy} package \citep{S+13}. This research has made use of NASA's Astrophysics Data System. \bibliographystyle{myaa}
\section{Clumping Factors and the Photon Budget for Reionization} \label{sec:ClumpingFactors} \subsection{Clumping Factor Analysis of Madau} \label{Madau} In this section we begin our examination of Equation \eqref{eq:ndot} from \cite{MadauEtAl1999} as an accurate predictor of when reionization completes, focusing on the clumping factor. While it is true that the Madau-type analysis was not designed to predict the precise redshift for reionization completion, only the ionization rate density needed to maintain the IGM in an ionized state after reionization has completed, it is effectively being used in this way when it is applied to galaxy populations at increasingly higher redshifts $z=6-7$ (cf. \cite{FanEtAl2006, RobertsonEtAl2013}). Our methodology is the following. The simulation supplies $\dot{N}_{sim}(z)$ ionizing photons, which increases with decreasing redshift because the SFRD increases with decreasing redshift. Equation \eqref{eq:ndot} poses a minimum requirement on the ionizing emissivity to maintain the IGM in an ionized state at given redshift z. This requirement decreases with decreasing redshift due to the strong z dependence. We look to see if the box becomes fully ionized when these two curves cross; i.e., when $\dot{N}_{sim} \geq \dot{\mathcal{N}}_{ion} $. In subsequent sections we do this for more recent definitions of the clumping factor that have been introduced by various authors, in roughly chronological order. The way the clumping factor is introduced and used, is to estimate the amount of recombination that radiation has to overcome, in order to keep the universe ionized \citep{GnedinOstriker1997,ValageasSilk1999,MadauEtAl1999,FanEtAl2006}. In a homogeneous universe, the hydrogen recombination rate is also homogeneous, and is a simple function of the mean density, ionization fraction, and temperature. The clumping factor is a correction factor to account for density inhomogeneities induced by structure formation, although in principle inhomogeneties in ionization fraction and temperature are also important. The most common definition for the clumping factor is: \begin{equation} C=\frac{\langle n_\mathrm{H\,II}^2 \rangle}{\langle n_\mathrm{H\,II} \rangle^2} \label{eq:clumpingfactor} \end{equation} Where the $\langle\rangle$ brackets denotes an average over the simulation volume. To see where this comes from lets look at the change of $n_\mathrm{H\,II}$ with respect to time due to recombinations: \begin{align} \label{eq:recombtime} \frac{\partial n_\mathrm{H\,II}}{\partial t} &= -n_\mathrm{e} n_\mathrm{H\,II}\alpha_B(T)\notag\\ \frac{\partial n_\mathrm{H\,II}}{n_\mathrm{H\,II}} &= -\partial t n_\mathrm{e}\alpha_B(T)\notag\\ \int^{n_f}_{n_i}\frac{\partial n_\mathrm{H\,II}}{n_\mathrm{H\,II}} &= -\int^{t_f}_{t_i}\partial t n_\mathrm{e}\alpha_B(T)\notag\\ ln\left(\frac{n_f}{n_i}\right) &= -(t_f-t_i)n_\mathrm{e}\alpha_B(T), \notag\\ \frac{n_f}{n_i} &= exp(-t_{rec}n_\mathrm{e}\alpha_B) \end{align} In the last step, we have set $(t_f-t_i)$ to be $t_{rec}$. This leads to \begin{equation} t_{rec} = [n_\mathrm{e}\alpha_B(T)]^{-1} \label{recombtime} \end{equation} being the characteristic time when the fraction $n_f/n_i = 1/e$. Using this expression for the recombination time, one can rewrite the right hand side of the equation as \begin{align} \label{eq:trec} \frac{\partial n_\mathrm{H\,II}}{\partial t} &= - n_\mathrm{H\,II}n_\mathrm{e}\alpha_B(T) = - n_\mathrm{H\,II} / t_{rec}\notag\\ &= - n_\mathrm{H\,II} (1+2\chi) n_\mathrm{H\,II} \alpha_B(T) \notag\\ &= - n_\mathrm{H\,II}^2 (1+2\chi) \alpha_B(T) \notag\\ \end{align} where in the last two steps, following \cite{MadauEtAl1999}, we replace $n_\mathrm{e}$ with $(1+2\chi)n_\mathrm{H\,II}$ assuming helium is fully ionized. Here $\chi$ is the cosmic fraction of helium. Taking the volume average we have: \begin{align} \label{eq:trecpart2} \langle \frac{\partial n_\mathrm{H\,II}}{\partial t} \rangle &= - \langle n_\mathrm{H\,II}^2 (1+2\chi) \alpha_B(T) \rangle \notag\\ &= - \langle n_\mathrm{H\,II}^2 \rangle (1+2\chi) \alpha_B \notag\\ &= - \langle n_\mathrm{H\,II} \rangle^2 (1+2\chi) \alpha_B C\notag\\ &= - \langle n_\mathrm{H\,II} \rangle /\bar{t}_{rec} \end{align} In the above we have made the oft-used assumption of a uniform IGM temperature of $10^4$K, making the Case B recombination coefficient, $\alpha_B$ a constant. Note this is not physically justified, but since the temperature of the IGM is not well determined observationally, it is a useful approximation, and one that is embedded in Equation \eqref{eq:ndot}. With this simplifying assumption, when taking the volume average on both sides of the equation, we may rewrite the result in the same form as the first line in Equation \eqref{eq:trec}. Therefore, the effective recombination time can be written as \begin{equation} \bar{t}_{rec} = t_\mathrm{Madau} \equiv [(1+2\chi)\langle n_\mathrm{H\,II} \rangle \alpha_B C]^{-1} \label{eq:tmadau} \end{equation} This expression is the same as Equation (20) of \cite{MadauEtAl1999} if we substitute $\langle n_\mathrm{H\,II} \rangle$ for $\bar{n}_\mathrm{H}$. In the case of a fully ionized universe these two quantities are equivalent. We note that $t_\mathrm{Madau}$ is not at all the volume average of $t_{rec}$ but is $\langle t_{rec}^{-1} \rangle ^{-1}C^{-1}$, which weights regions with the {\em shortest} recombination times; i.e. regions at the mean density and above. If we now make the {\em ansatz} $\dot{\mathcal{N}}_{ion} \times \bar{t}_{rec} = \bar{n}_\mathrm{H}(0)$, we may derive Equation (26) in \cite{MadauEtAl1999}, updated by \cite{FanEtAl2006}, repeated here for convenience: \begin{equation} \label{eq:updatedNdot} \dot{\mathcal{N}}(z)=10^{51.2}s^{-1}Mpc^{-3}\left(\frac{C}{30}\right)\left(\frac{\Omega_\mathrm{b} h^2}{0.02}\right)^{2}\left(\frac{1+z}{6}\right)^{3}. \end{equation} This equation gives an estimate of the ionizing photon production rate density (in units of s$^{-1}$Mpc$^{-3}$comoving) that is needed to balance the recombination rate density (the right-hand-side of Equation \eqref{eq:updatedNdot}) in a completely ionized universe. Values for $C$ ranging $\sim$10-30 are often quoted from earlier hydrodynamical simulations such as \cite{GnedinOstriker1997}, and $\sim 3$ for more recent work following \cite{PawlikEtAl2009, RaicevicTheuns2011, ShullEtAl2012, FinlatorEtAl2012} and the methods there. \begin{figure} \includegraphics[width=0.5\textwidth]{fig8-eps-converted-to.pdf} \caption{Ionizing photon production rate density and various estimates of the recombination rate density versus redshift. The blue curve labeled ``$\dot{N}_{sim}$'' is the measured photon production rate density averaged over the entire simulation volume. The green curve labeled ``$\dot{R}_\mathrm{H\,II}$'' is the recombination rate density estimate from using the clumping factor calculated with Equation \eqref{eq:clumpingfactor} substituted in Equation \eqref{eq:updatedNdot}. The red curve labeled ``$\dot{R}_b$'' is Equation \eqref{eq:updatedNdot} evaluated using a clumping factor calculated from the baryon density. The black curve labeled ``$\dot{R}_\mathrm{dm}$'' is using a clumping factor calculated with dark matter density.} \label{unthresholded} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig9-eps-converted-to.pdf} \caption{Unthresholded clumping factors used in Fig. \ref{unthresholded}. $C_{HII}, C_b, C_{dm}$ are calculated from the unthresholded H \footnotesize{II}, baryon, and dark matter densities, respectively.} \label{unthreshclumping} \end{figure} We follow these earlier studies using our own simulation data. In Figure \ref{unthresholded} we plot the ionizing photon production rate density and recombination rate density from our fiducial simulation. The curve in blue labeled $\dot{N}_{sim}$ is the photon production rate density from the simulation, calculated using a time average of the volume integrated ionizing emissivity $\eta$ (Equation \eqref{eq:emissivity}) divided by the average energy per photon which we obtain directly from the SED. The other three curves plot Equation \eqref{eq:updatedNdot} for three methods for calculating $C$: green uses the H {\footnotesize II} density directly (Equation \eqref{eq:clumpingfactor}); red uses the baryon density $C=\langle \rho^2_b \rangle / \langle \rho_b \rangle^2$; and black uses the dark matter density $C=\langle \rho^2_{dm} \rangle / \langle \rho_{dm} \rangle^2$. In all cases no thresholding is being applied (the effect of threholding is examined in the next section); the averages are done over every cell in the simulation including those inside the virial radii of galaxies. The H {\footnotesize II} curve drops sharply with decreasing redshift because $C$ is large when the H {\footnotesize II} distribution is patchy. The baryon and dark matter curves track one another for $z > 6$ because the clumping factors are nearly the same, but begin to separate after overlap as the baryon clumping factor drops due to Jeans smoothing. Where the ionization and recombination rate density lines cross is roughly when we expect the universe to become highly ionized. If we define the end of the EoR as when 99.9\% of the volume has reached the Well Ionized level, then our simulation reaches that point around $z\sim5.8$ according to Figure \ref{linearIonized}. The $\dot{N}_{sim}$ curve crosses the $\dot{R}_\mathrm{H\,II}$ curve at $z\sim6.2$. This is somewhat reassuring since we are counting every ionizing photon emitted and every recombination, at lease insofar as Equation \eqref{eq:updatedNdot} provides a good estimate of that. The recombination rate density curves using clumping factors computed from the baryon and dark matter densities curves cross the $\dot{N}_{sim}$ curve at a somewhat higher redshift of $z \approx 6.6$. By following the original methodology of using the clumping factor to estimate recombinations, we find that the clumping factor calculated with the H {\footnotesize II} density field to be the closest predictor for the end of EoR in our simulation. The photon budget that enabled us to reach different levels of ionization is plotted in Figure \ref{unthreshphotonbudget}. Here we plot the evolution of the ionized volume fraction versus $\gamma_{ion}/H=\int dt \dot{N}_{sim} / \bar{n}_\mathrm{H}(0)$. So, for the same definition for the end of EoR, we see that we need $\sim$4 photons per hydrogen atom to achieve. This cannot be considered a converged result because this estimate includes the dense gas inside galaxies, which is not well resolved in our simulation. Even though a small fraction of the baryons reside inside galaxies, due to the short recombination time many ionizing photons are required to keep the gas ionized. Since we have not resolved the internal structure of galaxies, and higher resolution would likely result in higher density gas, we must consider $\gamma_{ion}/H=4$ a lower bound. We eliminate this issue in the next subsection by excluding the dense gas in halos from the calculation. \begin{figure} \includegraphics[width=0.5\textwidth]{fig10-eps-converted-to.pdf} \caption{Ionized volume fraction as a function of the number of ionizing photons emitted per H atom averaged over the entire simulation volume (including inside halos) for three different ionization levels: $f_i \geq 0.1$ (blue line); $f_i \geq 0.999$ (green line); $f_i \geq 0.99999$ (red line). Compare with Fig. \ref{threshphotonbudget} which excludes gas inside halos.} \label{unthreshphotonbudget} \end{figure} \subsection{Quantitative Analysis of Recombinations} As the clumping factor method grew in popularity, various authors have applied thresholds of one form or another to improve upon its accuracy in predicting the recombination rate density needed to maintain an ionized universe. When thresholds are applied, parts of the volume are excluded from the photon counting analysis. \cite{PawlikEtAl2009, RaicevicTheuns2011} and others, limit the calculation of the clumping factor to the low density IGM by using $\Delta_b$ thresholds, usually set at 100. They threshold out gas in virialized halos and the self-shielded collapsed objects, because radiation does not penetrate these objects, or they recombine too fast, which leaves them neutral and not contributing to recombinations in the IGM. More recently \cite{ShullEtAl2012} has also thresholded out void regions ($\Delta_b < 1$), arguing that they do not contribute appreciably to the total recombinations due to their long recombination times. To investigate the contribution of gas of different density to the total recombination rate density, we plot in Figure \ref{recomb}, three quantities dealing with recombinations in our simulation. In the left column we have a 2D distribution plot of recombination rate density $\dot{R} = n_\mathrm{H\,II}n_e\alpha_B(T)$ divided by ionization rate density $\Gamma_\mathrm{H\,I}^{ph}n_\mathrm{H\,I}$ versus baryon overdensity $\Delta_b$, where \begin{align} \label{eq:photoionization} \Gamma_\mathrm{H\,I}^{ph} &= \frac{c E}{h} \left[\int_{\nu_\mathrm{H\,I}}^{\infty} \frac{\sigma_\mathrm{H\,I}(\nu) \chi_E(\nu)}{\nu}\,\mathrm d\nu \right] \bigg / \left[\int_{\nu_\mathrm{H\,I}}^{\infty} \chi_E(\nu)\,\mathrm d\nu\right]. \end{align} Here, $\sigma_\mathrm{H\,I}(\nu)$ and $\nu_\mathrm{H\,I}$ are the ionization cross section and ionization threshold for H {\footnotesize I}, respectively, and $h$ is Planck's constant (Paper I). In the middle column we plot the relative bin contribution to the total recombination rate density versus $\Delta_b$. We draw vertical lines at $\Delta_b$=1 and 100, and in the legend box calculate the cumulative contribution to total reionizations to those thresholds. In the right column, we plot the cell recombination time divided by the Hubble time versus $\Delta_b$. All three columns evolve with descreasing redshift from top to bottom. \begin{figure*}[!tp] \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11a-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11b-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11c-eps-converted-to.pdf} \end{minipage} \\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11d-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11e-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11f-eps-converted-to.pdf} \end{minipage} \\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11g-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11h-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11i-eps-converted-to.pdf} \end{minipage} \\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11j-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11k-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11l-eps-converted-to.pdf} \end{minipage} \\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11m-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11n-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 5mm 8mm 0mm 0mm, clip, width=1.0\textwidth]{fig11o-eps-converted-to.pdf} \end{minipage} \\ \caption{Quantifying recombination information. Left column is a 2D distribution of recombination rate density divided by ionization rate density versus overdensity. Middle column is plot relative bin contribution to the total recombination rate density versus overdensity bins. The lines show the cumulative of all previous bins. Blue line is at $\Delta_b$=100, red line is at $\Delta_b$=1. Right column is plot of recombination time divide by Hubble time versus overdensity. All three columns evolve with decreasing redshift from top to bottom.} \label{recomb} \end{figure*} At $z\sim9$, in the left column of Figure \ref{recomb}, we see that even though there are regions of the volume that are in approximate ionization equilibrium (indicated by the horizontal distribution near 10$^0$), there is a wide distribution of cells far out of equilibrium, some even off by $\sim120$ orders of magnitude. The middle column shows that about 37\% of all recombinations happen below a $\Delta_b$ of 100, and about 3.2\% happen below $\Delta_b$ of 1. The phase diagram in the right column shows that there is a bimodal distribution of cells in terms of their recombination time normalized by Hubble time. The top concentration of cells are more neutral, having long recombination times, and the lower concentration of cells are photoionized, having smaller recombination times. The recombination time is lower for the ionized cells simply because there are more free electrons available to recombine with protons. The blue cloud at low $\Delta_b$ and high $t_{rec}/t_\mathrm{Hubble}$ are the small number of cells that are shock heated to $T >$10$^6$K by supernova feedback. Due to this high temperature, even though there are more free electrons their recombination times remain long. At $z\sim7$, more of the volume has reached the Well Ionized level, and we see the size of the out of equilibrium distribution shrink in the left column. Now the maximum is only $\sim37$ orders of magnitude higher compared to equilibrium. The middle column shows about 40\% of total recombinations are happening below $\Delta_b$ of 100, and about 4.2\% happens below $\Delta_b$ of 1. In the right column, we see roughly equal numbers of cells in the upper (more neutral) distribution as compared to the lower (more ionized) distribution, whereas the top was much greater in numbers before. As more cells become ionized to a high degree, their recombination time will decrease and their cell counts will shift to the lower distribution. At $z\sim6$, looking at the left column, most of the cells are now in equilibrium. This is indicated by the peak of the distribution in red, being near zero on the y-axis. The maximum of the distribution is now less than 19 orders of magnitude apart from equilibrium. The middle column showing 30\% to 3.8\% recombinations below $\Delta_b$ of 100 and 1, respectively. The right column shows that the majority of the cells are now in the more ionized distribution and have a low recombination time. This can be verified by looking at the same redshift in Figure \ref{NeutralPhase}, where most of the cells are at the Well Ionized level compared to fewer before. At $z\sim5.5$, after the entire volume has become Well Ionized, and the vertical spread of the distribution has collapsed to about an order of magnitude away from equilibrium with the vast majority of the cells in equilibrium. The fraction of recombinations are 25\% and 4\% below $\Delta_b$ of 100 and 1, respectively. Looking at the recombination time to Hubble time, we no longer see the bimodal distribution of neutral cells and highly ionized cells, we only see the bottom distribution of highly ionized cells now. The small distribution of shock heated gas is still present, but now seem more prominent with the absence of the neutral distribution. At $z\sim5$, on the left column, the few cells that are in the low density void, which were recombining slower than ionizing are now all near equilibrium. Cells that are higher in $\Delta_b$ are more likely to be above equilibrium. In the middle column, we see the fraction of recombinations are 16\% and 2.9\% for region below $\Delta_b$ of 100 and 1, respectively. Not much has changed in the recombination time column except there are fewer cells above the $\Delta_b$ of 10$^4$, possibly due to effect of Jeans smoothing. We see that there is no real one-to-one correspondence between overdensity and the quantities we show on the y-axis. That is because in a given panel, we are only seeing two dimensions of a multidimensional physical process that depends on locality to sources of radiation, the behavior of said sources at a given moment, the local density of neutral and ionized gas, temperature, among others. It is helpful to speak about the average behavior in any given overdensity as we have done, but we should always keep in mind that the average may not be as representative of the wider distribution as we may think. \subsection{Investigating Thresholded Clumping Factor Analyses} \subsubsection{Excluding Halos} \label{ExcludingHalos} We saw in \S\ref{Madau} that using the unthresholded H {\footnotesize II} density field to calculate $C$ via Equation \eqref{eq:clumpingfactor} yields a reasonably good estimate of when reionization completes (Figure \ref{unthresholded}). This is perhaps not surprising since we count every ionizing photon emitted and every recombination to the accuracy of Equation \eqref{eq:updatedNdot}. Possible sources of disagreement between theory and simulation are: (1) inaccuracies in estimating the recombination rate density using Equation \eqref{eq:updatedNdot}; (2) breakdown of the ``instantaneous approximation'' used to derive Equation \eqref{eq:updatedNdot} due to history-dependent effects; (3) finite propagation time for I-fronts to cross voids; and (4) numerical inaccuracies. Regarding possibility (4) we note that our mathematical formalism is photon conserving, and that our I-front tests in Paper I show that I-fronts propagate at the correct speed, which is an indication that numerical photon conservation is good. To investigate whether improved estimates of the recombination rate density will improve the agreement, we follow the practice of some recent investigators \citep{PawlikEtAl2009, RaicevicTheuns2011} and threshold out dense gas bound to halos, leaving only the diffuse IGM to consider. The motivation for this is that since we are only interested in the photon budget required to maintain the diffuse IGM in an ionized state, by excluding the complicated astrophysics within halos we have a simpler problem to model and resolve numerically. To proceed we must calculate the ionization and recombination rate densities outside of collapsed objects. We estimate the number of ionizing photons escaping halos by multiplying $\dot{N}_{sim}(z)$ by a global escape fraction $\bar{f}_{esc}(z)$ derived in \S\ref{escape} and plotted in Figure \ref{RadEscFraction}: \begin{equation} \dot{N}_{IGM}(z)=\bar{f}_{esc}(z)\dot{N}_{sim}(z) \label{eq:ndot_igm} \end{equation} \noindent The recombination rate density outside of halos is calculated using Equation \eqref{eq:updatedNdot} where now the clumping factor is thresholded such that only cells for which $\Delta_b < 100$ contribute to the sum. As in Figure \ref{unthresholded} we plot three curves for the recombination rate density calculated using Equation \eqref{eq:updatedNdot} using H {\footnotesize II}, baryons, and dark matter density fields. These are plotted in Figure \ref{thresholded} as green, red, and black curves, respectively. We see that the recombination rate density based on the singly thresholded H {\footnotesize II} (labeled $\dot{R}_\mathrm{tH\,II}$) and on the thresholded dark matter (labeled $\dot{R}_\mathrm{tdm}$) curve cross the ionizing emissivity curve labeled ``$\dot{N}_\mathrm{IGM}$'' at $z \approx 6.7$ in Figure \ref{thresholded}, whereas the thresholded baryon density curve (labeled $\dot{R}_{tb}$) crosses ``$\dot{N}_\mathrm{IGM}$'' at $z \sim 7.2$. Taking the doubly-thresholded H {\footnotesize II} curve as the best estimate for the recombination rate density, we find that restricting the analysis to only IGM gas yields poorer agreement than the simpler, global model of Madau, which at first blush is a perplexing result. By thresholding out the gas in galaxies we have isolated the thing we care about: the ionization balance of the IGM. Why then should the implied redshift of reionization completion become worse compared to the analysis in \S\ref{Madau}? We defer addressing this question until later sections. \begin{figure} \includegraphics[width=0.5\textwidth]{fig12-eps-converted-to.pdf} \caption{Same quantities as Figure \ref{unthresholded}, except now the ``$\dot{N}_\mathrm{IGM}$'' curve is the number of ionizing photons which escape into the IGM (see \S\ref{escape}). The recombination rate densities with a subscript that begins with ``t" are calculated as described in the caption for Figure \ref{unthresholded}, except that the clumping factors are computed excluding regions satisfying $\Delta_b > 100$. The curve labelled $\dot{R}_{ttHII}$ is calculated from Equation (22) using the doubly-thresholded clumping factor $C_{ttHII}$ defined in Figure \ref{threshclumping}.} \label{thresholded} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig13-eps-converted-to.pdf} \caption{Thresholded clumping factors used in Fig. \ref{thresholded}. $C_{tHII}, C_{tb}, C_{tdm}$ are calculated using thresholded H II, baryon, and dark matter density fields, respectively, where only cells satisfying $\Delta_b < 100$ contribute. $C_{ttHII}$ is calculated from the H II density where only cells satisfying $\Delta_b < 100$ and $f_i > 0.1$ contribute.} \label{threshclumping} \end{figure} Finally, we ask how many ionizing photons per H atom are required to convert the neutral gas residing outside halos to a well ionized state. We repeat the analysis of Figure \ref{unthreshphotonbudget} and show the result in Figure \ref{threshphotonbudget}. We see that the effect of counting only escaped photons on the photon budget is significant. Previously, we summed $\dot{N}_{sim}(z)$ and divided by the total number of hydrogen atoms in the simulation volume, and used that as our progress variable. In Figure \ref{threshphotonbudget} we sum $\dot{N}_{IGM}(z)$ and divide by the number of hydrogen atoms in the thresholded volume, and use that as our progress variable. Instead of needing $\sim$4 to ionize the IGM, now we only need $\sim$2 photons per hydrogen atom for 99.9\% of the universe to reach Well Ionized level. This result supports the ``photon starved'' reionization scenario discussed by \cite{BoltonHaehnelt2007}. \begin{figure} \includegraphics[width=0.5\textwidth]{fig14-eps-converted-to.pdf} \caption{Ionized volume fraction as a function of the number of ionizing photons emitted per H atom averaged over the entire simulation volume (excluding gas inside halos) for three different ionization levels: $f_i \geq 0.1$ (blue line); $f_i \geq 0.999$ (green line); $f_i \geq 0.99999$ (red line). Compare with Fig. \ref{unthreshphotonbudget} which includes gas inside halos.} \label{threshphotonbudget} \end{figure} \subsubsection{Including Temperature Corrections} \label{IncludingTemperatureCorrections} During the preparation of this paper, a new way of estimating the recombinations in the IGM appeared in the literature. The authors \citep{ShullEtAl2012,FinlatorEtAl2012} reformulated the expression for the clumping factor taking the temperature dependence of the recombination rate into account. We briefly investigate their methods here. In order for the calculation of the clumping factor to take only IGM gas that is ionized but recombining, several additional thresholds were applied. Equation (15) in \cite{ShullEtAl2012} is a new expression for the clumping factor, similar in form to \cite{Gnedin2000}, \begin{equation} C_\mathrm{RR}=\frac{\langle n_e n_\mathrm{H\,II}\alpha_B(T) \rangle}{\langle n_e \rangle \langle n_\mathrm{H\,II} \rangle \langle \alpha_B(T) \rangle} \label{eq:CRR} \end{equation} with the following thresholds applied: 1$<\Delta_b<$100, 300K$<$$T$$<10^5$K, Z$<10^{-6}$Z$_\odot$, $x_e$$>$0.05. Here, Z is metalicity and $x_e$ is the ionized fraction. The reason that a lower limit threshold is applied to the baryon overdensity, the authors argued, is because very little recombinations happen there, due to the low density. \cite{ShullEtAl2012} also provide a new formulation for ionizing photon rate density that uses this definition of the clumping factor, in their Equation (10), \begin{align} \frac{dN}{dt}=4.6\times 10^{50}\mathrm{s}^{-1}\mathrm{Mpc}^{-3}\notag\\ \times \(\frac{(1+z)}{8}\)^3 T_4^{-0.845}\(\frac{C}{3}\) \label{eq:ShullNdot} \end{align} Here, T$_4$ is mean IGM temperature measured in units of 10$^4$K. Equation \eqref{eq:ShullNdot} is proposed as an improvement over Equation \eqref{eq:ndot}. To see if this is the case we used our data to evaluate the clumping factor C$_\mathrm{RR}$ and then used Equation \eqref{eq:ShullNdot} to calculate ionizing photon rate density versus redshift needed to maintain an ionized IGM. The result is shown in Figure \ref{Shull}. The curve labeled $\dot{R}_\mathrm{RR,T4}$ in green uses the average temperature, in units of 10$^4$K, of the region that satisfies the C$_\mathrm{RR}$ thresholds for T$_4$ in Equation \eqref{eq:ShullNdot}. The curve $\dot{R}_\mathrm{RR}$ uses 1 in place of T$_4$ in Equation \eqref{eq:ShullNdot}, essentially fixing the IGM temperature to a constant 10$^4$K. The green curve is lower than the red curve because the average temperature in the simulation is higher than $10^4$K. The blue curve labeled $\dot{N}_{IGM}$ is as defined previously. We see that Equation \eqref{eq:ShullNdot} predicts that reionization completes at significantly higher redshifts than exhibited by the simulation, calling into question the validity of the analysis. We find it curious that as the clumping factor analysis is refined through physically well-motivated modifications, it yields predictions for the redshift of reionization completion that become worse and worse, moving to higher redshift rather than lower redshift. This suggests that there is something fundamentally wrong with the whole approach, and that the seemingly good agreement found in \S\ref{Madau} was fortuitous. One worrisome aspect about the utility of Equation \eqref{eq:ShullNdot} is that the fraction of simulation volume included in the C$_\mathrm{RR}$ thresholds is actually quite small. This is illustrated in Figure \ref{volumefracCRR}. The included volume grows from 3\% at $z=9$ to only 23\% of the simulation volume by overlap. One wonders about the validity of making global statements about reionization based on such a restricted sample of the IGM. It is also unclear how we should interpret the redshift at which lines across in Figure \ref{Shull}. Should we interpret it as the redshift below which an ionization rate given by Equation \eqref{eq:ShullNdot} can keep the whole volume ionized, or only the fraction of the volume satisfying the thresholds? If it is the former, how do we account for the time it takes for I-fronts to cross neutral voids? At this point the reader may rightfully claim that the Madau-type analysis was never meant to predict the precise redshift for reionization completion, only the ionization rate density needed to maintain the IGM in an ionized state after reionization has completed. We would agree with that. However it is effectively being used in this way when it is applied to galaxy populations at increasingly higher redshifts $z=6-7$ (cf. \cite{FanEtAl2006, RobertsonEtAl2013}). Our investigations indicate that formulae such as Equation \eqref{eq:ndot} and \eqref{eq:ShullNdot} are not reliable estimates of when reionization completes. In \S\ref{Discussion} we examine whether they can be usefully applied at lower redshifts, as originally intended. \begin{figure} \includegraphics[width=0.5\textwidth]{fig15-eps-converted-to.pdf} \caption{Ionizing photon injection rate density in the IGM from the simulation $\dot{N}_{IGM}$ versus the predictions of Equation \eqref{eq:ShullNdot}, evaluated with two choices for the clumping factor which take temperature corrections into account. The curve labeled ``$\dot{R}_\mathrm{RR,T4}$'' is from Equation \eqref{eq:ShullNdot}, with T$_4$ being the average temperature in C$_\mathrm{RR}$ region in units of 10$^4$K. The curve ``$\dot{R}_\mathrm{RR}$'' is calculated the same way as $\dot{R}_\mathrm{RR,T4}$ except now T$_4$ is set to 1 in Equation \eqref{eq:ShullNdot}, for an effective IGM temperature of 10$^4$K.} \label{Shull} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig16-eps-converted-to.pdf} \caption{Evolution of the volume filling fraction with redshift of regions satisfying the C$_\mathrm{RR}$ thresholding criteria.} \label{volumefracCRR} \end{figure} \subsection{Comparing Clumping Factors} \label{ClumpingFactorEvolution} For ease of comparison we collect into one plot all the H {\footnotesize II} clumping factors used in the previous sections. The unthresholded H {\footnotesize II} calculated using Equation \eqref{eq:clumpingfactor} is denoted C$_\mathrm{H\,II}$. The singly thresholded clumping factor is denoted C$_\mathrm{tH\,II}$, in which the threshold $\Delta_b<100$ is being applied. The curve labeled C$_\mathrm{RR}$ plots the evolution of Equation \eqref{eq:CRR} with the following thresholds: 1$<\Delta_b<$100, 300K$<$$T$$<10^5$K, $x_e$$>$0.05. For comparison we also plot a doubly thresholded H {\footnotesize II} clumping factor denoted C$_\mathrm{ttH\,II}$ with thresholds $\Delta_b<100$ and $x_e>0.05$, which can be thought of as the clumping factor inside H {\footnotesize II} regions excluding the dense gas in halos. We see a clear trend that as more thresholds are applied the lower the value of the clumping factor goes. This is because as more regions of the volume are excluded from the averaging process the remaining regions are more homogeneous exhibiting less variations. If no thresholds are applied, the H {\footnotesize II} clumping factor starts around 200 at $z\sim9$ (Figure \ref{unthresholded}). Such high values arise because when the first couple of ionizing sources created high H {\footnotesize II}, they are localized and spread far apart, making the H {\footnotesize II} density very clumpy. As more of the universe is ionized, the H {\footnotesize II} density becomes more homogeneous. We see the single and double thresholded H {\footnotesize II} clumping factors become the same after overlap with a value of $\sim 4.5$ because the second threshold $x_e>0.05$ is satisfied everywhere. The clumping factor that is not based on the H {\footnotesize II} density alone is C$_\mathrm{RR}$. We see from Equation \eqref{eq:CRR}, C$_\mathrm{RR}$ depends on electron number density, H {\footnotesize II} number density, and the case B hydrogen recombinationation coefficient $\alpha_B(T)$, which is itself dependent on the gas temperature T (fit to Table 2.7 in \cite{OsterbrockFerland2006} implemented in Enzo). $\alpha_B(T)$ depends on T to a negative power and this causes Equation \eqref{eq:CRR} to sometimes have a very low numerator compared to the denominator. This as well as the exclusion of gas in the voids leads to the low clumping factor value of $\sim 2$ we see in the graph. It is very possible to have a value that is smaller than unity, which can lead to even more confusion with the original definition of the clumping factor in Equation \eqref{eq:clumpingfactor}. There, the clumping factor can only have a value of greater than 1, and 1 occurs only in the case of homogeneous distribution of the gas number density. \begin{figure} \includegraphics[width=0.5\textwidth]{fig17-eps-converted-to.pdf} \caption{Various clumping factors versus redshift. C$_\mathrm{H\,II}$ is Equation \eqref{eq:clumpingfactor} used in $\dot{R}_\mathrm{H\,II}$ curve in Figure \ref{unthresholded}, C$_\mathrm{tH\,II}$ is used in $\dot{R}_\mathrm{tH\,II}$ curve in Figure \ref{thresholded}, C$_\mathrm{ttH\,II}$ is clumping factor with two thresholds applied, $\Delta_b < 100$ and $f_i>0.1$, shown here solely for comparison. C$_\mathrm{RR}$ is the value of recombination rate clumping factor from Equation \eqref{eq:CRR} with the 5 thresholds applied.} \label{ClumpingFactors} \end{figure} \section{Summary and Conclusions} \label{Conclusions} We now summarize our main results. \begin{enumerate} \item We use a fully self-consistent simulation including self-gravity, dark matter dynamics, cosmological hydrodynamics, chemical ionization and flux limited diffusion radiation transport, to look at the epoch of hydrogen reionization in detail. By tuning our star formation recipe to approximately match the observed high redshift star formation rate density and galaxy luminosity function, we have created a fully coupled radiation hydrodynamical realization of hydrogen reionization which begins to ionize at $z \approx 10$ and completes at $z \approx 5.8$ without further tuning. While our goal is not the detailed prediction of the redshift of ionization completion, the simulation is realistic enough to analyze in detail the role of recombinations in the clumpy IGM on the progress of reionization. \item We find that roughly 2 ionizing photons per H atom are required to convert the neutral IGM to a well ionized state ($f_i>0.999$), which supports the ``photon starved'' reionization scenario discussed by \cite{BoltonHaehnelt2007}. \item Reionization proceeds initially ``inside-out", meaning that regions of higher mean density ionize first, consistent with previous studies. However the late stages of reionization are better characterized as ``outside-in" as isolated neutral islands are swept over by externally driven I-fronts. Intermediate stages of reionization exhibit both characteristics as I-fronts propagate from dense regions to voids to filaments of moderate overdensity. In general, the appropriateness of a given descriptor depends on the level of ionization of the gas, and the reionization process is rather more complicated that these simple descriptions imply. \item The evolution of the ionized volume fraction with time $Q_{H {\footnotesize II}~}(z)$ depends on the level of ionization chosen to define a parcel of gas as ionized. The curves for ionization fractions $f_i = 0.1$ and $f_i =0.999$ are very similar, but the curve for $f_i =0.99999$ is significantly lower at a given redshift, amounting to a delay of $\Delta z \approx 1$ relative to the other curves for $Q_{H {\footnotesize II}~} \ll 1$, smoothly decreasing to 0 as the redshift of overlap is approached. \item Before overlap, 30-40\% of the total recombinations occur outside halos in our simulation, where this refers to gas with $\Delta_b < 100$. After overlap, this fraction decreases to 20\% and continues to decrease to lower redshifts. \item Before and after overlap, 3-4\% of the total recombinations occur in voids (defined as $\Delta_b < 1$.) While this is a small fraction of all recombinations, it is about 10\% of the recombinations in the IGM before overlap, increasing to about 20\% by $z=5$. The contribution of voids to the ionization balance of the IGM is therefore not negligible. \item The formula for the ionizing photon production rate needed to maintain the IGM in an ionized state derived by \cite{MadauEtAl1999} (Eq. \ref{eq:ndot}) should not be used to predict the epoch of reionization completion because it ignores history-dependent terms in the global ionization balance which are not ignorable. While not originally intended for this purpose, it is being used by observers to assess whether increasingly higher redshift populations of star forming galaxies can account for the ionized state of the IGM. A direct application of the formula to our simulation predicts an overlap redshift of $z=7.4$ compared to the actual value of $z=5.8$. \item Estimating the recombination rate density in the IGM before overlap through the use of clumping factors based on density alone is unreliable because it ignores large variations in local ionization state and temperature which increase the effective recombination time compared to density-based estimates. For a currently popular value of the clumping factor $C=3$ \citep{ShullEtAl2012}, the formula for $\bar{t}_{rec}$ from \cite{MadauEtAl1999}(Eq. \ref{eq:tmadau}) understimates by $2\times$ at all redshifts the effective recombination time measured directly from the simulation. If we adjust $C$ downward so that Eq. \ref{eq:tmadau} matches $t_{rec,eff}$ from the simulation, then it is too low by 60\% at $z=6$ due to the aforementioned effects. \item The assumption that $\bar{t}_{rec}/t \ll 1$ which underlies the derivation of Eq. \ref{eq:ndot} is never valid over the range of reionization redshifts explored by our simulation (Fig. \ref{treceffhubble}). Depending on how $\bar{t}_{rec}$ is evaluated, $\bar{t}_{rec}/t$ increases from $0.3-0.4$ at $z=9.7$ to $\geq 1$ at overlap. This means that an instantaneous analysis of the ionization balance in the IGM post overlap is invalid because recombination times are so long. \item Retaining time-dependent effects is important for the creation of analytic models of global reionization. The analytic model for the evolution of $Q_{H {\footnotesize II}~}$ introduced by \cite{MadauEtAl1999}(Eq. \ref{eq:dQdt}) retains important time-dependent effects, and predicts well the shape of our simulated curve, but overpredicts $Q_{H {\footnotesize II}~}$ at all redshifts because it does not take into account that reionization begins in overdense regions consistent with the inside-out paradigm. It also assumes every emitted ionizing photon results in a prompt photoionization, which is not true in our simulation at late times $Q_{H {\footnotesize II}~}>0.5$. The Madau model, which ignores these effects, predicts a universe which reionizes too soon by $\Delta z \approx 1$. When we introduce correction factors for these effects into Eq. \ref{eq:dQdtdbg} the simulation and model curves agree to approximately 1\% accuracy. We recommend researchers use Eq. \ref{eq:dQdtdbg} for future analytic studies of reionization. \item Finally, we present in Figs. \ref{deltabvsQfit5}, \ref{treceffvszfit}, and \ref{RatiovsQfit} fitting functions for the overdensity correction $\delta_b(Q)$, the effective recombination time derived from our simulation, and the ionization efficiency parameter $\gamma(Q)$ which may be useful for other researchers in the field. \end{enumerate} This research was partially supported by National Science Foundation grants AST-0808184 and AST-1109243 and Department of Energy INCITE award AST025 to MLN and DRR. Simulations were performed on the {\em Kraken} supercomputer operated for the Extreme Science and Engineering Discovery Environment (XSEDE) by the National Institute for Computational Science (NICS), ORNL with support from XRAC allocation MCA-TG98N020 to MLN. MLN, DRR and GS would like to especially acknowledge the tireless devotion to this project by our co-author Robert Harkness who passed away shortly before this manuscript was completed. \section{Discussion} \label{Discussion} \subsection{Significance of our Main Results} We have carried out a fully-coupled radiation hydrodynamic cosmological simulation of hydrogen reionization by stellar sources using an efficient flux-limited diffusion radiation transport solver coupled to the Enzo code (Paper I). This method has the virtue of a high degree of scalability with respect to the number of sources, which allows us to simulate reionization in large cosmological volumes including hydrodynamic and radiative feedback effects self-consistently. In this paper we have presented first results from a simulation in a cosmological volume of modest size--20 Mpc comoving--to investigate the detailed radiative transfer, nonequilibrium photoionization, photoheating and recombination processes that operate during reionization and dictate its progress. In a future paper we apply our method to larger volumes to examine the large scale structure of reionization, evolution of the bubble size distribution, etc. The simulation presented here is carried out on a uniform mesh of $800^3$ cells and with an equivalent number of dark matter particles. As such, the mass resolution is sufficiently high to evolve a dark matter halo population which is complete down to ($M_{halo} \approx 10^8 M_{\odot}$) which cools via H and He atomic lines. However, a spatial resolution of 25 kpc comoving poorly resolves internal processes within early galaxies, but does an excellent job of resolving the Jeans length in the photoionized IGM \citep{BryanEtAl1999}. Our simulation is most appropriately thought of as a high redshift IGM simulation which evolves an inhomogeneous ionizing radiation field sourced by star-forming early galaxies. Star formation is modeled using a modified version of the Cen \& Ostriker (1992) recipe that can be tuned to reproduce the observed star formation rate density (SFRD) \citep{SmithEtAl2011}. We have tuned our simulation to roughly match the observed SFRD \citep{BouwensEtAl2011,RobertsonEtAl2013} for $z\geq 7$, but due to the small boxsize, it somewhat underpredicts the SFRD for $z < 7$. Our simulation also matches the observed $z=6$ galaxy luminosity function well, which gives us some confidence that our ionizing souce population is representative of the real universe. However a substantial fraction of our ionizing flux comes from sources that are too faint to be observed; we defer a discussion of this topic to Paper III in this series (So et al., {\em in prep.}) Our goal was not to predict the precise redshift of ionization completion, as this would depend on details such as escape fraction of ionizing radiation from galaxies and their stellar populations that we do not model directly. Rather our goal was to examine the mechanics of reionization in its early, intermediate, and late phases within a model which is calibrated to the observed source population. Nonetheless, we present a model in which reionization completes at $z\approx 6$, consistent with observations. At early and intermediate times we find that reionization proceeds ``inside-out", confirming the results of many previous investigations \citep{Gnedin2000,RazoumovEtAl2002,SokasianEtAl2003,FurlanettoEtAl2004,IlievEtAl2006,TracCen2007,TracEtAl2008}. However, at late times isolated islands of neutral gas are ionized from the outside-in as they have no internal sources of ionization. Even this characterization is somewhat oversimplified when {\em degree of ionization} is considered, as we discussed in Sec. \ref{IOOI}. It accurately depicts how reionization proceeds for a low degree of ionization (> 0). However for high degrees of ionization, ``inside-out-middle" is more appropriate, as filaments lag behind low and high density regions, as discussed by \cite{FinlatorEtAl2009}. Our most interesting findings concerns the widely used analytic model of reionization introduced by \cite{MadauEtAl1999}. Both the instantaneous (Equation \ref{eq:ndot}) and time-dependent (Equation \ref{eq:dQdt}) versions of this model underpredict the time (overpredict the redshift) when reionization completes, when applied to our simulation. There are two reasons for this having to do with the detailed mechanics of reionization at early and late times respectively. At early times, I-fronts are propagating in regions of higher density than the cosmic mean since the first sources are highly biased. Higher densities translate into slower bubble expansion rates, retarding $Q_{H {\footnotesize II}~}(z)$ relative to a solution which assumes the cosmic mean density (Figure \ref{Qeffv2}). At late times, which we loosely define as $Q_{H {\footnotesize II}~} > 0.5$, conversion of ionizing photons into new ionized hydrogen atoms becomes inefficient. This can be seen by forming this ratio directly from the simulation data (Figure \ref{Ndot_Ratio}), or by defining a global H {\footnotesize I}~ ionization parameter (Equation \eqref{eq:IP} and Figure \ref{IP}). The consequence of this dropping ionization efficiency, which is as low as 0.05 at overlap in our simulation, is to further retard $Q_{H {\footnotesize II}~}(z)$ relative to a solution which assumes an ionization efficiency of unity (Figure \ref{Qeffv3}). We have introduced a modified version of \cite{MadauEtAl1999}'s time-dependent analytic reionization model in Equation \eqref{eq:dQdtdbg}. Modifications which correct for the above-mentioned effects apply to the source term only, {\em not to the recombination term}. These corrections are therefore totally independent of issues like clumping factors and the temperature of the IGM, which enter into the characteristic recombination time of the IGM. The modifications are introduced as correction factors to the mean density of baryons in the vicinity of ionizing sources at early times ($\delta_b$), and the conversion efficiency of ionizing photons emitted to H {\footnotesize I}~ photoionization rate at late times ($\gamma$). Fits of these two correction factors versus $Q_{H {\footnotesize II}~}$ are presented in Figures \ref{deltabvsQfit5} and \ref{RatiovsQfit} for consumption by other researchers. At this point we do not know how general these results are. However we have indications based on another simulation we have analyzed with a softer source SED that the functional forms are representative of this class of reionization model. The significance of these results to high redshift galaxy observers is the following. Setting $Q_{H {\footnotesize II}~} = 1$ and $\delta_b = 1$ in Equation \eqref{eq:dQdtdbg}, we derive \begin{equation} \dot{n}_{ion} = \frac{1}{\gamma} \frac{\bar{n}_H}{\bar{t}_{rec}}. \label{eq:ndotgamma} \end{equation} This differs from the usual expression used to assess whether a given ionizing photon injection rate can maintain an ionized IGM by the factor $1/\gamma$, which is a factor of $\sim 20$ at overlap in our simulation. If this result is correct, then it means that the required UV luminosity density to maintain an ionized IGM has been underestimated by a factor of approximately 20. However, a more precise statement would be that the UV luminosity density required to maintain the IGM in a {\em highly ionized state; $f_n =10^{-5}$} is 20 times higher than what has been previously estimated. Lower levels of UV luminosity density than that specified in Equation \eqref{eq:ndotgamma} could still maintain the IGM in an ionized state, but one with a higher neutral fraction. As we showed in Figure \ref{treceffhubble}, the effective recombination time at and after overlap in our model is comparable to the Hubble time, whether we use the Madau formula to evaluate it for reasonable values for the clumping factor, or we evaluate it directly from our simulation data. This fact casts in doubt the entire instantaneous photon counting argument which is the basis of Equation \ref{eq:ndot}, and the equation becomes less useful for the purposes to which it has been applied (e.g., Robertson et al. 2013). It means that the ionization state of the IGM has a memory on the timescale of $\bar{t}_{rec}$ which is always a significant fraction of $t_{Hubble}$ before overlap, and of order the Hubble time after overlap. We therefore recommend observers use the time-dependent version Equation \eqref{eq:dQdtdbg} in future assessments of high redshift galaxy populations and their role in reionization. \subsection{Limitations of the Simulation} We conclude this section with a brief discussion of the known limitations of our simulation and a comparison of our results with others in the published literature. First the limitations. The principal limitation is the use of a uniform grid, which prevents us from resolving processes occuring inside galaxy halos. The main defect this introduces is an inability to calculate the ionizing escape fraction directly, as is done in some high resolution simulations; e.g., \cite{WiseCen2009,FernandezShull2011}. In our simulation, we calibrate our star formation recipe to match the observed SFRD, and then use that that to calculate UV feedback cell-by-cell via Equation \eqref{eq:emissivity}. We use a value for $\epsilon_{UV}$ taken from \cite{RicottiEtAl2002} for an unattenuated low metallicity stellar population. We underestimate the amount of internal attenuation of ionizing flux due to our limited resolution within halos, and we do not incorporate an explicit escape fraction parameter in Equation \eqref{eq:emissivity}. Effectively, we assume $f_{esc}(ISM)=1$. Using a lower value for $f_{esc}$ would result in a lower overlap redshift \citep{PetkovaSpringel2011a}. Clearly, it would be desirable to vary this parameter in future studies. A second limitation of our simulation is that we have presented only one realization in a relatively small box. Previous studies have shown that H {\footnotesize II}~ bubbles reach a characteristic size of $\sim 10$ Mpc comoving in the lates stages of reionization \citep{FurlanettoEtAl2004,ZahnEtAl2007,ShinEtAl2008}. At 20 Mpc on a side, our box is scarcely larger than this. Therefore one can ask how robust our results are to boxsize. We have addressed this by carrying out a simulation of identical physics, spatial, and mass resolution in a volume 64 times as large as the one described in this paper. The simulation is carried out in a box 80 Mpc on a side on a uniform mesh of $3200^3$ cells, and with an equivalent number of dark matter particles. Results of this simulation will be presented in a forthcoming paper (So et al., in preparation). For the present we merely state that the $Q_{H {\footnotesize II}~}(z)$ curve for the $800^3$ simulation falls within the $\pm 1 \sigma$ band for the larger simulation, where this band is obtained by subdividing the large simulation into 64 cubes of size 20 Mpc on a side, and calculating the mean and standard deviation. While the larger box begins to ionize at a slightly earlier redshift, due to the presence of higher sigma peaks forming galaxies, both simulations complete reionization at the same redshift, $z_{reion} = 5.8$. The $Q_{H {\footnotesize II}~}(z)$ curve for the $800^3$ simulation is near the lower edge of the band, which means that at intermediate redshifts ($7 \leq z \leq 8$), where the difference is largest, the small box simulation underestimates the fraction of the volume that is ionized by about 20\%, with differences smoothly decreasing to lower and higher redshift. A third limitation is that our SFRD systematically deviates from observations below $z \sim 7$, flattening and then decreasing slightly, rather than continuing to rise (Figure \ref{SFR}). The large box simulation does not show this effect, but rather tracks the observed SFRD over the entire range of redshifts. The difference in the mean SFRD between the large and small box simulations increases smoothly from 0.1 dex at $z=9$ to 0.3 dex at $z=6$. The higher levels of star formation in the large box simulation accounts for the higher ionized volume fraction at intermediate redshifts. Nonetheless, the two simulations complete reionization at virtually the same redshift, which is a curious result which we address in a subsequent paper. Another limitation of our method is the use of flux-limited diffusion (FLD) to transport radiation. It is well known that FLD does not cast shadows behind opaque blobs. This could potentially overestimate how rapidly the IGM ionizes, and hence overestimate $z_{reion}$. In Paper I we showed through a direct comparison between FLD and an adaptive ray tracing method incorporated in the {\em Enzo} code on a standard test problem that the differences in the volume- and mass-weighted ionized volume fraction are small. This was for a rather small volume with a small number of ionizing sources. The differences will likely be even smaller as larger volumes containing larger numbers of sources are considered. At the present time, no fully-coupled radiation hydrodynamic simulations of reionization using ray tracing in large volumes are available to compare our method against, to confirm or deny this conjecture. \subsection{Comparison with Other Self-Consistent Simulations} Finally, we compare our results to the results of several recent fully-coupled simulations of reionization including hydrodynamics, star formation, and radiative transfer. \cite{PetkovaSpringel2011a} simulated a (10 Mpc/h)$^3$ volume with the {\tt Gadget-2} code coupled to a variable tensor Eddington factor moment method for the ionizing radiation field sourced by star forming galaxies. They carried out a suite of simulations with $2 \times 128^3$ gas and dark matter particles, varying the ionizing escape fraction and the mean energy per photon from hot, young stars. The also performed one simulation at $2 \times 256^3$ resolution to check for convergence. Our simulation has 80/10 times superior mass resolution as their $128^3/256^3$ simulations. Because {\tt Gadget} is a Lagrangian code, our Eulerian simulation has 8/16 times lower resolution in the highest density regions, but 4.46/2.23 times higher resolution at mean density, and even higher resolution compared to the {\tt Gadget} simulations in low density voids. Our method also has a more accurate adaptive subcycling timestepping scheme for the coupled radiation-ionization-energy equations, obviating the need to model nonequilibrium effects by means of a gas heating parameter $\epsilon$. Morphologically, our results are qualitatively similar, as are the neutral hydrogen fraction versus overdensity phase diagrams. As might be expected from the two methods, the phase diagrams show some differences at the highest and lowest overdensities which is likely a resolution effect. The SFRD in the \cite{PetkovaSpringel2011a} simulation is about an order of magnitude higher than observed, making a direct comparison on $Q_{H {\footnotesize II}~}(z)$ somewhat problematic. However, since they vary the ionizing escape fraction, we can roughly compare their $f_{esc}=0.1$ case with our results. Their model completes reionization at $z \approx 5$ compared to our own which completes at $z \approx 5.8$. They plot the quantity $log[1-Q_{H {\footnotesize II}~}(z)]$, which makes the end of reionization look abrupt. We plot $Q_{H {\footnotesize II}~}(z)$, which makes the end of reionization look slow. When we plot $log[1-Q_{H {\footnotesize II}~}(z)]$ using our data, it looks very similar to their curves, and shows a rapid plunge in the average neutral fraction at late times. \cite{PetkovaSpringel2011a} do not compare with the predictions of the \cite{MadauEtAl1999} model, nor do they investigate the evolution of clumping factors, recombination times, or the number of photons per H atom to achieve overlap as we do. We do not investigate the properties of the $z=3$ IGM via Lyman $\alpha$ forest statistics, as they do. Therefore further comparisons are not possible at this time. \cite{FinlatorEtAl2012} examined some of the same issues we have, hence a comparison with their results is informative. They carried out a suite of {\tt Gadget-2} simulations in small volumes (3, 6)Mpc/h coupled to a variable tensor Eddington factor moment method. Unlike \cite{PetkovaSpringel2011a}, the radiation transport is solved on a uniform Cartesian grid, rather than evaluated using the SPH formalism. The results presented in \cite{FinlatorEtAl2012} use $2 \times 256^3$ dark matter and gas particles, which given their small volumes, yields a similar mass resolution to our simulation, superior spatial resolution in high density regions, and slightly coarser spatial resolution at mean density and below. However, their radiation transport is done on coarse $16^3$ mesh, which in their fiducial run is $536$ comoving kpc $\approx 20 \times$ as coarse as ours. Their simulation thus coarse-grains the radiation field relative to the density field, which necessitates the introduction of a sub (radiation) grid model for unresolved self-shielded gas (i.e., Lyman limit systems). The effect of their subgrid model is to remove some gas in the overdensity regime $1 \leq \Delta_b \leq 50$ in the calculation of the H {\footnotesize II}~ clumping factor, thereby lowering it. Since our radiation field is evolved on the same grid as the density field, we have not included an explicit subgrid model for unresolved self-shielded gas. Lyman limit systems, with neutral column densities of $\sim 10^{17}$ cm$^{-2}$, have a characteristic size of 10 physical kpc \citep{Schaye2001,McQuinnEtAl2011}. At $z=6$ this is 70 comoving kpc, which is resolved by 3 grid cells in our simulation. While this is lower than one would ideally like (5-10 cells), we believe we can make an apples-to-apples comparison between our resolution-matched simulation results and Finlator et al.'s results. Our results are in broad agreement with those of \cite{FinlatorEtAl2012}, with some minor quantitative differences. We both find that the unthresholded baryon clumping factor $C_b$ significantly overestimates the clumping in ionized gas at redshifts approaching overlap, and therefore that it should not be used to estimate the mean recombination rate in the IGM. We confirm their findings that properly accounting for the ionization state and temperature of gas of moderate overdensities lowers the clumping factor to less than $\approx 6$ (in our case less than 5; see Figure \ref{ClumpingFactors}). Finlator et al. quote a value for $C_{H {\footnotesize II}~}$ of 4.9 at $z=6$ taking self-shielding into account, which is in good agreement with our value of $C_{ttH {\footnotesize II}~} \approx 4.8$. However, they favor a lower value for $C$ of 2.7-3.3 taking temperature corrections into account. This can be compared with our value for $C_{RR} \approx 2.3$, which includes temperature corrections but also excludes gas with $\Delta_b<1$. Including this low density gas, as Finlator et al. do, would raise this value somewhat since a larger range of densities enter into the average. We conclude therefore that clumping factors derived from our simulation are in good agreement with those reported by \cite{FinlatorEtAl2012}. We find that approximately 2 photons per hydrogen atom ($\gamma/H\approx 2$) are required to reionize gas satisfying $\Delta_b<100$--our proxy for the fluctuating IGM. \cite{FinlatorEtAl2012} quote a model-dependent value for $\gamma/H$ which depends on the redshift at which the IGM becomes photoheated and thereby Jeans smoothed (their Fig. 7). For $z=6$, $\gamma/H \approx 5$, significantly higher than our number evaluated directly from the simulation. However, for $z=8$, when our box is already significantly ionized, $\gamma/H \approx 3$. Because there are many model-dependent assumptions that go into the Finlator et al. estimate, we consider this reasonably good agreement. However we point out that our estimate is the first to be derived from a self-consistent simulation of reionization with no subgrid models aside from the star formation/radiative feedback recipe. Finally, \cite{FinlatorEtAl2012} compare $Q_{H {\footnotesize I}~}(z)=1-Q_{H {\footnotesize II}~}(z)$ for their fiducial model with the time-dependent model of \cite{MadauEtAl1999}. They point out the sensitivity of the redshift of overlap on the choice of clumping factor, which enters into the recombination time, and showed that $C_{H {\footnotesize II}~}$ provides better agreement with theory at early times than $C_b$, consistent with our findings. Since small discrepancies in $Q_{H {\footnotesize II}~}(z)$ at early times are masked by plotting $Q_{H {\footnotesize I}~}(z)$, Finlator et al. did not discover the need for our overdensity correction $\delta_b$. Similar to us, they found that even with the best clumping factor estimate the analytic model predicts that reionization completes earlier than the simulation by $\Delta z \approx 1$. They ascribe this delay to finite speed-of-light effects (which can only account for $\Delta z =0.1$), while we ascribe it to nonequilibrium ionization effects. \cite{FinlatorEtAl2012} did not propose modifications to the \cite{MadauEtAl1999} model to improve agreement with simulation, as we do in Equation \eqref{eq:dQdtdbg}. \section{A Global Estimate for Circumgalactic Absorption of Ionizing Radiation} \label{escape} The ionizing escape fraction from galaxies is an important parameter in models of reionization. Typically, one thinks about the escape fraction as a property of individual galaxies, determined by the absorption of ionizing radiation on small scales in the ISM. However it is interesting to ask whether there is significant absorption in the denser Circumgalactic Medium (CGM) surrounding galaxies. If we write the total escape fraction as the product of escape fractions, then $f_{esc}=f_{esc}(ISM)f_{esc}(CGM)$. Here we use our simulation to derive an estimate of the globally averaged escape fraction as a function of redshift due to the circumgalactic medium $\bar{f}_{esc}(CGM)$. Recall from Sec. \S\ref{Method} that the halo escape fraction is not a model input parameter, but is rather an ouput since the equation of radiative transfer is solved throughout the computational domain. Our halos are not well resolved internally, and so we are underestimating the amount absorption of ionizing radiation on galaxy ISM scales. However if significant absorption occurs on scales of the virial radius or larger, then that would be simulated reasonably accurately. In the following we assume this is the case, and present results that can be taken to be an upper limit on the total escape fraction (ISM+CGM). Rather than measure the escape fraction halo by halo and take the average over all halos, we use a simpler method. Since we know every ionization requires an ionizing photon, and we have the ionization rate density as a field defined at every grid cell, then we can estimate $\bar{f}_{esc}(CGM)$ as follows (hereafter we drop the CGM modifier with the reader's understanding that this is what we are estimating): \begin{equation} \bar{f}_{esc}(I_t) = \int_{V_t} n_\mathrm{H\,I}\Gamma_\mathrm{H\,I}^{ph} d^3x \bigg / \int_{V} n_\mathrm{H\,I}\Gamma_\mathrm{H\,I}^{ph} d^3x , \label{eq:fesc} \end{equation} \\where $\Gamma_\mathrm{H\,I}^{ph}$ is evaluated cell by cell via Equation \eqref{eq:photoionization}, $V$ is the simulation volume and $V_t$ denotes the integration includes only cells which satisfy $\Delta_b < 100.$ In other words, $\bar{f}_{esc}$ is the ratio of the number of ionizations in the IGM, as defined by the overdensity threshold, to the total number of ionizations in the volume. The modifier $I_t$ refers to this method of estimating $\bar{f}_{esc}$ (a superior method is presented below). \begin{figure} \includegraphics[width=0.5\textwidth]{fig18-eps-converted-to.pdf} \caption{Estimate of the globally averaged ionizing radiation escape fraction due to circumgalactic absorption $\bar{f}_{esc}(I_t)$ computed as the ratio of the volume integrated ionization rate in the IGM ($\Delta_b < 100$) divided by the total ionization rate (Eq. \eqref{eq:fesc}). } \label{EscFraction} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig19-eps-converted-to.pdf} \caption{Evolution of the volume averaged rate densities for: (1) ionizing photons injected into the IGM ($\dot{N}_\mathrm{IGM}$), (2) gas photoionization ($\dot{N}_t$), and (3) gas recombination ($\dot{R}_t$) integrated over the singly thresholded volume $V_t$ defined as $\Delta_b<100$. The ionization rate density curve tracks the photon injection rate density curve in the photon starved regime at high redshifts, but begins to fall below it as the globally averaged ionization parameter approaches unity (Fig. \ref{IP}). After overlap, in the photon abundant regime, the ionization rate density is $\sim 20\times$ the photon injection rate density, but comes into balance with the recombination rate density.} \label{sanitycheckrate} \end{figure} The result is plotted in Fig. \ref{EscFraction}. At high redshifts the escape fraction is high and relatively constant at $\bar{f}_{esc} \sim 0.65-0.7$. As overlap is approached $\bar{f}_{esc}$ drops considerably, reaching values of $\sim 0.2$ by $z=5.$ There is no obvious reason why the escape fraction should drop so dramatically at the epoch of overlap. To investigate this properly would require a statisical analysis of individual halo escape fractions, which we defer to a subsequent paper. Perhaps this is an artifact of how we are estimating $\bar{f}_{esc}$. While it is true that every ionization requires and ionizing photon in the photon starved regime (i.e., before overlap), after overlap the volume becomes optically thin to ionizing radiation, and it is not true that every ionizing photon causes an ionization in the box. This is illustrated in Fig. \ref{sanitycheckrate}. The curve labeled $\dot{N}_t$ is the actual ionization rate density measured in the simulation averaged over the entire 20 Mpc cubic volume satisfying the overdensity threshold $\Delta_b < 100$; i.e. precisely the numerator of Eq. \eqref{eq:fesc} divided by 20$^3$. The curve labeled $\dot{R}_t$ is the recombination rate density averaged over the same volume; i.e. \begin{equation} \dot{R}_t = \int_{V_t} n_e n_\mathrm{H\,II}\alpha_B(T) d^3x . \label{eq:totrecombst} \end{equation} We see that ionization rate density $\dot{N}_t$ grows with redshift and reaches a maximum at $z \approx 6.5$, and then drops by roughly 0.8 dex by overlap completion at $z=5.8$. It continues to decrease thereafter. The reason for this sudden drop is that after overlap there are very few neutral atoms left to ionize ($n_\mathrm{H\,I}/n_\mathrm{H} \sim 10^{-5}$). This can be illustrated by considering the {\em global ionization parameter}, which is the number of ionizing photons per neutral H atom $\Gamma_{IP} = \langle n_{ph}\rangle/\langle n_{H I} \rangle$ averaged over the entire volume. Specifically, we integrate the grey radiation energy density divided by the mean photon energy $\bar{\epsilon}$ over the singly thresholded volume, and divide by the number of H {\footnotesize I} atoms in the same volume: \begin{equation} \Gamma_{IP} = \int_{V_t} (E/\bar{\epsilon}) d^3x \bigg / \int_{V_t} n_{HI} d^3x. \label{eq:IP} \end{equation} We see from Fig. \ref{IP} that $\Gamma_{IP}$ grows from $\sim 10^{-3}$ at $z=10$ to unity at $z\approx 6.5$ just before overlap. Thereafter $\Gamma_{IP}$ grows very rapidly, reaching a value around $10^5$ at the overlap redshift, and leveling off at around $10^6$ below that. From the standpoint of the global ionization parameter, reionization begins photon starved but completes photon abundant. \begin{figure} \includegraphics[width=0.5\textwidth]{fig20-eps-converted-to.pdf} \caption{Redshift evolution of the global H {\footnotesize I}~ ionization parameter as defined in Eq. \eqref{eq:IP}. } \label{IP} \end{figure} Returning to Fig. \ref{sanitycheckrate} we see that the recombination rate density $\dot{R}_t$ curve tracks the ionization rate density curve to $z\sim 7$, but is about 0.7 dex lower in magnitude, as it must be if the ionized volume filling fraction is to grow. As overlap is approached ionizations and recombinations come into balance, but the recombination rate density has dropped considerably since it reached its maximum value at $z \approx 6.5$. This is also the redshift at which the ionization rate achieves a maximum, and when the global ionization parameter reaches unity. We also observe that the $f_{esc}$ curve in Fig. \ref{EscFraction} begins its precipitous drop at this redshift. We believe all of these events signal the rapid rise in the global ionization parameter below $z=6.5$, and not some change in the escape fraction of young galaxies. Counting the fraction of all ionizations occuring outside halos is not a reliable estimate of the escape fraction for $\Gamma_{IP} \gg 1$ because it does not count the photons in the radiation field that have nothing to ionize. Therefore we need to modify Eq. \eqref{eq:fesc} to include photons which build up of the radiation field: \begin{equation} \bar{f}_{esc} = \int_{V_t} (n_\mathrm{H\,I}\Gamma_\mathrm{H\,I}^{ph} + \frac{1}{\bar{\epsilon}}\frac{dE}{dt}) d^3x \bigg / \int_{V} (\eta/\bar{\epsilon}) d^3x . \label{eq:fescimproved} \end{equation} \\Here the numerator is the rate at which ionizing photons are causing ionizations in the IGM and building up the UV background, and the denominator is volume integrated ionizing photon production rate. Fig. \ref{RadEscFraction} plots $\bar{f}_{esc}$ calculated according to Eq. \eqref{eq:fescimproved}. Each contribution to $\bar{f}_{esc}$ is plotted separately, as well as the sum. We see that $\bar{f}_{esc}$ is roughly constant with redshift with a value of around 0.6. We see that as the contribution due to ionizations declines below $z\sim 7$, the contribution due to the change in radiation background intensity increases in a compensating fashion. This confirms our earlier suspicions and gives us a better estimate of the mean circumgalactic attenuation of ionizing radiation from young galaxies. \begin{figure} \includegraphics[width=0.5\textwidth]{fig21-eps-converted-to.pdf} \caption{Redshift evolution of the globally averaged escape fraction contribution from circumgalactic absorption as estimated by the number of ionizations occuring in the IGM and the buildup of the ionizing radiation background. The curves labeled $\bar{f}_{esc}(I_t), \bar{f}_{esc}(\dot{E})$ plot the contributions of the first and second terms in Eq. \eqref{eq:fescimproved}, while the curve labeled $\bar{f}_{esc}$ plots their sum.} \label{RadEscFraction} \end{figure} To complete the picture we plot in Fig. \ref{sanitycheckrate} the number density of ionizing photons escaping into the IGM, calculated as $\dot{N}_{IGM}=\bar{f}_{esc}\dot{N}_{sim}$, where $\dot{N}_{sim}$ is the ionizing photon production rate in the simulation, and $\bar{f}_{esc}$ is the improved estimate for the escape fraction calculated using Equation \eqref{eq:fescimproved}. We see that at high redshifts the $\dot{N}_{IGM}$ and $\dot{N}_t$ track each other closely. This tells us two things. First, that reionization at high redshifts when $Q_{\mathrm{H\,II}} \ll 1$ is photon starved, in the sense that every ionizing photon emitted results in an ionization. And second that our estimate of $\bar{f}_{esc}$ is reasonably accurate at these redshifts. However, as redshift decreases, the two curves systematically begin to deviate from one another in the sense that $\dot{N}_t < \dot{N}_{IGM}$. Beginning at $z = 6.5$ the ionization rate density begins to decrease while the ionizing photon production rate into the IGM continues to rise. After overlap the large disparity between the $\dot{N}_{IGM}$ and $\dot{N}_t$ curves can then be understood as saying that the IGM becomes photon abundant. The ratio of ionization rate density and the photon injection rate into the IGM is plotted in Fig. \ref{Ndot_Ratio}. The ratio is unity initially, and slowly decreases until $z\approx 7$, and then drops rapidly as overlap is approached. After overlap the ratio is about 0.05. In other words, after overlap, the photon production rate is about 20$\times$ the ionization rate in a volume averaged sense. Since the ionization and recombination rates are in balance after overlap, we conclude that the volume averaged photon injection rate is about 20$\times$ the recombination rate. \begin{figure} \includegraphics[width=0.5\textwidth]{fig22-eps-converted-to.pdf} \caption{Ratio of the volume integrated photoionization rate in the IGM $\dot{N}_t$ to the integrated photon injection rate into the IGM $\dot{N}_{IGM}$, where the IGM is defined as cells with $\Delta_b < 100$. The ratio is near unity initilly, remains high until $z\approx 7$ ($Q_{HII} \approx 0.5$), and then drops rapidly as overlap is approached and the IGM becomes highly ionized.} \label{Ndot_Ratio} \end{figure} \section{Introduction} \label{sec:introduction} The Epoch of Reionization (EoR) is an active area of research observationally, theoretically, and computationally. Observations constrain the tail end of hydrogen reionization to the redshift range $z=6-8$ \citep{RobertsonEtAl2010}. These observations include the presence of Gunn-Peterson troughs in the Ly $\alpha$ absorption spectra of high redshift quasars \citep{FanEtAl2006}, and the strong evolution of Lyman $\alpha$ emitter luminosity function (Robertson et al. 2010 and references therein.) Observations from the WMAP and Planck satellites tell us that the universe was substantially ionized by $z \approx 10$ but can say little about the reionization history or topology \citep{JarosikEtAl2011,Planck2013}. High redshift 21cm observations hold forth great promise of elucidating the details of this transition \citep{BarkanaLoeb2007, PritchardLoeb2012}, but these results are still in the future. It is believed that early star forming galaxies provided the bulk of the UV photons responsible for reionization \citep{RobertsonEtAl2010,RobertsonEtAl2013}, but early QSOs may have also contributed \citep{MadauEtAl1999, BoltonHaehnelt2007, HaardtMadau2012}. The ``galaxy reionizer" hypothesis has been greatly strengthened by the recent advances in the study of high redshift galaxies afforded by the IR-sensitive Wide Field Camera 3 (WFC3) aboard the Hubble Space Telescope \citep[e.g.][]{RobertsonEtAl2010, RobertsonEtAl2013, BouwensEtAl2011, BouwensEtAl2011b, OeschEtAl2013}. Within uncertainties, the luminosity function of $z=6$ Lyman break galaxies (LBGs) appears to be sufficient to account for reionization at that redshift from a photon counting argument \citep{BoltonHaehnelt2007, RobertsonEtAl2010, BouwensEtAl2012}. Among the observational uncertainties are the faint-end slope of the galaxy luminosity function \citep{WiseCen2009,LabbeEtAl2010,BouwensEtAl2012}, the spectral energy distribution of the stellar population \citep{CowieEtAl2009,WillotEtAl2010,HaardtMadau2012}, and the escape fraction of ionizing photons \citep{WyitheEtAl2010, YajimaEtAl2011, MitraEtAl2013}. Among the theoretical uncertainties are the number of ionizing photons per H atom required to bring the neutral IGM to its highly ionized state by $z=6$, the clumping factor correction to the mean IGM recombination time \citep{PawlikEtAl2009, RaicevicTheuns2011, FinlatorEtAl2012, ShullEtAl2012, RobertsonEtAl2013}, and the contribution of Pop III stars and accreting black holes to the early and late stages of reionization \citep{BoltonHaehnelt2007,TracGnedin2011,AhnEtAl2012}. When assessing whether an observed population of high-z galaxies is capable of reionizing the universe (e.g., Robertson et al. 2013), observers often use the criterion derived by \cite{MadauEtAl1999} for the ionzing photon volume density $\dot{\mathcal{N}}_{ion}$ necessary to maintain the clumpy IGM in an ionized state: \begin{align} \label{eq:ndot} \dot{\mathcal{N}}_{ion}(z) &= \frac{\bar{n}_\mathrm{H}(0)}{\bar{t}_{rec}(z)} = (10^{51.2}s^{-1}Mpc^{-3})\left(\frac{C}{30}\right) \notag\\ &\times \left(\frac{1+z}{6}\right)^3 \left(\frac{\Omega_b h_{50}^2}{0.08}\right)^2, \end{align} where $\bar{n}_\mathrm{H}(0)$ is the mean comoving number density of H atoms, $C \equiv \langle n^2_\mathrm{H\,II} \rangle /\langle n_\mathrm{H\,II} \rangle ^2$ is the H {\footnotesize II} clumping factor (angle brackets denote volume average over a suitably large volume that the average is globally meaningful), and the rest of the symbols have their usual meaning. The origin of this formula is a simple photon counting argument, which says that in order to maintain ionization at a given redshift $z$, the number of ionizing photons emitted in a large volume of the universe multiplied by a characteristic recombination time, denoted $\bar{t}_{rec}$, must equal the number of hydrogen atoms: $\dot{\mathcal{N}}_{ion} \times \bar{t}_{rec} = \bar{n}_\mathrm{H}(0)$. The clumping factor enters as a correction factor to account for the density inhomogeneties in the IGM induced by structure formation. We note that $\bar{t}_{rec}$ is not the volume average of the local recombination time of the ionized plasma, as this would heavily weight regions with the {\em longest} recombination times; i.e. voids. A proper derivation of Equation \eqref{eq:ndot} shows that $\bar{t}_{rec} \propto \langle t_{rec}^{-1} \rangle ^{-1}$, which weights regions with the {\em shortest} recombination times; i.e. regions at the mean density and above. Equation \eqref{eq:ndot} is based on a number of simplifying assumptions discussed by \cite{MadauEtAl1999}, including the assumption $\bar{t}_{rec} \ll t$. It is this assumption that allows history-dependent effects to be ignored, and a quasi-instantaneous analysis of the photon budget for reionization to be done. The validity of this assumption is naturally redshift dependent, but it is also dependent upon the adopted definition of $\bar{t}_{rec}$. A second comment about Equation \eqref{eq:ndot} is that it does not ask how many ionizing photons per H atom are required to convert a neutral IGM to a fully ionized one, only how many are required to {\em maintain} the IGM in an ionized state. Because the recombination time is short at high redshifts, it is expected that this number is greater than one. In this paper we examine these and related topics within the context of a direct numerical simulation of cosmic reionization based on a new flux-limited diffusion radiation transport solver installed in the {\em Enzo} code \citep{NormanEtAl2013} (hereafter Paper I). Our approach self-consistently couples all the relevant physical processes (gas dynamics, dark matter dynamics, self-gravity, star formation/feedback, radiative transfer, nonequilibrium ionization/recombination, heating and cooling) and evolves the system of coupled equations on the same high resolution mesh. We refer to this approach as {\em direct numerical simulation} or {\em resolution matched}, in contrast to previous approaches which decouple and coarse-grain the radiative transfer and ionization balance calculations relative to the underlying dynamical calculation. Our method is scalable with respect to the number of radiation sources, size of the mesh, and the number of computer processors employed. This scalability permits us to simulate cosmological reionization in large cosmological volumes (L $\sim100$ Mpc) while directly modeling the sources and sinks of ionizing radiation, including radiative feedback effects such as photoevaporation of gas from halos, Jeans smoothing of the IGM, and enhanced recombination due to small scale clumping. In this the first of several application papers, we investigate in a volume of modest size (L=$20$ Mpc) the mechanics of reionization from stellar sources forming in high-$z$ galaxies, the role of gas clumping, recombinations, and the photon budget required to complete reionization. By analyzing this simulation we are able to critically examine the validity of Equation \eqref{eq:ndot} as a predictor of when the end of EoR will occur, and we can calculate the integrated number of ionizing photons per H atom needed to ionize the simulated volume $\gamma_{ion}/H=\int dt \dot{\mathcal{N}}_{ion} / \bar{n}_\mathrm{H}(0)$. Ignoring recombinations within the virial radii of collapsed halos, we find $\gamma_{ion}/H \approx 2$. This result supports the ``photon starved'' reionization scenario discussed by \cite{BoltonHaehnelt2007}. We also examine whether modern revisions to Equation \eqref{eq:ndot} using alternatively defined clumping factors \citep{PawlikEtAl2009, RaicevicTheuns2011, FinlatorEtAl2012, ShullEtAl2012} are improvements over the original. We find they systematically overestimate the redshift of reionization completion $z_{reion}$ because the condition $\bar{t}_{rec}/t \ll 1$ is never obeyed. We study the accuracy and validity of the time-dependent analytic model of \cite{MadauEtAl1999}, and find that while it is in better agreement with the simulation, it also overestimates $z_{reion}$ because it ignores important corrections to the ionization term at early and late times. This paper is organized as follows: in \S\ref{Method} we discuss the design criteria for the simulation and briefly outline the basic equations and implementation of the FLD radiation transport model, referring the reader to Paper I for a more complete description of the numerical algorithms and tests. In \S\ref{GeneralResults}, we present some general features of the simulation and demonstrate its broad consistency with observed star formation rate density and high redshift galaxy luminosity function. In \S\ref{sec:ClumpingFactors} we examine the accuracy of different clumping factor approaches to estimating the redshift of complete reionization. In \S\ref{escape} we derive a global estimate for the circumgalactic absorption of ionizing radiation from our simulation. In \S\ref{Qdot} we test a simple analytic model for the evolution of the ionized volume fraction $Q_\mathrm{H\,II}$ and present an improvement to the model which better agrees with our simulation. In \S\ref{Discussion} we discuss implications of our results on the current understanding of reionization. And finally, in \S\ref{Conclusions} we end with a summary of our main results and conclusions. \section{Method} \label{Method} \subsection{Simulation Goals and Parameters} We use the Enzo code \citep{TheEnzoCollaboration}, augmented with a flux-limited diffusion radiative transfer solver and a parameterized model of star formation and feedback \citep{NormanEtAl2013} to simulate inhomogeneous hydrogen reionization in a 20 Mpc comoving box in a WMAP7 $\Lambda$CDM cosmological model. Details of the numerical methods and tests are provided in Paper I. Here we briefly describe the simulation's scientific goals and design considerations to put it into perspective with other reionization simulations. For completeness, the physical equations we solve and the treatment of the ionizing sources and radiation field are included below. Our principle goal is to simulate the physical processes occuring in the IGM outside the virial radii of high redshift galaxies in a {\em representative} realization of inhomogenous reionization. We wish to simulate the early, intermediate, and late phases of reionization in a radiation hydrodynamic cosmological framework so that we may study the nonequilibrium ionization/recombination processes in the IGM at reasonably high resolution self-consistently coupled to the dynamics. In this way we can study such effects as optically thick heating behind the I-fronts \citep{AbelHaehnelt1999}, Jeans smoothing \citep{ShapiroEtAl1994,Gnedin2000b}, photoevaporation of dense gas in halos \citep{ShapiroEtAl2004}, and nonequilibrium effects in the low density voids. Because we carry out our simulation on a fixed Eulerian grid, we do not resolve the internal processes of protogalaxies very well. In this sense, our simulation is not converged on all scales. Nonetheless Equations \eqref{eq:gravity} to \eqref{eq:cons_radiation} are solved everywhere on the mesh self-consistently, including ionization/recombination and radiative transfer inside protogalaxies. The escape of ionizing radiation from galaxies to the IGM is thus simulated directly, and not introduced as a parameter. We use a star formation recipe that can be tuned to closely reproduce the observed high-$z$ galaxy luminosity function (LF), star formation rate density (SFRD), and redshift of reionization completion. This gives us confidence that we are simulating IGM processes in a realistic scenario of reionization. We simulate a WMAP7 \citep{JarosikEtAl2011} $\Lambda$CDM cosmological model with the following parameters: $\Omega_{\Lambda} = 0.73$, $\Omega_m = 0.27$, $\Omega_b = 0.047$, $h = 0.7$, $\sigma_8 = 0.82$, $n_s = 0.95$, where the symbols have their usual meanings. A Gaussian random field is initialized at $z=99$ using the {\em Enzo} initial conditions generator {\em inits} using the \cite{EisensteinHu1999} fits to the transfer functions.The simulation is performed in a comoving volume of (20 Mpc)$^3$ with a grid resolution of $800^3$ and the same number of dark matter particles. This yields a comoving spatial resolution of 25 kpc and dark matter particle mass of $4.8 \times 10^5 M_{\odot}$. This resolution yields a dark matter halo mass function that is complete down to $M_h = 10^8 M_{\odot}$, which is by design, since this is the mass scale below which gas cooling becomes inefficient. However, due to our limited boxsize, our halo mass function is incomplete above $M_h \approx 10^{11} M_{\odot}$ (see Figure \ref{HMF}). In a forthcoming paper we will report on a simulation of identical design and resolution as this one, but in a volume 64 times as large, which contains the rarer, more massive halos. With regard to resolving the diffuse IGM, our $25$ kpc resolution equals the value recommended by \cite{BryanEtAl1999} to converge on the properties of the Ly $\alpha$ forest at lower redshifts, is $3\times$ better than the optically thin high resolution IGM simulation described in \cite{ShullEtAl2012}, and nearly $4\times$ better than the inhomogeneus reionization simulation described in \cite{TracEtAl2008}. As described below in \S\ref{starformationandfeedback}, we use a parameterized model of star formation calibrated to observations of high redshift galaxies. The star formation efficiency parameter $f_*$ is adjusted to match the observed star formation rate density in the interval $6 \leq z \leq 10$ from \cite{BouwensEtAl2011}. The simulation consumed 255,000 core-hrs running on 512 cores of the Cray XT5 system {\em Kraken} operated by the National Institute for Computational Science at ORNL. \subsection{Governing Equations} \label{GoverningEquations} The equations of cosmological radiation hydrodynamics implemented in the Enzo code used for this research are given by the following system of partial differential equations (Paper I): \begin{align} \nabla^2 \phi &= \frac{4\pi g}{a}(\rho_b + \rho_\mathrm{dm} - \langle \rho \rangle), \label{eq:gravity}\\ \partial_\mathrm{t} \rho_b + \frac1a {\bf v}_b \cdot \nabla \rho_b &= -\frac1a \rho_b \nabla\cdot{\bf v}_b -\dot{\rho}_{SF}, \label{eq:cons_mass}\\ \partial_\mathrm{t} {\bf v}_b + \frac1a\({\bf v}_b\cdot\nabla\){\bf v}_b &= -\frac{\dot{a}}{a}{\bf v}_b - \frac{1}{a\rho_b}\nabla p - \frac1a \nabla\phi, \label{eq:cons_momentum}\\ \partial_\mathrm{t} e + \frac1a{\bf v}_b\cdot\nabla e &= - \frac{2\dot{a}}{a}e - \frac{1}{a\rho_b}\nabla\cdot\left(p{\bf v}_b\right) \nonumber\\ &- \frac1a{\bf v}_b\cdot\nabla\phi + G - \Lambda + \dot{e}_{SF} \label{eq:cons_energy}\\ \partial_\mathrm{t} {\tt n}_\mathrm{i} + \frac{1}{a}\nabla\cdot\({\tt n}_\mathrm{i}{\bf v}_b\) &= \alpha_\mathrm{i,j} {\tt n}_\mathrm{e} {\tt n}_\mathrm{j} - {\tt n}_\mathrm{i} \Gamma_\mathrm{i}^{ph}, \qquad \nonumber\\ &i=1,\ldots,N_\mathrm{s} \label{eq:chemical_ionization}\\ \partial_\mathrm{t} E + \frac1a \nabla\cdot\(E {\bf v}_b\) &= \nabla\cdot\(D\nabla E\) - \frac{\dot{a}}{a}E \nonumber\\ &- c \kappa E + \eta \label{eq:cons_radiation} \end{align} Equation \eqref{eq:gravity} describes the modified gravitational potential $\phi$ due to baryon density $\rho_\mathrm{b}$ and dark matter density $\rho_\mathrm{dm}$, with $a$ being the cosmological scale factor, $g$ being the gravitational constant, and $\langle \rho \rangle$ being the cosmic mean density. The collisionless dark matter density $\rho_\mathrm{dm}$ is evolved using the Particle Mesh method (equation not shown above), as described in \citealt{HockneyEastwood1988, TheEnzoCollaboration}. Equations \eqref{eq:cons_mass}, \eqref{eq:cons_momentum} and \eqref{eq:cons_energy} are conservation of mass, momentum and energy, respectively, in a comoving coordinate system \citep{BryanEtAl1995,TheEnzoCollaboration}. In the above equations, ${\bf v}_b\equiv a(t)\dot{{\bf x}}$ is the proper peculiar baryonic velocity, $p$ is the proper pressure, $e$ is the total energy per unit mass, and $G$ and $\Lambda$ are the heating and cooling coefficients. Equation \eqref{eq:chemical_ionization} describes the chemical balance between the different ionization species (in this paper we used H {\footnotesize I}, H {\footnotesize II}, He {\footnotesize I}, He {\footnotesize II}, He {\footnotesize III} densities) and electron density. Here, ${\tt n}_\mathrm{i}$ is the comoving number density of the $i^{th}$ chemical species, ${\tt n}_\mathrm{e}$ is the electron number density, ${\tt n}_\mathrm{j}$ is the ion that reacts with species $i$, and $\alpha_\mathrm{i,j}$ are the reaction rate coefficient between species $i$ and $j$ \citep{AbelEtAl1997, HuiGnedin1997}, and finally $\Gamma^{ph}_\mathrm{i}$ is the photoionization rate for species $i$. \subsection{Radiation Transport} \label{RadiationTransport} Equation \eqref{eq:cons_radiation} describes radiation transport in the Flux Limited Diffusion (FLD) approximation in an expanding cosmological volume \citep{ReynoldsEtAl2009,NormanEtAl2013}. $E$ is the comoving grey radiation energy density. The {\em flux limiter} $D$ is a function of $E$, $\nabla E$, and the opacity $\kappa$ \citep{Morel2000}, and has the form: \begin{align} D &= \mbox{diag}\left(D_1, D_2, D_3\right), \quad\mbox{where} \\ D_\mathrm{i} &= c \(9\kappa^2 + R_\mathrm{i}^2\)^{-1/2},\quad\mbox{and} \\ R_\mathrm{i} &= \max\left\{\frac{|\partial_\mathrm{x_i} E|}{E},10^{-20}\right\} \end{align} In the calculation of the grey energy density $E$, we assume $E_\nu(\mathbf{x},t,\nu)=\tilde{E}(\mathbf{x},t)\,\chi_E(\nu)$, therefore: \begin{align} \label{eq:grey_definition} E(\mathbf{x},t) &= \int_{\nu_1}^{\infty} E_\nu(\mathbf{x},t,\nu)\,\mathrm d\nu \nonumber \\ &=\tilde{E}(\mathbf{x},t) \int_{\nu_1}^{\infty} \chi_E(\nu)\,\mathrm d\nu, \end{align} Which separates the dependence of $E$ on coordinate $\mathbf{x}$ and time $t$ from frequency $\nu$. Here $\chi_E$ is the spectral energy distribution (SED) taken to be that of a Pop II stellar population similiar to one from \citep{RicottiEtAl2002}. \subsection{Star Formation and Feedback} \label{starformationandfeedback} Because star formation occurs on scales not resolved by our uniform mesh simulation, we rely on a subgrid model which we calibrate to observations of star formation in high redshift galaxies. The subgrid model is a variant of the \cite{CenOstriker1992} prescription with two important modifications as described in \cite{SmithEtAl2011}. In the original \cite{CenOstriker1992} recipe, a computational cell forms a collisionless ``star particle" if a number of criteria are met: the baryon density exceeds a certain numerical threshold; the gas velocity divergence is negative, indicating collapse; the local cooling time is less than the dynamical time; and the cell mass exceeds the Jeans mass. In our implementation, the last criterion is removed because it is always met in large scale, fixed-grid simulations, and the overdensity threshold is taken to be $\rho_b/(\rho_{c,0}(1+z)^3) > 100$, where $\rho_{c,0}$ is the critical density at $z=0$. If the three remaining criteria are met, then a star particle representing a large collection of stars is formed in that timestep and grid cell with a total mass \begin{equation} m_* = f_* m_{cell} \frac{\Delta t}{t_{dyn}}, \end{equation} where $f_*$ is an efficiency parameter we adjust to match observations of the cosmic star formation rate density (SFRD) \citep{BouwensEtAl2011}, $m_{cell}$ is the cell baryon mass, $t_{dyn}$ is the dynamical time of the combined baryon and dark matter fluid, and $\Delta t$ is the hydrodynamical timestep. An equivalent amount of mass is removed from the grid cell to maintain mass conservation. Although the star particle is formed instantaneously (i.e., within one timestep), the conversion of removed gas into stars is assumed to proceed over a longer timescale, namely $t_{dyn}$, which more accurately reflects the gradual process of star formation. In time $\Delta t$, the amount of mass from a star particle converted into newly formed stars is given by \begin{equation} \Delta m_{SF} = m_* \frac{\Delta t}{t_{dyn}} \frac{t-t_*}{t_{dyn}} e^{-(t-t_*)/t_{dyn}}, \end{equation} where $t$ is the current time and $t_*$ is the formation time of the star particle. To make the connection with Equation \eqref{eq:cons_momentum}, we have $\dot{\rho}_{SF} =\Delta m_{SF}/(V_{cell}\Delta t)$, where $V_{cell}$ is the volume of the grid cell. Stellar feedback consists of the injection of thermal energy, gas, and radiation to the grid, all in proportion to $\Delta m_{SF}$. The thermal energy $\Delta e_{SF}$ and gas mass $\Delta m_g$ returned to the grid are given by \begin{equation} \Delta e_{SF} = \Delta m_{SF} c^2 \epsilon_{SN}, \qquad \Delta m_g = \Delta m_{SF} f_{m*}, \end{equation} where $c$ is the speed of light, $\epsilon_{SN}$ is the supernova energy efficiency parameter, and $f_{m*}=0.25$ is the fraction of the stellar mass returned to the grid as gas. Rather than add the energy and gas to the cell containing the star particle, as was done in the original \cite{CenOstriker1992} paper, we distribute it evenly among the cell and its 26 nearest neighbors to prevent overcooling. As shown by \cite{SmithEtAl2011}, this results in a star formation recipe which can be tuned to reproduce the observed SFRD. This is critical for us, as we use the observed high redshift SFRD to calibrate our reionization simulations. To calculate the radiation feedback, we define an emissivity field $\eta(x)$ on the grid which accumulates the instantaneous emissivities $\eta_i(t)$ of all the star particles within each cell. To calculate the contribution of each star particle $i$ at time $t$ we assume an equation of the same form for supernova energy feedback, but with a different energy conversion efficiency factor $\epsilon_{UV}$. Therefore \begin{equation} \label{eq:emissivity} \eta= \sum_\mathrm{i}\epsilon_\mathrm{uv}\frac{\Delta m_\mathrm{SF} c^2}{V_\mathrm{cell}\Delta t} \end{equation} Emissivity $\eta$ is in units of erg s$^{-1}$cm$^{-3}$. The UV efficiency factor $\epsilon_\mathrm{uv}$ is taken from \cite{RicottiEtAl2002} as 4$\pi\times 1.1 \times 10^{-5}$, where the factor $4\pi$ comes from the conversion from mean intensity to radiation energy density. \subsection{Data Analysis} \label{DataAnalysis} Due to the enormous amount of data produced by the simulation (one output file is about 100 GB), we needed a scalable tool suited to the task of organizing and manipulating the data into human readable form. We use the analysis software tool \texttt{yt} \citep{TurkEtAl2011} specifically created for doing this type of vital task. It is a python based software tool that does ``Detailed data analysis and visualizations, written by working astrophysicists and designed for pragmatic analysis needs." \texttt{yt} is open source and publicly available at http://yt-project.org. \section{An Improved Model for the Evolution of $Q_\mathrm{H\,II}$} \label{Qdot} In this section we compare the evolution of the ionized volume fraction $Q_\mathrm{H\,II}$ from our simulation with the analytic model introduced by \cite{MadauEtAl1999}. We are motivated to do this because as we have seen from \S\ref{sec:ClumpingFactors}, Equation \eqref{eq:ndot} is not a useful predictor of when $Q_\mathrm{H\,II}$ reaches unity. We therefore want to investigate the accuracy of the time dependent model from which Equation \eqref{eq:ndot} is derived as a limiting case. \cite{MadauEtAl1999} derived the following ODE for the evolution of $Q_\mathrm{H\,II}$ (their Equation 20): \begin{equation} \label{eq:dQdt} \frac{dQ_\mathrm{H\,II}}{dt} = \frac{\dot{n}_{ion}}{\bar{n}_\mathrm{H}}-\frac{Q_\mathrm{H\,II}}{\bar{t}_{rec}} \end{equation} where $\dot{n}_{ion}$ is ionizing photon injection rate, $\bar{n}_\mathrm{H}$ is the mean density of H atoms in the universe, and $\bar{t}_{rec}$ is some characteristic recombination time taking the clumpiness of the IGM into account. For a constant clumping factor and comoving emissivity \cite{MadauEtAl1999} show that \begin{equation} Q_\mathrm{H\,II}(t) \approx \frac{\dot{n}_{ion}}{\bar{n}_\mathrm{H}} \bar{t}_{rec} \end{equation} Setting $Q=1$ one arrives at $\dot{n}_{ion}\bar{t}_{rec}=\bar{n}_\mathrm{H}$, the basis for deriving Equation \eqref{eq:ndot}. \cite{MadauEtAl1999} state that this relation should still be valid provided the clumping factor and comoving emissivity are slowly varying on a timescale of $\bar{t}_{rec}$. We utilize the differential form for our comparison because our emissivity is not a constant value, nor is it slowly varying on a recombination time as $Q \rightarrow 1$, as we show below. A practical issue when testing Equation \eqref{eq:dQdt} is how $\bar{t}_{rec}$ should be evaluated when $Q<1$, and in particular when $Q\ll 1$. In the limit $Q\ll 1$ one is dealing with isolated H {\footnotesize II} regions evolving under the influence of local conditions. Yet the definition for $\bar{t}_{rec}$ in Equation \eqref{eq:tmadau} invokes {\em global} values for $C$ and $\langle n_\mathrm{H\,II} \rangle$. Should these quantitles be evaluated locally only within ionized regions? Or are global estimates good enough? In particular, since \cite{MadauEtAl1999}'s Equation (20) uses $\bar{n}_\mathrm{H}$ as a proxy for $\langle n_\mathrm{H\,II} \rangle$, what is the appropriate value for $C$ to use? A second practical issue is what to take for $\dot{n}_{ion}$. This is commonly understood to be the rate at which ionizing photons are injected into the IGM (e.g., Haardt \& Madau 2012, \S9.3), which in our parlance is $\dot{N}_{IGM}$. Or should we take the actual ionization rate density measured in the simulation $\dot{N}_t$? As we saw in the previous section, these two rates diverge as overlap is approached, and differ by more than an order of magnitude after overlap (Fig. \ref{Ndot_Ratio}). To examine these issues we plot in Figure \ref{Qeffv1} $Q(z)$ from our simulation, as well as theoretical curves obtained by integrating Equation \eqref{eq:dQdt} under various assumptions. The curve labelled $Q(sim)$ is the ionized volume fraction from our simulation that is at least 99.9\% ionized (Well Ionized). The other four curves are obtained by integrating Equation \eqref{eq:dQdt} setting $\dot{n}_{ion}=\dot{N}_t$ for various choices for $\bar{t}_{rec}$ (we investigate the $\dot{n}_{ion}=\dot{N}_{IGM}$ case at the end of this section.) The integral is approximated by summing a piecewise linear interpolation of the two terms on the RHS of Equation \eqref{eq:dQdt} using the trapezoidal rule: \begin{align} Q(t) &= \int_{t*}^{t} \frac{dQ}{dt}dt \approx \sum \frac{dQ}{dt} \Delta t \notag\\ &= \sum_{i}(\mathrm{Term_1} - \mathrm{Term_2})_i \Delta t_i \label{eq:integration} \end{align} where $t*$ is the time when the first star forms in the simulation. The curve labeled $Q(\langle t_{rec}\rangle)$ uses the volume averaged recombination time (volume average of Equation \ref{recombtime}). The two curves labeled $Q(t_\mathrm{Madau})$ use Equation \eqref{eq:tmadau} to evaluate $\bar{t}_{rec}$ for $C=2$ and $3$, substituting $\bar{n}_\mathrm{H}$ for $\langle n_\mathrm{H\,II} \rangle$ and assuming a constant T=10$^4$K for the IGM. The curve labeled $Q(t_{rec,eff})$ uses the effective recombination time definition \begin{equation} \label{eq:treceff} \bar{t}_{rec}=t_{rec,eff}\equiv \frac{\langle n_\mathrm{H\,II}\rangle}{\langle n_\mathrm{H\,II} n_e \alpha_B(T)\rangle} \end{equation} This particular definition makes the last line of Equation \eqref{eq:trecpart2} true trivially, with no assumption about the IGM temperature or ionization state of the hydrogen. It involves no {\em ad hoc} clumping factors, and represents the actual appropriately averaged recombination time in the simulation. All the above volume averaged quantities have the threshold of $\Delta_b<100$ applied, and thus exclude dense gas bound to halos. Several of the curves derived from integrating $\frac{dQ}{dt}$ reach values above unity at the end of the overlapping phase. While it is physically impossible to have $Q>1$ it is not mathematically forbidden, and so we show the complete curves because they give us some insight about the relative contribution of the recombination term (Term$_2$) as compared to the ionization term (Term$_1$). The $Q(\langle t_{rec}\rangle)$ curve ionizes the quickest, reaching $Q=1$ at $z\sim 6.5$, which is substantially before the simulation which achieves it at $z\approx 5.8$. The reason for this, as we will analyze shortly, is that recombinations play essentially no role in this model. The $Q(t_{rec,eff})$ curve has the same shape as the $Q(sim)$, but is everywhere higher, and crosses $Q=1$ at $z\sim 6.1$. Given that this integration uses the actual ionization rate density and effective recombination time in the simulation, this discrepancy demands an explanation. We address this below. Finally the $Q(t_\mathrm{Madau})$ curves do not match the shape of the $Q(sim)$ curve, ionizing more quickly at early times, and exhibiting a maximum value for $Q$ at $z \sim 6$. \begin{figure} \includegraphics[width=0.5\textwidth]{fig23a-eps-converted-to.pdf} \includegraphics[width=0.5\textwidth]{fig23b-eps-converted-to.pdf} \caption{{\em Top}: Comparison of the evolution of the ionized volume fraction Q from our simulation with the analytic model introduced by \cite{MadauEtAl1999}. Q(sim) is calculated directly from counting the cells satisfying the Well Ionized threshold of $f_i>0.999$. The other curves are calculated from integrating Equation \eqref{eq:integration} with the different expressions for $\bar{t}_{rec}$ in Term$_2$, as described in the text. {\em Bottom}: Plot of Term$_1$ and Term$_2$ individually using the different expressions for $\bar{t}_{rec}$.} \label{Qeffv1} \end{figure} To understand this behavior more fully we plot in Figure \ref{Qeffv1} {\em bottom} the values for Term$_1$ and Term$_2$ in Equation \eqref{eq:dQdt}. The blue curve is Term$_1$ of Equation \eqref{eq:dQdt}. The other four curves plot Term$_2$ with their respective values for $\bar{t}_{rec}$. The ionization curve dominates all the recombination curves at high redshifts, and reaches a maximum at $z\sim 6.5$. This is a partial reflection of the plateauing and subsequent decline of the SFRD shown in Figure \ref{SFR}. More fundamentally, it is a reflection of the rapid drop in the neutral fraction of the IGM as overlap is approached. The curve using the volume averaged recombination time $\langle t_{rec}\rangle$ yields such low values compared to the others that we multiply it by 100 to make it more visible. Although this is not the relevant recombination time to use, since it weights low density regions, it is effectively the limiting case $\bar{t}_{rec}\rightarrow \infty$. We can therefore interpret the blue curve in Figure \ref{Qeffv1}a as an integration of the ionization term only. It is significantly higher than the $Q(sim)$ curve, suggesting that recombinations are important in the simulation at some level. The ionization term dominates the recombination term by factors of $6-10$ in the $t_{rec,eff}$ curve until just before overlap, and the two terms come into balance after overlap. The two $t_\mathrm{Madau}$ recombination curves are subdominant to the ionization term until $z \sim 6$, and at lower redshifts they become dominant. This explains the turnaround in the corresponding $Q$ curves in Figure \ref{Qeffv1}a. The differences in the magnitude of the recombination curves in Figure \ref{Qeffv1}b, especially at higher redshifts, is directly attributable to the magnitude of $\bar{t}_{rec}$. For completeness we plot $\bar{t}_{rec}$ versus redshift in Figure \ref{treceffhubble}, both unnormalized and normalized by $t_\mathrm{Hubble}$. In addition to the three curves for $t_{rec,eff}$ and $t_\mathrm{Madau}$ for $C=2, 3$, we also plot $t_\mathrm{Madau}$ for $C=C_\mathrm{ttH\,II}$ and $C=C_\mathrm{tdm}$. We see that all the curves with the exception of the Madau formula curve using the thresholded dark matter clumping factor exhibit an increasing recombination time with decreasing redshift, in line with our expections. The latter curve shows the opposite trend, which is due to the fact that the dark matter clumping factor increases with decreasing redshift, even if it is thresholded to exclude halos (see Figure \ref{thresholded} bottom). Among the remaining curves the $t_{rec,eff}$ has the highest values, and increases more sharply than the $t_\mathrm{Madau}$ curves due to the temperature of the IGM. To demonstrate that, we plot one additional curve (dashed curve) for $t_{rec,eff}$ evaluated assuming a constant $T=10^4$K in the recombination rate coefficient. We now comment on the often-made assumption in reionization models that $\bar{t}_{rec} \ll t$. \cite{MadauEtAl1999} make this assumption in order to derive Equation \eqref{eq:ndot}. It is this assumption that allows for an instantaneous analysis of the photon budget to maintain the universe in an ionized state while ignoring history dependent effects. Referring to Figure \ref{treceffhubble}b we see this is never true for $t_{rec,eff}$ and it is not true for $t_\mathrm{Madau}$ at redshifts approaching overlap for any sensible value of $C$. We therefore conclude that history-dependent effects cannot be ignored, and that this is the reason Equations \eqref{eq:ndot}, \eqref{eq:updatedNdot} and \eqref{eq:ShullNdot} mis-predict the epoch of reionization completion. For the same reason applying these formulae at lower redshifts is highly suspect. \begin{figure} \includegraphics[width=0.5\textwidth]{fig24a-eps-converted-to.pdf} \includegraphics[width=0.5\textwidth]{fig24b-eps-converted-to.pdf} \caption{{\em Top}: Recombination time versus redshift, for various expressions for $\bar{t}_{rec}$ as described in the text. Curve labeled $t_{rec,eff}$ is the characteristic recombination time measured directly in the simulation. Curves labeled $t_{Madau}$ evaluate Eq. \eqref{eq:tmadau} for various choices for the clumping factor C. {\em Bottom}: Recombination time versus redshift normalized by the Hubble time, for various expressions for $\bar{t}_{rec}$.} \label{treceffhubble} \end{figure} Returning to the discrepancy between the $Q(sim)$ and $Q(t_{rec,eff})$ curves in Figure \ref{Qeffv1}a, since the most sensible choice for $t_{rec}$ did not give us satisfactory agreement, we wondered what the origin of the discrepancy could be. Since we have shown that recombinations are relatively unimportant at high redshifts, but that the discrepancy is already present at high redshifts, the only possibility is that there is something wrong with the first term of Equation \eqref{eq:integration}. When looking at the derivation for Equation \eqref{eq:dQdt} in \cite{MadauEtAl1999}, it is stated that it ``approximately holds for every isolated source of ionizing photon in the IGM.'' That got us to think that our calculation of $\bar{n}_\mathrm{H}$ may be off from what is originally intended if it is a global average over the entire simulation box. Since the original $\frac{dQ}{dt}$ is derived from the analytical Str\"{o}mgren sphere model, it assumed a single ionizing source at the center of the volume, and the the average density of the box is just the uniform density everywhere, we thought that might be the discrepancy. In an Inside-out model, I-fronts are not initially propagating in a gas with an average density given by $\bar{n}_H$, but somewhat higher density. Would agreement improve if instead of using $\bar{n}_H$ in the first term of Equation \eqref{eq:dQdt}, we used the local average density? We therefore modify Equation \eqref{eq:dQdt} as follows: \begin{equation} \frac{dQ}{dt} = \frac{\dot{n}_{ion}}{\delta_b\bar{n}_\mathrm{H}}-\frac{Q}{\bar{t}_{rec}} \label{eq:dQdtdb} \end{equation} where we have introduced in the denominator of the first term a factor $\delta_b \geq 1$ which corrects for the higher mean density within ionized bubbles. We measure $\delta_b$ from each redshift output as follows: $\delta_b = \langle \rho_b\rangle_{tt}/\langle \rho_b\rangle_{t}$. The volume average $\langle\rangle$ with subscript $t$ is the usual $\Delta_b<100$ threshold, the double subscript $tt$ indicates the additional threshold of $x_e>0.1$. Thus $\delta_b$ is the average baryon overdensity within Ionized regions excluding gas inside halos. Figure \ref{deltabvsQfit5} shows a plot of $\delta_b$ versus $Q$ together with a simple fitting formula which fits the data extremely well over the domain $0.01 \leq Q \leq 1$. To see if this formulation improves agreement with our simulated data, in Figure \ref{Qeffv2} we integrate Equation \eqref{eq:dQdtdb} again setting $\dot{n}_{ion}=\dot{N}_t$ and using $t_{rec,eff}$ to evaluate the second term. For comparison we show the curve obtained setting $\delta_b=1$, which repeats a curve already presented in Figure \ref{Qeffv1}. Although the simulated and integrated analytic model curves do not agree exactly, the $Q(\delta_b,t_{rec,eff})$ curve shows much better agreement with the simulation, with error on the order of 1\% instead of 10\%. \begin{figure} \includegraphics[width=0.5\textwidth]{fig25-eps-converted-to.pdf} \caption{Improved agreement between theory and simulation. Green and blue curves are as in Fig. \ref{Qeffv1}. Red curve is obtained by integrating modified evolution equation for Q taking into account the overdensity effect of Inside-out reionization (Equation \eqref{eq:dQdtdb}).} \label{Qeffv2} \end{figure} By not assuming a constant emissivity and using the modified differential form in determining the volume filling fraction of Equation \eqref{eq:dQdtdb}, we are able to more accurately model the evolution of the simulated volume filling fraction of H {\footnotesize II} to the Well Ionized level. For completeness we plot in Figure \ref{treceffvszfit} the evolution of $t_{rec,eff}$ used in the above integration, including a reasonably good fit to the data. \begin{figure} \includegraphics[width=0.5\textwidth]{fig26-eps-converted-to.pdf} \caption{Mean baryon overdensity of ionized gas as a function of the ionized volume filling fraction Q. Blue points are measured in the simulation by averaging over the doubly thresholded cells obeying $\Delta_b<100$ and $x_e > 0.1$. Red curve is a fit to the data.} \label{deltabvsQfit5} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig27-eps-converted-to.pdf} \caption{Analytic fit to $t_{rec,eff}$ (red line) , evaluated using simulation data (blue points) via Equation \eqref{eq:treceff}.} \label{treceffvszfit} \end{figure} Finally, we return to the question of what is the appropriate choice for $\dot{n}_{ion}$ in Equation \eqref{eq:dQdtdb}. This is commonly taken to be the rate at which ionizing photons are injected into the IGM (e.g., Haardt \& Madau 2012, \S9.3), because this can be connected to the observed UV luminosity density $\rho_{UV}$ by the formula $\dot{n}_{ion}=f_{esc}\xi_{ion}\rho_{UV}$, where $f_{esc}$ is the escape fraction for ionizing radiation, and $\xi_{ion}$ is the rate of ionizing photons per unit UV (1500 \AA{}) luminosity for the stellar population \citep{RobertsonEtAl2013}. However we have obtained excellent agreement between simulation and Equation \eqref{eq:dQdtdb} using the mean ionization rate density in the IGM $\dot{N}_t$, which differs from the ionizing photon injection rate density $\dot{N}_{IGM}$ as $Q \rightarrow 1$. In Figure \ref{Qeffv3} we show the result of integrating Equations \eqref{eq:dQdt} and \eqref{eq:dQdtdb} with the choice $\dot{n}_{ion}=\dot{N}_{IGM}$, as originally proposed by \citep{MadauEtAl1999}. Also plotted in Figure \ref{Qeffv3} is $Q(sim)$ (blue line) and our best agreeing model (green line). The red line ignores the $\delta_b$ correction, and deviates to the high side of $Q(sim)$ almost immediately, for reasons we discussed earlier. It crosses $Q=1$ at $z\approx 6.6$, which is too early by $\Delta z =0.8$. The teal line includes the $\delta_b$ correction, and tracks the $Q(sim)$ closely to $z \approx 7$, and thereafter deviates on the high side. It crosses $Q=1$ at $z\approx 6.4$, which is too early by $\Delta z =0.6$. Both curves show an accelerated change in $Q$ as z decreases, which is characteristic of standard analytic ionization models (e.g., Haardt \& Madau 2012, Fig.14). By contrast, the simulation and our best fit model using $\dot{n}_{ion}=\dot{N}_t$ show a decelerated change in $Q(z)$ as $Q \rightarrow 1$. This is clearly due to the fact that the ration of ionizations to emitted photons decreases as $Q \rightarrow 1$, as illustrated in Figure 22. The consequence of this flattening in the $Q(z)$ curve is a delay in redshift of overlap of $\Delta z=0.6-0.8$, relative to the predictions of Equations \eqref{eq:dQdtdb} and \eqref{eq:dQdt}, respectively, using the photon injection rate as the source term. We have seen above that the ionization rate density is the appropriate quantity to use to source the $dQ/dt$ equation, independent of $\delta_b$ corrections. Because the ionization rate density is not directly observable, but since $\dot{n}_{ion}$ can be derived from observables, we introduce a correction factor to convert from one to the other. Defining \begin{equation} \gamma \equiv \frac{\langle n_{HI}\Gamma_{HI}^{ph}\rangle}{\dot{n}_{ion}} = \frac{\dot{N}_t}{\dot{n}_{ion}} \label{gamma} \end{equation} \\where the angle brackets denote an average over the singly thresholded volume (IGM), then we can recast Equation \eqref{eq:dQdtdb} into a form useful for observers: \begin{equation} \frac{dQ}{dt} = \frac{\gamma\dot{n}_{ion}}{\delta_b\bar{n}_\mathrm{H}}-\frac{Q}{\bar{t}_{rec}}, \label{eq:dQdtdbg} \end{equation} \\where $\gamma$ and $\delta_b$ are functions of $Q$. In Fig. \ref{RatiovsQfit} we plot data values for $\gamma(Q)$ taken from our simulation, as well as a simple powerlaw fit. The fit is not meant to be definitive, but merely illustrative. More simulations need to be performed under various circumstances, and better fits made, to see whether our $\gamma(Q)$ is approximately universal, or merely anecdotal. \begin{figure} \includegraphics[width=0.5\textwidth]{fig28-eps-converted-to.pdf} \caption{Dependence of analytic models on the choice for $\dot{n}_{ion}$. Red and teal curves assume $\dot{n}_{ion}=\dot{N}_{IGM}$; i.e., the photon injection rate into the IGM. Green curve assumes $\dot{n}_{ion}=\dot{N}_{t}$; i.e., the measured photoionization rate in the IGM. Blue curve is $Q(sim)$--the measured ionized volume filling fraction in the simulation. The green and teal curves take into account the overdensity effect of inside-out reionization (Equation \eqref{eq:dQdtdbg}), while the red curve assumes $\delta_b=1$. All models assume $\bar{t}_{rec}=t_{rec,eff}$ as measured in the simulation (Fig. \ref{treceffvszfit}.} \label{Qeffv3} \end{figure} \begin{figure} \includegraphics[width=0.5\textwidth]{fig29-eps-converted-to.pdf} \caption{Ratio of the volume averaged H {\footnotesize I}~ photoionization rate to photon injection rate in the IGM as a function of $Q$. Data points are measured from the simulation; line is a simple powerlaw fit. } \label{RatiovsQfit} \end{figure} \section{General Results} \label{GeneralResults} \begin{figure*}[!tp] \begin{minipage}[h]{0.5\linewidth} \centering \includegraphics[trim = 15mm 5mm 0mm 15mm, clip, width=1.0\textwidth]{fig1a-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.5\linewidth} \centering \includegraphics[trim = 15mm 5mm 0mm 15mm, clip, width=1.0\textwidth]{fig1b-eps-converted-to.pdf} \end{minipage} \\ \begin{minipage}[h]{0.5\linewidth} \centering \includegraphics[trim = 15mm 5mm 0mm 15mm, clip, width=1.0\textwidth]{fig1c-eps-converted-to.pdf} \end{minipage} \hspace*{-4.00mm} \begin{minipage}[h]{0.5\linewidth} \centering \includegraphics[trim = 15mm 5mm 0mm 15mm, clip, width=1.0\textwidth]{fig1d-eps-converted-to.pdf} \end{minipage} \caption{H {\footnotesize I} density on slices through the 20 Mpc volume showing the growth, percolation, and final overlap of H II regions. Panels show $z=9.18, 8.0, 7.0, 6.1$. The box becomes fully ionized at $z=5.8$ as the last neutral islands are overrun by the I-fronts. Regions of extremely low H {\footnotesize I} density are shock-heated bubbles due to supernova feedback.} \label{HI_slices} \end{figure*} Here we first present the basic properties of the simulation before delving into specific topics in subsequent sections. The star formation and feedback parameters for this simulation are $f_* =0.1, f_{m*}=0.25, \epsilon_{SN}=10^{-5}, \epsilon_{UV}=1.38 \times 10^{-4}$. Figure \ref{HI_slices} shows the reionization process as it proceeds through growth, percolation, and final overlap of ionized hydrogen (H {\footnotesize II}) regions driven by ionizing radiation from star forming galaxies. We plot the neutral hydrogen (H {\footnotesize I}) density on a slice through the densest cell in the volume at redshifts $z=9.18, 8.0, 7.0, 6.1$. At $z=9.18$ several isolated quasi-spherical I-fronts are intersected by the slice plane. These grow and have begun to merge by $z=8.0$. By $z=7.0$ the toplogy is beginning to invert, in that there are now isolated peninsula of H {\footnotesize I} gas embedded in an otherwise ionized IGM. By $z=6.1$ the remaining neutral island has almost disappeared as it is being irradiated from all sides. We can also see in the figure small patches of extremely low H {\footnotesize I} density; these correspond to bubbles of shock heated gas near galaxies heated to above $10^6$K in temperature by supernova feedback. \begin{figure} \includegraphics[width=0.5\textwidth]{fig2-eps-converted-to.pdf} \caption{Evolution of the ionized volume fraction versus redshift for hydrogen ionized to less than 1 neutral in 10$^3$ atoms. As redshift decreases, the volume filling fraction grows rapidly until around redshift of 6, at which time the rate of growth slows significantly as the last neutral island is ionized . The sensitivity of this curve to ionization level is discussed in \S\ref{QuantitativeLanguage}.} \label{Ion1E3} \end{figure} Figure \ref{Ion1E3} plots the evolution of the ionized volume fraction $Q_\mathrm{H\,II}$ versus redshift. Here a cell is called ionized if $\rho_\mathrm{H\,II}/\rho_\mathrm{H} \geq 0.999$ (In \S\ref{QuantitativeLanguage} we discuss the sensitivity of this curve to level of ionization.) The first ionizing sources turn on at $z \sim 10$ in this simulation. The ionized volume fraction rises rapidly, reaching 0.5 at $z \approx 6.8$, 0.95 at $z \approx 6.0$, and near unity at $z \approx 5.8$. We compare this evolution with the predictions of the simple analytic model introduced by \cite{MadauEtAl1999} in \S\ref{Qdot}. For now we only draw attention to the flattening of the curve in the redshift interval $5.8 \leq z \leq 6$. This is the signature of neutral islands being ionized by I-fronts converging in 3D, as opposed to being ionized by internal sources. Our simulation was not designed to complete reionization by a certain fiducial redshift. Rather we adjusted our star formation efficiency parameter $f_*$ so that we can approximately match the star formation rate density (SFRD) in \citep{BouwensEtAl2011}. Our SFRD is shown in Figure \ref{SFR}, along with the Bouwens data, plotted without error bars. For reference we also include the fitting function described in \citep{HaardtMadau2012}. This shows that our simulated universe is one that produces approximately the same amount of stars in a given comoving volume, albeit a bit low relative to the data. We also note that the SFRD begins to flatten out at $z \approx 6.5$, and even turns over after overlap at $z \approx 5.8$, rather than continue to rise as indicated by the data points. This is an artifact of the small box size as a simulation completed in a 80 Mpc comoving on a side box with identical physics, mass, and spatial resolution and star formation/feedback parameters does not show this slowing down of the SFRD. This will be reported on in a future paper. \begin{figure} \includegraphics[width=0.5\textwidth]{fig3-eps-converted-to.pdf} \caption{A comparison of simulated and observed star formation rate densities (SFRD) in units of M$_\odot$yr$^{-1}$Mpc$^{-3}$ comoving. Blue curve labeled ``This Work'' is from our 20 Mpc / $800^3$ simulation, and ``Bouwens et al 2011'' are observationally derived data points from \cite{BouwensEtAl2011b} plotted without error bars. The leveling off of the simulated SFRD is an artifact of the small volume as a simulation carried out with identical physics, mass, and spatial resolution but in 64 times the volume does not show this effect.} \label{SFR} \end{figure} To check and make sure that our simulation is giving us a fair representation of the universe, we plot several more quantities and look for any anomalies. In Figure \ref{HMF}, we see that our halo mass function at redshift of $z \sim 6$ matches well with the Warren fit implemented in \texttt{yt} \citep{WarrenEtAl2006,TurkEtAl2011}. The mass function captures haloes down to $\sim$10$^8$M$_\odot$, which as previously stated was a simulation design criterion. The haloes are found by first running the parallelHOP halo finder installed in \texttt{yt} \citep{SkoryEtAl2010}, then taking the linked list of dark matter particles for each halo and wrapping the region around them in an ellipsoidal 3D container introduced in \texttt{yt} 2.4. The 3D container enables the query of the fluid quantities of the haloes, such as baryonic, emissivity, radiation contents in addition to the particle information. Since the dark matter particles used are $\sim 5 \times$ 10$^5$M$_\odot$, the 10$^8$M$_\odot$ dark matter haloes are considered to be resolved \citep{TrentiEtAl2010}. \begin{figure} \includegraphics[width=0.5\textwidth]{fig4-eps-converted-to.pdf} \caption{The dark matter halo mass function from our simulation (blue line). Green line is the fit from \citep{WarrenEtAl2006}. Our low-mass HMF is reasonably complete down to $M_{halo} \approx 10^8 M_{\odot}$; i.e. halos believed to form stars efficiently due to atomic line cooling. Incompleteness at the high mass end is due to the limited volume sampled.} \label{HMF} \end{figure} As a final check that our ionizing source population is not wildly unrepresentative of the observed universe, in Figure \ref{scaledLF} we plot the luminosity function of our simulated galaxies at $z=6.1$ along side the observational data points from Table 5 of \citep{BouwensEtAl2007}. The points in red are the bolometric luminosities for our galaxy population calculated directly from the $z=6.1$ halo catalogue. To calculate the luminosity of a given halo we sum the emissivity field within the 3D ellipsoidal containers defined by the halos' dark matter particles. Our error bars are taken using one standard deviation of luminosity in the mass bins. Although this is not proof that our simulation is matching observations exactly, it does lend support that our realization of reionization is being driven by sources not too dissimilar to those observed and is sufficient for the purposes of this study. \begin{figure} \includegraphics[width=0.5\textwidth]{fig5-eps-converted-to.pdf} \caption{Bolometric luminosity function derived from our simulation data (red), compared with observational data points (blue) from \citep{BouwensEtAl2007}.} \label{scaledLF} \end{figure} \subsection{Quantitative Language} \label{QuantitativeLanguage} Earlier works on reionization such as \cite{ValageasSilk1999,Gnedin2000,MiraldaEscudeEtAl2000,IlievEtAl2006} speak of a two phase medium composed of completely neutral and completely ionized hydrogen gas, while more recent works \citep{CiardiEtAl2003,ZahnEtAl2007,ShinEtAl2008,PetkovaSpringel2011a,FinlatorEtAl2012} begin to consider the {\em degree of ionization} within ionized gas. The simplification of considering a two phase medium helps reduce the simulation complexity and the language needed to describe the results. However, as simulations become more sophisticated, the two phase paradigm becomes ill-suited to convey the wealth of information contained in the larger and more detailed simulations. As people begin to describe the new simulations, the old paradigm lingers and causes ambiguities. As a case in point, consider the ionized volume filling fraction versus redshift, one of the simplest quantitative metrics of any reionization simulation. Within the framework of a two-phase medium, this is uniquely defined at any redshift. For a simulation such as ours which tracks the ionization state in every cell, the volume filling fraction depends on the degree of ionization, as illustrated in Figure \ref{linearIonized}. This figure shows the evolution of the volume filling fraction of ionized gas which exceeds a minimum local ionization fraction $f_i \equiv \rho_\mathrm{H\,II}/\rho_\mathrm{H}$. The three thresholds are $f_i=$ 0.1, 0.999, and 0.99999 and are labelled 10\%, 1E3, 1E5, respectively in Figure \ref{linearIonized}. We choose three specific levels not because we think they are more important than others, but because it suits our later narrative and gives a range values. With the ionization state tracked by the simulation, we see that it is now ambiguous to ask at what redshift 50\% of the volume is ionized. In our simulation this occurs at $z \approx$ 7, 6.8 and 6.5 for $f_i$=0.1, 0.999, and 0.99999, respectively. \begin{figure} \includegraphics[width=0.5\textwidth]{fig6a-eps-converted-to.pdf} \includegraphics[width=0.5\textwidth]{fig6b-eps-converted-to.pdf} \caption{Volume filling fraction of ionized gas versus redshift for three ionized fraction thresholds. {\em Top} linscale; {\em Bottom} logscale. The three ionization levels are ``10\%'' in blue: fractional volume that have more than 1 ionized hydrogen atom per 10 hydrogen atoms. ``1E3'' in green: fractional volume that have less than 1 neutral hydrogen atom per 10$^3$ hydrogen atoms. ``1E5'' in red: fractional volume that have less than 1neutral hydrogen atom per 10$^5$ hydrogen atoms.} \label{linearIonized} \end{figure} In the rest of this paper we will often report results as a function of these three ionization fraction thresholds. To make the text easier to read we will use the terms ``Ionized'' to designate $f_i$=0.1, ``Well Ionized'' to designate $f_i$=0.999, and ``Fully Ionized'' to designate $f_i$=0.99999 ionization levels. \subsection{Inside-out or Outside-in} \label{IOOI} Besides specifying the amount of ionized volume and levels of ionization, another area where quantitative language is useful is in the description of the reionization history. Since the Outside-in model was proposed by \cite{MiraldaEscudeEtAl2000}, there is gathering support for the opposing view of the Inside-out model by \cite{SokasianEtAl2003,FurlanettoEtAl2004,IlievEtAl2006} to name a few. In \cite{FinlatorEtAl2009}, the authors go even further and add to the lexicon ``Inside-outside-middle'', trying to describe the rich detail in a reionization scenario. The basic Inside-out picture is that galaxies form in the peaks of the dark matter density field and drive expanding H {\footnotesize II} regions into their surroundings ({\em expansion phase}). These H {\footnotesize II} regions are initially isolated, but begin to merge into larger, Mpc-scale H {\footnotesize II} regions due to the clustering of the galaxy distribution ({\em percolation phase}). Driven by a steadily increasing global star formation rate and recombination time (due to cosmic expansion) this process goes on until H {\footnotesize II} regions completely fill the volume ({\em overlap phase}). In this picture, rare peaks in the density field ionize first while regions of lower density ionize later from local sources that themselves formed later. To investigate how reionization progresses in regions of different density, we plot in Figure \ref{NeutralPhase} the hydrogen neutral fraction ($\rho_\mathrm{H\,I}/\rho_\mathrm{H}$) versus overdensity $\Delta_b\equiv\rho_b/\langle\rho_b\rangle$ in the left column, and in the right column a slice of the gas temperature, with redshift decreasing from top to bottom. One would expect if inside-out ionization is the case, that the neutral fraction of higher density region should drop down more quickly than lower density regions. Below, we will describe each row of the figure in more detail. Looking at the redshift $z=10$ row, we see in the gas temperature slice that two isolated regions of ionization appear due to UV feedback from new stars, indicated by the $\sim$10$^4$K gas . These regions correspond to places on the neutral fraction vs. overdensity phase plot where a small amount of volume emerges around $\Delta_b$ of $10^{-1}-10^1$, reaching Well Ionized to Fully Ionized levels. The T $\sim$10$^7$K region corresponds to the extended tail of very low neutral fraction gas in the left column, and indicates gas shock heated by supernova feedback. Although the cell count of shock heated gas will grow, it remains orders of magnitude smaller compared to the photoionized regions that we will emphasize. Even at this early stage, there are high density regions above $\Delta_b$ of 10$^2$-10$^3$ that are Well Ionized; this is due to their close proximity to the ionizing sources, supporting the Inside-out paradigm. Looking at the next row of figures at a redshift of $z=7$, we see that the volume of Well Ionized regions has increased greatly, and so has the shock heated region in the phase plot. We also see that most, but not all the $\Delta_b > 10^2$ cells have reached the Well Ionized level. Although a large portion of the volume is in the Well Ionized regime, the majority of the volume (the red pixels) is still neutral, as we can see in the corresponding temperature slice plot. Most of the volume is still well under 10$^4$K, where we expect the temperature to hover around once the ionization front has passed through the region and the gas has had time to come into photoionization thermal eqilibrium. By a redshift of $z=6.1$, we see from the left column that the region that is ionized beyond the Fully Ionized level (an irony in terms, which means there is definitely room for improvement in the naming convention), dominates the simulation volume. There are still some regions not yet consumed by the ionization front, that is seen on the top of the neutral fraction plot and on the right according to the temperature slice. The next row at redshift of $z=5.5$ is after the entire volume has been swept over by ionization fronts. Most of the volume is beyond the Well Ionized level, except for a few cells around $\Delta_b \sim 10^2$. There are also some cells that are still neutral around $\Delta_b \sim 10^4$. They remain neutral because their densities are so high, leading to high recombination rates. Over time these cells will shift up and down the neutral fraction plot with waves of star formation and supernova explosions since they are likely close to the source of the radiation and kinetic energy. The last row of Figure \ref{NeutralPhase} is at redshift $z\sim5$, where we can see that the previous few cells that have yet to reach Well Ionized levels around $\Delta_b \sim 10^2-10^3$ have now disappeared. The cells that have not reached Well Ionized level before are cells where either the radiation is not strong enough due to shielding effects or the density is so high the gas recombines quickly even after being ionized. After the ionization front has passed though and highly ionized the IGM, there is little material left to shield against the radiation background and we see all but the densest few cells become Well Ionized. The high density region reaching the same ionization level after the under dense void, would fit well with the description for the Outside-in model. Note, that the remaining cells that finally reached Well Ionized levels, are orders of magnitude smaller in total volume compare to the rest of the cells at the same density. So if we call cells of $\Delta_b \sim 10^2$ filaments, not all dense filaments get Well Ionized until late in the EoR. Before the volume is filled with radiation, these dense filaments are able to remain relatively neutral. Unfortunately, the evolution of these redshift panels is not enough to capture the propagation of radiation fronts from the initial sources, but they do convey the overall ionization history of the universe. The panels suggest that the region surrounding the ionization sources, whether they are dense cores, filaments, or voids, are all affected by the radiation on roughly the same time scale. However, the degree to which they are ionized is different. It is this difference, that is the key to answering the original question, whether the universe ionize inside-out or outside-in. When focusing on the ionization of the IGM, lets for a moment neglect the $\Delta_b \sim10^4$ cells that shift ionization level with waves of star formation which comprise a tiny fraction of the volume. If we use the ``Ionized'' level to characterize something as completely ionized and draw the line for neutral fraction at 10\%, then the universe reaches end of EoR before $z \sim 5.5$. Since radiation propagates from sources outward, that would correspond to the Inside-out picture. If we were to instead draw the completely ionized line at ``Well Ionized'' level, then we can see that even at $z \sim 5.5$, there is a small peak in the dense region of the phase diagram ($\Delta_b \sim 2.4\times10^2$) that has yet to reach below the line to be considered completely ionized. This would correspond to the Outside-in picture which reaches end of EoR sometime before $z \sim 5$ (or Inside-outside-middle if one uses the \cite{FinlatorEtAl2009} terminology and considers the neutral peak to be a part of the filaments). And finally, if we were to draw the line at the ``Fully Ionized'' level, the universe has yet to ionize even for regions that are only 10$\times$ over dense. Thus the ionization history is a story with many perspectives, and it really depends on how the story teller draws the line as to whether Inside-out, Outside-in, or Inside-outside-middle is a better qualitative description. \begin{figure*}[!tp] \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 7mm 9mm 1mm 7mm, clip, width=1.0\textwidth]{fig7a-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7b-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7c-eps-converted-to.pdf} \end{minipage}fi \vspace*{-2.00mm}\\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 7mm 9mm 1mm 7mm, clip, width=1.0\textwidth]{fig7d-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7e-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7f-eps-converted-to.pdf} \end{minipage} \vspace*{-2.00mm}\\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 7mm 9mm 1mm 7mm, clip, width=1.0\textwidth]{fig7g-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7h-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7i-eps-converted-to.pdf} \end{minipage} \vspace*{-2.00mm}\\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 7mm 9mm 1mm 7mm, clip, width=1.0\textwidth]{fig7j-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7k-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7l-eps-converted-to.pdf} \end{minipage} \vspace*{-2.00mm}\\ \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 7mm 9mm 1mm 7mm, clip, width=1.0\textwidth]{fig7m-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7n-eps-converted-to.pdf} \end{minipage} \hspace*{-2.00mm} \begin{minipage}[h]{0.33\linewidth} \centering \includegraphics[trim = 10mm 0mm 7mm 7mm, clip, width=1.0\textwidth]{fig7o-eps-converted-to.pdf} \end{minipage} \caption{{\em Left}: Phase diagram of neutral hydrogen fraction versus baryon overdensity with decreasing redshift from top to bottom. {\em Middle}: Slices of Log Temperature [K] through a region that remained mostly neutral until just before overlap at redshift of $\sim$5.8. {\em Right}: Slices of neutral hydrogen fraction through the same region as before. Please refer to \S\ref{IOOI} for detailed description.} \label{NeutralPhase} \end{figure*}
\section{Introduction} Faddeev and Niemi introduced new variables for the SU(2) Yang-Mills theory \cite{Faddeev_Niemi_SU2}, which reveal a structure of Faddeev-Skyrme non-linear sigma model \cite{sigma} in the Lagrangian. The non-linear sigma model is expected to have topologically non-trivial excitations: in the static case, one assumes boundary conditions that compactify the domain to $\mathbb{S}^3$, and therefore the sigma field, being a map from $\mathbb{S}^3$ to $\mathbb{S}^2$, is characterized by the third homotopy group of the sphere $\pi_3(\mathbb{S}^2)$. The corresponding topological charge is given by the Hopf integral. It bounds from below the energy functional of the Skyrme-Faddeev model \cite{VK}. This property suggested, that there are topologically non-trivial solutions of the model. Numerical investigations \cite{num} showed that such solutions indeed may exist. We proposed a Faddeev-Niemi type decomposition for the static SU(3) Yang-Mills theory that reveals a structure of a non-linear sigma model in the Lagrangian \cite{FNK}. The non-linear sigma model has been introduced by Faddeev and Niemi as a generalization of the Faddeev-Skyrme model that may be relevant to infrared limit of SU(3) Yang-Mills theory \cite{Faddeev_Niemi_SUN}. In this model the sigma fields take values in ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ and SU(3)/(SU(2)$\times$U(1)) \cite{Faddeev_Niemi_SUN}. Given a map $g:\mathbb{S}^3\to {\rm SU(3)}$, the sigma fields are $$ \mathfrak{n}=g\kappa_3 g^\dagger,\quad \mathfrak{m}= g\kappa_8 g^\dagger, $$ where $\kappa_a=-\frac{\mathrm i}{\sqrt{2}}\lambda_a, a\in\{1,2,\ldots,8\}$ form a basis of su(3), $\lambda_a$ are the Gell-Mann matrices. On the orbits ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ and SU(3)/(SU(2)$\times$U(1)) there are natural symplectic forms, called the Kirillov-Kostant symplectic forms. The pull-backs of those forms via the maps $\mathfrak{n}$ and $\mathfrak{m}$, respectively, are: $$ F=\frac{1}{4\sqrt{2}}(\mathfrak{n}, [d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]),\quad G=\frac{1}{4\sqrt{2}}(\mathfrak{m}, [d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]), $$ where $(C,D)=\Tr{C^\dagger D}$ for $C,D\in \rm{su(3)}$, $[d\mathfrak{n},d\mathfrak{n}]=d\mathfrak{n}^a\wedge d\mathfrak{n}^{b}[\kappa_a,\kappa_b],\, [d\mathfrak{m},d\mathfrak{m}]=d\mathfrak{m}^a\wedge d\mathfrak{m}^{b}[\kappa_a,\kappa_b]$, $\mathfrak{n}=\mathfrak{n}^a\kappa_a,\,\mathfrak{m}=\mathfrak{m}^a \kappa_a$. The forms $F$ and $G$ are closed and since $H^2(\mathbb{S}^3)=0$ they are also exact, i.e. there exist one-forms $A,B \in \Omega^1(\mathbb{S}^3)$ such that $F=dA,\,G=dB$. Shabanov showed, that the energy functional of the non-linear sigma model is bounded from below by a functional $Q_{\rm S}$ \cite{Shabanov}: $$ Q_{\rm S}[\mathfrak{n}]=\frac{1}{4 \pi^2}\int_{\mathbb{S}^3} (A\wedge F + B\wedge G). $$ Note, that we do not write the dependence on $\mathfrak{m}$. It is because $\mathfrak{m}$ is completely determined by $\mathfrak{n}$. For more details on this issue, see the beginning of the section \ref{sc:charge_chiral_fields}. The problem is that the functional $Q_{\rm S}$ is not homotopy invariant and its value can be an arbitrary real number. The problem is illustrated on the following examples. Let $U:\mathbb{S}^3\to \rm{SU(2)}$ be a map whose degree is 1, i.e. $$ \frac{1}{24\pi^2}\int_{\mathbb{S}^3}\Tr{U^{-1}dU\wedge U^{-1}dU\wedge U^{-1}dU}=1. $$ Consider a field $g_1:\mathbb{S}^3\to \rm{SU(3)}$ defined by $$ (g_1)\indices{^i_j}=\frac{1}{2}\Tr{\sigma_i U \sigma_j U^\dagger}, $$ where $\sigma_i$ are the Pauli matrices. It is easy to calculate, that the Wess-Zumino invariant for this map is: $$ Q_{\rm WZ}[g]=\frac{1}{24\pi^2}\int_{\mathbb{S}^3}\Tr{g^{-1}dg\wedge g^{-1}dg\wedge g^{-1}dg}=4. $$ It means in particular that the map $g_1$ is not homotopic to a constant map. On the other hand $$ Q_{\rm S}[\mathfrak{n}_1]=0,{\rm\ where\ } \mathfrak{n}_1=g_1\kappa_3 g_1^\dagger\,(\mathfrak{m}_1=g_1\kappa_8g_1^\dagger). $$ The Shabanov functional takes the value zero also on the constant map. From the long exact sequence for fibration (see e.g. \cite{BT}): \begin{eqnarray*} \ldots\to\pi_3({\rm U(1)}\times {\rm U(1)})\to \pi_3({\rm SU(3)}) \xrightarrow{\pi_*} \pi_3({\rm SU(3)/(U(1)}\times{\rm U(1))})\to \pi_2({\rm U(1)}\times {\rm U(1)})\to \ldots \\ \ldots\to \pi_0({\rm SU(3)}) \to \pi_0({\rm SU(3)/(U(1)}\times{\rm U(1))})\to 0 \end{eqnarray*} it is easy to infer, that the homotopy groups $\pi_3({\rm SU(3)})$ and $\pi_3({\rm SU(3)/(U(1)}\times{\rm U(1))})$ are isomorphic, and the isomorphism is given by the projection $\pi:{\rm SU(3)}\to {\rm SU(3)/(U(1)}\times{\rm U(1))}$. This means that $\mathfrak{n}_1$ is not homotopic to a constant map, however the value of the Shabanov functional on both maps is equal 0. Therefore the Shabanov functional is not homotopy invariant. Moreover, a value of the Shabanov functional can be an arbitrary real number. To see this, consider a map $U':\mathbb{S}^3\to \rm{SU(2)}$ of degree 4: $$ \frac{1}{24\pi^2}\int_{\mathbb{S}^3}\Tr{{U'}^{-1}dU'\wedge {U'}^{-1}dU'\wedge {U'}^{-1}d{U'}}=4. $$ Consider also a map $g_{\frac{1}{2}}':\mathbb{S}^3\to \rm{SU(3)}$, $$ g_{\frac{1}{2}}'=\left(\begin{array}{cc} U'&0\\0&1 \end{array} \right). $$ Let $\mathfrak{n}'_{\frac{1}{2}}=g_{\frac{1}{2}}'\kappa_3(g_{\frac{1}{2}}')^\dagger\,(\mathfrak{m}'_{\frac{1}{2}}=g_{\frac{1}{2}}'\kappa_8(g_{\frac{1}{2}}')^\dagger)$. It is easy to check, that $ Q_{\rm WZ}[g_{\frac{1}{2}}']=4$ as well as $Q_{\rm S}[\mathfrak{n}'_{\frac{1}{2}}]=4$. Let us recall, that also $Q_{\rm S}[\mathfrak{n}_1]=4$. Therefore there exists a homotopy $f_t:\mathbb{S}^3 \to \rm{SU(3)}$, $t\in [0,1]$ such that $f_0=g_1,\,f_1=g_{\frac{1}{2}}'$. Calculating the corresponding Shabanov charges at each instant $t$ we obtain a continuous function $t\mapstoQ_{\rm S}^t$ such that $Q_{\rm S}^0=0,\, Q_{\rm S}^1=4$. Therefore for every $r\in [0,4]$ there exists $t_r\in [0,1]$ such that $Q_{\rm S}^{t_r}=r$. This shows that $Q_{\rm S}$ can take any value in the interval $[0,4]$. Obviously, the reasoning can be repeated for a map $U:\mathbb{S}^3\to \rm{SU(2)}$ with arbitrary degree and therefore the possible values of $Q_{\rm S}$ is the whole set of real numbers. Our goal is to construct a functional $Q$, such that for each $g:\mathbb{S}^3\to{\rm SU(3)}$ it satisfies: $$ Q[\pi(g)]=Q_{\rm WZ}[g], $$ where $Q_{\rm WZ}[g]=\frac{1}{24\pi^2}\int_{\mathbb{S}^3}\Tr{g^{-1}dg\wedge g^{-1}dg\wedge g^{-1}dg}$ is the Wess-Zumino invariant. Such functional takes only integer values and defines an isomorphism of groups $\pi_3({\rm SU(3)/(U(1)}\times{\rm U(1))})$ and $\mathbb{Z}$. We obtain this goal by using Novikov's procedure \cite{Novikov}. \section{The cohomology ring of SU(3)/(U(1)$\times$U(1))} A preparatory step in Novikov's procedure \cite{Novikov} is to calculate the cohomology ring of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$. In this section we find the cohomology ring of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ using the theorem 22.1 of the seminal paper by Chevalley and Eilenberg on cohomology of Lie groups \cite{CE}. Let us fix some notation: let $\Omega^p=\bigwedge^p {\rm su(3)}^*$ be the space of skew-symmetric linear functions on su(3), and $\Omega=\bigoplus_p \Omega^p$. From theorem 13.1. of \cite{CE} follows, that the invariant differential k-forms on SU(3)/(U(1)$\times$U(1)) are in 1-1 correspondence with the forms $\omega\in\Omega^p$, such that: \begin{enumerate} \item $\omega(X_1,\ldots, X_p)=0$ if at least one $X_i$ lies in u(1)$\oplus$u(1), \item $\omega([X_1,X],X_2,\ldots, X_p)+\ldots+\omega(X_1,X_2,\ldots, [X_p,X])=0$ for $X\in {\rm u(1)}\oplus {\rm u(1)}$, $X_1,\ldots, X_p\in {\rm su(3)}$. \end{enumerate} A form $\omega\in \Omega^p$ satisfying those conditions is called orthogonal to u(1)$\oplus$u(1). Denote by $\Omega^p_{\rm ort}$ the subspace of such forms. We denote by $d:\Omega^p\to \Omega^{p+1}$ a differential $$ (d\omega) (X_1,\ldots,X_{p+1})=\frac{1}{p+1}\sum_{k<l} (-1)^{k+l} \omega([X_k,X_l],X_1,\ldots,\hat{X_k},\ldots, \hat{X_l},\ldots, X_{p+1}). $$ Let $Z^p=\ker d:\Omega^p\to \Omega^{p+1}$ be the cocycles of dimension $p$, and $B^p=\im d:\Omega^{p-1}\to\Omega^p$ the coboundaries of dimension $p$. Let $Z^p_{\rm ort}=Z^p\cap \Omega^p_{\rm ort}$ and $B^p_{\rm ort}=d \Omega^{p-1}_{\rm ort}$. The relative cohomology group of su(3) mod u(1)$\oplus$u(1) is defined as the quotient space $H^q_{\rm ort}=Z^p_{\rm ort}/ B^p_{\rm ort}$. From theorem 22.1 of \cite{CE} now follows that the cohomology groups $H^p_{\rm ort}$ and the p-th de Rham cohomology group of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$, $H^p_{\rm dR}\left({\rm SU(3)/(U(1)}\times{\rm U(1))}\right)$, are isomorphic. We will now calculate explicitly $H^p_{\rm ort}$. Since $\Omega^1_{\rm ort}=0$, it immediately follows that $H^1_{\rm ort}=0$. We look now for $H^2_{\rm ort}$. Denote by $\mathcal L^a$ the basis dual to $\kappa_a=-\frac{\mathrm i}{\sqrt{2}}\lambda_a$: $\mathcal L^a(\kappa_b)=\delta^{a}_b$. It will be convenient for us to use the Cartan-Weyl basis: $$ \kappa_1^{\pm}:=\frac{1}{\sqrt{2}}(\kappa_1\pm \mathrm i\kappa_2),\quad \kappa_2^{\pm}:=\frac{1}{\sqrt{2}}(\kappa_4\mp \mathrm i\kappa_5),\quad \kappa_3^{\pm}:=\frac{1}{\sqrt{2}}(\kappa_6\pm \mathrm i\kappa_7),\quad \kappa_3,\quad \kappa_8. $$ The dual basis will be denoted analogously by: $\mathcal L^\pm_{j},\ j\in\{1,2,3\}$, $\mathcal L^3$, $\mathcal L^8$. Let us underline, that although we use complex basis, we consider cohomology with real coefficients, i.e. we require that the forms considered are real-valued. Any two-form $\omega\in \Omega^2_{\rm ort}$ is of the following form: $$ \omega= \omega_1\, \mathrm i \mathcal L_1^+\wedge \mathcal L_1^-+\omega_2\, \mathrm i \mathcal L_2^+\wedge \mathcal L_2^- + \omega_3\,\mathrm i \mathcal L_3^+\wedge \mathcal L_3^-, $$ where $\omega_1,\,\omega_2,\,\omega_3\in \mathbb{R}$. The cocycle condition $d \omega = 0$ leads to a condition on $\omega_1, \omega_2, \omega_3$: $$ \omega_1+\omega_2+\omega_3=0. $$ We choose a basis in the cocycle space $Z^2_{\rm ort}$: \begin{equation}\label{eq:basis} \alpha= -\mathrm i \mathcal L_1^+\wedge \mathcal L_1^-+ \frac{\mathrm i}{2}\mathcal L_2^+\wedge \mathcal L_2^-+\frac{\mathrm i}{2}\mathcal L_3^+\wedge \mathcal L_3^-,\quad \beta=\frac{\sqrt{3}}{2} \mathrm i \mathcal L_2^+\wedge \mathcal L_2^- -\frac{\sqrt{3}}{2}\mathrm i \mathcal L_3^+\wedge \mathcal L_3^-. \end{equation} The corresponding forms on SU(3)/(U(1)$\times$U(1)) and SU(3)/(SU(2)$\times$U(1)) are the Kirillov-Kostant symplectic forms. Since $B^2_{\rm ort}=0$, it follows that $H^2_{\rm ort}= Z^2_{\rm ort}=\mathbb{R}\times\mathbb{R}$. The space $\Omega^3_{\rm ort}$ is 2-dimensional. The basis elements are: $$ \rho=\frac{1}{2}\mathcal L_1^+\wedge \mathcal L_2^+ \wedge \mathcal L_3^+ + \frac{1}{2}\mathcal L_1^-\wedge \mathcal L_2^- \wedge \mathcal L_3^- {\rm \ and\ } \sigma=\frac{\mathrm i}{2} \mathcal L_1^+\wedge \mathcal L_2^+ \wedge \mathcal L_3^+ -\frac{\mathrm i}{2} \mathcal L_1^-\wedge \mathcal L_2^- \wedge \mathcal L_3^-. $$ The form $\rho$ is exact $\rho=d(-\frac{\mathrm i}{2} \mathcal L_1^+\wedge \mathcal L_1^-)$ and the form $\sigma$ is not closed: $$ d \sigma = \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_2^+\wedge \mathcal L_2^- + \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_3^+\wedge \mathcal L_3^- +\mathcal L_2^+\wedge \mathcal L_2^-\wedge \mathcal L_3^+\wedge \mathcal L_3^-. $$ Therefore $H^3_{\rm ort}=0$. A 4-form $\omega\in \Omega^4_{\rm ort}$, can be written in the following way: $$ \omega = \omega_{12}\, \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_2^+\wedge \mathcal L_2^- + \omega_{23}\, \mathcal L_2^+\wedge \mathcal L_2^-\wedge \mathcal L_3^+\wedge \mathcal L_3^- + \omega_{31}\, \mathcal L_3^+\wedge \mathcal L_3^-\wedge \mathcal L_1^+\wedge \mathcal L_1^-, $$ where $\omega_{12},\,\omega_{23},\,\omega_{31}\in \mathbb{R}$. It is easy to verify, that $d \omega =0$ for any $\omega\in \Omega^4_{\rm ort}$. Therefore $Z^4_{\rm ort}=\Omega^4_{\rm ort}$, and the basis of $Z^4_{\rm ort}$ is: \begin{eqnarray*} &\alpha\wedge \alpha = \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_2^+\wedge \mathcal L_2^- + \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_3^+\wedge \mathcal L_3^- - \frac{1}{2}\mathcal L_2^+\wedge \mathcal L_2^-\wedge \mathcal L_3^+\wedge \mathcal L_3^-,\\ &\beta\wedge \beta = \frac{3}{2}\mathcal L_2^+\wedge \mathcal L_2^-\wedge \mathcal L_3^+\wedge \mathcal L_3^-,\\ &\alpha\wedge \beta= \frac{\sqrt{3}}{2} \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_2^+\wedge \mathcal L_2^- - \frac{\sqrt{3}}{2} \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_3^+\wedge \mathcal L_3^-. \end{eqnarray*} In particular $d \sigma$ can be expressed in this basis giving: $$ d\sigma= \alpha\wedge \alpha + \beta\wedge \beta. $$ The space $B^4_{\rm ort}$ is 1-dimensional, and spanned by $\alpha\wedge \alpha+\beta\wedge \beta$. Therefore $H^4_{\rm ort}=\mathbb{R}\times\mathbb{R}$. Since $\Omega^5_{\rm ort}=0$, it follows that $H^5_{\rm ort}=0$. The space $\Omega^6_{\rm ort}$ is one-dimensional, spanned by: $$ \alpha\wedge \alpha \wedge \alpha= \mathcal L_1^+\wedge \mathcal L_1^-\wedge \mathcal L_2^+\wedge \mathcal L_2^-\wedge \mathcal L_3^+\wedge \mathcal L_3^-. $$ Therefore $H^6_{\rm ort}= Z^6_{\rm ort}= \Omega^6_{\rm ort}=\mathbb{R}$. As a result the cohomology ring $H_{\rm ort}^*:=\bigoplus_{p=1}^6 H_{\rm ort}^p$ is generated by $[\alpha]$ and $[\beta]$ satisfying: $$ [\alpha]\wedge [\alpha]+[\beta]\wedge[\beta] = [d \sigma]=[0], \quad [\beta]\wedge[\beta]\wedge[\beta]=[0]. $$ The cohomology ring can be also inferred from the fact, that SU(3)/(U(1)$\times$U(1)) is a flag manifold \cite{BT}: $$ H_{\rm dR}({\rm SU(3)/(U(1)}\times{\rm U(1))})=\mathbb{R}\left[[x_1],[x_2],[x_3]\right]\big/\left((1+[x_1])(1+[x_2])(1+[x_3])=1\right). $$ This coincides with the ring we found under the identification: \begin{equation} x_1= \frac{1}{2\pi}\left(\alpha+\frac{\sqrt{3}}{3}\beta \right),\quad x_2= \frac{1}{2\pi}\left(-\alpha+\frac{\sqrt{3}}{3}\beta \right),\quad x_3=-\frac{1}{2\pi} \frac{2\sqrt{3}}{3}\beta. \end{equation} \section{The integral expression for the topological charge} Having found the cohomology ring of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ we can now construct the minimal model of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ and apply Novikov's procedure \cite{Novikov} to obtain the integral expression for the topological charge. \subsection{Sullivan's minimal model} Let us recall the definition of a minimal model of a manifold $M$ \cite{BT}. Let $\mathcal{M}=\bigoplus_{p\geq 0} \mathcal{M}^p$ be a differential graded algebra, and $\Omega^*(M)$ the algebra of differential forms on $M$. We say that an element in $\mathcal{M}$ is decomposable if it is a sum of products of positive elements in $\mathcal{M}$, i.e. $m\in \mathcal{M}^+\cdot \mathcal{M}^+$, where $\mathcal{M}^+=\bigoplus_{p>0}\mathcal{M}^p$. A differential graded algebra $\mathcal{M}$ is called a minimal model of a manifold $M$ if: \begin{itemize} \item $\mathcal{M}$ is free; \item there is a chain map $\Psi:\mathcal{M}\to \Omega^*(M)$ which induces an isomorphism in cohomology; \item the differential of a generator is either zero or decomposable. \end{itemize} The minimal model $\mathcal{M}$ of the manifold SU(3)/(U(1)$\times$U(1)) has: \begin{itemize} \item two generators in dimension 2: $\widetilde{\alpha}, \widetilde{\beta}$, \item one generator in dimension 3: $ \widetilde{\sigma}$, \item one generator in dimension 5: $\tau$. \end{itemize} The generators satisfy: $$ d\widetilde{\sigma}=(\widetilde{\alpha})^2+(\widetilde{\beta})^2,\quad d\tau = (\widetilde{\beta})^3. $$ Let $\overline{\alpha},\ \overline{\beta}, \overline{\sigma}$ denote the differential forms on ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ corresponding to $\alpha,\, \beta, \sigma$. The map $\Psi:\mathcal{M}\to \Omega^*({\rm SU(3)/(U(1)}\times{\rm U(1)})$ is given by: $$ \widetilde{\alpha}\mapsto \overline{\alpha},\quad \widetilde{\beta}\mapsto \overline{\beta},\quad \widetilde{\sigma}\mapsto \overline{\sigma},\quad \tau\mapsto 0. $$ \subsection{Novikov's construction} Since $\dim {\rm SU(3)/(U(1)}\times{\rm U(1))} > \dim \mathbb{S}^3$, the Novikov's construction reduces to the following procedure. Consider an extension $\overline{C}^q(\mathcal{M})$ of $\mathcal{M}$ obtained by adding new free generators $v_\kkt,\, v_\kke \in \overline{C}^q(\mathcal{M})$ whose $d$-operator lies in $\mathcal{M}$: $$ dv_\kkt=\widetilde{\alpha},\quad dv_\kke=\widetilde{\beta}.$$ Let $\widetilde{z}=v_\kkt\,\widetilde{\alpha}+v_\kke\,\widetilde{\beta}-\widetilde{\sigma}$. Clearly $\widetilde{z}$ is a cocycle and therefore represents a cohomology class $[\widetilde{z}]\in H^3(\overline{C}^q(\mathcal{M}))$. Consider a map $\mathbf{n}:\mathbb{S}^3\to{\rm SU(3)/(U(1)}\times{\rm U(1))}$. The map induces a homomorphism: $$ \mathbf{n}^* \psi : \mathcal{M} \to \Omega^*(\mathbb{S}^3). $$ Since $H^q_{\rm dR}(\mathbb{S}^3)=0$ for $q=1,2$, the homomorphism $\mathbf{n}^*\psi$ extends to a homomorphism $$ \psi_\mathbf{n}: \overline{C}^q(\mathcal{M})\to \Omega^*(\mathbb{S}^3), $$ called geometric realization. The extension is not unique, however the resulting topological charge does not depend on the choice of geometric realization. We denote images of $v_\kkt,\,v_\kke,\,\widetilde{\alpha},\widetilde{\beta},\widetilde{z}$ under $\psi_\mathbf{n}$ by $A,B,F,G,Z$. The topological charge is proportional to the integral: $$ Q[\mathbf{n}]\propto\int_{\mathbb{S}^3}( A\wedge F + B\wedge G - Z ). $$ The proportionality factor will be calculated in the next subsection by the requirement, that for each $g:\mathbb{S}^3\to{\rm SU(3)}$ it satisfies: $$ Q[\pi(g)]=\frac{1}{24\pi^2}\int_{\mathbb{S}^3} \Tr{g^{-1}dg\wedge g^{-1}dg \wedge g^{-1}dg}.$$ \subsection{The topological charge} In order to fix the proportionality factor, let us first re-write the Wess-Zumino invariant using the basis $\kappa_j^\pm,\,\kappa_3,\,\kappa_8$ of su(3): $$ Q_{\rm WZ}[g]=\frac{1}{24 \pi^2} \int_{\mathbb{S}^3}\left( \mathfrak L^3\wedge d \mathfrak L^3 + \mathfrak L^8\wedge d \mathfrak L^8 + \mathfrak L^+_1\wedge d \mathfrak L^-_1+\mathfrak L^-_1\wedge d \mathfrak L^+_1+\mathfrak L^+_2\wedge d \mathfrak L^-_2+\mathfrak L^-_2\wedge d \mathfrak L^+_2+\mathfrak L^+_3\wedge d \mathfrak L^-_3+\mathfrak L^-_3\wedge d \mathfrak L^+_3\right), $$ where $\mathfrak L^3:=\Tr{(\kappa_3)^\dagger g^{-1}dg},\,\mathfrak L^8:=\Tr{(\kappa_8)^\dagger g^{-1}dg},\,\mathfrak L^\pm_j:=\Tr{(\kappa^\pm_j)^\dagger g^{-1}dg},\,j=1,2,3$. The proportionality factor is fixed using the following equality: $$ Q_{\rm WZ}[g]=\frac{1}{8 \pi^2} \int_{\mathbb{S}^3}\left( \mathfrak L^3\wedge d\mathfrak L^3+\mathfrak L^8\wedge d\mathfrak L^8-\mathrm i \mathfrak L^+_1\wedge \mathfrak L^+_2\wedge \mathfrak L^+_3 +\mathrm i \mathfrak L^-_1\wedge \mathfrak L^-_2\wedge \mathfrak L^-_3\right)=\frac{1}{4\pi^2} \int_{\mathbb{S}^3}\left( A\wedge F + B\wedge G - Z\right). $$ Therefore we define the topological charge to be: $$ Q[\mathbf{n}]=\frac{1}{4\pi^2} \int_{\mathbb{S}^3} \left( A\wedge F + B\wedge G - Z \right), $$ where $\mathbf{n}$ is a map from $\mathbb{S}^3$ to ${\rm SU(3)/(U(1)}\times{\rm U(1))}$, $$ F= -\mathrm i \mathfrak L_1^+\wedge \mathfrak L_1^-+ \frac{\mathrm i}{2}\mathfrak L_2^+\wedge \mathfrak L_2^-+\frac{\mathrm i}{2}\mathfrak L_3^+\wedge \mathfrak L_3^-,\quad G=\frac{\sqrt{3}}{2} \mathrm i \mathfrak L_2^+\wedge \mathfrak L_2^- -\frac{\sqrt{3}}{2}\mathrm i \mathfrak L_3^+\wedge \mathfrak L_3^- $$ are pull-backs of the forms $\overline{\alpha},\,\overline{\beta}$ with the map $\mathbf{n}$, $dA = F,\, dB=G$ and $$ Z=\frac{\mathrm i}{2} \mathfrak L_1^+\wedge \mathfrak L_2^+ \wedge \mathfrak L_3^+ -\frac{\mathrm i}{2} \mathfrak L_1^-\wedge \mathfrak L_2^- \wedge \mathfrak L_3^- $$ is the pull-back of the form $\overline{\sigma}$ with the map $\mathbf{n}$. \section{Topological charge in terms of sigma fields}\label{sc:charge_chiral_fields} The field variables in the Faddev-Niemi non-linear sigma model \cite{Faddeev_Niemi_SUN, Shabanov,FNK} are two sigma fields: $$ \mathfrak{n}=g \kappa_3 g^\dagger ,\quad \mathfrak{m}= g \kappa_8 g^\dagger. $$ It is therefore instructive to express the forms $F,G,Z$ in terms of those fields. Note, that the field $\mathfrak{m}$ is auxiliary --- it is completely determined by the field $\mathfrak{n}$. Indeed, the field $g$ is determined by $\mathfrak{n}$ up to the U(1)$\times$U(1) phase factor. The columns $g_1(x),g_2(x),g_3(x)\in \mathbb{C}^3$ of the SU(3) matrix $g(x)$ are determined by the eigenvalue equations for the anti-hermitian matrix $\mathfrak{n}(x)$: $\mathfrak{n} g_1=-\frac{\mathrm i}{\sqrt{2}} g_1$, $\mathfrak{n} g_2=\frac{\mathrm i}{\sqrt{2}} g_2$, $g_3 = \overline{g_1}\times \overline{g_2}$ ($g^i_3=\epsilon^{ijk}\overline{g^j_1}\overline{g^j_2}$). The forms $F$ and $G$ coincide with the Kirillov-Kostant symplectic forms on SU(3)/(U(1)$\times$U(1)) and SU(3)/(SU(2)$\times$U(1)) pulled-back with the maps $\mathfrak{n}$ and $\mathfrak{m}$: $$ F=\frac{1}{4\sqrt{2}}(\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]),\quad G=\frac{1}{4\sqrt{2}}(\mathfrak{m},[d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]). $$ We now calculate the form $Z=\mathfrak{n}^* \overline{\sigma}$. We decompose the form $\mathfrak L$ into a form with values in the Cartan subalgebra of su(3) spanned by $\kappa_3$ and $\kappa_8$ and a form with values in a space orthogonal to this subalgebra: $$ \mathfrak L=\mathfrak L_\perp+\mathfrak L_\parallel. $$ One can show, that (see e.g. \cite{FNK}) $$ g \mathfrak L_\perp g^\dagger= \frac{1}{2} \left([\mathfrak{n},d\mathfrak{n}]+[\mathfrak{m},d\mathfrak{m}] \right) $$ Using this formula we have: \begin{eqnarray*} \Tr{g \mathfrak L_\perp g^\dagger \wedge d \left(g \mathfrak L_\perp g^\dagger \right)}=\frac{1}{4} \Tr{\left([\mathfrak{n},d\mathfrak{n}]+[\mathfrak{m},d\mathfrak{m}] \right)\wedge \left([d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}] \right)}=\\=-\frac{1}{4}\left(\mathfrak{n}, [d\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{n},[d\mathfrak{m},d\mathfrak{m}]]\right)-\frac{1}{4}\left(\mathfrak{m}, [d\mathfrak{m},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{m},[d\mathfrak{m},d\mathfrak{m}]]\right), \end{eqnarray*} where $(C,D)=\Tr{C^\dagger D}$ is the SU(3) invariant scalar product in su(3), $[d\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]]=d\mathfrak{n}_a\wedge d\mathfrak{n}_b\wedge d\mathfrak{n}_c [\kappa_a,[\kappa_b,\kappa_c]]$ and similarly for other terms of this form. On the other hand: \begin{eqnarray*} \Tr{g \mathfrak L_\perp g^\dagger \wedge d \left(g \mathfrak L_\perp g^\dagger \right)}=\Tr{g\mathfrak L_\perp g^\dagger\wedge g d \mathfrak L_\perp g^\dagger}+\Tr{g\mathfrak L_\perp g^\dagger\wedge dg g^\dagger \wedge g \mathfrak L_\perp g^\dagger}+ \Tr{g\mathfrak L_\perp g^\dagger\wedge g \mathfrak L_\perp g^\dagger\wedge dg g^\dagger}=\\=\Tr{\mathfrak L_\perp \wedge d \mathfrak L_\perp}+ 2\Tr{\mathfrak L_\perp\wedge \mathfrak L_\perp \wedge \mathfrak L}=\Tr{\mathfrak L_\perp \wedge d \mathfrak L}+ 2\Tr{\mathfrak L_\perp\wedge \mathfrak L_\perp \wedge \mathfrak L}=\Tr{\mathfrak L_\perp\wedge\mathfrak L_\perp\wedge \mathfrak L_\perp}= 6\, \mathfrak{n}^* \overline{\sigma}=6 Z. \end{eqnarray*} Therefore: $$ Z=-\frac{1}{24}\left(\mathfrak{n}, [d\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{n},[d\mathfrak{m},d\mathfrak{m}]]\right)-\frac{1}{24}\left(\mathfrak{m}, [d\mathfrak{m},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{m},[d\mathfrak{m},d\mathfrak{m}]]\right). $$ \section{Summary and discussion} \subsection{Summary} The maps $\mathbf{n}:\mathbb{S}^3\to {\rm SU(3)/(U(1)}\times{\rm U(1))}$ can be characterised by the third homotopy group of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$. We applied Novikov's procedure \cite{Novikov} to obtain the corresponding integral expression for this topological charge. The construction is as follows. We choose specific generators $[\overline{\alpha}]$ and $[\overline{\beta}]$ of the cohomology ring of ${\rm SU(3)/(U(1)}\times{\rm U(1))}$. They are represented by closed forms $\overline{\alpha}$ and $\overline{\beta}$ given in equation \eqref{eq:basis}. The chosen generators satisfy $$[\overline{\alpha}]\wedge[\overline{\alpha}]+[\overline{\beta}]\wedge [\overline{\beta}]=0.$$ Therefore there exists a 3-form $\overline{\sigma}\in\Omega^3({\rm SU(3)/(U(1)}\times{\rm U(1))})$ such that $$ d\overline{\sigma}=\overline{\alpha}\wedge \overline{\alpha}+\overline{\beta}\wedge \overline{\beta}. $$ Let $F,\,G,\,Z$ be pull-backs of the forms $\overline{\alpha},\, \overline{\beta},\, \overline{\sigma}$ with a map $\mathbf{n}:\mathbb{S}^3\to {\rm SU(3)/(U(1)}\times{\rm U(1))}$: $$ F=\mathbf{n}^* \overline{\alpha},\quad G=\mathbf{n}^* \overline{\beta}, \quad Z=\mathbf{n}^* \overline{\sigma}. $$ Those pull-backs can be expressed in terms of sigma fields $\mathfrak{n}=g \kappa_3 g^\dagger,\, \mathfrak{m}=g \kappa_8 g^\dagger$, where $g:\mathbb{S}^3\to \rm{SU(3)}$. The forms $F$ and $G$ coincide with the pull-backs of the Kirillov-Kostant symplectic forms on ${\rm SU(3)/(U(1)}\times{\rm U(1))}$ and SU(3)/(SU(2)$\times$U(1)) via the maps $\mathfrak{n}$ and $\mathfrak{m}$ respectively: $$ F=\frac{1}{4\sqrt{2}}(\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]),\quad G=\frac{1}{4\sqrt{2}}(\mathfrak{m},[d\mathfrak{n},d\mathfrak{n}]+[d\mathfrak{m},d\mathfrak{m}]). $$ The form $Z$ expressed in terms of the fields $\mathfrak{n}$ and $\mathfrak{m}$ is: $$ Z=-\frac{1}{24}\left(\mathfrak{n}, [d\mathfrak{n},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{n},[d\mathfrak{m},d\mathfrak{m}]]\right)-\frac{1}{24}\left(\mathfrak{m}, [d\mathfrak{m},[d\mathfrak{n},d\mathfrak{n}]]+[d\mathfrak{m},[d\mathfrak{m},d\mathfrak{m}]]\right). $$ Being pull-backs of closed forms, the forms $F$ and $G$ are also closed. Since $H^2_{\rm dR}(\mathbb{S}^3)=0$ there exist global one-forms $A,\, B\in \Omega^1(\mathbb{S}^3)$, such that: $$ F=d A,\quad G=d B. $$ Following Novikov \cite{Novikov} we define the topological charge to be $$ Q[\mathbf{n}]=\frac{1}{4\pi^2}\int_{\mathbb{S}^3} \left( A\wedge F + B\wedge G -Z \right). $$ The charge differs from the Shabanov functional by a term $-\frac{1}{4\pi^2}\int_{\mathbb{S}^3}Z$. The charge $Q$ is homotopy invariant and takes only integer values --- in fact, it is an isomorphism of groups $\pi_3({\rm SU(3)/(U(1)}\times{\rm U(1))})$ and $\mathbb{Z}$. \subsection{Discussion} Shabanov showed that the functional $Q_{\rm S}$ bounds from below the energy functional of the Faddeev-Niemi non-linear sigma model \cite{Faddeev_Niemi_SUN,Shabanov,FNK}: $$ E[\mathfrak{n},\mathfrak{m}]=E_2[\mathfrak{n},\mathfrak{m}]+E_4[\mathfrak{n},\mathfrak{m}],\ E_2[\mathfrak{n},\mathfrak{m}]=\int_{\mathbb{R}^3} d^3x \left((\partial_i \mathfrak{n},\partial_i \mathfrak{n})+(\partial_i \mathfrak{m},\partial_i \mathfrak{m})\right),\ E_4[\mathfrak{n},\mathfrak{m}]=\frac{1}{e^2}\int_{\mathbb{R}^3}d^3x\left(F_{ij}F_{ij}+G_{ij}G_{ij}\right). $$ It would be very interesting to check, if the topological charge, that we obtained from the Novikov's procedure, still bounds from below the energy functional. Let us note, that the considerations from \cite{Shabanov} cannot be easily extended to the charge we study here. The bound found by Shabanov (as well as the Vakulenko-Kapitansky bound) strongly relies on an estimate of the form \cite{VK,Shabanov}: $$ |Q_{\rm S}|\leq c (E_2 E_4)^{\frac{2}{3}}. $$ Consider a map $U:\mathbb{S}^3\to \rm{SU(2)}$ of, say, degree 1, and the map $g_1:\mathbb{S}^3\to \rm{SU(3)}$ studied in the introduction: $(g_1)\indices{^i_j}=\frac{1}{2}\Tr{\sigma_i U \sigma_j U^\dagger }$. A simple calculation shows that $E_4=0$, however $Q=4$. Therefore there will be no such estimate for the charge $Q$. On the other hand $E_2\not=0$ and $E_4=0$ means that the configuration of the sigma fields defined by $(g_1)\indices{^i_j}=\frac{1}{2}\Tr{\sigma_i U \sigma_j U^\dagger }$ is not a stationary point of the energy functional (it follows from the virial theorem, see e.g. corollary 2.8 of \cite{JPhD}). Therefore a possible solution to this problem could be to look for an estimate for stationary points only. Another important problem is to interpret this topological charge in the language of knotted strings. In the Faddeev-Skyrme model the sigma field $n$ is a map $n:\mathbb{S}^3\to \mathbb{S}^2$. A pre-image of a regular value $p\in\mathbb{S}^2$ is a closed curve. The Hopf integral can be interpreted as the linking number of pre-images $n^{-1}(p),n^{-1}(q)$ of any two different regular values $p,q\in \mathbb{S}^2$. Those pre-images are also lines of force of the magnetic field corresponding to the pull-back with the map $n$ of the Kirillov-Kostant symplectic form on $\mathbb{S}^2$. Interestingly, a pre-image of a point with the map $\mathfrak{n}_{\frac{1}{2}}=g_{\frac{1}{2}}\kappa_3 g_{\frac{1}{2}}^\dagger$ is a closed curve (or empty set) whereas a pre-image of a point with the map $\mathfrak{n}_1=g_1\kappa_3 g_1^\dagger$ is a set of points (or empty set). Therefore in the Faddeev-Niemi non-linear sigma model the topologically non-trivial excitations could be in principle either knot-like (as in the case of map $\mathfrak{n}_{\frac{1}{2}}$) or point-like (as in the case of map $\mathfrak{n}_1$) or could be a mixture of them. It would be also of interest to extend the ideas to the $\rm{SU(N)}$ case (especially SU(4) since it may be important to non-static Euclidean SU(3) Yang-Mills theory --- see discussion section in \cite{FNK}). We expect that the calculation will be completely analogous to the calculation in the SU(3) case, a technical difficulty will be in finding the form $Z$ and expressing it in terms of sigma fields. \section*{Acknowledgements} The author gratefully acknowledges discussions with L. D. Faddeev. The author would like to thank for the hospitality he received at St. Petersburg Department of Steklov Mathematical Institute. This work was supported by the Foundation for Polish Science International PhD Projects Programme co-financed by the EU European Regional Development Fund.
\chapter{Continuous-time quantum error correction} \section{Introduction}\label{introduction} In the standard theory of quantum error correction, both the noise and the error-correcting operations are represented by discrete transformations. If $\mathcal{B}(\mathcal{H})$ denotes the space of bounded operators over a Hilbert space $\mathcal{H}$, and $\mathcal{H^S}$ is the (finite-dimensional) Hilbert space of the controlled system, we say that the code subsystem $\mathcal{H}^A$ in the decomposition \begin{equation} \mathcal{H^S}=\mathcal{H^A}\otimes\mathcal{H}^B\oplus \mathcal{K}\label{decomposition} \end{equation} is correctable under the completely positive trace-preserving (CPTP) noise map $\mathcal{E}:\mathcal{B}(\mathcal{H}^S)\rightarrow \mathcal{B}(\mathcal{H}^S)$, if there exists a CPTP error-correcting map $\mathcal{R}:\mathcal{B}(\mathcal{H}^S)\rightarrow \mathcal{B}(\mathcal{H}^S)$, such that \begin{gather} \textrm{Tr}_B\{(\mathcal{P}^{AB}\circ\mathcal{R}\circ\mathcal{E})(\sigma)\}=\textrm{Tr}_B\{\sigma\},\label{correctablesystem}\\ \hspace{0.1cm} \textrm{for all }\sigma\in \mathcal{B}(\mathcal{H}^S), \sigma=\mathcal{P}^{AB}(\sigma)\hspace{0.1cm} ,\notag \end{gather} where $\mathcal{P}^{AB}(\cdot)$ denotes the superoperator projector on $\mathcal{B}(\mathcal{H^A}\otimes\mathcal{H}^B)$. This formalism is fundamental for the understanding of preserved information under CPTP dynamics, but it depicts an idealized version of the error-correction process. It represents both the noise and the error-correcting operations as discrete CPTP maps, and assumes that error correction is applied \textit{after} the noise. Such a picture is a good approximation for the case when we are concerned with error correction at a single instant via an operation which is fast on the time scale of the noise, or in the case of repeated error correction with fast operations in a regime where the accumulation of uncorrectable errors can be ignored. In general, however, a full error-correcting operation takes a finite time interval during which the noise process is on. Furthermore, even if we assume that error-correcting operations are instantaneous, deviations from perfect correctability between repeated corrections are unavoidable in any real situation. Thus in the case of non-Markovian dynamics, the system may develop correlations with the environment and the effective error maps between successive corrections need not be completely positive. Therefore, a complete description must take into account the continuous nature of both the decoherence and the error-correction processes. Situations in which both these processes are regarded as continuous in time are the subject of continuous-time quantum error correction (CTQEC). The first CTQEC model was proposed by Paz and Zurek (PZ) \cite{Paz:1998:355} as a method of studying the performance of repeated error correction with fast operations in the presence of Markovian decoherence. Rather than describing the overall evolution as a continuous decoherence process interrupted by instantaneous error-correcting operations at discrete intervals, the authors proposed to model the error-correcting procedure as a continuous quantum-jump process, which allows a description of the evolution of the system in terms of a continuous master equation in the Lindblad form \cite{Lindblad:1976:119}. In this model, the infinitesimal error-correcting transformation that the density matrix of the controlled system undergoes during a time step $dt$ is \begin{equation}\label{basicequation} \rho\rightarrow (1-\kappa dt)\rho + \kappa dt \mathcal{R}(\rho), \end{equation} where $\mathcal{R}(\rho)$ is the completely positive trace-preserving (CPTP) map describing a full error-correcting operation, and $\kappa$ is the error-correction rate. The full error-correcting operation $\mathcal{R}(\rho)$ can be thought of as consisting of a syndrome detection, followed (if necessary) by a unitary operation conditioned on the syndrome. The master equation describing the evolution of a system subject to Markovian decoherence plus error correction is then \begin{equation} \frac{d\rho}{dt}=\mathcal{L}(\rho)+\kappa\mathcal{J}(\rho),\label{errorcorrectionequation} \end{equation} where $\mathcal{L}(\rho)$ is the Lindblad generator describing the noise process, and \begin{gather} \mathcal{J}(\rho)=\mathcal{R}(\rho)-\rho \end{gather} is the quantum-jump error-correction generator. The Lindblad generator has the form \begin{equation} \mathcal{L}(\rho)=-i[H,\rho]+\frac{1}{2}\underset{j}{\sum}\lambda_j(2L_j\rho L_j^{\dagger}-L_j^{\dagger}L_j\rho-\rho L_j^{\dagger}L_j),\label{firstLindblad} \end{equation} where $H$ is a system Hamiltonian and the $\{L_j\}$ are suitably normalized Lindblad operators describing different error channels with decoherence rates $\lambda_j$. For example, the Lindbladian \begin{equation} \mathcal{L}(\rho)= \underset{j}{\sum}\lambda_j(X_j\rho X_j - \rho),\label{Lbitflip} \end{equation} where $X_j$ denotes a local bit-flip operator acting the $j^{\textrm{th}}$ qubit, describes independent Markovian bit-flip errors. The quantum-jump model can be viewed as a smoothed version of the discrete scenario of repeated error correction, in which instantaneous full error-correcting operations are applied at random times with rate $\kappa$. It can also be looked upon as arising from a continuous sequence of infinitesimal CPTP maps of the type \eqref{basicequation}. In practice, such a weak map is never truly infinitesimal, but rather has the form \begin{equation} \rho \rightarrow (1-\epsilon^2)\rho + \epsilon^2 \mathcal{R}(\rho),\label{wm} \end{equation} where $\epsilon \ll 1$ is a small but finite parameter, and the weak operation takes a small but finite time $\tau_c$. For times $t$ much greater than $\tau_c$, the weak error-correcting map (\ref{wm}) is well approximated by the infinitesimal form \eqref{basicequation}, where the rate of error correction is \begin{equation} \kappa = \epsilon^2 /\tau_c. \label{tauc} \end{equation} A weak map of the form \eqref{wm} could be implemented, for example, by a weak coupling between the system and an ancilla via an appropriate Hamiltonian, followed by discarding of the ancilla. The continuous process in such a case corresponds to coupling the system to a stream of fresh ancillas which continuously pump out the entropy accumulated due to correctable errors. A closely related scenario, where the ancilla is continuously cooled in order to reset it to its initial state, was studied by Sarovar and Milburn in Ref.~\cite{Sarovar:2005:012306}. Another possible implementation of the above scheme is via weak measurements and weak unitary operations, as we will see in this chapter. If the set of errors $\{L_j\}$ are correctable by the code, the effect of the described CTQEC procedure is to slow down the rate at which information is lost, and in the limit of infinite error-correction rate (strong error-correcting operations applied continuously often) the state of the system freezes and is protected from errors at all times \cite{Paz:1998:355}. The effect of freezing can be understood by noticing that the transformation arising from decoherence during a short time step $\Delta t$, is \begin{equation} \rho\rightarrow \rho + \mathcal{L}(\rho)\Delta t +\textit{O}(\Delta t^2), \end{equation} i.e., the weight of correctable errors emerging during this time interval is proportional to $\Delta t$, whereas uncorrectable errors (higher-order terms) are of order $\textit{O}(\Delta t^2)$. Thus, if errors are constantly corrected, in the limit $\Delta t \rightarrow 0$ uncorrectable errors cannot accumulate and the evolution stops. The idea of using continuous weak operations for error correction was developed further by Ahn, Doherty and Landahl (ADL) who proposed a scheme for CTQEC based on continuous measurements of the error syndromes and feedback operations conditioned on the measurement record \cite{Ahn:2002:042301}. A continuous measurement is one resulting from the continuous application of weak measurements, i.e., measurements whose outcomes change the state by a small amount \cite{Aharonov:1988:1351, Leggett:1989:2325, Peres:1989:2326, Aharonov:1989:2327, Aharonov:1990:11, Brun:2002:719, Oreshkov:2005:110409, Oreshkov:0812.4682}. As shown in Refs.~\cite{Oreshkov:2005:110409, Oreshkov:0812.4682}, weak measurements can be used to generate any quantum operation and therefore provide a natural tool for approaching the problem of error correction in continuous time. In the ADL scheme, the evolution of the density matrix of the system subject to Markovian noise with Lindbladian $\mathcal{L}$ and continuous-time quantum error correction is described by the stochastic differential equation \begin{gather} d\rho(t)=\mathcal{L}(\rho(t))dt+\frac{\kappa}{4}\sum_l\mathcal{D}[M_l](\rho(t))dt+\frac{\sqrt{\kappa}}{2}\sum_l\mathcal{F}[M_l](\rho(t))dW_l(t)\notag\\ -i\sum_r\lambda_r(\rho(t))[H_r,\rho(t)]dt,\label{ADL} \end{gather} where $\mathcal{D}[A](\rho) = A\rho A^{\dagger} - \frac{1}{2}(A^{\dagger}A\rho+ \rho A^{\dagger}A)$, $\mathcal{F}[A](\rho) = A\rho + \rho A^{\dagger} - \rho \textrm{Tr}[A\rho + \rho A^{\dagger}]$, $M_l$ are the stabilizer generators of the code, $W_l$ are Wiener processes (see Sec.~\eqref{indirectsq}), and $H_r$ are correcting Hamiltonians that are turned on with strength $\lambda_r(\rho)$ dependent on the state of the system. Note that the encoded information is in principle unknown, but the feedback is not conditioned on properties of the state related to the encoded information. Thus in order to estimate the state of the system at the present moment for the purpose of applying feedback, one can assume that the encoded state was initially the maximally mixed state. The parameters $\lambda_r(\rho)$ are chosen so as to maximize the instantaneous increase of the code-space fidelity, and are given by $\lambda_r(\rho)=\lambda\textrm{sgn}\textrm{Tr}([\Pi_c,H_r]\rho)$, where $\lambda$ is the maximum strength of the control Hamiltonians and $\Pi_c$ is the projector on the code subspace. (Here the code is assumed to be a standard stabilizer code.) Following the ADL scheme, a number of variations of this approach were proposed (see, e.g., Refs.~\cite{Ahn:2003:052310, Sarovar:2004:052324,Chase:2008:032304}). All these schemes are to a large extent heuristic, and their workings are not thoroughly understood. The difficulty in rigorously motivating the construction of error-correction protocols based on weak measurements and feedback is that stochastic evolutions are generally too complicated to study analytically. This is further complicated by the large dimension of the Hilbert space of all qubits participating in the code (note that even the problem of controlling a single qubit generally requires numerical treatment \cite{Jacobs:2004:355}). However, numerical simulations have shown that these schemes often lead to a better performance in the presence of continuous noise than the application of strong operations at finite time intervals. Therefore, the use of continuous measurements and feedback seems to offer a promising tool for decoherence control. In this chapter, we will try to understand CTQEC and how to approach the problem of constructing CTQEC protocols by looking at the evolution of the state of the system in an encoded basis in which the subsystem containing the protected information is explicit. We will see that this point of view reduces the problem to that of protecting a known state, and allows for designing CTQEC procedures from protocols for the protection of a single qubit. We will show how the PZ quantum-jump model and the ADL and similar schemes with indirect feedback can be obtained from strategies for the protection of a single qubit based on weak measurements and weak unitary operations. We will also study the performance of CTQEC of the quantum-jump type in the case of Markovian and non-Markovian decoherence. We will show that due to the existence of a Zeno regime in non-Markovian dynamics, the performance of CTQEC can exhibit a quadratic improvement if the time resolution of the weak error-correcting operations is sufficiently high to reveal the non-Markovian character of the noise process. \section{CTQEC in an encoded basis}\label{sectionENCODEDBASIS} As discussed in Chapter~6, correctable information is always contained in subsystems of the system's Hilbert space \cite{Knill:2006:042301, Blume-Kohout:2008:030501}. This means, in particular, that if the information initially encoded in the subsystem $\mathcal{H}^A$ in Eq.~\eqref{decomposition} is correctable after the noise map $\mathcal{E}$, it is \textit{unitarily recoverable} \cite{Kribs:2006:042329}, i.e., there exists a unitary map $\mathcal{U}(\cdot)=U(\cdot)U^{\dagger}$, $U\in\mathcal{B}(\mathcal{H}^S)$, such that \begin{gather} \mathcal{U}\circ\mathcal{E}(\rho\otimes\tau)=\rho\otimes\tau^{\prime } \hspace{0.4cm} \tau^{\prime }\in\mathcal{B}(\mathcal{H}^{B^{\prime }}), \label{unitarilyrecoverable2} \\ \hspace{0.2cm} \text{for all }\rho\in \mathcal{B}(\mathcal{H}^A), \hspace 0.1cm} \tau\in \mathcal{B}(\mathcal{H}^B), \notag \end{gather} where the subsystem $\mathcal{H}^{B'}$ can be different from $\mathcal{H}^{B}$. Complete correction generally requires an additional CPTP map that transforms the operators on $\mathcal{H}^{B^{\prime }}$ into operators on $\mathcal{H}^{B}$. As shown in Ref.~\cite{Oreshkov:2008:022333}, Eq.~\eqref{unitarilyrecoverable2} is equivalent to the condition that the Kraus operators $M_{\alpha}$ of $\mathcal{E}$ satisfy \begin{gather} M_{\alpha}P^{AB}=U^{\dagger}I^{A}\otimes C^{B\rightarrow B'}_{\alpha},\hspace{0.2cm}C^{B\rightarrow B'}_{\alpha}: \mathcal{H}^B\rightarrow \mathcal{H}^{B'}, \hspace{0.2cm} \forall \alpha. \label{conditionKraus1} \end{gather} Observe that if a particular set of error operators $\{M_{i}\}$ is correctable by the code, that is, if any CPTP map whose Kraus operators are linear combinations of $\{M_{i}\}$ is correctable, then there is a common recovery unitary $U$ for all such CPTP maps. Note also that if the identity is among the correctable errors for which the code is designed (this is the case, in particular, for all stabilizer codes), from condition \eqref{conditionKraus1} it follows that the unitary ${U}$ must leave the subsystem $\mathcal{H}^A$ in $\mathcal{H}^A\otimes\mathcal{H}^B$ invariant up to a transformation of the co-subsystem, $\mathcal{H}^B\rightarrow \mathcal{H}^{\tilde{B}}$ ($\textrm{dim}\mathcal{H}^{\tilde{B}}=\textrm{dim}\mathcal{H}^B$). This means that if we change the basis by the unitary map ${U}$, the effect of the error operators $M'_{\alpha}=UM_{\alpha}U^{\dagger}$ in the new basis is \begin{gather} M'_{\alpha}P^{A\tilde{B}}=UM_{\alpha}P^{AB}U^{\dagger}=I^{A}\otimes C^{\tilde{B}\rightarrow B'}_{\alpha},\hspace{0.2cm}C^{\tilde{B}\rightarrow B'}_{\alpha}: \mathcal{H}^{\tilde{B}}\rightarrow \mathcal{H}^{B'}, \hspace{0.2cm} \forall \alpha, \label{conditionKraus} \end{gather} i.e., the errors leave the code subsystem invariant up to a transformation of the co-subsystem. A method of obtaining $U$ can be found in Ref.~\cite{Kribs:2006:042329}. In what follows, we will imagine for concreteness the case of an $[[n,1,r,d]]$ operator stabilizer code. This is a code that encodes $1$ qubit into $n$, has $r$ gauge qubits, and has distance $d$. In the encoded basis defined above, the Hilbert space of all $n$ qubits can be written as \begin{gather} \mathcal{H}^S=\mathcal{H}^A\otimes\bigotimes_{i=1}^{n-r-1}\mathcal{H}^s_i \otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j,\label{decomp} \end{gather} where $\mathcal{H}^A$ is a subsystem which corresponds to the logical qubit, $\mathcal{H}^s_i$ are the subsystems of the \textit{syndrome} qubits, and $\mathcal{H}^g_j$ are the subsystems of the \textit{gauge} qubits. Up to a redefinition of the basis of the syndrome qubits, we can assume that the subspace $\mathcal{H}^A\otimes\mathcal{H}^B$ in Eq.~\eqref{decomposition} corresponds to \begin{gather} \mathcal{H}^A\otimes\mathcal{H}^B=\mathcal{H}^A\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle^s_i \otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j.\label{codespace} \end{gather} We will refer to this subspace loosely as the \textit{code space}, since this is where the state of the system is initialized, but we must keep in mind that the information of interest is contained in the tensor factor $\mathcal{H}^A$ in Eq.~\eqref{decomp}. If each of the syndrome qubits is initialized in the state $|0\rangle$, any correctable error will leave the subsystem $\mathcal{H}^A$ invariant and will only affect the co-subsystem, most generally transforming density operators on $\bigotimes_{i=1}^{n-r-1}|0\rangle^s_i \otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$ into density operators on $\bigotimes_{i=1}^{n-r-1}\mathcal{H}^s_i \otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$. In this basis, an error-correcting operation is simply a map on the syndrome qubits, which returns them to the state $|00...0\rangle$. In the language of stabilizer codes, a measurement of the syndrome is a measurement of the state of all syndrome qubits in the $\{|0\rangle, |1\rangle\}$ basis, and a correcting operation is any operation that effectively realizes a bit flip to those qubits which are in the state $|1\rangle$. If the syndrome qubits are not properly initialized (as for example, after the occurrence of an error), a subsequent error generally would not leave the code subsystem invariant. Most generally, after a system subject to decoherence and error correction evolves for a given time $t$, the state of the system becomes \begin{gather} \rho^S(t)=\alpha(t)\rho^{Ag}(t)\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle\langle 0|^{s}_i+(1-\alpha(t))\widetilde{\rho}^{Ags}(t)+\textrm{cross terms}.\label{densmat} \end{gather} Here $\rho^{Ag}(t)$ is a density matrix on the Hilbert space $\mathcal{H}^A\otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$, $\widetilde{\rho}^{Ags}(t)$ is a density matrix with support on the orthogonal complement $\widetilde{\mathcal{H}}$ of the code space ($\mathcal{H}^A\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle^s_i \bigotimes_{j=1}^{r}\mathcal{H}^g_j\oplus \widetilde{\mathcal{H}}=\mathcal{H}^S$), $\alpha(t)\in[0,1]$ is the code-space fidelity, and ``cross terms'' refers to linear combinations of terms of the form $|\psi_i\rangle\langle\widetilde{\phi}_j|$ and $|\widetilde{\phi}_j\rangle\langle\psi_i|$, where $\{|\psi_i\rangle\}$ is an orthonormal basis of $\mathcal{H}^A\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle^s_i \otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$ and $\{|\widetilde{\phi}_i\rangle\}$ is an orthonormal basis of $\widetilde{\mathcal{H}}$. The density matrix of the logical subsystem is $\mathcal{\rho}^A(t)=\alpha(t)\textrm{Tr}_g(\rho^{Ag}(t))+(1-\alpha(t))\textrm{Tr}_{gs}(\widetilde{\rho}^{Ags}(t))$, where $\textrm{Tr}_g$ denotes partial tracing over the gauge qubits and $\textrm{Tr}_{gs}$ denotes partial tracing over the gauge qubits and the syndrome qubits. This density matrix is a transformed version of the state initially encoded in the code subsystem, where the transformation is the result of accumulation of uncorrectable errors. (Note that any transformation inside the subsystem $\mathcal{H}^A$ in Eq.~\eqref{decomp} is by definition uncorrectable.) Let us see how the density matrix $\rho^A$ changes as a result of the action of the generator of noise during a time step $\Delta t$. Since by assumption the action of the noise generator leaves the code subsystem invariant up to a transformation of the co-subsystem, its effect on the term $\alpha\rho^{Ag}(t)\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle\langle 0|^{s}_i$ in Eq.\eqref{densmat} during a time step $\Delta t$ does not give rise to a non-trivial change in $\rho^A(t)$, but only to a decrease in the code-space fidelity, \begin{gather} \alpha(t)\rightarrow \alpha(t)-\gamma(t)\alpha(t)\Delta t + O(\Delta t^2), \end{gather} where $\gamma(t)\geq 0$ is a parameter which depends on the characteristics of the noise process, such as the rates of different errors, and possibly on the current density matrix of the gauge qubits inside the code space, $\textrm{Tr}_A(\rho^{Ag}(t))$. Note that if the noise is non-Markovian, the leading-order correction to $\alpha(t)$ due to the action of the noise on $\alpha\rho^{Ag}(t)\otimes\bigotimes_{i=1}^{n-r-1}|0\rangle\langle 0|^{s}_i$ is $O(\Delta t^2)$, i.e., $\gamma(t)=0$ (see Sec.~\ref{nM}). The only way errors can arise inside the subsystem $\mathcal{H}^A$ is by the action of the noise mechanism on the other terms in Eq.~\eqref{densmat}. The weight of the second term is $(1-\alpha(t))$, and during a single time step the noise generator can give rise to a change in $\rho^A(t)$ \begin{gather} \rho^A(t)\rightarrow \rho^A(t)+\delta \rho^A(t), \end{gather} where \begin{gather} \parallel\delta \rho^A(t)\parallel \leq B(1-\alpha(t))\Delta t+O(\Delta t^2), \hspace{0.2cm} B\geq 0.\end{gather} The constant $B$ depends on the rate of the noise process, its characteristics and the characteristics of the code. From the positivity of the density matrix $\rho^S$ one can show that the coefficients in front of the cross terms $|\psi_i\rangle\langle\widetilde{\phi}_j|$ and $|\widetilde{\phi}_j\rangle\langle\psi_i|$ are at most $\sqrt{\alpha(1-\alpha)}$ in magnitude, and therefore the change that can result in $\mathcal{\rho}^A$ due to the action of the noise generator on the third term in Eq.\eqref{densmat} is limited by \begin{gather} \parallel\delta \rho^A(t)\parallel \leq C\sqrt{\alpha(1-\alpha(t))}\Delta t+O(\Delta t^2), \end{gather} where $C\geq 0$ is another constant dependent on the characteristics of the noise and the code. Thus we see that the rate of change of the density matrix $\rho^A$ is upper bounded as follows: \begin{gather} \parallel \frac{d\rho^A}{dt}\parallel \leq B(1-\alpha(t))+C\sqrt{\alpha(1-\alpha(t))}. \end{gather} In other words, if we manage to keep $(1-\alpha(t))$ small, we will suppress the rate of accumulation of uncorrectable errors. The goal of continuous-time quantum error correction can thus be understood as that of keeping the state of every syndrome qubit close to the state $|0\rangle$. Notice that a strong error-correcting operation in this basis can be realized by bringing each of the syndrome qubits to the state $|0\rangle$ independently. Therefore, the problem of implementing a strong error-correcting operation in terms of weak operations can be reduced to the problem of implementing the corresponding single-qubit operations via weak single-qubit operations. Of course, this is not the most general way of realizing collective initialization of the syndrome qubits, but it is appealing because it reduces the task to that of addressing several independent qubits individually. We will see, however, that the performance can be enhanced if instead of addressing each of the syndrome qubits individually, we address each syndrome which can be associated with a qubit subspace in the space of the syndrome qubits. This will be discussed in the next section. Here we note that the operations in the original basis can be obtained by applying the inverse of the basis transformation to the operations in the encoded basis. To get an idea of what the transformation between bases looks like, let us consider as an example the three-qubit bit-flip code with stabilizer generated by $\{IZZ, ZZI \}$. This code has logical codewords $|0_L\rangle=|000\rangle$ and $|1_L\rangle=|111\rangle$ and even though it only corrects bit-flip errors and does not have gauge qubits, it captures all the characteristics of non-trivial codes which are pertinent to our discussion. It can be verified that a correcting unitary for this code is $U=U_cCX_{1,2}CX_{1,3}$, where \begin{gather} U_c=X_1\otimes|11\rangle\langle 11|_{23} + I_1\otimes (I_2\otimes I_3-|11\rangle\langle 11|_{23}),\label{Uc} \end{gather} and $CX_{i,j}$ denotes the ``controlled not'' with qubit $i$ being the control and qubit $j$ the target. This unitary transforms the single-qubit bit-flip error operators as \begin{eqnarray} XII&\rightarrow & I\otimes(|00\rangle\langle 11|+|11\rangle\langle 00|)+X\otimes(|01\rangle\langle 10|+|10\rangle\langle 01|),\notag\\ IXI&\rightarrow & I\otimes X\otimes |0\rangle\langle 0|+X\otimes X\otimes |1\rangle\langle 1|,\notag\\ IIX&\rightarrow & I\otimes|0\rangle\langle 0|\otimes X+X\otimes|1\rangle\langle 1|\otimes X. \end{eqnarray} In this basis, when the second and third qubits are in the state $|0\rangle$, the error operators leave the state of the first qubit invariant. Going back to the original basis is achieved by applying the basis transformation backwards, i.e., by applying the unitary $CX_{1,3}CX_{1,2}U_c$. \section{Quantum-jump CTQEC with weak measurements}\label{sectionQJ} \subsection{The single-qubit problem}\label{sqp} In this section we will show how to implement the PZ quantum-jump error-correction scheme (Eq.~\eqref{basicequation}) using weak measurements in the encoded basis. We start with the problem of protecting a single qubit in the state $|0\rangle$ from noise using weak measurements. The state $|0\rangle$ can be thought of as a trivial stabilizer code with stabilizer generated by $Z$. We will first consider the case of Markovian bit-flip decoherence, since this model is simple and provides a good intuition. Later, we will extend the result to general noise models. A Markovian bit-flip process is described by the master equation \begin{equation} \frac{d\rho(t)}{dt} = \gamma ( X \rho X - \rho). \end{equation} where $\gamma$ is the bit-flip rate. The general solution to this equation is \begin{equation} \rho(t) =\frac{1+e^{-2\gamma t}}{2}\rho(0)+\frac{1-e^{-2\gamma t}}{2}X\rho(0)X. \label{solu1} \end{equation} If the system starts in the state $|0\rangle\langle 0|$, without error correction it will decay down the $Z$-axis towards the maximally mixed state. In the language of stabilizer codes, an error-correcting operation for this code consists of a measurement of the stabilizer generator $Z$ followed by a unitary correction. If the result is $|1\rangle$, we apply a bit-flip operation $X$, and if the result is $|0\rangle$, we do nothing. The completely positive map corresponding to this strong error-correcting operation is \begin{equation} \mathcal{R}(\rho)= X |1\rangle \langle 1| \rho |1\rangle \langle 1| X + |0\rangle \langle 0| \rho |0\rangle \langle 0| =|0\rangle \langle 1| \rho |1\rangle \langle 0|+|0\rangle \langle 0| \rho |0\rangle \langle 0|.\label{singlequbitstrongmap} \end{equation} One heuristic approach to making the above procedure continuous is to consider weak measurements of the stabilizer generator $Z$ and weak rotations around the $X$-axis of the Bloch sphere conditioned on the measurement record. This is exactly the approach considered in the feedback procedures of the ADL type, and we will discuss it in Sec.~\ref{indirectsq}. Observe that the transformation \eqref{singlequbitstrongmap} can also be written as \begin{gather} \mathcal{R}(\rho)= |0\rangle \langle +| \rho |+\rangle \langle 0|+|0\rangle \langle -| \rho |-\rangle \langle 0|=\notag\\ ZW |+\rangle \langle +| \rho |+\rangle \langle +|WZ + XW|-\rangle \langle -| \rho |-\rangle \langle -|WX, \end{gather} where $|\pm\rangle = (|0\rangle\pm |1\rangle)/\sqrt{2}$ and $W$ is the Hadamard gate. Therefore the same error-correcting operation can be implemented as a measurement in the $\{ |+\rangle, |-\rangle\}$ basis (measurement of the operator $X$), followed by a unitary conditioned on the outcome: if the outcome is $|+\rangle$, we apply $ZW$; if the outcome is $|-\rangle$, we apply $XW$. This choice of unitaries is not unique---for example, we could apply just $W$ instead of $ZW$ after outcome $|+\rangle$. But this particular choice has a convenient geometric interpretation---the unitary $ZW$ corresponds to a rotation around the $Y$-axis by an angle $\pi/2$, $ZW = e^{i\frac{\pi}{2}\frac{Y}{2}}$, and $XW$ corresponds to a rotation around the same axis by an angle $-\pi/2$, $ZW = e^{-i\frac{\pi}{2}\frac{Y}{2}}$. A weak version of the above error-correcting operation can be constructed by taking the corresponding weak measurement of the operator $X$, followed by a weak rotation around the $Y$-axis, whose direction is conditioned on the outcome: \begin{equation} \begin{split} \rho \rightarrow \frac{I+i\epsilon'Y}{\sqrt {1+{\epsilon'}^2}}\sqrt{\frac{I+\epsilon X}{2}}\rho\sqrt{\frac{I+\epsilon X}{2}}\frac{I-i\epsilon'Y}{\sqrt {1+{\epsilon'}^2}}+ \\ +\frac{I-i\epsilon'Y}{\sqrt {1+{\epsilon'}^2}}\sqrt{\frac{I-\epsilon X}{2}}\rho\sqrt{\frac{I-\epsilon X}{2}}\frac{I+i\epsilon'Y}{\sqrt {1+{\epsilon'}^2}}.\label{singlequbitweakmap} \end{split} \end{equation} Here $\epsilon$ and $\epsilon'$ are small parameters. Note that the fact that we describe the net result of the transformation by a CPTP map means that after we apply feedback, we discard information about the outcome of the measurement, or rather, we do not condition any future operations on that information and therefore the transformation of the average density matrix during a single time step is given by Eq.~\eqref{singlequbitweakmap}. Such a scheme is said to be based on \textit{direct} feedback, i.e., the feedback Hamiltonian depends only on the outcome of the most recent measurement, which does not require information processing of the measurement record. Generally, discarding information leads to suboptimal protocols, and we will discuss the possibility of improving that scheme in Sec.~\ref{indirectsq}. From the symmetry of the map \eqref{singlequbitweakmap} it can be seen that if the map is applied to a state which lies on the $Z$-axis, it will keep the state on the $Z$-axis. Whether the state will move towards $|0\rangle\langle 0|$ or towards $|1\rangle\langle 1|$, depends on the relation between $\epsilon$ and $\epsilon'$. Since our goal is to protect the state from drifting away from $|0\rangle\langle 0|$ due to bit-flip decoherence, for now we will assume that the state lies on the $Z$-axis in the northern hemisphere. We would like, if possible, to choose the relation between the parameters $\epsilon$ and $\epsilon'$ in such a way that the effect of this map on any state on the $Z$-axis to be to move that state towards $|0\rangle\langle 0|$. In order to calculate the effect of this map on a given state, it is convenient to write the state in the $\{|+\rangle,|-\rangle\}$ basis. For a state on the $Z$-axis, $\rho = \alpha |0\rangle\langle 0|+(1-\alpha) |1\rangle\langle 1|$, we have \begin{equation} \rho = \frac{1}{2}|+\rangle\langle +|+ \frac{1}{2}|-\rangle\langle -| + (2\alpha -1)\left(\frac{1}{2}|+\rangle\langle -|+ \frac{1}{2}|-\rangle\langle +|\right).\label{rhopm} \end{equation} For the action of our map on the state \eqref{rhopm} we obtain: \begin{gather} \rho \rightarrow \frac{1}{2}|+\rangle\langle +|+ \frac{1}{2}|-\rangle\langle -| \notag\\+ \frac{(1-{\epsilon'}^2)\sqrt{1-\epsilon^2}(2\alpha -1) + 2\epsilon\epsilon'}{1+{\epsilon'}^2}\left(\frac{1}{2}|+\rangle\langle -|+ \frac{1}{2}|-\rangle\langle +|\right).\label{transf} \end{gather} Thus we can think that upon this transformation the parameter $\alpha$ transforms to $\alpha'$, where \begin{equation} 2\alpha'-1 =\frac{(1-{\epsilon'}^2)\sqrt{1-\epsilon^2}(2\alpha -1) + 2\epsilon\epsilon'}{1+{\epsilon'}^2}.\label{alpha'} \end{equation} If it is possible to choose the relation between $\epsilon$ and $\epsilon'$ in such a way that $\alpha'\geq \alpha$ for every $0\leq \alpha \leq 1$, then clearly the state must remain invariant when $\alpha = 1$. Imposing this requirement, we obtain \begin{equation} \epsilon = \frac{2\epsilon'}{1+{\epsilon'}^2}, \end{equation} or equivalently \begin{equation} \epsilon'=\frac{1-\sqrt{1-\epsilon^2}}{\epsilon}.\label{varepsilon'} \end{equation} Substituting back in \eqref{alpha'}, we can express \begin{equation} \alpha'-\alpha = \frac{4{\epsilon'}^2}{(1+{\epsilon'}^2)^2}(1-\alpha)\geq 0. \end{equation} We see that the coefficient $\alpha$ (which is the fidelity of the state with $|0\rangle\langle 0|$) indeed increases after every application of our weak completely positive map. The amount by which it increases for fixed $\epsilon'$ depends on $\alpha$ and becomes smaller as $\alpha$ approaches 1. Since we will be taking the limit $\epsilon\rightarrow 0$, we can write Eq.~\eqref{varepsilon'} as \begin{equation} \epsilon'=\frac{\epsilon}{2}+\textit{O}(\epsilon^3).\label{epsilonprime} \end{equation} If we define the relation between the time step $\tau_c$ and $\epsilon$ as in Eq.~\eqref{tauc}, for the effect of the CPTP map \eqref{singlequbitweakmap} on an arbitrary state of the form $\rho = \alpha|0\rangle \langle 0 | +\beta |0\rangle \langle 1| + \beta^* |1\rangle \langle 0| + (1-\alpha)|1\rangle \langle 1|$, $\alpha\in R$, $\beta\in C$, we obtain \begin{gather} \alpha \rightarrow \alpha + (1-\alpha) \kappa \tau_c,\\ \beta \rightarrow \sqrt{1-\kappa \tau_c} \beta = \beta - \frac{1}{2}\kappa \beta \tau_c + O({\tau_c}^2).\label{effectofmapgen} \end{gather} This is exactly the map \eqref{wm} for $\mathcal{R}(\rho)$ given by Eq.~\eqref{singlequbitstrongmap}. We see that for an infinitesimal time step $dt$, the effect of the noise is to decrease $\alpha(t)$ by the amount $\lambda (2\alpha(t)-1) dt$ and that of the correcting operation is to increase it by $\kappa (1-\alpha(t)) dt$. Combining both effects, we obtain the net master equation that describes the evolution of the qubit subject to Markovian bit-flip errors and the quantum-jump error-correction scheme: \begin{equation} \label{equation1} \frac{d\alpha(t)}{dt}=-(\kappa+2\lambda)\alpha(t)+(\kappa + \lambda). \end{equation} The solution is \begin{equation} \alpha(t)=(1-\alpha_*)e^{-(\kappa+2\lambda)t}+\alpha_*, \label{MSQS} \end{equation} where \begin{equation} \alpha_*=1-\frac{1}{2+r}, \label{attractor} \end{equation} and $r=\kappa/\lambda$ is the ratio between the rate of error correction and the rate of decoherence. We see that the fidelity decays, but it is confined above its asymptotic value $\alpha_*$ which can be made arbitrarily close to 1 for sufficiently large $r$. Finally, let us show that this procedure works for any kind of decoherence where the state need not remain on the $Z$-axis at all times. From Eq.~\eqref{effectofmapgen} we see that the effect of a single application of the map to a general state is to transfer a small portion of the $|1\rangle \langle 1|$-component to $|0\rangle \langle 0|$, and to decrease the magnitude of the off-diagonal components by multiplying them by $\sqrt{1-\kappa \tau_c}$. If there is noise, the most general negative effect of a single step of the noise process is to increase the magnitude of $\beta$ and decrease $\alpha$. For a realistic physical map, the amounts by which these components change during a time step $\Delta t$ should tend to zero when $\Delta t \rightarrow 0$. Since ultimately any noise process is driven by a Hamiltonian acting on the system and its environment, this means that for small $\Delta t$, each of these amounts can be upper-bounded by $\gamma_{max} \Delta t$, where $\gamma_{max}$ is some finite positive number. Therefore, if the system is simultaneously subject to decoherence and error correction, $|\beta|$ and $(1-\alpha)$ will not increase above certain values for which the single-step effects of decoherence and error-correction exactly cancel each other. We can upper-bound these quantities by \begin{gather} (1-\alpha)_{max} = \frac{\gamma_{max}}{\kappa},\\ |\beta|_{max} = \frac{2\gamma_{max}}{\kappa}. \end{gather} This means that the state can be kept arbitrarily close to $|0\rangle \langle 0|$ for sufficiently high rates of error correction $\kappa$. In Sec.~\ref{exact} we will see that if the noise is non-Markovian, $(1-\alpha)_{max}$ scales as $\frac{1}{\kappa^2}$ for large $\kappa$! We remark that one way of implementing the weak measurement of the $X$ operator used in this scheme, is by coupling the system qubit to an ancilla qubit prepared in the state $|+\rangle\langle +|$ for a short time, via the Hamiltonian $H_X=-X\otimes Y$ where $X$ acts on the system qubit and $Y$ acts on the ancilla, followed by a measurement of the ancilla in the $\{|0\rangle, |1\rangle\}$ basis (the latter can be destructive). It can be verified that if we first apply the unitary transformation $U_X(\epsilon)=\exp{i\frac{\epsilon}{2}X\otimes Y}$ followed by a measurement of the ancilla, up to second order in $\epsilon$ the resulting measurement on the system is \begin{gather} \rho\rightarrow \frac{\sqrt{\frac{I\pm\epsilon X}{2}}\rho\sqrt{\frac{I\pm\epsilon X}{2}}}{p_{\pm}}, \end{gather} with probabilities $p_{\pm}=\frac{1}{2}(1\pm\epsilon\textrm{Tr}(X\rho))$. Since we are interested in the limit where $\epsilon\rightarrow 0$, only the lowest-order nontrivial contributions to the error-correcting CPTP map are important, and they are of order $\epsilon^2$. \subsection{General codes} How do we extend this approach to general codes? As we mentioned earlier, one way is to simply apply the described operation to each of the syndrome qubits in the encoded basis. According to the argument in the previous subsection, no matter what the exact form of the noise process on the syndrome qubits is, this scheme will keep each of them close to the state $|0\rangle\langle 0|$ within some distance that can be made arbitrarily small for sufficiently large error-correction rates. This in turn would ensure that the code-space fidelity is close to $1$, which would suppress the rate of accumulation of uncorrectable errors as argued in Sec.~\ref{sectionENCODEDBASIS}. This approach is particularly attractive because of its conceptual simplicity and the fact that it involves operations only on each of the syndrome qubits whose number $n-r-1$ is smaller than the number of different nontrivial correctable errors which can be up to $2^{n-r-1}-1$. Furthermore, it is obvious that the operations on the different qubits commute and therefore can be applied simultaneously. However, it is not difficult to see that even though the equivalent infinitesimal map has the form \eqref{basicequation}, the effective $\mathcal{R}(\rho)$ is not equal to the error correcting map for this code, where the latter acts as \begin{gather} \mathcal{R}(\rho^s)= \bigotimes_{i=1}^{n-r-1}|0\rangle\langle 0|^{s}_i \label{Phiofrho} \end{gather} for any state $\rho^s$ of all syndrome qubits. This is because, if we apply error correction separately on the different qubits, up to first order in $dt$ only those terms in which there is one qubit in the state $|1\rangle$ and all the rest are in the state $|0\rangle$ (such as, e.g., $|10...0\rangle\langle 10...0|$) will get mapped to $|00...0\rangle\langle 00...0|$. The full error-correcting map, however, maps all states to the state $|00...0\rangle\langle 00...0|$ and therefore it is more powerful. Is there a way to construct the full map based on the single-qubit operations described in the previous subsection? It turns out that the answer is yes. The idea is to associate an abstract qubit to each non-trivial error syndrome in the code as follows. As was mentioned earlier, each syndrome corresponds to a state of the syndrome qubits of the form $|\nu_1\nu_2...\nu_{n-r-1}\rangle$, where $\nu_i$ can be either $0$ or $1$. Let us label these different syndrome states by $|i_s\rangle$, $i_s=0,...,2^{n-r-1}-1$, with $|0_s\rangle=|00...0\rangle$ being the trivial syndrome corresponding to ``no error''. The density matrix of the entire system can then be written \begin{gather} \rho^S=\alpha(t)\rho^{Ag}(t)\otimes |0_s\rangle\langle 0_s| +\sum_{i_s\geq 1}\beta_{i_s}\widetilde{\rho}^{Ag}_{i_s}(t)\otimes |i_s\rangle\langle i_s|\notag\\ +\sum_{i_s\neq j_s}\sigma^{Ag}_{i_sj_s}(t)\otimes |i_s\rangle \langle j_s|,\label{dmsb} \end{gather} where $\widetilde{\rho}^{Ag}_{i_s}(t)$ are density matrices on $\mathcal{H}^A\otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$, $\beta_{i_s}\geq 0$ are the weights of the state inside the different error subspaces, and $\sigma^{Ag}_{i_sj_s}(t)$ are operators on $\mathcal{H}^A\otimes\bigotimes_{j=1}^{r}\mathcal{H}^g_j$. To each nontrivial syndrome we can associate a qubit subspace of the space of all syndrome qubits, which is spanned by the state $|0_s\rangle$ and the state $|i_s\rangle$ corresponding to that syndrome. Let us take for concreteness one of these qubits---the subspace spanned by $|0_s\rangle$ and $|1_s\rangle$. If we apply the single-qubit operations described in the previous subsection to this subspace while acting trivially on its orthogonal complement, the effect of the resulting operation on the terms $\beta_{1_s}\widetilde{\rho}^{Ag}_{1_s}(t)\otimes |1_s\rangle\langle 1_s|+\sigma^{Ag}_{0_s1_s}(t)\otimes |0_s\rangle \langle 1_s|+\sigma^{Ag}_{1_s0_s}(t)\otimes |1_s\rangle \langle 0_s|$ in Eq.~\eqref{dmsb} will be the same as that of the quantum-jump error correcting map \eqref{basicequation} with $\mathcal{R}(\rho)$ given by Eq.~\eqref{Phiofrho}. At the same time, the effect on the rest of the terms will be trivial. Therefore, if we apply the analogous operation to each of the qubit subspaces spanned by $|i_s\rangle$ and $|0_s\rangle$, we will effectively realize the desired quantum-jump error correcting map. Observe that all these single-qubit maps commute and so do the generators they give rise to in the corresponding continuous quantum-jump equation. If we think of the resulting processes as being driven by the action of the quantum-jump generators, then it is obvious that all of them can be implemented simultaneously. However, if we think of each of these maps as resulting from weak measurements and weak unitary operations as described in the previous subsection, the measurements and unitaries do \textit{not} commute. For example, the $X$ operator for the $j_s^{\textrm{th}}$ qubit has the form $X_{j_s}=|j_s\rangle\langle 0_s|+|0_s\rangle\langle j_s|$, and therefore $[X_{i_s},X_{j_s}]=|i_s\rangle\langle j_s|-|j_s\rangle\langle i_s|$. This means that the measurements of the $X$ operators cannot be implemented simultaneously on all qubits. The same holds for the rotations around the $Y$-axes. Does this mean that we have to apply the different operations in series? This would require the ability to precisely turn on and off, on a very short time scale, the couplings to the external fields needed for the different measurements, which does not correspond to a continuous measurement. It turns out that alternating the different couplings is not needed---the same couplings that one would use for implementing the weak measurements on the individual qubits can be turned on simultaneously, and so can the feedback Hamiltonians that one would use depending on the outcomes of the different measurements. This is because all extra terms that arise from the fact that the operations on the different qubits do not commute, cancel out when we average over the outcomes. We outline how this can be verified using the implementation of the weak measurement via a qubit ancilla described at the end of Sec.~\ref{sqp}. For each of the qubits corresponding to different syndromes, we will need to turn on a different Hamiltonian that couples that qubit to a separate ancilla initially prepared in the state $|+\rangle$. Let us label the ancilla corresponding to the ${j_s}^{\textrm{th}}$ qubit also by $j_s$. If we turn on all of these Hamiltonians simultaneously, the overall Hamiltonian is \begin{gather} H_{meas}=-\sum_{j_s}X_{j_s}\otimes Y^a_{j_s}, \end{gather} where the $Y^a_{j_s}$ act on the different ancilla systems but the $X_{j_s}$ do not act on different systems and do not commute. Imagine that this Hamiltonian acts for time $\frac{\epsilon}{2}$, i.e., it gives rise to the unitary $U=\exp(\frac{i\epsilon}{2}\sum_{j_s}X_{j_s}\otimes Y^a_{j_s})$. At this point we can measure projectively each of the ancillas in the $\{|0\rangle, |1\rangle\}$ basis and turn on the corresponding single-qubit correction Hamiltonians $\xi_{j_s}Y_{j_s}$ where $\xi_{j_s}=\pm 1$ is the sign of the Hamiltonian which depends on the outcome of the measurement, and $Y_{j_s}=i|j_s\rangle\langle0_s|-i|0_s\rangle\langle j_s|$. The overall feedback Hamiltonian is \begin{gather} H_{fb}=\sum_{j_s}\xi_{j_s}Y_{j_s}. \end{gather} One can verify that up to second order in $\epsilon$, the resulting operation after averaging over the outcomes is exactly equal to the quantum jump operation \eqref{basicequation} with $\mathcal{R}(\rho)$ given by Eq.~\eqref{Phiofrho}. The easiest way to see this is to observe that all unwanted terms in the resulting density matrix are proportional to $\xi_{i_s}\xi_{j_s}$, $i_s\neq j_s$, and therefore when we sum over all different outcomes, these terms disappear. To get an idea of what the weak measurements and feedback unitaries mean in the original basis, let us look again at the three-qubit bit-flip code. Observe that the syndrome states $|i\rangle^s$ in the encoded basis are $|1_s\rangle=|10\rangle$, $|2_s\rangle=|01\rangle$, $|3_s\rangle=|11\rangle$, i.e., the three abstract qubits corresponding to these syndromes have $X$ and $Y$ operators \begin{eqnarray} X_{1_s}&=&I\otimes (|10\rangle\langle 00|+|00\rangle\langle 10|+|01\rangle\langle 01|+|11\rangle\langle 11|),\notag\\ X_{2_s}&=&I\otimes (|01\rangle\langle 00|+|00\rangle\langle 01|+|10\rangle\langle 10|+|11\rangle\langle 11|),\notag\\ X_{3_s}&=&I\otimes (|11\rangle\langle 00|+|00\rangle\langle 11|+|01\rangle\langle 01|+|10\rangle\langle 10|), \end{eqnarray} \begin{eqnarray} Y_{1_s}&=&I\otimes (i|10\rangle\langle 00|-i|00\rangle\langle 10|+|01\rangle\langle 01|+|11\rangle\langle 11|),\notag\\ Y_{2_s}&=&I\otimes (i|01\rangle\langle 00|-i|00\rangle\langle 01|+|10\rangle\langle 10|+|11\rangle\langle 11|),\notag\\ Y_{3_s}&=&I\otimes (i|11\rangle\langle 00|-i|00\rangle\langle 11|+|01\rangle\langle 01|+|10\rangle\langle 10|). \end{eqnarray} By applying the inverse basis transformation $CX_{1,3}CX_{1,2}U_c$ with $U_c$ given by Eq.~\eqref{Uc}, we obtain these operators in the original basis: \begin{gather} X^{\prime}_{1_s}=\frac{1}{2}ZXZ+\frac{1}{2}IXI+\frac{1}{2}III-\frac{1}{2}ZIZ,\notag\\ X^{\prime}_{2_s}=\frac{1}{2}ZZX+\frac{1}{2}IIX+\frac{1}{2}III-\frac{1}{2}ZZI,\notag\\ X^{\prime}_{3_s}=\frac{1}{2}XZZ+\frac{1}{2}XII+\frac{1}{2}III-\frac{1}{2}IZZ,\label{Xsprime} \end{gather} \begin{gather} Y^{\prime}_{1_s}=\frac{1}{2}ZYI+\frac{1}{2}IYZ+\frac{1}{2}III-\frac{1}{2}ZIZ,\notag\\ Y^{\prime}_{2_s}=\frac{1}{2}IZY+\frac{1}{2}ZIY+\frac{1}{2}III-\frac{1}{2}ZZI,\notag\\ Y^{\prime}_{3_s}=\frac{1}{2}YZI+\frac{1}{2}YIZ+\frac{1}{2}III-\frac{1}{2}IZZ. \label{Ysprime} \end{gather} We see that implementing the PZ scheme using weak measurements and unitary operations requires the ability to apply Hamiltonians which are complicated sums of different elements of the Pauli group. We will postpone the analysis of the performance of that scheme in the presence of decoherence until Sec.~\ref{exact}. We now turn to look at alternative methods for protecting a single qubit from noise using weak measurements, and their corresponding generalizations to multi-qubit codes. \section{Schemes with indirect feedback}\label{indirect} \subsection{The single-qubit problem}\label{indirectsq} We already mentioned that another way of ``continuization'' of the discrete single-qubit error-correcting map \eqref{singlequbitstrongmap} is to apply continuous measurements of the stabilizer generator $Z$ and rotations around the $X$-axis conditioned on the measurement record. A continuous measurement of the operator $Z$ can be achieved by an infinite repetition of a weak measurement with measurement operators \begin{equation} M^Z_{\pm}(\epsilon)= \sqrt{\frac{I\pm\tanh(\epsilon) Z}{2}}. \end{equation} The evolution of the state of the system under such observation can be described by a random walk along a curve parameterized by $x\in R$. The state at any moment during the procedure can be written in the form \begin{equation} \rho(x) =\frac{M^Z(x)\rho(0)M^Z(x)}{\textrm{Tr}(M^Z(x)\rho(0) M^Z(x))} \end{equation} for some value of $x$, where $M^Z(x) = \sqrt{(I+\tanh(x) Z)/2}$ and $\rho(0)$ is the initial state. After every application of the weak measurement $M^Z_{\pm}(\epsilon)$, the parameter $x$ changes to $x\pm\epsilon$ depending on the outcome. The two projective measurement outcomes of the strong measurement of $Z$ correspond to $x=\pm\infty$. The procedure is continued until $|x|\geq X$ for some $X$ which is sufficiently large that $M^Z(X)\approx |0\rangle\langle 0|$ and $M^Z(-X)\approx |1\rangle\langle 1|$ to any desired precision \cite{Oreshkov:0812.4682}. In the limit when $\epsilon\rightarrow 0$, the evolution of the state of the system can be described by a continuous stochastic differential equation. We can introduce a time step $\delta t$ and a rate \begin{equation} \kappa=\epsilon^2/\delta t. \end{equation} Then we can define a mean-zero increment $\delta W$ as follows: \begin{gather} \delta W=(\delta x-M[\delta x])/\sqrt{\kappa}, \end{gather} where $\delta x=\pm\epsilon$ and $M[\delta x]$ is the mean of $\delta x$, \begin{equation} M[\delta x]=\epsilon (p_+(x)-p_-(x)). \end{equation} Here $p_{\pm}(x)$ are the probabilities for the two outcomes of the weak measurement $M^Z_{\pm}(\epsilon)$ at the point $x$, \begin{equation} p_{\pm}(x)=\frac{1}{2}(1\pm \epsilon\langle Z\rangle_x ), \end{equation} with $\langle Z\rangle_x=\textrm{Tr}(Z\rho(x))$. Note that $M[(\delta W)^2]=\delta t+\textit{O}(\delta t^2)$. Expanding the change of the state under the measurement $M^Z_{\pm}(\epsilon)$ up to second order in $\delta W$, and taking the limit $\delta W\rightarrow 0$ while keeping the rate $\kappa$ fixed, it can be shown that the evolution of the state of the system subject to such a continuous observation is described by the following stochastic differential equation: \begin{equation} d\rho(t) = \frac{\kappa}{4} \mathcal{D}[Z](\rho(t)) dt + \frac{\sqrt{\kappa}}{2}\mathcal{F}[Z](\rho(t)) dW(t). \end{equation} Here $dW(t)$ is a Wiener increment, i.e., a mean-zero normally distributed random variable with variance $dt$. The evolution of the parameter $x$ is given by \begin{equation} dx(t)= \kappa \langle Z \rangle_t dt + \sqrt{\kappa} dW(t), \end{equation} where $\langle Z \rangle_t=\textrm{Tr}(Z\rho(t))$. From $x(t)$ one can define the average measurement current as the mean of $dx(t)/dt$, \begin{gather} I^{ave}_x(t)= \kappa \langle Z\rangle_t.\label{meascurr} \end{gather} If we apply no error correction to our qubit (initially in the state $|0\rangle\langle 0|$), under bit-flip decoherence its state will drift down the $Z$-axis of the Bloch sphere towards the center of the sphere (the maximally mixed state). According to the scheme proposed in Ref.~\cite{Ahn:2002:042301}, at a given moment we apply a weak measurement of the stabilizer $Z$ and a weak rotation around the X-axis, which depends on the state of the system at that moment. In the simplified version of that scheme in Ref.~\cite{Sarovar:2004:052324}, the feedback is condition only on an estimate of the average measurement current. If at a given moment the state is somewhere along the $Z$-axis, i.e., $\rho = \alpha |0\rangle \langle 0| + (1-\alpha) |1\rangle \langle 1|$, $0\leq\alpha\leq 1$, the effect of a weak measurement would be to move the state slightly up or down along the axis depending on the outcome. It is easy to see that the result of such a measurement does not change the value of $\alpha$ on average, because $M^Z_+(\epsilon)\rho M^Z_+(\epsilon) +M^Z_-(\epsilon)\rho M^Z_-(\epsilon) =\rho$. One is then led to ask whether including feedback could improve the average fidelity. The answer depends on whether the state lies in the northern or the southern hemisphere of the Bloch sphere. If the state lies on the $Z$-axis in the northern hemisphere, it is not possible to improve its fidelity by feedback. Assuming that the measurement is sufficiently weak so that the negative outcome $M^Z_-(\epsilon)\rho M^Z_-(\epsilon)/\textrm{Tr}(M^Z_-(\epsilon)\rho M^Z_-(\epsilon))$ is still in the northern hemisphere, no unitary operation can bring any of the two outcomes closer to the north pole since unitary operations preserve the distance from the center. On the contrary, a unitary rotation around the X-axis would move both outcomes away from the $Z$-axis and therefore away from the target. In the ADL scheme there is no risk for the feedback to decrease the fidelity with the target state because the feedback is conditioned on the current state and always tends to increase the fidelity with the code space; if the state lies on the $Z$-axis in the northern hemisphere, no rotation would be applied. However, during initial times that scheme would not be helpful for increasing the average value of $\alpha$ either, because a weak measurement keeps the state on the $Z$-axis in the northern hemisphere. If we go to the continuous limit, $\epsilon\rightarrow 0$, the Wiener parameter is normally distributed and during an infinitesimal time step the state may enter the southern hemisphere, but with a negligible probability. Thus during initial times, the scheme would not be helpful with respect to the average fidelity, and only after the probability for the state to enter the southern hemisphere becomes significant will it start to have an effect. This intuition is confirmed by the numerical simulations of a generalization of this protocol to multi-qubit codes presented in Ref.~\cite{Ahn:2002:042301}. In the scheme in Ref.~\cite{Sarovar:2004:052324}, the feedback is not conditioned on the state but on an estimate of the average measurement current \eqref{meascurr}. The idea is that by filtering the noisy measurement data obtained during some short time interval before a given moment $t$, we can try to obtain an estimate of the average change of $x(t)$ with time at that moment, i.e., an estimate of $\langle Z\rangle_t$. But clearly such an estimate cannot be precise, because it would mean that we could measure the expectation value of an observable almost without disturbing the state. Therefore, any such estimate inevitably carries imprecision. For example, it could be that the state of the system is $|0\rangle\langle 0|$ but we obtain a sequence of negative outcomes which give rise to the effective measurement operator $M^Z(x) = \sqrt{I+\tanh(x) Z}$ with $x<0$. This can occur with finite probability and it would suggest that the state lies in the southern hemisphere, while the state will remain $|0\rangle\langle 0|$ under this measurement. In such a case, this scheme would apply a rotation which would take the state away from the target state, i.e., during short initial times this scheme could have a negative effect. Nevertheless, as time progresses, more and more trajectories enter the southern hemisphere and the scheme may lead to an improvement of the average fidelity with the target state at later times. Indeed, numerical simulations have confirmed the efficiency of this scheme and its generalization to multi-qubit codes in certain parameter regimes \cite{Sarovar:2004:052324}. We point out that the two general strategies for the protection of a qubit that we considered---the one involving continuous measurement of the $X$ operator and direct feedback (the quantum-jump scheme), and the one involving continuous measurements of the $Z$ operator and indirect feedback (the ADL and similar schemes)---strongly resemble two optimal protocols for the purification of a qubit discussed in Refs.~\cite{Jacobs:2004:355} and \cite{Wiseman:2006:90}. In Ref.~\cite{Jacobs:2004:355} it was shown that the fastest increase on average of the purity of a single qubit using weak measurements is achieved if the qubit is measured in a basis perpendicular to the axis in the Bloch sphere that connects the current state with the center of the sphere. If we assume that we can apply fast unitary rotations on the time scale of the measurements, the fastest preparation of a qubit in the state $|0\rangle\langle 0|$ can be achieved by measuring the state in the eigenbasis of $X$, and after every weak measurement apply a rotation around the $Y$-axis that brings this state to the $Z$-axis. This is almost the same as the quantum-jump scheme, except that we did not assume that we can apply an arbitrarily strong and precise rotation that brings each outcome on the $Z$-axis, but only a rotation which would bring the state to the north pole if it was there before the measurement. In Ref.~\cite{Jacobs:2004:355}, on the other hand, it was shown that if we are interested in the \textit{average time} that it would take to purify the qubit to a certain degree, we have to measure it along the axis that connects it with the center of the Bloch sphere. Again, if we assume that we can apply arbitrarily fast rotations, the optimal average time for preparing a qubit in the state $|0\rangle\langle 0|$ with some precision can be achieved if we measure the qubit in the eigenbasis of $Z$ and whenever the qubit enters the southern hemisphere, apply rotations around the $X$-axis that bring it to the northern half of the $Z$-axis. The difference of the ADL scheme from this approach is again that the ADL scheme does not assume infinitely fast and precise rotations. Thus we see that the two competitive error-correction schemes we discussed can be regarded as originating from two optimal protocols for the preparation of a qubit in a known state---one that optimizes the average fidelity with the target state, and another that optimizes the average time to reach the target state. Of course, this does not mean that the two schemes we described are optimal for the resources they use. In the quantum jump scheme, for example, we discard information about the outcome of the measurement after every feedback operation. If we keep this information and estimate the current state, we can in principal improve the performance of the scheme. Let us say that the state is somewhere far from the $Z$-axis. Since each of the outcomes of the weak measurement change the state by a small amount, after either outcome we will have to apply rotations in the same direction in order to bring the state closer to the $Z$-axis. If we do not keep track of the actual state, however, we would apply rotations in opposite directions after the two different outcomes. But it turns out that the improvement we can gain by keeping track of the actual states is small. It can be verified that even if we assume that we are able to apply infinitely fast and precise rotations, i.e., that we can bring the state on the $Z$-axis after every weak measurement outcome, if the measurement strength is fixed, the correction to the quantity $(1-\alpha_*^{\rm M})$ (Eq.~\eqref{attractor}) we can obtain is of order $O((1-\alpha_*^{\rm M})^2)$. But as we argued in Sec.~\ref{sectionENCODEDBASIS} and will discuss further in Sec.~\ref{exact}, this is the quantity that is responsible for the effective decrease of the error rate in a general code. In that sense, the performance of the quantum-jump scheme is very close to optimal when $(1-\alpha_*^{\rm M})$ is small, even though the scheme requires no side information processing. Note, however, that we assumed that at the level of a single weak operation we can ensure a particular relation between the measurement strength and the strength of the correcting rotation---Eq.~\eqref{epsilonprime}. If we cannot apply a sufficiently strong rotation to keep the state $|0\rangle\langle 0|$ invariant, the equilibrium fidelity with the target state $\alpha_*$ would be lower. \subsection{Generalizations to multi-qubit codes} A natural extension of the single-qubit schemes with indirect feedback to non-trivial codes can be obtained simply by applying these schemes to the syndrome qubits in the encoded basis with the purpose of keeping each of them close to the state $|0\rangle\langle 0|$. It is not hard to see that the operators $Z_i^s$ on the gauge qubits in the encoded basis are actually the stabilizer generators for the code. For example, by applying the inverse of the basis transformation for the bit-flip code described in Sec.~\ref{sectionENCODEDBASIS}, one can see that the operators $I^A\otimes Z^s_1\otimes I^s_2$ and $I^A\otimes I^s_1\otimes Z^s_2$ correspond to the generators $ZZI$ and $ZIZ$, respectively. The Hamiltonians $X_i^s$ needed for the feedback, however, do not have simple forms in the original basis. In particular, for the bit-flip code, the operators $I^A\otimes X^s_1\otimes I^s_2$ and $I^A\otimes I^s_1\otimes X^s_2$ correspond to $\frac{1}{2}XIX+\frac{1}{2}YIY+\frac{1}{2}ZXZ+\frac{1}{2}IXI$ and $\frac{1}{2}XXI+\frac{1}{2}YYI+\frac{1}{2}ZZX+\frac{1}{2}IIX$, respectively. The models considered in Refs.~\cite{Ahn:2002:042301, Sarovar:2004:052324, Chase:2008:032304} also measure continuously the stabilizer generators of the code, but the feedback Hamiltonians are assumed to be single-qubit operators in the original basis. However, note that in the general formulation of the ADL scheme---Eq.~\eqref{ADL}---the correcting Hamiltonians $H_r$ are not specified, and in that sense the possibility we discuss here can be regarded as a special case of the ADL scheme. In the case of the bit-flip code, the authors in Refs.~\cite{Ahn:2002:042301, Sarovar:2004:052324} take the correcting Hamiltonians to be $XII$, $IXI$ and $IIX$. This choice is motivated one hand by its analogy with the strong version of the error-correcting operation for this code, and on the other by its simplicity. In the encoded basis, however, these operators are correlated and act on subsystem $\mathcal{H}^A$ as well. More precisely, $XII$, $IXI$ and $IIX$ are equal to $\frac{1}{2}IXX-\frac{1}{2}IYY+\frac{1}{2}XXX+\frac{1}{2}XYY$, $\frac{1}{2}IXZ+\frac{1}{2}IXI+\frac{1}{2}XXI-\frac{1}{2}XXZ$, and $\frac{1}{2}IZX+\frac{1}{2}IIX+\frac{1}{2}XIX-\frac{1}{2}XZX$, respectively. Naturally, since the code is designed to correct single-qubit bit flips, these operators leave the factor $\mathcal{H}^A$ in the code space $\mathcal{H}^A\otimes |0\rangle^s_1\otimes |0\rangle^s_2$ invariant by definition. A similar property holds for codes that can correct arbitrary single-qubit errors. But these operators can introduce errors to the code subsystem through their non-trivial action on the orthogonal complement of the code space. In particular, imagine that the system undergoes just a single perfectly correctable error, say, a single bit flip. Then a strong error correcting operation must be able to correct it. But if we apply a continuous scheme in which the correcting Hamiltonians act non-trivially on the complement of the code space, this scheme would generally apply non-trivial transformations to the subsystem $\mathcal{H}^A$ in the error subspace, which are by definition uncorrectable. (Note that this cannot occur with a scheme which uses operations acting locally on the syndrome qubits.) Nevertheless, in the case of continuous decoherence where uncorrectable errors inevitably arise, this property is not of crucial significance. As we argued earlier, the way CTQEC works is by keeping the weight outside the code space small, which suppresses the effective accumulation of uncorrectable errors. As long as the scheme is able to keep that weight small, it will still have an effect according to our earlier arguments. Indeed, numerical simulations show that with the use of single-qubit feedback Hamiltonians one can achieve a significant improvement of the codeword fidelity with respect to that of an unprotected qubit and outperform the approach of single-shot error correction in various regimes. For details about the numerical results, we refer the reader to Refs.~\cite{Ahn:2002:042301, Sarovar:2004:052324, Chase:2008:032304}. \section{Quantum jumps for Markovian and non-Markovian noise}\label{exact} In this section we will look at the performance of the quantum-jump scheme in the cases of Markovian and non-Markovian decoherence. We will consider the bit-flip code in the case of simple noise models for which the evolution is exactly solvable. The conclusions we obtain, however, hold for general codes and noise models. \subsection{Markovian decoherence} The model described by Eq.~\eqref{errorcorrectionequation} represents the noise as driven by a Lindblad generator, which is valid under the Markovian assumption of bath correlation times that are much shorter than any characteristic time scale of the system \cite{Breuer:2002:Oxford}. In the case of protecting a single qubit from Markovian bit-flip decoherence, we already found the solution for this model---Eq.~\eqref{MSQS}. We saw that the equilibrium fidelity to which the qubit decays scales as $1/\kappa$ for large error-correction rates $\kappa$. For the bit-flip code, we will assume that all qubits decohere through identical independent bit-flip channels, i.e., $\mathcal{L}(\rho)$ is of the form \eqref{Lbitflip} with $\lambda_1=\lambda_2=\lambda_3=\lambda$. Then one can verify that the density matrix at any moment can be written \begin{equation} \rho(t) = a(t)\rho(0)+b(t)\rho_{1}+c(t)\rho_{2}+d(t)\rho_{3}, \label{rhooft} \end{equation} where \begin{eqnarray} \rho_{1}&=&\frac{1}{3}(X_1\rho(0)X_1+X_2\rho(0)X_2 + X_3\rho(0)X_3),\notag\\ \rho_{2}&=&\frac{1}{3}(X_1X_2\rho(0)X_1X_2+ X_2X_3\rho(0)X_2X_3+X_1X_3\rho(0)X_1X_3),\notag\\ \rho_{3} &=& X_1X_2X_3\rho(0) X_1X_2X_3, \end{eqnarray} are equally-weighted mixtures of single-qubit, two-qubit and three-qubit errors on the original state. The evolution of the system subject to decoherence plus error correction is described by the following system of first-order linear differential equations: \begin{eqnarray} \frac{da(t)}{dt}&=&-3\lambda a(t) + (\lambda+\kappa)b(t),\notag\\ \frac{db(t)}{dt}&=&3\lambda a(t)- (3\lambda+\kappa)b(t) + 2 \lambda c(t),\notag\\ \frac{dc(t)}{dt}&=&2\lambda b(t)- (3\lambda+\kappa)c(t)+3\lambda d(t),\notag\\ \frac{dd(t)}{dt}&=&(\lambda+\kappa)c(t)-3\lambda d(t).\label{equations} \end{eqnarray} The exact solution was found in \cite{Paz:1998:355} and we will not present it here. We only note that for the initial conditions $a(0)=1, b(0)=c(0)=d(0)=0$, the exact solution for the weight outside the code space is \begin{equation} b(t)+c(t)=\frac{3}{4+r}(1-e^{-(4+r)\gamma t}), \end{equation} where $r=\kappa/\lambda$. We see that similarly to what we obtained for the single-qubit code, the weight outside the code space quickly decays to its asymptotic value $\frac{3}{4+r}$ which scales as $1/r$. But note that this value is roughly three times greater than that for the single-qubit model. This corresponds to the fact that there are three single-qubit channels. More precisely, it can be verified that if for a given $\kappa$ the uncorrected weight by the single-qubit scheme is small, then the uncorrected weight by a multi-qubit code using the same $\kappa$ and the same kind of decoherence for each qubit, scales approximately linearly with the number of qubits. Similarly, the ratio $r$ required to preserve a given overlap with the code space scales linearly with the number of qubits in the code. The most important difference from the single-qubit model is that in this model there are non-correctable errors that cause a decay of the state inside the code space. Due to the finiteness of the resources employed by our scheme, there always remains a finite portion of the state outside the code space, which gives rise to non-correctable three-qubit errors. To understand how the state decays inside the code space, one can ignore terms of the order of the weight outside the code space in the exact solution. The result is \begin{gather} a(t)\approx \frac{1+e^{-\frac{6}{r}2\gamma t}}{2}, \hspace{0.3cm} b(t)\approx 0,\hspace{0.3cm} c(t) \approx 0, \hspace{0.3cm} d(t)\approx \frac{1-e^{-\frac{6}{r}2\gamma t}}{2}. \end{gather} Comparing with the expression for the fidelity of a single decaying qubit without error correction which can be seen from \eqref{MSQS} for $\kappa=0$, we see that the encoded qubit decays roughly as if subject to bit-flip decoherence with rate $\frac{6}{r}\gamma$. Therefore, for large $r$ this error-correction scheme can reduce the rate of decoherence approximately $\frac{r}{6}$ times. In the limit $r \rightarrow \infty$, it leads to perfect protection of the state for all times. \subsection{Non-Markovian decoherence}\label{nM} \subsubsection{The Zeno effect. Error correction versus error prevention} The effect of freezing of the evolution in the limit of infinite error-correction rate bears a strong similarity to the quantum Zeno effect \cite{Mishra:1977:756}, where frequent measurements slow down the evolution of a system and freeze the state in the limit where they are applied continuously. The Zeno effect arises when the system and its environment are initially decoupled and they undergo a Hamiltonian-driven evolution, which leads to a quadratic change with time of the state during the initial moments \cite{Nakazato:1996:247} (the so-called Zeno regime). Let the initial state of the system plus the bath be $\rho^{SB}(0)=|0\rangle \langle 0|^S\otimes\rho^B(0)$. For small times, the fidelity $\alpha=\textrm{Tr}\{(|0\rangle\langle0|^S\otimes I^B)\rho^{SB}(t)\}$ of the system's density matrix with the initial state can be approximated as \begin{equation} \alpha(t)= 1-C t^2+\mathcal{O}(t^3).\label{Zeno} \end{equation} In terms of the Hamiltonian $H^{SB}$ acting on the entire system, the coefficient $C$ is \begin{eqnarray} C&=&\textrm{Tr}\{(H^{SB})^2|0\rangle\langle0|^S\otimes \rho^B(0)\}\notag\\ &&-\textrm{Tr}\{H^{SB}|0\rangle\langle0|^S\otimes I^B H^{SB} |0\rangle\langle0|^S\otimes \rho^B(0)\}.\label{C} \end{eqnarray} According to \eqref{Zeno}, if after a time step $\Delta t$ the state is measured in an orthogonal basis which involves the initial state, the probability for not projecting it on the initial state is of order $\mathcal{O}(\Delta t^2)$. Thus if the state is continuously measured ($\Delta t \rightarrow 0$), this prevents the system from evolving. It has been proposed to utilize the quantum Zeno effect in schemes for error prevention \cite{Zurek:1984:391, Barenco:1997:1541, Vaidman:1996:1745}, in which an unknown encoded state is protected from errors simply by frequent measurements that keep it inside the code space. From the point of view of the encoded basis, this approach can be understood as measuring the operators $Z^s_i$ which prevents the syndrome qubits from leaving the state $|00...0\rangle$. The approach is similar to error correction, in that the errors for which the code is designed send a codeword to a space orthogonal to the code space. The difference is that the subsystem containing the protected information generally does not remain invariant under the errors, since the procedure does not involve correction of errors but only their prevention. In \cite{Vaidman:1996:1745} it was shown that with this approach it is possible to use codes of smaller redundancy than those needed for error correction and a four-qubit encoding of a qubit was proposed, which is capable of preventing arbitrary independent errors arising from Hamiltonian interactions. The workings of this approach are based on the existence of a Zeno regime, and fail if we assume Markovian decoherence for all times. This is because the errors emerging during a time step $dt$ in a Markovian model are proportional to $dt$ and they accumulate with time if not corrected. By the above observations, error correction can achieve results in noise regimes where error prevention fails. Of course, this advantage is at the expense of a more complicated procedure---in addition to the measurements that constitute error prevention, error correction involves correcting unitaries, and in general is based on codes with higher redundancy. At the same time, we see that in the Zeno regime it is possible to reduce decoherence using weaker resources than those needed for Markovian noise. This suggests that in this regime error correction may exhibit higher performance than it does for Markovian decoherence. In many situations of practical significance, the memory of the environment cannot be neglected, and the evolution is highly non-Markovian \cite{Breuer:2002:Oxford,Quang:1997:5238, Breuer:2004:045323, Krovi:2007:052117}. Furthermore, no evolution is strictly Markovian and for a system initially decoupled from its environment a Zeno regime is always present, short though it may be \cite{Nakazato:1996:247}. Therefore, if the time resolution of error-correcting operations is high enough so that they ``see'' the Zeno regime, this could give rise to different behavior. One important difference between Markovian and non-Markovian noise is that, in the latter case, the error correction and the effective noise on the reduced density matrix of the system cannot be treated as independent processes. One could derive an equation for the effective evolution of the system alone subject to interaction with the environment, such as the Nakajima-Zwanzig \cite{Nakajima:1958:948, Zwanzig:1960:1338} or the time-convolutionless (TCL) \cite{Shibata:1977:171, Shibata:1980:891} master equations, but the generator of transformations at a given moment in general will depend (implicitly or explicitly) on the entire history up to this moment. Therefore, adding error correction can affect the effective error model nontrivially. This means that in order to describe the evolution of a system subject to non-Markovian decoherence and error correction, one either has to derive an equation for the effective evolution taking into account error correction from the very beginning, or one has to look at the evolution of the entire system including the bath, where the error generator and the generator of error correction can be considered independent. In the latter case, for sufficiently small $\tau_c$, the evolution of the entire system including the bath can be described by \begin{equation} \frac{d \rho}{dt}=-i[H, \rho(t)]+\kappa \mathcal{J}(\rho),\label{NMerrorcorrectionequation} \end{equation} where $\rho$ is the density matrix of the system plus the bath, $H$ is the total Hamiltonian, and the error-correction generator $\mathcal{J}$ acts locally on the encoded system. We will consider a description in terms of Eq.~\eqref{NMerrorcorrectionequation} for a sufficiently simple bath model which allows us to find a solution for the evolution of the entire system. To gain understanding of how the scheme works, we will again look at the single-qubit model first. \subsubsection{The single-qubit code}\label{sectionSQNM} We choose the simple scenario of a system coupled to a single bath qubit via the Hamiltonian \begin{equation} H=\gamma X\otimes X, \end{equation} where $\gamma$ is the coupling strength. This can be a good approximation for situations in which the coupling to a single spin from the bath dominates over other interactions \cite{Krovi:2007:052117}. We assume that the bath qubit is initially in the maximally mixed state, which can be thought of as an equilibrium state at high temperature. From \eqref{NMerrorcorrectionequation} one can verify that if the system is initially in the state $|0\rangle$, the state of the system plus the bath at any moment will have the form \begin{eqnarray} \rho(t) = \left(\alpha (t) |0\rangle \langle 0| +(1-\alpha(t))|1\rangle \langle 1|\right)\otimes \frac{I}{2}- \beta(t)Y \otimes \frac{X}{2}. \end{eqnarray} In the tensor product, the first operator belongs to the Hilbert space of the system and the second to the Hilbert space of the bath. We have $\alpha(t) \in [0,1]$, and $|\beta(t)|\le\sqrt{\alpha(t)(1-\alpha(t))}, \beta(t)\in R$. The reduced density matrix of the system has the same form as the one for the Markovian case. The part proportional to $\beta(t)$ can be thought of as a ``hidden'' part, which nevertheless plays an important role in the error-creation process, since errors can be thought of as being transferred to the ``visible'' part from the ``hidden'' part (and vice versa). This can be seen from the fact that during an infinitesimal time step $dt$, the Hamiltonian changes the parameters $\alpha$ and $\beta$ as follows: \begin{gather} \alpha\rightarrow \alpha-2\beta \gamma dt ,\notag\\ \beta \rightarrow \beta +(2\alpha-1)\gamma dt .\label{sqe} \end{gather} The effect of an infinitesimal error-correcting operation is \begin{gather} \alpha \rightarrow \alpha + (1-\alpha)\kappa dt,\notag\\ \beta\rightarrow \beta-\beta\kappa dt. \end{gather} Note that the ``hidden'' part is also being acted upon. Putting it all together, we get the system of equations \begin{gather} \frac{d \alpha(t)}{dt}=\kappa(1-\alpha(t))-2\gamma \beta(t),\notag\\ \frac{d \beta(t)}{dt}=\gamma(2\alpha-1)-\kappa\beta(t)\label{equation2}. \end{gather} The solution for the fidelity $\alpha(t)$ is \begin{gather} \alpha(t)=\frac{2\gamma^2 + \kappa^2}{4\gamma^2+\kappa^2} +e^{-\kappa t}\left(\frac{\kappa\gamma}{4\gamma^2+\kappa^2}\sin{2\gamma t}+\frac{2\gamma^2}{4\gamma^2+\kappa^2}\cos{2\gamma t}\right).\label{singlequbitsolution} \end{gather} We see that as time increases, the fidelity stabilizes at the value \begin{equation} \alpha_*^{NM}= \frac{2+R^2}{4+R^2}=1-\frac{2}{4+R^2}, \end{equation} where $R=\kappa/\gamma$ is the ratio between the error-correction rate and the coupling strength. Fig.~\ref{fig1} shows the fidelity as a function of the dimensionless parameter $\gamma t$ for three different values of $R$. For error-correction rates comparable to the coupling strength ($R=1$), the fidelity undergoes a few partial recurrences before it stabilizes close to $\alpha_*^{NM}$. For larger $R=2$, however, the oscillations are already heavily damped and for $R=5$ the fidelity seems confined above $\alpha_*^{NM}$. As $R$ increases, the evolution becomes closer to a decay like the one in the Markovian case. \begin{figure}[h] \begin{center} \includegraphics[width=4in]{fig1} \caption{Fidelity of the single-qubit code with continuous bit-flip errors and correction, as a function of dimensionless time $\gamma t$, for three different values of the ratio $R=\kappa/\gamma$.} \label{fig1} \end{center} \end{figure} A remarkable difference, however, is that the asymptotic weight outside the code space ($1-\alpha_*^{NM}$) decreases with $\kappa$ as $1/\kappa^2$, whereas in the Markovian case the same quantity decreases as $1/\kappa$. The asymptotic value can be obtained as an equilibrium point at which the infinitesimal weight flowing out of the code space during a time step $dt$ is equal to the weight flowing into it. The latter corresponds to vanishing right-hand sides in equations \eqref{equation1} and \eqref{equation2}. In Sec.~\ref{sectionROLEOFZENO} we will see that the difference in that quantity for the two different types of decoherence arises from the difference in the corresponding evolutions during initial times. \subsubsection{The three-qubit bit-flip code}\label{sectionMQNM} We will consider a model where each qubit independently undergoes the same kind of non-Markovian decoherence as the one we studied for the single-qubit code. Here the system we look at consists of six qubits---three for the codeword and three for the environment. We assume that all system qubits are coupled to their corresponding environment qubits with the same coupling strength, i.e., the Hamiltonian is \begin{equation} H=\gamma\overset{3}{\underset{i=1}{\sum}}X^S_i\otimes X^B_i,\label{Hamiltonian} \end{equation} where the operators $X^S$ act on the system qubits and $X^B$ act on the corresponding bath qubits which are initially in the maximally mixed state. The subscripts label on which particular qubit they act. Obviously, the types of effective single-qubit errors on the density matrix of the system that can result from this Hamiltonian at any time, CP or not, will have operator elements which are linear combinations of the identity and $X^S$. According to the error-correction conditions for non-CP maps obtained in Ref.~\cite{Shabani:07081953}, these errors are correctable by the code. Considering the form of the Hamiltonian \eqref{Hamiltonian} and the error-correcting map, one can see that the density matrix of the entire system at any moment is a linear combination of terms of the type \begin{equation} \varrho_{lmn,pqr}\equiv X_1^lX_2^mX_3^n\rho(0) X_1^pX_2^qX_3^r\otimes \frac{X_1^{l+p}}{2}\otimes \frac{X_2^{m+q}}{2} \otimes \frac{X_3^{n+r}}{2}. \end{equation} Here the first term in the tensor product refers to the Hilbert space of the system, and the following three refer to the Hilbert spaces of the bath qubits that couple to the first, the second and the third qubits from the code respectively. The power indices $l,m,n,p,q,r$ take values $0$ and $1$ in all possible combinations, and $X^1=X$, $X^0=X^2=I$. (Note that $\varrho_{lmn,pqr}$ should not be mistaken with the components of the density matrix in the computational basis.) More precisely, we can write the density matrix in the form \begin{eqnarray} \rho(t)&=&\underset{l,m,n,p,q,r}{\sum}(-i)^{l+m+n}(i)^{p+q+r}C_{lmn,pqr}(t)\times \varrho_{lmn,pqr},\label{fullDM} \end{eqnarray} where the coefficients $C_{lmn,pqr}(t)$ are real. The coefficient $C_{000,000}$ is less than or equal to the codeword fidelity (with equality when $\rho(0)=|\bar{0}\rangle\langle \bar{0}|$ or $\rho(0)=|\bar{1}\rangle\langle \bar{1}|$). Since the scheme aims at protecting an unknown codeword, we will be interested in its performance in the worst case and we will assume that the codeword fidelity is $C_{000,000}$. The exact equations for the coefficients $C_{lmn,pqr}(t)$ and their solutions were obtained in Ref.~\cite{Oreshkov:2007:022318}. Here we will present an approximation which can be obtained by perturbation theory for $\gamma \delta t \ll 1 \ll \kappa \delta t$ \cite{Oreshkov:2007:022318}. The approximate system of equation reads \begin{gather} \frac{d C_{000,000}}{dt}=\frac{24}{R^2}\gamma C_{111,000},\notag\\ \frac{d C_{111,000}}{dt}=-\frac{12}{R^2}\gamma (2C_{000,000}-1).\label{approxeqn} \end{gather} Comparing with \eqref{sqe}, we see that the encoded qubit undergoes approximately the same type of evolution as that of a single qubit without error correction, but the coupling constant is effectively decreased $R^2/12$ times. The solution of \eqref{approxeqn} yields for the codeword fidelity \begin{equation} C_{000,000}(t)=\frac{1+\cos (\frac{24}{R^2}\gamma t)}{2} \label{firstapproxsoln}. \end{equation} This solution is valid only with precision $\mathcal{O}(\frac{1}{R})$ for times $\gamma t \ll R^3$. If one carries out the perturbation to fourth order in $\gamma$, one obtains the approximate equations \begin{gather} \frac{d C_{000,000}}{dt}=\frac{24}{R^2}\gamma C_{111,000}-\frac{72}{R^3}\gamma (2 C_{000,000}-1),\notag\\ \frac{d C_{111,000}}{dt}=-\frac{12}{R^2}\gamma (2C_{000,000}-1) -\frac{144}{R^3}\gamma C_{111,000},\label{approxeqn2} \end{gather} which yield for the fidelity \begin{equation} C_{000,000}(t)=\frac{1+e^{-\frac{144}{R^3}\gamma t}\cos (\frac{24}{R^2}\gamma t)}{2}. \end{equation} We see that in addition to the effective error process which is of the same type as that of a single qubit, there is an extra Markovian bit-flip process with rate $\frac{72}{R^3}\gamma$. This Markovian behavior is due to the Markovian character of our error-correcting procedure which, at this level of approximation, is responsible for the direct transfer of weight between $\varrho_{000,000}$ and $\varrho_{111,111}$, and between $\varrho_{111,000}$ and $\varrho_{000,111}$. The exponential factor explicitly reveals the range of applicability of solution \eqref{firstapproxsoln}---with precision $\mathcal{O}(\frac{1}{R})$, it is valid only for times $\gamma t$ of up to order $R^2$. For times of the order of $R^3$, the decay becomes significant and cannot be neglected. The exponential factor may also play an important role for short times of up to order $R$, where its contribution is bigger than that of the cosine. But in the latter regime the difference between the cosine and the exponent is of order $\mathcal{O}(\frac{1}{R^2})$, which is negligible for the precision that we consider. Fig.~\ref{fig3} presents the exact solution for the codeword fidelity $C_{000,000}(t)$ as a function of the dimensionless parameter $\gamma t$ for $R=100$. For very short times after the beginning ($\gamma t \sim 0.1$), one can see a fast but small in magnitude decay (Fig.\ref{fig4}). The maximum magnitude of this quickly decaying term obviously decreases with $R$, since in the limit of $R\rightarrow \infty$ the fidelity should remain constantly equal to $1$. \begin{figure}[h] \begin{center} \includegraphics[width=4in]{fig3} \caption{Long-time behavior of three-qubit system with bit-flip noise and continuous error correction. The ratio of correction rate to decoherence rate is $R=\kappa/\gamma=100$.} \label{fig3} \end{center} \end{figure} \begin{figure}[h] \begin{center} \includegraphics[width=4in]{fig4} \caption{Short-time behavior of three-qubit system with bit-flip noise and continuous error correction. The ratio of correction rate to decoherence rate is $R=\kappa/\gamma=100$.} \label{fig4} \end{center} \end{figure} We see that in the limit $R\rightarrow \infty$, the evolution approaches an oscillation with an angular frequency $\frac{24}{R^2}\gamma$. This is the same type of evolution as that of a single qubit interacting with its environment, but the coupling constant is effectively reduced $R^2/12$ times. While the coupling constant can serve to characterize the decoherence process in this particular case, such a description is not valid in general. As a general measure of the effect of noise one can use the instantaneous rate of decrease of the codeword fidelity $F_{cw}$ (in our case $F_{cw}=C_{000,000}$): \begin{equation} \Lambda(F_{cw}(t)) = -\frac{dF_{cw}(t)}{dt}.\label{errorrate} \end{equation} This quantity does not coincide with the decoherence rate in the Markovian case (which can be defined naturally from the Lindblad equation), but it is a good estimate of the rate of loss of fidelity and can be used for any decoherence model. We will refer to it simply as an error rate. Since the goal of error correction is to preserve the codeword fidelity, the quantity \eqref{errorrate} is a useful indicator of the performance of a given scheme. Note that $\Lambda (F_{cw})$ is a function of the codeword fidelity and therefore it makes sense to use it for a comparison between different cases only for identical values of $F_{cw}$. For our example, the fact that the coupling constant is effectively reduced approximately $R^2/12$ times implies that the error rate for a given value of $F_{cw}$ is also reduced $R^2/12$ times. Similarly, the reduction of $\lambda$ by the factor $r/6$ in the Markovian case implies a reduction of $\Lambda$ by the same factor. We see that the effective reduction of the error rate increases quadratically as $\kappa^2$ in the non-Markovian case, whereas it increases only linearly as $\kappa$ in the Markovian case. \subsection{The role of the Zeno regime}\label{sectionROLEOFZENO} The effective continuous evolution \eqref{approxeqn} is derived under the assumption $\gamma \delta t \ll 1 \ll \kappa \delta t $. The first inequality implies that $\delta t$ can be considered within the Zeno time scale of the system's evolution without error correction. On the other hand, from the relation between $\kappa$ and $\tau_c$ in \eqref{tauc} we see that $\tau_c\ll\delta t$. Therefore, the time for implementing a weak error-correcting operation has to be sufficiently small so that on the Zeno time scale the error-correction procedure can be described approximately as a continuous Markovian process. This suggests a way of understanding the quadratic enhancement in the non-Markovian case based on the properties of the Zeno regime. Let us consider again the single-qubit code from Sec.~\ref{sectionSQNM}, but this time let the error model be any Hamiltonian-driven process. We assume that the qubit is initially in the state $|0\rangle\langle 0|$, i.e., the state of the system including the bath has the form $\rho(0)=|0\rangle \langle 0|\otimes\rho_B(0)$. For times smaller than the Zeno time $\delta t_Z$, the evolution of the fidelity without error correction can be described by \eqref{Zeno}. Equation \eqref{Zeno} naturally defines the Zeno regime in terms of $\alpha$ itself: \begin{equation} \alpha\geq \alpha_Z \equiv 1-C\delta t_Z^2. \end{equation} For a single time step $\Delta t \ll \delta t_Z$, the change in the fidelity is \begin{equation} \alpha\rightarrow \alpha-2\sqrt{C}\sqrt{1-\alpha}\Delta t+\mathcal{O}(\Delta t^2).\label{singlestepdecoh} \end{equation} On the other hand, the effect of error correction during time $\Delta t$ is \begin{equation} \alpha \rightarrow \alpha+\kappa (1-\alpha)\Delta t +\mathcal{O}(\Delta t^2),\label{singlestepcorr} \end{equation} i.e., it tends to oppose the effect of decoherence. If both processes happen simultaneously, the effect of decoherence will still be of the form \eqref{singlestepdecoh}, but the coefficient $C$ may vary with time. This is because the presence of error-correction opposes the decrease of the fidelity, and consequently can lead to an increase in the time for which the fidelity remains within the Zeno range. If this time is sufficiently long, the state of the environment could change significantly under the action of the Hamiltonian, thus giving rise to a different value for $C$ in \eqref{singlestepdecoh} according to \eqref{C}. Note that the strength of the Hamiltonian puts a limit on $C$, and therefore this constant can vary only within a certain range. The equilibrium fidelity $\alpha_*^{NM}$ that we obtained for the error model in Sec.~\ref{sectionSQNM} can be thought of as the point at which the effects of error and error correction cancel out. For a general model, where the coefficient $C$ may vary with time, this leads to a quasi-stationary equilibrium. From \eqref{singlestepdecoh} and \eqref{singlestepcorr}, one obtains the equilibrium fidelity \begin{equation} \alpha_*^{NM}\approx 1-\frac{4C}{\kappa^2}. \end{equation} In agreement to the result in Sec.~\ref{sectionSQNM}, the equilibrium fidelity differs from $1$ by a quantity proportional to $1/\kappa^2$. If one assumes a Markovian error model, for short times the fidelity changes linearly with time which leads to $1-\alpha_*^{M}\propto 1/\kappa$. Thus the difference can be attributed to the existence of a Zeno regime in the non-Markovian case. This argument readily generalizes to the case of non-trivial codes if we look at the picture in the encoded basis. There each syndrome qubit undergoes a Zeno-type evolution and so do the abstract qubits associated with each error syndrome. Then using only the properties of the Zeno behavior as we did above, we can conclude that the weight outside the code space will be kept at a quasi-stationary value of order $1/\kappa^2$. As we argued in Sec.~\ref{sectionENCODEDBASIS}, this in turn would lead to an effective decrease of the uncorrectable error rate at least by a factor proportional to $1/\kappa^2$. Finally, let us make a remark about the resources needed to achieve the effect of quadratic reduction of the error rate. As it was pointed out, there are two conditions involved---one concerns the magnitude of the error-correction rate, the other concerns the time resolution of the weak error-correcting operations. Both of these quantities should be sufficiently large. There is, however, an interplay between the two, which involves the strength of the interaction required to implement the weak error correcting map \eqref{wm}. Let us imagine that the weak map is implemented by making the system interact weakly with an ancilla in a given state, after which the ancilla is discarded. The error correction procedure consists of a sequence of such interactions and can be thought of as a cooling process. If the time for which a single ancilla interacts with the system is $\tau_c$, one can verify that the parameter $\varepsilon$ in \eqref{wm} would be proportional to $g^2\tau_c^2$, where $g$ is the coupling strength between the system and the ancilla. From \eqref{tauc} we then obtain that \begin{equation} \kappa \propto g^2\tau_c. \end{equation} The two parameters that can be controlled are the interaction time and the interaction strength, and they determine the error-correction rate. Thus, if $g$ is kept constant, a decrease in the interaction time $\tau_c$ leads to a proportional decrease in $\kappa$ which may be undesirable. Therefore, in order to achieve a good working regime, one generally may need to adjust both $\tau_c$ and $g$. But in some situations decreasing $\tau_c$ alone can prove advantageous, since this may lead to a time resolution that reveals the non-Markovian character of an error model that was previously treated as Markovian. Then the quadratic enhancement of the performance as a function of $\kappa$ may compensate the decrease in $\kappa$, thus leading to a seemingly paradoxical result---better performance with a lower error-correction rate. \section{Outlook}\label{discussion} In this chapter we saw that the subsystem principle can be useful for understanding various aspects of the workings of CTQEC and its performance under different noise models, as well as for the design of CTQEC protocols using protocols for the protection of a known state. However, further research is needed to understand how to construct optimal CTQEC protocols. In the case of the quantum-jump model, the code-space fidelity reaches a quasi-equilibrium value which can be used to estimate the performance of the scheme. It would be interesting to see whether an analogue of the equilibrium fidelity exists for schemes with indirect feedback. This could prove useful since stochastic evolutions are generally too complicated for analytical treatment. The equilibrium code-space fidelity can be useful also in assessing the performance of CTQEC under non-Markovian decoherence, where the description of the evolution of a system subject to CTQEC can be difficult due to the large number of environment degrees of freedom. We discussed two main methods for obtaining CTQEC protocols from protocols for the protection of a single qubit: one based on the application of single-qubit protocols to the separate syndrome qubits, and another based on the application of single-qubit protocols to qubit subspaces associated with the different syndromes. An interesting question is whether the performance of CTQEC protocols obtained by these methods can be related to the performance of the underlying single-qubit protocols. A difficulty in the case of indirect feedback is that the noise in the encoded basis is correlated, and the effective noise on a given qubit can depend on the outcomes of the measurements on the rest of the qubits. Another interesting direction for future investigation is to explore CTQEC for specific physical models and limitations of the control parameters (for a recent work, see Ref.~\cite{Kerckhoff:0812:1246}). We saw that applying single-qubit schemes to the syndrome qubits in the encoded basis generally requires multi-qubit operations in the original basis, but numerical simulations show that single-qubit feedback Hamiltonians in the original basis are also efficient. It would be interesting to see whether it is possible to construct efficient CTQEC protocols for non-trivial codes assuming only one- and two-qubit Hamiltonians. The ability to apply CTQEC with Hamiltonians of limited locality would be important for the scalability of this approach. So far, CTQEC has been considered only as a method of protecting quantum memory. A natural next step is to combine this approach with universal quantum computation. An important question in this respect is whether CTQEC can be made fault tolerant. In the theory of quantum fault tolerance, logical operations and error correction are implemented mainly in terms of transversal operations between physical qubits from different blocks, where the basic operations are assumed to be discrete. Is something similar possible for CTQEC? One way of approaching this problem could be to look for fault-tolerant implementations of a universal set of weak operations using only weak transversal unitary operations and projective ancilla measurements. Undoubtedly, the area of CTQEC offers a variety of interesting problems for future investigation. As quantum operations with limited strength or limited rate are likely to be the tools available in many quantum computing architectures in the near term, developing further the approach to protecting quantum information from noise via continuous-time feedback seems a promising direction for research. \acknowledgements{O.O. acknowledges the support of the European Commission under the Marie Curie Intra-European Fellowship Programme (PIEF-GA-2010-273119). This research was supported in part by the Spanish MICINN (Consolider-Ingenio QOIT). } \bibliographystyle{plain}
\section{Introduction} \label{sec:intro} Magnetic flux ropes play a key role in the models of various solar activities, such as flares, filament/prominence eruptions, and coronal mass ejections (CMEs). They have been thoroughly studied by numerical simulations \citep[e.g.,][]{2000Amari,2004Fan,2007Fan,2004Torok,2005Torok,2010Aulanier} and observations \citep[e.g.,][]{2011Cheng,2013Cheng,2012Zhang,2013Li1,2013Li2}. To drive an eruption, there must be sufficient free energy stored in the magnetic field and forces acted on the erupted object. However, a current-free (potential) magnetic field contains the least amount of magnetic energy in all the possible magnetic configurations given the same magnetic fluxes on the bottom boundary. Therefore, a flux rope with large electric currents is required to exist in the pre-eruptive magnetic field or form during the eruption. The field lines are highly twisted around each other in such a configuration. The loss of equilibrium \citep{1991Forbes} or the torus instability \citep{2006Kliem} ejects the flux rope, whose eruption stretches the overlying magnetic field lines. \citet{2010Demoulin} pointed out that the loss of equilibrium and the torus instability are two different views of the same physical mechanism, which is the Lorentz repulsion force of electric currents with different paths. Or in the magnetohydrodynamic (MHD) point of view, it is the result of the force imbalance between the magnetic pressure in the flux rope and the magnetic tension of the overlying magnetic field. There are still some key problems awaiting to be answered for the magnetic flux rope eruption. For example, can a magnetic flux rope exist long time before a major flare or CME? How is the non-potential state built up before the flux rope eruption? Does the magnetic reconnection occur in a flux rope? It matters whether a flux rope exists before an eruption because it determines the initiation mechanism of an eruption. For example, \citet{2005Torok} found that the helical kink instability of a pre-existing twisted flux rope could trigger and initially drive an eruption. Magnetic flux ropes have been found before the onset of flares or CMEs by nonlinear force-free field (NLFFF) extrapolations \citep{2009Canou,2010Canou,2009Savcheva,2010Guo2,2010Guo1,2010Cheng,2011Su} and indicated by extreme-ultraviolet (EUV) and X-ray observations \citep{2009Green,2011Green,2010Liu,2013Patsourakos}. On the other hand, a flux rope can be formed from magnetic arcades during the eruption by magnetic reconnection \citep{1992Moore,2001Moore,1999Antiochos}. A magnetic flux rope is built up by the line-tied photospheric motions, such as the magnetic flux emergence or the horizontal flows. This process injects the magnetic helicity into the higher solar atmosphere, which increases the twist and kink of a flux rope (self-helicity) and the linkage between different flux ropes (mutual helicity). The magnetic helicity is conserved in an ideal MHD process and changes very slowly in a resistive process. Thus, a flux rope with continuous injection of magnetic helicity inevitably erupts to remove the accumulated helicity. Magnetic helicity injection can be inferred from magnetic field and velocity field observations in the photosphere. Theory and techniques to measure the magnetic helicity and the velocity field have been developed in the past years. \citet{2005Pariat} proposed a new expression $G_\theta$ for the flux density of magnetic helicity. Various optical flow techniques have been proposed and tested with analytical velocity field models and MHD simulations \citep{2002Kusano,2004Welsch,2007Welsch,2004Longcope,2006Georgoulis,2006Schuck,2008Schuck,2008Chae}. However, it is not easy to estimate the reliability of these calculations when they are applied to observations. Therefore, we cross-check the helicity injection in a fast evolving active region with two methods, one from the magnetic field and the velocity field in the photosphere and the other from the NLFFF extrapolation. With NLFFF extrapolations, one could compute the total relative helicity in a three dimensional volume and the self-helicity of an elementary magnetic flux rope. The comparison between magnetic helicity injection and total relative helicity from force-free fields has been done in a few cases \citep{2007Lim,2010Park,2012Jing}. They all found overall good agreement between the two helicity estimation methods. \citet{2007Lim} assumed linear force-free field models, for which the total relative helicity measurement is mathematically straightforward and well posed. However, there are still problems in dealing with the boundaries in the direct helicity measurement with NLFFF models as did in \citet{2010Park} and \citet{2012Jing}. Four methods to deal with these problems have been proposed by \citet{2011Rudenko}, \citet{2011Thalmann}, \citet{2012Valori}, and \citet{2013Yang1,2013Yang2}. Here, our main purpose is not to quantify the relationship between the magnetic helicity injection from the photosphere and the total relative helicity from NLFFF models. Instead, we compare in detail the former with the self-helicity contained in a magnetic flux rope. Neither the total relative magnetic helicity nor the self-helicity measures the same quantity as the magnetic helicity injection from the photosphere, because the coronal helicity may see variations related to a CME that could not be seen in the helicity flux on the photosphere. However, compared to the total relative helicity, magnetic helicity contained in a magnetic flux rope has a more direct relationship with the trigger mechanism of the eruption. A traditional view is that magnetic reconnection in a flare occurs in the current sheet tracing behind an erupting flux rope. There is more and more evidence suggesting that it could also occur in the leading edge of a flux rope, both from numerical simulations \citep{2003Amari,2003roussev,2005Torok} and from observations \citep{2003Ji,2009Wang,2011Huang}. \citet{2012Guo} proposed an alternative possibility that the magnetic reconnection could occur inside a flux rope by the internal kink instability. The occurrence of the magnetic reconnection requires the formation of current sheets, which are prone to form at locations where the magnetic linkages change drastically, namely the quasi-separatrix layers (QSLs; \citealt{1995Priest,1996Demoulin,2006Demoulin,2007Demoulin}). In the extreme case, the magnetic linkages are discontinuous, and QSLs degenerate into the separatrix. \citet{1996Demoulin} proposed a norm, $N$, of the Jacobian matrix of the field line mapping to compute the locations of QSLs. However, the norm $N$ is not constant along a magnetic field line. \citet{2002Titov} proposed a new parameter (squashing degree, $Q$) to define QSLs. The squashing degree $Q$ is invariant along a magnetic field line. QSLs are three dimensional volumes where $N$ or $Q$ are large. \citet{2012Pariat} analyzed several methods and proposed the best one to compute two-dimensional (2D) Q maps in the three-dimensional (3D) domain. This magnetic field topology analysis method has been applied to studying coronal sigmoids \citep{2012Savcheva750,2012Savcheva744}. To solve the above discussed problems, we have to know the 3D kinematic, thermal, and magnetic parameters in the solar atmosphere. The kinematic and thermal parameters can be derived from multi-wavelength imaging and spectral observations. However, the magnetic field information is difficult to observe directly in the higher solar atmosphere above the photosphere, where vector magnetic field has been observed routinely with both ground-based and space-borne instruments. Although big efforts have been made to observe magnetic fields in the chromosphere and corona \citep{1998Judge,2000Lin,2009Kuckein,2012Kuckein}, much information of the 3D magnetic field in the solar atmosphere can only be derived by various magnetic models. Due to the low plasma $\beta$ condition in the solar corona \citep{2001Gary}, the magnetic field, $\mathbf{B}$, is usually modeled by the force-free field model that obeys the equations $\nabla \times \mathbf{B} = \alpha \mathbf{B}$ and $\nabla \cdot \mathbf{B} = 0$. If the torsional parameter $\alpha$ is a constant, the equations describe a linear force-free field. A special case is when $\alpha = 0$, where the model degenerates to a potential field. From observations, it is found that $\alpha$ usually changes in solar active regions \citep{2002Wiegelmann,2002Regnier}. Therefore, the NLFFF model is necessary to model such cases more realistically than the potential or linear force-free field models. The force-free field equations are thus nonlinear and needed to be solved numerically \citep[see the review by][]{2012Wiegelmann}. Several numerical methods to solve the NLFFF equations have been thoroughly tested against analytical, numerical, and observations \citep{2006Schrijver,2008Metcalf,2009DeRosa}. In this paper, we study the magnetic helicity injection and 3D magnetic field evolutions before an X2.6 class flare that peaked at 23:02 UT on 2005 January 15. An evolving magnetic flux rope extrapolated by the NLFFF model has been reported in \citet{2010Cheng}. Here, we further compute the twist of the flux rope and compare it with the accumulated magnetic helicity. In particular, we study the build up phase of the magnetic flux rope. Since the velocities on the photosphere are much lower than that in the corona, the evolution can be regarded as a quasi-static process. We use a series of NLFFF to approximate this process. In most cases, only one major flux rope exists in a solar active region, and we could use the twist of the flux rope to approximate the self-helicity if it is not highly kinked. Besides, we analyze the magnetic topology, namely the 3D squashing degree (Q) maps, of the evolving magnetic flux rope to study the magnetic reconnection locations. Observations and data analysis are presented in Section~\ref{sec:data}. Results on the comparison between the injected magnetic helicity and the twist evolution and QSLs of the flux rope are described in Section~\ref{sec:resu}. We finally summarize and discuss our findings in Section~\ref{sec:disc}. \section{Observations and Data Analysis} \label{sec:data} \subsection{X-ray and EUV Observations} Two halo CMEs and associated flares were observed in active region NOAA 10720 on 2005 January 15 by the Large Angle and Spectrometric Coronagraph (LASCO; \citealt{1995Brueckner}) and the Extreme-ultraviolet Imaging Telescope (EIT; \citealt{1995Delaboudiniere}), respectively, on board the \textit{Solar and Heliospheric Observatory} (\textit{SOHO}). \citet{2010Cheng} studied the evolution of the flare loops after the peak of the M8.6 flare at 06:38 UT. The second flare (X2.6) peaked at 23:02 UT. The authors found that the flare loops after the first flare were accelerated again and became the envelope field of the second flare and CME, which were driven by the fast evolving flux rope. Between the two CMEs and flares, there was no CME detected by \textit{SOHO}/LASCO. However, many flares were observed in active region NOAA 10720 during this period. Here, we only focus on the period between 18:00 to 24:00 UT on 2005 January 15 as shown by the \textit{GOES} soft X-ray fluxes in Figure~\ref{fig:goes}. There were at least five flares before the X2.6 flare in this period in active region NOAA 10720, i.e., the C4.4 (18:16 UT), C8.8 (18:53 UT), C3.5 (19:49 UT), C5.4 (20:11 UT), and M1.0 (22:08 UT) class flares. The time in the parenthesis indicates the peak time of that flare. \textit{SOHO}/EIT recorded the aforementioned flares in the 195~\AA \ band. The spatial resolution and temporal cadence of EIT are $2.6''$ per pixel and 12 minutes, respectively. Figure~\ref{fig:195} and an online only movie attached to it displays the evolution of the 195~\AA \ images. We can get two pieces of important information from Figure~\ref{fig:195} and the movie. First, some brightenings appeared intermittently in the core field along the polarity inversion line, as indicated by the five C and M class flares. Secondly, the core field always stayed in the low corona and did not erupt into a higher place. This is because no evidence for a flux rope eruption was found until the X2.6 flare, when the magnetic flux rope eruption led to the second CME. \subsection{Magnetic Helicity Injection} \label{sec:dhdt} In the case of the ideal condition that the conductivity approaches infinity, the time ($t$) variation of the magnetic helicity ($H$) can be written as \citep{1984Berger} \begin{equation} \frac{\mathrm{d} H}{\mathrm{d} t} = -2 \int\limits_{S} (\mathbf{A}_\mathrm{p} \cdot \mathbf{u}) B_n \mathrm{d} S, \label{eqn:dhdt1} \end{equation} where $S$ denotes the boundary surface, $\mathbf{A}_\mathrm{p}$ is the vector potential of the potential field, $\mathbf{u}$ denotes the velocity of the footpoints of flux tubes (namely the flux transport velocity), and $B_n$ is the normal component of the magnetic field. In practical applications to solar events, the magnetic helicity flux through the photosphere dominates those from other surfaces. Therefore, we can integrate the magnetic helicity flux only in the photosphere. The flux transport velocity is defined as \citep{2003Demoulin} \begin{equation} \mathbf{u} = \mathbf{v}_t - \frac{v_n}{B_n}\mathbf{B}_t, \label{eqn:u} \end{equation} where $\mathbf{v}_t$ and $v_n$ are the transverse and normal components of the velocity, respectively, and $\mathbf{B}_t$ is the transverse component of the magnetic field. The integrand in Equation~(\ref{eqn:dhdt1}) can be recognized as the helicity flux density, \begin{equation} G_A(\mathbf{x}) = -2(\mathbf{A}_\mathrm{p} \cdot \mathbf{u}) B_n, \label{eqn:ga} \end{equation} where $\mathbf{x}$ is the postion vector. $G_A$ has been adopted to compute the magnetic helicity flux distribution in active regions by many authors \citep{2001Chae,2001Chae2,2004Chae,2002Kusano,2004Kusano,2002Moon1,2002Moon2,2002Nindos,2003Nindos}. \citet{2005Pariat} found that $G_A$ creates artificial polarities even with simple flows without magnetic helicity injection into the corona. They proposed a new expression for the time variation of the magnetic helicity: \begin{equation} \frac{\mathrm{d} H}{\mathrm{d} t} = -\frac{1}{2\pi} \displaystyle\int\limits_{S} \displaystyle\int\limits_{S'} \frac{\mathrm{d} \theta(\mathbf{r})}{\mathrm{d} t} B_n B'_n \; \mathrm{d} S' \; \mathrm{d} S, \label{eqn:dhdt2} \end{equation} where \begin{equation} \frac{\mathrm{d} \theta(\mathbf{r})}{\mathrm{d} t} = \frac{1}{r^2} \left(\mathbf{r} \times \frac{\mathrm{d} \mathbf{r}}{\mathrm{d} t} \right)_n = \frac{1}{r^2} [\mathbf{r} \times ( \mathbf{u} - \mathbf{u'})]_n , \label{eqn:dthdt} \end{equation} where $\mathbf{r} = \mathbf{x} - \mathbf{x'}$ represents the position vector pointing from $\mathbf{x'}$ to $\mathbf{x}$. Equation~(\ref{eqn:dhdt2}) tells us that the magnetic helicity changes due to the rotation between each pair of two infinitesimal magnetic flux tubes \citep{1984Berger2}. The helicity flux density $G_\theta(\mathbf{x})$ is then defined as \begin{equation} G_\theta(\mathbf{x}) = -\frac{B_n}{2\pi} \displaystyle\int\limits_{S'} \frac{\mathrm{d} \theta(\mathbf{r})}{\mathrm{d} t} B'_n \; \mathrm{d} S' . \label{eqn:gth} \end{equation} \citet{2005Pariat} found that $G_\theta=0$ for the non-rotating motion of a single footpoint. For two magnetic regions with opposite polarities, both $G_A$ and $G_\theta$ have two artificial polarities, but $G_\theta$ is lower by a factor of 10 than $G_A$. From Equations~(\ref{eqn:dhdt1}) and (\ref{eqn:dhdt2}), we find that it is necessary to measure the velocity field in the photosphere in order to compute the helicity injection. The velocity field can be inferred from a series of magnetic fields by the so-called optical flow techniques, for instance, the local correlation tracking (LCT) method. Using only the normal component of magnetic fields, LCT has been extensively applied to derive the photospheric velocity, $\mathbf{u}_\mathrm{LCT}$, which is assumed to be the horizontal plasma velocity, $\mathbf{v}_t$ \citep{2001Chae,2001Chae2,2004Chae,2002Moon1,2002Moon2,2002Nindos,2003Nindos}. The normal magnetic fields can be constructed from the line-of-sight magnetic fields by assuming that all magnetic fields are vertical. LCT tries to maximize the correlation coefficient of the intensity ($I$) between two images using a prescribed window, which corresponds to \citep{2005Schuck} \begin{equation} I(\mathbf{x},t_2) \equiv I[\mathbf{x}-\mathbf{u}_0(t_2-t_1),t_1] . \end{equation} If the image intensity is differentiable, $I(\mathbf{x},t)$ satisfies the advection equation: \begin{equation} \frac{\partial I}{\partial t} + \mathbf{u}_0 \cdot \nabla I = 0 . \end{equation} The normal component of the magnetic induction equation in the ideal condition is \citep{2002Kusano,2004Kusano,2003Demoulin,2004Welsch,2004Longcope} \begin{equation} \frac{\partial B_n}{\partial t} + \nabla_t \cdot (B_n \mathbf{v}_t - v_n \mathbf{B}_t) = 0 , \label{eqn:dbdt1} \end{equation} or \begin{equation} \frac{\partial B_n}{\partial t} + \nabla_t \cdot (\mathbf{u} B_n) = 0, \label{eqn:dbdt2} \end{equation} where we have used the expression for the flux transport velocity defined by Equation~(\ref{eqn:u}). Equations~(\ref{eqn:dbdt1}) or (\ref{eqn:dbdt2}) govern the evolution of the magnetic fields on the photosphere. All the newly developed optical flow techniques have considered the induction equation, such as the inductive local correlation tracking method (ILCT; \citealt{2004Welsch}), the minimum energy fit method (MEF; \citealt{2004Longcope}), and the minimum structure reconstruction method (MSR; \citealt{2006Georgoulis}). \citet{2005Schuck} pointed out that the magnetic induction equation is a continuity equation, and the LCT method is inconsistent with it. Therefore, an appropriate method should start with the continuity Equations~(\ref{eqn:dbdt1}) or (\ref{eqn:dbdt2}). There are two ambiguities in deriving the photospheric velocities from Equations~(\ref{eqn:dbdt1}) or (\ref{eqn:dbdt2}). First, the flux transport vector, which is defined as the product of the flux transport velocity and the normal magnetic field, can be decomposed to \citep{2004Welsch,2004Longcope} \begin{equation} \mathbf{u} B_n = B_n \mathbf{v}_t - v_n \mathbf{B}_t = -(\nabla_t \phi + \nabla_t \psi \times \hat{\mathbf{n}}) , \end{equation} where $\hat{\mathbf{n}}$ is the unit vector of the normal direction. The scalar functions $\phi$ and $\psi$ are the inductive and electrostatic potentials. Only the inductive potential $\phi$ can be determined by Equations~(\ref{eqn:dbdt1}) or (\ref{eqn:dbdt2}), while the electrostatic potential $\psi$ could be arbitrary functions. The velocity field cannot be determined uniquely without additional assumptions. For example, \citet{2005Schuck} assumed an affine velocity profile for the velocity model in a small window, where the velocity is linearly dependent on the local coordinates around the central position. With the affine velocity model, the continuity equation has an analytical solution, thus the velocities can be determined via a least-square method. This optical flow technique is termed as the differential affine velocity estimator (DAVE; \citealt{2006Schuck}). DAVE not only removes the first ambiguity but also guarantees that the dynamics of magnetic fields is consistent with the continuity equation. The second ambiguity is that the plasma velocity along a field line, $\mathbf{v}_\parallel$, cannot be constrained only by the evolution of the normal magnetogram. The DAVE method has been extended for vector magnetic fields (DAVE4VM; \citealt{2008Schuck}). The plasma velocity is assumed to be a three dimensional affine velocity profile. Under such an assumption, the first and the second ambiguities are removed at the same time. A vector velocity field including the component along a field line can be determined through a time sequence of vector magnetic fields. The Digital Vector Magnetograph (DVMG) at Big Bear Solar Observatory (BBSO) observed a series of vector magnetic fields with a cadence of about 1 minute and a spatial resolution of about $0.6''$. The details of the data process can be found in \citet{2009Jing} and \citet{2010Cheng}. The $180^\circ$ ambiguity of the transverse component of the vector magnetic field is removed by the minimum energy method that minimizes the electric current density and the magnetic field divergence simultaneously \citep{1994Metcalf,2006Metcalf,2009Leka}. Since the active region 10720 was crossing the central meridian and close to the solar disk center, the projection effect is of little influence. The finally processed vector magnetic field at 18:27 UT is shown in Figures~\ref{fig:bvh}a and \ref{fig:bvh}b. We adopt the DAVE4VM method developed by \citet{2008Schuck} to compute the vector velocity fields through a time sequence of magnetic fields. The window size is selected to be 23 pixels, which is determined by examining the Pearson correlation and slope between $\nabla_t \cdot (\mathbf{u} B_n)$ and $\Delta B_n / \Delta t$ \citep{2008Schuck}. To restrict the amount of computation and resolve small velocities, we select the following vector magnetic fields to analyze in detail at the time of 18:27, 19:04, 19:30, 20:05, 20:38, 21:02, 21:33, and 22:20 UT on 2005 January 15. With each pair of consecutive magnetic fields, we compute the velocity field at the middle time. For example, Figure~\ref{fig:bvh}c shows the velocity derived by DAVE4VM at the middle time (18:46 UT) of 18:27 and 19:04 UT. An averaged vector magnetic field is also constructed by the same pair of magnetic fields. The magnetic helicity flux density, $G_\theta$, is thus obtained by Equation~(\ref{eqn:gth}) and the computed vector magnetic field and velocity field at 18:46 UT. Figure~\ref{fig:bvh}d displays the magnetic helicity flux density at 18:46 UT. The magnetic helicity flux (namely the helicity injection rate or the time variation of the helicity), $\mathrm{d} H / \mathrm{d}t$, is derived by the integration of $G_\theta$ over the photosphere, i.e., by Equation~(\ref{eqn:dhdt2}). \subsection{Twist of the Flux Rope} \label{sec:twist} \citet{2006Berger} defined the twist between an axis curve and a secondary curve for arbitrary geometries as long as they do not intersect with themselves and are smooth. As shown in Figure~\ref{fig:twist1}, $\mathbf{x}(s)$ denotes a smooth axis curve and $\mathbf{y}(s)$ a secondary curve. $\mathbf{T}(s)$ is a unit vector tangent to $\mathbf{x}(s)$, where $s$ is the arc length from an reference starting point on the axis curve. $\mathbf{V}(s)$ denotes a unit vector normal to $\mathbf{T}(s)$ and pointing from $\mathbf{x}(s)$ to $\mathbf{y}(s)$. Then, the twist density of the secondary curve around the axis is defined by \begin{equation} \frac{\mathrm{d} \Phi}{\mathrm{d} s} = \frac{1}{2 \pi} \mathbf{T}(s) \cdot \mathbf{V}(s) \times \frac{\mathrm{d} \mathbf{V}(s)}{\mathrm{d} s} . \label{eqn:twist} \end{equation} The total twist is derived by the integration along the axis curve. Equation~(\ref{eqn:twist}) can be applied to the twisted field lines of the flux ropes, which are constructed by the NLFFF extrapolations with the optimization method \citep{2000Wheatland,2004Wiegelmann}. As an example, Figure~\ref{fig:bvh}b shows the vector magnetic field and its field of view on the photosphere that are used for the NLFFF extrapolation. The flux balance parameter, which is defined as $\int_S B_z \mathrm{d}S/\int_S |B_z| \mathrm{d}S$, is -0.02. It indicates that the bottom boundary is very well flux balanced. Then, we preprocess the vector magnetic field with the method of \citet{2006Wiegelmann} to remove the net magnetic force and torque on the boundary. The preprocessed vector magnetic field is finally used as the bottom boundary for the NLFFF extrapolation. To compute the twist of the reconstructed flux rope, we have to determine the axis. From the extrapolated results (as shown in \citet{2010Cheng} and in the following analysis), it is found that the middle section of the flux rope was always horizontal and close to the polarity inversion line. Similar to what has been done in \citet{2010Guo2}, we assume the section of the flux rope between points 1 and 3 in Figure~\ref{fig:twist2}a is horizontal and tangent to the polarity inversion line. An objective function is defined to measure how well the above assumptions are met: \begin{equation} F = \int\limits_\mathrm{P1}^\mathrm{P3} |y(x) - y_\mathrm{p}(x)| \mathrm{d} x + \int\limits_\mathrm{P1}^\mathrm{P3} |z(x) - z_\mathrm{a}| \mathrm{d} x, \label{eqn:f} \end{equation} where $y(x)$ and $z(x)$ are the Cartesian coordinates of a magnetic field line, $y_\mathrm{p}(x)$ is the $y$-component of the polarity inversion line, and $z_\mathrm{a}$ is the average height of a magnetic field line. The integration is done between points 1 and 3. The axis field line should have a minimum objective function $F$. To calculate the starting points for integrating the sample magnetic field lines and determining the axis, we fit the polarity inversion line between points 1 and 3 as shown in Figure~\ref{fig:twist2}a with a third order polynomial. Points 1 and 3 are selected in the middle part of the magnetic flux rope with a certain degree of freedom. We will test the errors in computating the twists caused by the positions of points 1 and 3 in the following analysis. A square surface of $16'' \times 16''$ is selected perpendicular to the third order polynomial in the middle of points 1 and 3 (at point 2) and perpendicular to the bottom surface. The area of the square is selected to be large enough to include all the magnetic field lines of the magnetic flux rope and small enough to miminize the computation efforts. The bottom edge of the square is on the photosphere, and the projection of the square is shown as the line segment in Figure~\ref{fig:twist2}a. Then, we integrate some field lines starting from $161 \times 161$ sample points uniformly distributed on the square. The field lines are computed in the computation box until they reach the boundary. Finally, the axis is determined as the magnetic field line with minimum objective function $F$ defined in Equation~(\ref{eqn:f}). The axis is shown as the dark field line in Figure~\ref{fig:twist2}b. The sample field lines of the magnetic flux rope is determined by two criteria. Namely, they all pass through the $16'' \times 16''$ square and are longer than the axis. The twist density between a sample magnetic field line and the axis is computed with Equation~(\ref{eqn:twist}). The total twist is then derived by the integration of the twist density along the axis. The twists of some sample field lines are noted in Figure~\ref{fig:twist2}b. We quantify the twist of the flux rope as the average twist of the sample field lines of the magnetic flux rope. The average twist is about $\sim 1.90 \pm 0.27$ turns, or $(3.80 \pm 0.54) \pi$ in radian, for the flux rope at 18:27 UT on 2005 January 15. The error is estimated as the standard deviation of the twists of the sample field lines. The error can also be caused by the determination of the axis, which depends on the locations of points 1 and 3 in Figure~\ref{fig:twist2}a. If we assume the $x$-coordinates of points 1 and 3 are $x_1$ and $x_3$, respectively, the following four cases are considered to compute the twist error caused by the determination of the axis. We chose four pairs of points 1 and 3 on the polarity inversion line and their $x$-coordinates are located at \{$x_1 + 5.0'', x_3 + 5.0''$\}, \{$x_1 - 5.0'', x_3 - 5.0''$\}, \{$x_1 + 5.0'', x_3 - 5.0''$\}, and \{$x_1 - 5.0'', x_3 + 5.0''$\}. The average and standard deviation of the twists of the four cases at 18:27 UT are -1.90 and 0.03 turns, respectively. Compared to $1.90 \pm 0.27$ turns derived from the original case, the average twist of the four test cases are the same, and the standard deviation is much smaller. We have also checked the results for the other 7 samples, whose results are similar to this one at 18:27 UT. Therefore, the method to determine the axis of the flux rope is very robust. The twist does not depends sensitively on the selection of the choice of points 1 and 3. \section{Results} \label{sec:resu} \subsection{Accumulated Helicity and Twist Evolution} Using the methods described in Sections~\ref{sec:dhdt} and \ref{sec:twist}, we compute the helicity injection rate and twist of the flux rope at the time of 18:27, 19:04, 19:30, 20:05, 20:38, 21:02, 21:33, and 22:20 UT on 2005 January 15. The injected magnetic helicity $\Delta H$ is evaluated by \begin{equation} \Delta H_{t+\Delta t} = \Delta H_t + \left. \frac{\mathrm{d} H}{\mathrm{d} t} \right |_{t+\frac{1}{2}\Delta t} \Delta t , \label{eqn:delh} \end{equation} with $\Delta H_{t_0} = 0$, where $t_0$ is a reference time. Here, we select $t_0$ to be 18:27 UT, i.e., the first moment we extrapolate the 3D magnetic field. The time evolution of $\Delta H$ is plotted in Figure~\ref{fig:ht}, which shows that the negative helicity is injected during the studied period. With a linear fitting, we find that the helicity injection rate is $(-16.47 \pm 3.52) \times 10^{40}~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$. The errors of the injected helicity, $\Delta H$, are estimated as follows. Equations~(\ref{eqn:dhdt2}) and (\ref{eqn:dthdt}) indicate that the errors come both from the magnetic field measurements and the velocity field computation. For the velocity field, the spatial alignment between two consecutive magnetic fields also affects the accuracy derived by DAVE4VM, besides the magnetic field measurement errors. For simplicity, we consider the two effects one by one. The first source of the errors of the helicity injection rate ($\mathrm{d}H/\mathrm{d}t$) can be estimated by considering the errors in the magnetic field measurements. Some artificial errors in the normal distribution with the standard deviation of 2 G for $B_z$ and 10 G for $B_x$ and $B_y$ are added to the vector magnetic field, with which the velocity field is derived using the DAVE4VM method. Here, the errors for $B_z$ and for $B_x$ and $B_y$ are determined by the DVMG sensitivity. Then, the helicity injection rate, $\mathrm{d}H/\mathrm{d}t$, is computed with Equations~(\ref{eqn:dhdt2}) and (\ref{eqn:dthdt}). We repeated the above process 10 times. The error for $\mathrm{d}H/\mathrm{d}t$ is estimated as the standard deviation of the 10 results. We computed the errors of $\mathrm{d}H/\mathrm{d}t$ for all the 7 middle time points of the 8 selected samples. The maximum and minimum errors are $0.95 \times 10^{40}$ and $0.23 \times 10^{40} ~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$, respectively. The second source of the errors of the helicity injection rate ($\mathrm{d}H/\mathrm{d}t$) is estimated by considering the spatial alignment between two consecutive magnetic fields. We assume the upper limit of the alignment error is one pixel. Then, the second magnetic field is shifted one pixel upward, downward, leftward, and rightward, respectively, each of which can be paired with the first magnetic field and yield a new velocity field. Therefore, we get four more velocity fields in addition to the one computed by the originally aligned magnetic fields. Next, the helicity injection rate, $\mathrm{d}H/\mathrm{d}t$, is computed with Equations~(\ref{eqn:dhdt2}) and (\ref{eqn:dthdt}) based on the five velocity fields. We further estimate the standard deviation of the five results as three times the error of the helicity injection rate. The errors are also computed for the other 6 middle time points. The maximum and minimum errors are $3.53 \times 10^{40}$ and $1.74 \times 10^{40} ~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$, respectively. Since the errors of $\mathrm{d}H/\mathrm{d}t$ caused by the magnetic field measurement errors are much smaller than that caused by the spatial alignment, we only consider the latter in estimating the errors of the injected helicity, $\Delta H$. The errors are finally computed with the error propagation formula for the addition when Equation~(\ref{eqn:delh}) is used to compute $\Delta H$. If we include the magnetic field measurement errors, it would only increase the errors of $\Delta H$ slightly. However, the following conclusion does not change. The twists of the flux rope at all the 8 selected time points are computed with the method described in Section~\ref{sec:twist}. The twist error at each time is estimated as the standard deviation of the twists of the sample field lines. Figure~\ref{fig:ht} shows the time evolution of the twist, whose absolute value increases with the time. The negative sign indicates that the twist is left-handed. A linear fitting to the twist time evolution suggests that the twist increasing rate is $-0.18 \pm 0.08$ Turns hr$^{-1}$. As described in \citet{2000Priest}, the magnetic helicity of a magnetic field can be divided into the twist and kink of elementary flux tubes and the linkage between them, which are deemed as the self-helicity ($H_s$) and mutual helicity ($H_m$), respectively: \begin{equation} H = \sum\limits_{i=1}^{N} H_{si} + \sum\limits_{i,j=1(i<j)}^{N} H_{mij}, \end{equation} where the summation runs over $N$ flux tubes. The self-helicity of an elementary flux tube is \begin{equation} H_{si} = T_i F_i^2, \label{eqn:hsi} \end{equation} where $T_i$ is the twist in the unit of turns and $F_i$ is the axial magnetic flux of the $i$th magnetic flux tube. In our case, there is only one major twisted flux tube (or flux rope). The ratio between the radius and the length of the flux rope is less than $1/15$ as shown in Figure~\ref{fig:twist2}, which shows that the radius is less than $8''$ and the length is $\sim 120''$. Therefore, the flux rope is relatively thin, and the magnetic helicity of the flux rope can be estimated as the twist since the axis of the flux rope is not highly kinked. The axial magnetic flux of the flux rope is estimated as follows. We cut a section perpendicular to the flux rope at point 2 as shown in Figure~\ref{fig:twist2} and as described in Section~\ref{sec:twist}. The magnetic flux of the flux rope is defined as the integration of the magnetic flux of the field lines belonging to the flux rope through the above defined cross section. The flux rope field lines have been defined in Section~\ref{sec:twist} as the axis field line and those sample field lines that are longer than it. We compute the axial magnetic field and the magnetic fluxes for all the 8 samples. The mean values of the 8 results are 520 G and $1.28 \times 10^{20}$ Mx, respectively. We have derived that the twist increasing rate is $-0.18 \pm 0.08$ Turns hr$^{-1}$. Following Equation~(\ref{eqn:hsi}), the corresponding self-helicity increasing rate is $(-0.29 \pm 0.13) \times 10^{40} ~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$, which is about 1.8\% of the total helicity flux, $(-16.47 \pm 3.52) \times 10^{40}~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$, injected from the bottom boundary. \subsection{QSLs of the Flux Rope} \citet{2012Pariat} analyzed three methods to compute the 2D squashing degree (Q) maps in the 3D domain. They proposed that the third method (their Equations (12) to (22)) is the most accurate, which we adopt to do the following analysis. The normal direction of the 2D cut points to the right. It moves from left to right to scan the 3D domain. To save the computation resource and to increase the spatial resolution, we only compute the QSLs in a sub-domain of the NLFFF computation box. For the time at 18:27 UT, the sub-domain is selected as $x \in [-41.5'',108.5'']$, $y \in [269.7'',325.1'']$, and $z \in [0.0'',25.4'']$, where $x$ and $y$ are the heliocentric coordinates and $z$ represents the height referred to the solar surface. This sub-domain is resolved by $325 \times 120 \times 55$ grid points. Therefore, the spatial resolution to compute the QSLs is 5 times the original one to resolve the NLFFF extrapolation. The 3D distribution of the squashing degree $Q$ is derived by the scanning of the 2D vertical cuts along the $x$-axis. The maximum value is about $10^{13}$. It is visualized by a 3D contour at $Q = 10^4$ as shown in Figure~\ref{fig:qsl_j1}. The most prominent feature of the QSLs is that they go along and wrap the flux rope. Similar to the helical structure of the twisted magnetic field lines, the hollow shell of the QSL contours along the flux rope also have a helical shape. These QSLs are associated with bald patches, where the magnetic field lines touch the bottom surface on the polarity inversion line and concave up. The magnetic field line mappings at the bald patches are discontinuous; therefore, the $Q$ values are extremely large in these regions. But due to the finite spatial resolutions, $Q$ cannot be infinite in practical numerical computations. Also, as a generalization to true separatrix surfaces, such as those at the bald patches, QSLs along the flux rope include more features. For example, the extended QSL contours at about $x \in [40'',60'']$ represent some magnetic field lines bifurcating from the flux rope. The magnetic field line mappings are not necessarily discontinuous, but change drastically. Besides the QSLs with large $Q$ values along the main flux rope, some are distributed in the eastern part detached from the flux rope at about $x \in [-40'',0'']$ as shown in Figure~\ref{fig:qsl_j1}a. They belong to some sheared and twisted field lines to the east of the main flux rope. In fact, the \textit{SOHO}/EIT 195~\AA \ image indicates the existence of these QSLs as shown in Figures~\ref{fig:195}a and \ref{fig:195}b, which shows that the brightening extended longer and more to the east than the main flux rope in Figure~\ref{fig:qsl_j1}a. However, due to the limitation of the spatial resolution of \textit{SOHO}/EIT, it is not clear if the brightening feature disconnected in the middle part where the magnetic field lines disconnected. In Figure~\ref{fig:qsl_j1}, we also plot the 3D contours of the electric current density, $J = |\mathbf{J}|$, where $\mathbf{J} = \frac{1}{\mu_0} \nabla \times \mathbf{B}$. Similar to the distribution of the QSLs, the electric currents are also along the magnetic flux rope. It is the basic property of the NLFFF model, which requires that the electric current density, $\mathbf{J}$, is parallel to the magnetic field, $\mathbf{B}$. However, there is one major difference between the distributions of the QSLs and the electric current density. It is a hollow shell for the QSLs, but a solid body for the electric current density. To clearly display this feature, we plot the distributions of the QSLs and the electric current density on three selected cuts, which are both perpendicular to the $x$-axis, as shown in Figure~\ref{fig:qsl_j2}. It is found that the QSLs, where the squashing degree is large, are self-closed or self-intersected at some places. But the electric current density is large at a center part, and the magnitude decreases from the center to the periphery. From Figure~\ref{fig:qsl_j2}, we also find that the places where $J$ is the largest is also where the QSLs appears. But there are also places where QSLs appear while $J$ is not necessarily large. If we plot the QSL cuts in the NLFFF as shown in Figure~\ref{fig:qsl}, we can find clearly what magnetic field structures are associated with the QSLs. An online only movie is attached to Figure~\ref{fig:qsl} to show the scanning of the QSL cut along the $x$-axis. It is found that the self-closed QSLs are located at the border of the magnetic flux rope. In addition to these QSLs, some other QSL sections are associated with highly sheared but not twisted field lines. A twisted field line points to the inverse direction as a potential field would do at the associated magnetic dips. This kind of QSLs can be found, for example, in Figures~\ref{fig:qsl_j2}b and \ref{fig:qsl}c as those QSL sections extend outside the flux rope. There are also some QSL sections inside the flux rope as shown in Figures~\ref{fig:qsl_j2}c and \ref{fig:qsl}d. They are surrounded by the self-closed QSL sections and the electric current density is large there. \section{Summary and Discussions} \label{sec:disc} Two major CMEs and the associated flares, M8.6 at 06:38 UT and X2.6 at 23:02 UT, occurred on 2005 January 15 in the fast evolving active region NOAA 10720. In a previous paper, \citet{2010Cheng} found that a magnetic flux rope existed about 5 hours before the X2.6 flare by the NLFFF model. In this paper, we further study the pre-flare brightening by the \textit{SOHO}/EIT and compute the magnetic helicity injection, the vector velocity field, the twist accumulation, and the topology structure of the magnetic flux rope. First, the helicity flux density is computed via the expression, $G_\theta$, proposed by \citet{2005Pariat} and the velocity field is derived by the DAVE4VM method \citep{2008Schuck} using the time series of vector magnetic fields observed by BBSO/DVMG. Next, we compute the twist of the flux rope using the NLFFF model and the twist density formula proposed by \citet{2006Berger}. We have got the NLFFF models at eight time points during about five hours before the X2.6 flare at 23:02 UT. Thus, the time evolution of the twist is also derived in this period. Finally, we compute the 3D distributions of the squashing degree, $Q$, via the scan of selected 2D cuts in the 3D domain using the method proposed in \citet{2012Pariat}. The QSLs are 3D volumes where the $Q$ values are large. With the above analysis, we have the following findings. First, there were five C and M class flares that appeared intermittently in the core field along the polarity inversion line in the five hour period before the X2.6 flare. The core field did not erupt to a CME until the one associated with the X2.6 flare. Secondly, NOAA 10720 was a fast evolving active region, where large horizontal flows and fast magnetic flux emergence existed. These photospheric motions injected negative magnetic helicity into the corona. The helicity injection rate was $(-16.47 \pm 3.52) \times 10^{40}~\mathrm{Mx}^2 ~\mathrm{hr}^{-1}$ from 18:27 UT to 22:20 UT on 2015 January 15. About 1.8\% of the injected magnetic helicity became the internal helicity of the magnetic flux rope, resulting in the twist increasing with a rate of $-0.18 \pm 0.08$ Turns hr$^{-1}$. The evolutions of the accumulated magnetic helicity and twist had a good correlation with each other. The correlation coefficient was about 0.6. It is expected that most part of the magnetic helicity in a given magnetic configuration is stored in the mutual helicity. From the theoretical analysis of \citet{2006Demoulin2}, the ratio of the self-helicity to the mutual helicity scales to $1/N$, where $N$ is number of the elementary flux tubes in the magnetic field and it is usually large. Besides, we only compute the self-helicity in one major flux rope, but do not count those in other elementary flux tubes. This may explain the large difference in quantity between the injected magnetic helicity and the self-helicity in the major flux rope, though they are highly correlated. As a test, we also estimate the total relative helicity contained in the computation box with the NLFFF models, using the method proposed by \citet{2013Yang1,2013Yang2}. The preliminary results show that it is overall consistent with the injected magnetic helicity from the photosphere. However, since there are still large errors in the computed total relative magnetic helicity, an affirmative conclusion asks for a detailed error analysis and computations in a larger time range of, say, several days. Thirdly, we find that the flux rope was wrapped by QSLs, which were associated with bald patches and highly sheared and twisted magnetic field lines. The main feature of the QSLs was a hollow shell, but there were also some QSLs in the shell. The magnitude of the electric current density, $J$, did not have a one to one correspondence to the QSLs. For example, QSLs with large $Q$ were not necessarily associated with large $J$, and the distributions of $Q$ and $J$ are different. The largest $Q$ was located on a shell and their distributions are highly intermittent. While the largest $J$ was located at a center and decreased to its periphery. The features of the QSL and electric current density distributions are similar for all the other 7 time points. \citet{2012Savcheva750,2012Savcheva744} made a detailed comparison between the electric current density and the QSL distributions for a sigmoid observed in 2007 February. The NLFFF extrapolation was made using the flux rope insertion method. Similar results are reached regarding the QSL distributions in \citet{2012Savcheva750,2012Savcheva744} and our studies. Both \citet{2012Savcheva750,2012Savcheva744} and this work find that QSLs wrap the boundary of a flux rope. QSLs and strong currents do not necessarily match each other. There are QSLs without strong electric currents. For example, QSLs could appear in a potential magnetic field with no electric currents, such as the initial configuration of \citet{2005Aulanier}. Strong currents can also exist where no QSLs appear. Here, we have to discriminate the difference between volume currents, which is aligned with the magnetic field, and current sheets, where the electric current is not parallel with the magnetic field. The electric current distribution found here represents the volume currents associated with the twisted magnetic flux rope. While QSls are preferential sites for non-field aligned currents, whose width is very thin. The current sheet is not found in the QSLs derived by the NLFFF model because of two reasons. First, the very small width of a current sheet is out of the resolution ability of the NLFFF model. Second, the NLFFF model intends to smear out the non-field aligned currents (it only allows field aligned currents). The current sheet formation can only be revealed by an MHD process, which has been shown in \citet{2005Aulanier}, \citet{2006Buechner}, and \citet{2009Wilmot-Smith}. The difference between \citet{2012Savcheva750,2012Savcheva744} and our work lies mainly in the electric current density distributions. While \citet{2012Savcheva750,2012Savcheva744} found that the large electric current is distributed on a hollow shell (see also \citealt{2008Bobra,2011Su}), we find that it is located in a center region and decreases to its periphery. Despite different NLFFF extrapolation methods adopted by \citet{2012Savcheva750,2012Savcheva744} and us, the different distributions of the electric current density may reflect the different nature of the active regions, the former of which is a decaying active region and the latter is a new flux emerging region (private communication with B. Kliem). \citet{2012Savcheva750,2012Savcheva744} also found a hyperbolic flux tube (HFT), where two QSLs intersect with each other and the $Q$ value is very large. We have made time series of the 3D magnetic field and topology analysis at 8 time points. The last time is about 40 minutes before the flux rope eruption. We do not find HFT along the flux rope in all the time series of the 3D magnetic fields. The difference may either reflect the real structures of the different active regions, or, come from the different extrapolation methods. In \citet{2012Savcheva750,2012Savcheva744}, on the one hand, there was a clear sigmoid in the active region; on the other hand, the flux rope is ejected (not in equilibrium) and relaxes to a solution more or less close to the boundary. Thus, the presence of a flux rope is an initial assumption. Therefore, HFT and bald patches are easy to appear. While with the optimization method, depending on the preprocessing parameters, it tends to smooth the currents and therefore reduce the size and strength of the flux rope. Based on the above results, we propose the following scenario for the occurrence of the five confined flares and the X2.6 class eruptive flare. The photospheric motions caused by the magnetic flux emergence injected magnetic helicity (and energy, which is not studied here) continuously into the solar corona. Although only a small part ($\sim$ 1.8\%) of the magnetic helicity was built into the magnetic twist of the flux rope, QSLs were created around and inside it. According to previous studies \citep[e.g.,][]{2005Aulanier,2006Buechner,2009Wilmot-Smith}, electric currents are preferentially built up in QSLs. Once the strength of the electric currents were large enough to trigger the resistive instability, magnetic energy in the current layers was released by the magnetic reconnection, which also redistributed the magnetic helicity in the active region. The intermittent confined flares were the manifestation of this reconnection process. We find that the twist of the flux rope always exceeded the critical value of 1.5 and 1.75 turns for the kink instability found by \citet{2003Fan} and \citet{2004Torok}, respectively. However, the flux rope touched the photosphere in bald patches, where the line tied effects prevented the flux rope from an eruption via the kink instability. Consequently, the flux rope did not rise to a height to trigger the torus instability that would lead to a full eruption. For the X2.6 eruptive flare, the mechanism was most probably the same as that proposed in \citet{2012Savcheva744}. An HFT would be formed below the flux rope. Tether-cutting magnetic reconnection \citep{2001Moore} is supposed to occur in the HFT. The flux rope would erupt into the interplanetary space successfully via the loss-of-equilibrium \citep{1991Forbes} or the torus instability \citep{2006Kliem,2010Olmedo}. \acknowledgments The authors thank B. Kliem and S. Yang for helpful discussions and thank the referee for constructive suggestions. YG, MDD, and XC were supported by the National Natural Science Foundation of China (NSFC) under the grant numbers 11203014, 10933003, 11373023, 11303016, and the grant from the 973 project 2011CB811402.
\section{Introduction} Connectivity is a fundamental graph theoretic property. Recently, the concept of rainbow connection was introduced by Chartrand, Johns, McKeon and Zhang in \cite{chartrand}. We say that a set of edges is {\em rainbow colored} if its every member has a distinct color. An edge colored graph $G$ is {\em rainbow edge connected} if any two vertices are connected by a rainbow colored path. Furthermore, the {\em rainbow connection} $rc(G)$ of a connected graph $G$ is the smallest number of colors that are needed in order to make $G$ rainbow edge connected. Notice, that by definition a rainbow edge connected graph is also connected. Moreover, any connected graph has a trivial edge coloring that makes it rainbow edge connected, since one may color the edges of a given spanning tree with distinct colors. Other basic facts established in \cite{chartrand} are that $rc(G)=1$ if and only if $G$ is a clique and $rc(G)=|V(G)|-1$ if and only if $G$ is a tree. Besides its theoretical interest, rainbow connection is also of interest in applied settings, such as securing sensitive information transfer and networking (see, e.g., \cite{chakraborty, lisun}). For instance, consider the following setting in networking \cite{chakraborty}: we want to route messages in a cellular network such that each link on the route between two vertices is assigned with a distinct channel. Then, the minimum number of channels to use is equal to the rainbow connection of the underlying network. Caro, Lev, Roditty, Tuza and Yuster \cite{caro} prove that for a connected graph $G$ with $n$ vertices and minimum degree $\delta$, the rainbow connection satisfies $rc(G)\leq \frac{\log{\delta}}{\delta}n(1+f(\delta))$, where $f(\delta)$ tends to zero as $\delta$ increases. The following simpler bound was also proved in \cite{caro}, $rc(G) \leq n \frac{4\log{n}+3}{\delta}$. Krivelevich and Yuster \cite{krivelevichyuster} removed the logarithmic factor from the upper bound in \cite{caro}. Specifically they proved that $rc(G) \leq \frac{20n}{\delta}$. Chandran, Das, Rajendraprasad and Varma \cite{CDRV} improved this upper bound to $\frac{3n}{\delta+1}+3$, which is close to best possible. As pointed out in \cite{caro} the random graph setting poses several intriguing questions. Specifically, let $G=G(n,p)$ denote the binomial random graph on $n$ vertices with edge probability $p$. Caro, Lev, Roditty, Tuza and Yuster \cite{caro} proved that $p=\sqrt{\log{n}/n}$ is the sharp threshold for the property $rc(G)\leq 2$. This was sharpened to a hitting time result by Heckel and Riordan \cite{HR}. He and Liang \cite{heliang} studied further the rainbow connection of random graphs. Specifically, they obtain a threshold for the property $rc(G) \leq d$ where $d$ is constant. Frieze and Tsourakakis \cite{gnprainbow} studied the rainbow connection of $G=G(n,p)$ at the connectivity threshold $p=\frac{\log{n}+{\omega}}{n}$ where $\omega\to\infty$ and $\omega=o(\log{n})$. They showed that w.h.p.\footnote{An event ${\cal E}_n$ occurs {\em with high probability}, or w.h.p.\ for brevity, if $\lim_{n\rightarrow\infty}\mbox{{\bf Pr}}({\cal E}_n)=1$.} $rc(G)$ is asymptotically equal to $\max\set{diam(G), Z_1(G)}$, where $Z_1$ is the number of vertices of degree one. For further results and references we refer the interested reader to the recent survey of Li, Shi and Sun \cite{lisun}. In this paper we study the rainbow connection of the random $r$-regular graph $G(n,r)$ of order $n$, where $r\geq 4$ is a constant and $n\to \infty$. It was shown in Basavaraju, Chandran, Rajendraprasad, and Ramaswamy~\cite{BCRR} that for any bridgeless graph $G$, $rc(G)\leq \rho(\rho+2)$, where $\rho$ is the radius of $G=(V,E)$, i.e., $\min_{x\in V}\max_{y\in V} dist(x,y)$. Since the radius of $G(n,r)$ is $O(\log n)$ w.h.p., we see that \cite{BCRR} implies that $rc(G(n,r))=O(\log^2n)$ w.h.p. The following theorem gives an improvement on this for $r\geq 4$. \begin{theorem}\label{thrm:mainthrm} Let $r\ge 4$ be a constant. Then, w.h.p.\ $rc(G(n,r))=O(\log n)$. \end{theorem} \noindent The rainbow connection of any graph~$G$ is at least as large as its diameter. The diameter of $G(n,r)$ is w.h.p.\ asymptotically $\log_{r-1}{n}$ and so the above theorem is best possible, up to a (hidden) constant factor. We conjecture that Theorem \ref{thrm:mainthrm} can be extended to include $r=3$. Unfortunately, the approach taken in this paper does not seem to work in this case. \section{Proof of Theorem~\ref{thrm:mainthrm}} \label{sec:regular} \subsection{Outline of strategy} Let $G=G(n,r)$, $r\ge 4$. Define \begin{equation}\label{eq:kr} k_r= \log_{r-1}(K_1\log n), \end{equation} where $K_1$ will be a sufficiently large absolute constant. Recall that the {\em distance between two vertices} in $G$ is the number of edges in a shortest path connecting them and the {\em distance between two edges} in $G$ is the number of vertices in a shortest path between them. (Hence, both adjacent vertices and incident edges have distance 1.) For each vertex $x$ let $T_x$ be the subgraph of $G$ induced by the vertices within distance $k_r$ of $x$. We will see (due to Lemma~\ref{density}) that w.h.p., $T_x$ is a tree for most $x$ and that for all $x$, $T_x$ contains at most one cycle. We say that $x$ is {\em tree-like} if $T_x$ is a tree. In which case we denote by $L_x$ the leaves of $T_x$. Moreover, if $u\in L_x$, then we denote the path from $u$ to $x$ by $P(u,x)$. We will randomly color $G$ in such a way that the edges of every path $P(u,x)$ is rainbow colored for all $x$. This is how we do it. We order the edges of $G$ in some arbitrary manner as $e_1,e_2,\ldots,e_m$, where $m=rn/2$. There will be a set of $q = \lceil K_1^2 r\log n\rceil$ colors available. Then, in the order $i=1,2,\ldots,m$ we randomly color $e_i$. We choose this color uniformly from the set of colors not used by those $e_j,j<i$ which are within distance $k_r$ of $e_i$. Note that the number of edges within distance $k_r$ of $e_i$ is at most \begin{equation}\label{eq:edges_k_r} 2\left( (r-1) +(r-1)^2 + \dots + (r-1)^{\lfloor k_r \rfloor-1} \right) \le (r-1)^{k_r} = K_1 \log n. \end{equation} So for $K_1$ sufficiently large we always have many colors that can be used for $e_i$. Clearly, in such a coloring, the edges of a path $P(u,x)$ are rainbow colored. Now consider a fixed pair of tree-like vertices $x,y$. We will show (using Corollary~\ref{cor1}) that one can find a partial 1-1 mapping $f=f_{x,y}$ between $L_x$ and $L_y$ such that if $u\in L_x$ is in the domain $D_{x,y}$ of $f$ then $P(u,x)$ and $P(f(u),y)$ do not share any colors. The domain $D_{x,y}$ of $f$ is guaranteed to be of size at least $K_2\log n$, where $K_2=K_1/10$. Having identified $f_{x,y},D_{x,y}$ we then search for a rainbow path joining $u\in D_{x,y}$ to $f(u)$. To join $u$ to $f(u)$ we continue to grow the trees $T_x,T_y$ until there are $n^{1/20}$ leaves. Let the new larger trees be denoted by $\widehat{T}_x,\widehat{T}_y$, respectively. As we grow them, we are careful to prune away edges where the edge to root path is not rainbow. We do the same with $T_y$ and here make sure that edge to root paths are rainbow with respect to corresponding $T_x$ paths. We then construct at least $n^{1/21}$ vertex disjoint paths $Q_1,Q_2,\ldots,$ from the leaves of $\widehat{T}_x$ to the leaves of $\widehat{T}_y$. We then argue that w.h.p.\ one of these paths is rainbow colored and that the colors used are disjoint from the colors used on $P(u,x)$ and $P(f(u),y)$. We then finish the proof by dealing with non tree-like vertices in Section \ref{nontree}. \subsection{Coloring lemmata} In this section we prove some auxiliary results about rainbow colorings of $d$-ary trees. Recall that a {\em complete $d$-ary tree $T$} is a rooted tree in which each non-leaf vertex has exactly $d$ children. The {\em depth} of an edge is the number of vertices in the path connecting the root to the edge. The set of all edges at a given depth is called a {\em level} of the tree. The {\em height} of a tree is the distance from the root to the deepest vertices in the tree (i.e. the leaves). Denote by $L(T)$ the set of leaves and for $v\in L(T)$ let $P(v,T)$ be the path from the root of $T$ to $v$ in $T$. \begin{lemma}\label{lemcol} Let $T_1,T_2$ be two vertex disjoint \emph{rainbow} copies of the complete $d$-ary tree with $\ell$ levels, where $d\geq 3$. Let $T_i$ be rooted at $x_i$, $L_i=L(T_i)$ for $i=1,2$, and \[ m(T_1,T_2)=\left| \{ (v,w)\in L_1\times L_2 : P(v,T_1) \cup P(w,T_2)\text{ is rainbow} \}\right|. \] Let $$\kappa_{\ell,d}=\min_{T_1,T_2}\set{m(T_1,T_2)}.$$ Then, \begin{equation}\label{eq:col_lem1} \kappa_{\ell,d}\geq d^{2\ell}/4. \end{equation} \end{lemma} \begin{proof} We prove that \[ \kappa_{\ell,d}\geq \brac{1-\sum_{i=1}^\ell \frac{i}{d^{i}}} d^{2\ell}\geq d^{2\ell}/4. \] We prove this by induction on $\ell$. If $\ell=1$, then clearly \[ \kappa_{1,d}=d(d-1). \] Suppose that \eqref{eq:col_lem1} holds for an $\ell \ge 2$. Let $T_1,T_2$ be rainbow trees of height $\ell+1$. Moreover, let $T_1' = T_1\setminus L(T_1)$ and $T_2' = T_2\setminus L(T_2)$. We show that \begin{equation}\label{eq:col_lem1:m} m(T_1,T_2) \ge d^2 \cdot m(T_1',T_2') - (\ell+1)d^{\ell+1}. \end{equation} Each $(v',w')\in L_1'\times L_2'$ gives rise to $d^2$ pairs of leaves $(v,w)\in L_1\times L_2$, where $v'$ is the parent of $v$ and $w'$ is the parent of $w$. Hence, the term $d^2 \cdot m(T_1',T_2')$ accounts for the pairs $(v,w)$, where $P_{v',T_1'}\cup P_{w',T_2'}$ is rainbow. We need to subtract off those pairs for which $P_{v,T_1}\cup P_{w,T_2}$ is not rainbow. Suppose that this number is $\nu$. Let $v\in L(T_1)$ and let $v'$ be its parent, and let $c$ be the color of the edge $(v,v')$. Then $P_{v,T_1}\cup P_{w,T_2}$ is rainbow unless $c$ is the color of some edge of $P_{w,T_2}$. Now let $\nu(c)$ denote the number of root to leaf paths in $T_2$ that contain an edge color $c$. Thus, \[ \nu \leq \sum_{c} \nu(c), \] where the summation is taken over all colors $c$ that appear in edges of $T_1$ adjacent to leaves. We bound this sum trivially, by summing over all colors in $T_2$ (i.e., over all edges in $T_2$, since $T_2$ is rainbow). Note that if the depth of the edge colored $c$ in $T_2$ is $i$, then $\nu(c) \le d^{\ell+1-i}$. Thus, summing over edges of $T_2$ gives us \[ \sum_c\nu(c)\leq\sum_{i=1}^{\ell+1}d^{\ell+1-i} \cdot d^i=(\ell+1)d^{\ell+1}, \] and consequently \eqref{eq:col_lem1:m} holds. Thus, by induction (applied to $T_1'$ and $T_2'$) \begin{align*} m(T_1,T_2) &\ge d^2 \cdot m(T_1',T_2') - (\ell+1)d^{\ell+1}\\ &\ge d^2 \brac{1-\sum_{i=1}^\ell \frac{i}{d^{i}}}d^{2\ell} - (\ell+1)d^{\ell+1}\\ &\ge \brac{1-\sum_{i=1}^{\ell+1} \frac{i}{d^{i}}}d^{2(\ell+1)}, \end{align*} as required. \end{proof} In the proof of Theorem~\ref{thrm:mainthrm} we will need a stronger version of the above lemma. \begin{lemma}\label{lemcol1} Let $T_1,T_2$ be two vertex disjoint edge colored copies of the complete $d$-ary tree with $L$ levels, where $d\geq 3$. For $i=1,2$, let $T_i$ be rooted at $x_i$ and suppose that edges $e,f$ of $T_i$ have a different color whenever the distance between $e$ and $f$ in $T_i$ is at most $L$. Let $\kappa_{\ell,d}$ be as defined in Lemma \ref{lemcol}. Then \[ \kappa_{L,d}\geq \brac{1-\frac{L^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}}}d^{2L}. \] \end{lemma} \begin{proof} Let $T_i^\ell$ be the subtree of $T_i$ spanned by the first $\ell$ levels, where $1\le \ell \le L$ and $i=1,2$. We show by induction on $\ell$ that \begin{equation}\label{eq:col_lem2} m(T_1^\ell, T_2^{\ell}) \ge \brac{1-\frac{\ell^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}}}d^{2\ell}. \end{equation} Observe first that Lemma \ref{lemcol} implies \eqref{eq:col_lem2} for $1\le \ell \le \lfloor L/2 \rfloor -1$, since in this case $T_1^\ell$ and $T_2^\ell$ must be rainbow. Suppose that $\lfloor L/2 \rfloor \leq\ell<L$ and consider the case where $T_1,T_2$ have height $\ell+1$. Following the argument of Lemma \ref{lemcol} we observe that color $c$ can be the color of at most $d^{\ell +1 - \lfloor L/2 \rfloor}$ leaf edges of $T_1$. This is because for two leaf edges to have the same color, their common ancestor must be at distance (from the root) at most $\ell-\lfloor L/2 \rfloor$. Therefore, \begin{align*} m(T_1^{\ell+1}, T_2^{\ell+1}) &\geq d^{2}\cdot m(T_1^\ell, T_2^\ell) -d^{\ell+1-\lfloor L/2 \rfloor} \sum_c\nu(c) \\ &\geq d^{2}\cdot m(T_1^\ell, T_2^\ell) -d^{\ell+1-\lfloor L/2 \rfloor }(\ell+1)d^{\ell+1}\\ &= d^{2}\cdot m(T_1^\ell, T_2^\ell) -(\ell+1)d^{2(\ell+1)-\lfloor L/2 \rfloor }. \end{align*} Thus, by induction \begin{align*} m(T_1^{\ell+1}, T_2^{\ell+1}) &\ge d^{2}\brac{1-\frac{\ell^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}}}d^{2\ell} -(\ell+1)d^{2(\ell+1)-\lfloor L/2 \rfloor }\\ &= \brac{1-\frac{\ell^2 + \ell+1}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}}}d^{2(\ell+1)}\\ &\ge \brac{1-\frac{(\ell+1)^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}}}d^{2(\ell+1)} \end{align*} yielding~\eqref{eq:col_lem2} and consequently the statement of the lemma. \end{proof} \begin{corollary}\label{cor1} Let $T_1,T_2$ be as in Lemma \ref{lemcol1}, except that the root degrees are $d+1$ instead of $d$. If $d\geq 3$ and $L$ is sufficiently large, then there exist $S_i\subseteq L_i, i=1,2$ and a bijection $f:S_1\to S_2$ such that \begin{enumerate}[(a)] \item $|S_i|\geq d^L/10$, and \item $x\in S_1$ implies that $P_{x,T_1}\cup P_{f(x),T_2}$ is rainbow. \end{enumerate} \end{corollary} \begin{proof} To deal with the root degrees being $d+1$ we simply ignore one of the subtrees of each of the roots. Then note that if $d\geq 3$ then \[ 1-\frac{L^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\lfloor L/2 \rfloor} \frac{i}{d^{i}} \ge 1-\frac{L^2}{d^{\lfloor L/2 \rfloor}}-\sum_{i=1}^{\infty} \frac{i}{d^{i}} = 1-\frac{L^2}{d^{\lfloor L/2 \rfloor}}-\frac{d}{(d-1)^2} \ge \frac{1}{5} \] for $L$ sufficiently large. Now we choose $S_1,S_2$ in a greedy manner. Having chosen a matching $(x_i,y_i=f(x_i))\in L_1\times L_2$, $i=1,2,\ldots,p$, and $p < d^L/10$, there will still be at least $d^{2L}/5-2pd^L>0$ pairs in $m(T_1,T_2)$ that can be added to the matching. \end{proof} \subsection{Configuration model}\label{sec:configuration} We will use the configuration model of Bollob\'as \cite{b1} in our proofs (see, e.g., \cite{bollobas, JLR, wormald} for details). Let $W=[2m=rn]$ be our set of {\em configuration points} and let $W_i=[(i-1)r+1,ir]$, $i\in [n]$, partition $W$. The function $\phi:W\to[n]$ is defined by $w\in W_{\phi(w)}$. Given a pairing $F$ (i.e. a partition of $W$ into $m$ pairs) we obtain a (multi-)graph $G_F$ with vertex set $[n]$ and an edge $(\phi(u),\phi(v))$ for each $\{u,v\}\in F$. Choosing a pairing $F$ uniformly at random from among all possible pairings $\Omega_W$ of the points of $W$ produces a random (multi-)graph $G_F$. Each $r$-regular simple graph $G$ on vertex set $[n]$ is equally likely to be generated as $G_F$. Here simple means without loops or multiple edges. Furthermore, if $r$ is a constant, then $G_F$ is simple with a probability bounded below by a positive value independent of $n$. Therefore, any event that occurs w.h.p.\ in $G_F$ will also occur w.h.p.\ in $G(n,r)$. \subsection{Density of small sets} Here we show that w.h.p.\ almost every subgraph of a random regular graph induced by the vertices within a certain small distance is a tree. Let \begin{equation}\label{eq:ell1} t_0= \frac{1}{10}\log_{r-1}n. \end{equation} \begin{lemma}\label{density} Let $k_r$ and $t_0$ be defined in \eqref{eq:kr} and \eqref{eq:ell1}. Then, w.h.p.\ in $G(n,r)$ \begin{enumerate}[(a)] \item\label{lem:a} no set of $s\leq t_0$ vertices contains more than $s$ edges, and \item\label{lem:b} there are at most $\log^{O(1)}n$ vertices that are within distance $k_r$ of a cycle of length at most~$k_r$. \end{enumerate} \end{lemma} \begin{proof} We use the configuration model described in Section~\ref{sec:configuration}. It follows directly from the definition of this model that the probability that a given set of $k$ disjoint pairs in $W$ is contained in a random configuration is given by \[ p_k = \frac{1}{(rn-1)(rn-3)\dots(rn-2k+1)} \le \frac{1}{(rn - 2k)^k} \le \frac{1}{r^k (n-k)^k}. \] Thus, in order to prove \eqref{lem:a} we bound: \begin{align*} \mbox{{\bf Pr}}(\exists S\subseteq [n],|S|\leq t_0,e[S]\geq |S|+1) &\leq \sum_{s=3}^{\lfloor t_0 \rfloor}\binom{n}{s}\binom{\binom{s}{2}}{s+1} r^{2(s+1)} p_{s+1}\\ &\leq \sum_{s=3}^{\lfloor t_0 \rfloor} \left( \frac{en}{s} \right)^s \left( \frac{es}{2} \right)^{s+1} \left(\frac{r}{n-(s+1)}\right)^{s+1}\\ &\leq \frac{et_0}{2} \cdot \frac{r}{n-(t_0+1)} \cdot \sum_{s=3}^{\lfloor t_0 \rfloor} \left( \frac{en}{s} \cdot \frac{es}{2} \cdot \frac{r}{n-(s+1)}\right)^{s}\\ &\leq \frac{et_0}{2} \cdot \frac{r}{n-(t_0+1)} \cdot \sum_{s=3}^{\lfloor t_0 \rfloor} \left( e^2 r \right)^{s}\\ &\leq \frac{et_0}{2} \cdot \frac{r}{n-(t_0+1)} \cdot t_0 \cdot \left( e^2 r \right)^{t_0}\\ &\leq \frac{e r t_0^2}{2(n-(t_0+1))} \cdot n^{\frac{\log_{r-1} (e^2 r)}{10}} = o(1), \end{align*} as required. We prove \eqref{lem:b} in a similar manner. The expected number of vertices within $k_r$ of a cycle of length at most $k_r$ can be bounded from above by \begin{align*} \sum_{\ell=0}^{\lfloor k_r \rfloor}\binom{n}{\ell} \sum_{k=3}^{\lfloor k_r \rfloor}\binom{n}{k}\frac{(k-1)!}{2} r^{2(k+\ell)} p_{k+\ell} & \leq \sum_{\ell=0}^{\lfloor k_r \rfloor} \sum_{k=3}^{\lfloor k_r \rfloor} n^{k+\ell} \left( \frac{r}{n-(k+\ell)} \right)^{k+\ell}\\ & \leq \sum_{\ell=0}^{\lfloor k_r \rfloor} \sum_{k=3}^{\lfloor k_r \rfloor} \left( 2r \right)^{k+\ell}\\ & \leq k_r^2 (2r)^{2k_r} = \log^{O(1)} n. \end{align*} Now \eqref{lem:b} follows from the Markov inequality. \end{proof} \subsection{Chernoff bounds} In the next section we will use the following bounds on the tails of the binomial distribution ${\rm Bin}(n,p)$ (for details, see, e.g., \cite{JLR}): \begin{align} &\mbox{{\bf Pr}}({\rm Bin}(n,p)\leq \alpha np)\leq e^{-(1-\alpha)^2np/2},\quad 0\leq \alpha\leq 1, \label{chernoff_lower}\\ &\mbox{{\bf Pr}}({\rm Bin}(n,p)\geq \alpha np)\leq \bfrac{e}{\alpha}^{\alpha np},\quad \alpha\geq 1. \label{chernoff_upper} \end{align} \subsection{Coloring the edges} We now consider the problem of coloring the edges of $G=G(n,r)$. Let $H$ denote the line graph of $G$ and let $\Gamma=H^{k_r}$ denote the graph with the same vertex set as $H$ and an edge between vertices $e,f$ of $\Gamma$ if there there is a path of length at most $k_r$ between $e$ and $f$ in $H$. Due to \eqref{eq:edges_k_r} the maximum degree $\Delta(\Gamma)$ satisfies \beq{maxH} \Delta(\Gamma)\leq K_1\log n. \end{equation} We will construct a proper coloring of $\Gamma$ using \beq{qcol} q= \lceil K_1^2r\log n \rceil \end{equation} colors. Let $e_1,e_2,\ldots,e_m$ with $m=rn/2$ be an arbitrary ordering of the vertices of $\Gamma$. For $i=1,2,\ldots,m$, color $e_i$ with a random color, chosen uniformly from the set of colors not currently appearing on any neighbor in $\Gamma$. At this point only $e_1,e_2,\ldots,e_{i-1}$ will have been colored. Suppose then that we color the edges of $G$ using the above method. Fix a pair of vertices $x,y$ of $G$. \subsubsection{Tree-like and disjoint} Assume first that $T_x,T_y$ are vertex disjoint and that $x,y$ are both tree-like. We see immediately, that $T_x,T_y$ fit the conditions of Corollary \ref{cor1} with $d=r-1$ and $L=k_r$. Let $S_x\subseteq L(T_x)$, $S_y\subseteq L(T_y)$, $f:S_x\to S_y$ be the sets and function promised by Corollary \ref{cor1}. Note that $|S_x|,|S_y|\geq K_2\log n$, where $K_2=K_1/10$. In the analysis below we will expose the pairings in the configuration as we need to. Thus an unpaired point of $W$ will always be paired to a random unpaired point in~$W$. We now define a sequence $A_0=S_x,A_1,\ldots,A_{t_0}$, where $t_0$ defined as in \eqref{eq:ell1}. They are defined so that $T_x\cup A_{\leq t}$ spans a tree $T_{x,t}$ where $A_{\leq t}=\bigcup_{j\leq t}A_j$. Given $A_1,A_2,\ldots,A_i=\set{v_1,v_2,\ldots,v_p}$ we go through $A_i$ in the order $v_1,v_2,\ldots,v_p$ and construct $A_{i+1}$. Initially, $A_{i+1}=\emptyset$. When dealing with $v_j$ we add $w$ to $A_{i+1}$ if: \begin{enumerate}[(a)] \item $w$ is a neighbor of $v_j$; \item\label{A:b} $w\notin T_x\cup T_y\cup A_{\leq i+1}$ (we include $A_{i+1}$ in the union because we do not want to add $w$ to $A_{i+1}$ twice); \item\label{A:c} If the path $P(v_j,x)$ from $v_j$ to $x$ in $T_{x,i}$ goes through $v\in S_x$ then the set of edges $E(w)$ is rainbow colored, where $E(w)$ comprises the edges in $P(x,v_j)+(v_j,w)$ and the edges in the path $P(f(v),y)$ in $T_y$ from $y$ to $f(v)$. \end{enumerate} We do not add neighbors of $v_j$ to $A_{i+1}$ if ever one of \eqref{A:b} or \eqref{A:c} fails. We prove next that \beq{grow} \mbox{{\bf Pr}}\left(|A_{i+1}|\leq (r-1.1)|A_i|\;\big{|}\, K_2\log n\leq |A_i|\leq n^{2/3}\right)=o(n^{-3}). \end{equation} Let $X_{\ref{A:b}}$ and $X_{\ref{A:c}}$ be the number of vertices lost because of case \eqref{A:b} and \eqref{A:c}, respectively. Observe that \begin{equation}\label{eq:Abounds} (r-1)|A_i|-X_{\ref{A:b}}-X_{\ref{A:c}} \le |A_{i+1}| \le (r-1)|A_i| \end{equation} First we show that $X_{\ref{A:b}}$ is dominated by the binomial random variable \[ Y_{\ref{A:b}} \sim (r-1){\rm Bin}\left((r-1)|A_i|,\frac{r|A_i|}{rn/2-rn^{2/3}}\right) \] conditioning on $K_2\log n\leq |A_i|\leq n^{2/3}$. This is because we have to pair up $(r-1)|A_i|$ points and each point has a probability less than $\frac{r|A_i|}{rn/2-rn^{2/3}}$ of being paired with a point in $A_i$. (It cannot be paired with a point in $A_{\leq i-1}$ because these points are already paired up at this time). We multiply by $(r-1)$ because one ``bad'' point ``spoils'' the vertex. Thus, \eqref{chernoff_upper} implies that \[ \mbox{{\bf Pr}}(X_{\ref{A:b}}\geq |A_i|/20) \le \mbox{{\bf Pr}}(Y_{\ref{A:b}}\geq |A_i|/20)\leq \bfrac{40er(r-1)^2|A_i|}{n}^{|A_i|/20}=o(n^{-3}). \] We next observe that $X_{\ref{A:c}}$ is dominated by \[ Y_{\ref{A:c}} \sim (r-1){\rm Bin}\brac{r|A_i|, \frac{4\log_{r-1}n}{q}}. \] To see this we first observe that $|E(w)|\leq 2\log_{r-1}n$, with room to spare. Consider an edge $e=(v_j,w)$ and condition on the colors of every edge other than $e$. We examine the effect of this conditioning, which we refer to as $\mathcal{C}$. We let $c(e)$ denote the color of edge $e$ in a given coloring. To prove our assertion about binomial domination, we prove that for any color $x$, \beq{colcond} \mbox{{\bf Pr}}(c(e)=x\mid\mathcal{C})\leq \frac{2}{q}. \end{equation} We observe first that for a particular coloring $c_1,c_2,\ldots,c_m$ of the edges $e_1,e_2,\ldots,e_m$ we have $$\mbox{{\bf Pr}}(c(e_i)=c_i,\,i=1,2,\ldots,m)=\prod_{i=1}^m\frac{1}{a_i}$$ where $q-\Delta\leq a_i\leq q$ is the number of colors available for the color of the edge $e_i$ given the coloring so far i.e. the number of colors unused by the neighbors of $e_i$ in $\Gamma$ when it is about to be colored. Now fix an edge $e=e_i$ and the colors $c_j,\,j\neq i$. Let $C$ be the set of colors not used by the neighbors of $e_i$ in $\Gamma$. The choice by $e_i$ of its color under this conditioning is not quite random, but close. Indeed, we claim that for $c,c'\in C$ $$\frac{\mbox{{\bf Pr}}(c(e)=c\mid c(e_j)=c_j,\,j\neq i)}{\mbox{{\bf Pr}}(c(e)= c'\mid c(e_j)=c_j,\,j\neq i)}\leq \bfrac{q-\Delta}{q-\Delta-1}^\Delta.$$ This is because, changing the color of $e$ only affects the number of colors available to neighbors of $e_i$, and only by at most one. Thus, for $c\in C$, we have \beq{uppcol} \mbox{{\bf Pr}}(c(e)=c\mid c(e_j)=c_j,\,j\neq i)\leq \frac{1}{q-\Delta}\bfrac{q-\Delta}{q-\Delta-1}^\Delta. \end{equation} Now from \eqref{maxH} and \eqref{qcol} we see that $\Delta\leq \frac{q}{K_1r}$ and so \eqref{uppcol} implies \eqref{colcond}. Applying \eqref{chernoff_upper} we now see that \[ \mbox{{\bf Pr}}(X_{\ref{A:c}}\geq |A_i|/20) \le \mbox{{\bf Pr}}(Y_{\ref{A:c}}\geq |A_i|/20) \leq \bfrac{80e(r-1)}{K_1^2}^{|A_i|/20}=o(n^{-3}). \] This completes the proof of \eqref{grow}. Thus, \eqref{grow} and \eqref{eq:Abounds} implies that w.h.p. \[ |A_{t_0}| \ge (r-1.1)^{t_0} \ge (r-1)^{\frac{1}{2}t_0} = n^{{1}/{20}} \] and \[ |A_{t_0}| \le (r-1)^{t_0} |A_0| \le K_1 n^{{1}/{10}}\log n, \] since trivially $|A_0| \le K_1\log n$. In a similar way, we define a sequence of sets $B_0=S_y,B_1,\ldots,B_{t_0}$ disjoint from $A_{\leq t_0}$. Here $T_y\cup B_{\leq t_0}$ spans a tree $T_{y,t_0}$. As we go along we keep an injection $f_i:B_i\to A_i$ for $0\leq i\leq t_0$. Suppose that $v\in B_i$. If $f_i(v)$ has no neighbors in $A_{i+1}$ because \eqref{A:b} or \eqref{A:c} failed then we do not try to add its neighbors to $B_{i+1}$. Otherwise, we pair up its $(r-1)$ neighbors $b_1,b_2,\ldots,b_{r-1}$ outside $A_{\leq i}$ in an arbitrary manner with the $(r-1)$ neighbors $a_1,a_2,\ldots,a_{r-1}$. We will add $b_1,b_2,\ldots,b_{r-1}$ to $B_{i+1}$ and define $f_{i+1}(b_j)=a_j,\,j=1,2,\ldots,r-1$ if for each $1\leq j\leq r-1$ we have $b_j\notin A_{\leq t_0}\cup T_x\cup T_y\cup B_{\leq i+1}$ and the unique path $P(b_j,y)$ of length $i+k_r$ from $b_i$ to $y$ in $T_{y,i}$ is rainbow colored and furthermore, its colors are disjoint from the colors in the path $P(a_j,x)$ in $T_{x,i}$. Otherwise, we do not grow from $v$. The argument that we used for \eqref{grow} will show that \[ \mbox{{\bf Pr}}\left(|B_{j+1}|\leq (r-1.1)|B_j| \; \big{|}\, K_2\log n\leq |B_j|\leq n^{2/3}\right)=o(n^{-3}). \] The upshot is that w.h.p.\ we have $B_{t_0}$ and $A'_{t_0}=f_{t_0}(B_{t_0})$ of size at least $n^{1/20}$. Our aim now is to show that w.h.p.\ one can find vertex disjoint paths of length $O(\log_{r-1}n)$ joining $u\in B_{t_0}$ to $f_{t_0}(u)\in A_{t_0}$ for at least half of the choices for $u$. Suppose then that $B_{t_0}=\set{u_1,u_2,\ldots,u_p}$ and we have found vertex disjoint paths $Q_j$ joining $u_j$ and $v_j=f_{t_0}(u_j)$ for $1\leq j<i$. Then we will try to grow breadth first trees $T_i,T_i'$ from $u_i$ and $v_i$ until we can be almost sure of finding an edge joining their leaves. We will consider the colors of edges once we have found enough paths. Let $R=A_{\leq t_0}\cup B_{\leq t_0}\cup T_x\cup T_y$. Then fix $i$ and define a sequence of sets $S_0=\set{u_i},S_1,S_2,\ldots,S_t$ where we stop when either $S_t=\emptyset$ or $|S_t|$ first reaches size $n^{3/5}$. Here $S_{j+1}= N(S_j)\setminus (R\cup S_{\leq j})$. ($N(S)$ will be the set of neighbors of $S$ that are not in $S$). The number of vertices excluded from $S_{j+1}$ is less than $O(n^{1/10}\log n)$ (for $R$) plus $O(n^{1/10}\log n \cdot n^{3/5})$ for $S_{\leq j}$. Since \[ \frac{O(n^{1/10}\log n \cdot n^{3/5})}{n} = O(n^{-3/10}\log n) = O(n^{-3/11}), \] $|S_{j+1}|$ dominates the binomial random variable \[ Z \sim {\rm Bin}\left((r-1)|S_j|,1-O(n^{-3/11})\right). \] Thus, by \eqref{chernoff_lower} \begin{multline}\label{AT1} \mbox{{\bf Pr}}\big{(}|S_{j+1}|\leq (r-1.1)|S_j|\; \big{|}\,100<|S_j|\leq n^{3/5}\big{)}\notag \\ \le \mbox{{\bf Pr}}\left(Z \leq (r-1.1)|S_j|\; \big{|}\, 100<|S_j|\leq n^{3/5}\right)\notag =o(n^{-3}). \end{multline} Therefore w.h.p., $|S_j|$ will grow at a rate $(r-1.1)$ once it reaches a size exceeding~100. We must therefore estimate the number of times that this size is not reached. We can bound this as follows. If $S_j$ never reaches 100 in size then some time in the construction of the first $\log_{r-1}100$ $S_j$'s there will be an edge discovered between an $S_j$ and an excluded vertex. The probability of this can be bounded by $100 \cdot O(n^{-3/11})=O(n^{-3/11})$. So, if $\beta$ denotes the number of $i$ that fail to produce $S_t$ of size $n^{3/5}$ then \[ \mbox{{\bf Pr}}(\beta\geq 20)\leq o(n^{-3})+\binom{n^{1/10}\log n}{20} \cdot O(n^{-3/11})^{20}=o(n^{-3}). \] Thus w.h.p.\ there will be at least $n^{1/20}-20>n^{1/21}$ of the $u_i$ from which we can grow a tree with $n^{3/5}$ leaves $L_{i,y}$ such that all these trees are vertex disjoint from each other and $R$. By the same argument we can find at least $n^{1/21}$ of the $v_i$ from which we can grow a tree $L_{i,x}$ with $n^{3/5}$ leaves such that all these trees are vertex disjoint from each other and $R$ {\em and the trees grown from the} $u_i$. We then observe that if $e(L_{i,x},L_{i,y})$ denotes the edges from $L_{i,x}$ to $L_{i,y}$ then $$\mbox{{\bf Pr}}(\exists i:e(L_{i,x},L_{i,y})=\emptyset)\leq n^{1/20}\brac{1-\frac{(r-1)n^{3/5}} {rn/2}}^{(r-1)n^{3/5}}=o(n^{-3}).$$ We can therefore w.h.p.\ choose an edge $f_i\in e(L_{i,x},L_{i,y})$ for $1\leq i\leq n^{1/21}$. Each edge $f_i$ defines a path $Q_i$ from $x$ to $y$ of length at most $2\log_{r-1}n$. Let $Q_i'$ denote that part of $Q_i$ that goes from $u_i\in A_{t_0}$ to $v_i\in B_{t_0}$. The path $Q_i$ will be rainbow colored if the edges of $Q_i'$ are rainbow colored and distinct from the colors in the path from $x$ to $u_i$ in $T_{x,t_0}$ and the colors in the path from $y$ to $v_i$ in $T_{y,t_0}$. The probability that $Q_i'$ satisfies this condition is at least $\brac{1-\frac{2\log_{r-1}n}{q}}^{2\log_{r-1}n}$. Here we have used \eqref{colcond}. In fact, using \eqref{colcond} we see that \begin{align*} \mbox{{\bf Pr}}(\not\exists i:Q_i\text{ is rainbow colored})&\leq \brac{1-\brac{1-\frac{2\log_{r-1}n}{q}}^{2\log_{r-1}n}}^{n^{1/21}}\\ &\leq \brac{1-\frac{1}{n^{4/(rK_1^2)}}}^{n^{1/21}}=o(n^{-3}). \end{align*} This completes the case where $x,y$ are both tree-like and $T_x\cap T_y=\emptyset$. \subsubsection{Tree-like but not disjoint} Suppose now that $x,y$ are both tree-like and $T_x\cap T_y\neq\emptyset$. If $x\in T_y$ or $y\in T_x$ then there is nothing more to do as each root to leaf path of $T_x$ or $T_y$ is rainbow. Let $a\in T_y\cap T_x$ be such that its parent in $T_x$ is not in $T_y$. Then $a$ must be a leaf of $T_y$. We now bound the number of leaves $\lambda_a$ in $T_y$ that are descendants of $a$ in $T_x$. For this we need the distance of $y$ from $T_x$. Suppose that this is $h$. Then $$\lambda_a=1+(r-2)+(r-1)(r-2)+(r-1)^2(r-2)+\cdots+(r-1)^{k_r-h-1}(r-2)=(r-1)^{k_r-h}+1.$$ Now from Lemma \ref{density} we see that there will be at most two choices for $a$. Otherwise, $T_x\cup T_y$ will contain at least two cycles of length less than $2k_r$. It follows that w.h.p.\ there at most $\lambda_0=2((r-1)^{k_r-h}+1)$ leaves of $T_y$ that are in $T_x$. If $(r-1)^h\geq 201$ then $\lambda_0\leq |S_y|/10$. Similarly, if $(r-1)^h\geq 201$ then at most $|S_x|/10$ leaves of $T_x$ will be in $T_y$. In which case we can use the proof for $T_x\cap T_y=\emptyset$ with $S_x,S_y$ cut down by a factor of at most $4/5$. If $(r-1)^h\leq 200$, implying that $h\leq 5$ then we proceed as follows: We just replace $k_r$ by $k_r+5$ in our definition of $T_x,T_y$, for these pairs. Nothing much will change. We will need to make $q$ bigger by a constant factor, but now we will have $y\in T_x$ and we are done. \subsubsection{Non tree-like}\label{nontree} We can assume that if $x$ is non tree-like then $T_x$ contains exactly one cycle $C$. We first consider the case where $C$ contains an edge $e$ that is more than distance 5 away from $x$. Let $e=(u,v)$ where $u$ is the parent of $v$ and $u$ is at distance 5 from $x$. Let $\widehat{T}_x$ be obtained from $T_x$ by deleting the edge $e$ and adding two trees $H_u,H_v$, one rooted at $u$ and one rooted at $v$ so that $\widehat{T}_x$ is a complete $(r-1)$-ary tree of height $k_r$. Now color $H_u,H_v$ so that Lemma \ref{lemcol1} can be applied. We create $\widehat{T}_y$ from $T_y$ in the same way, if necessary. We obtain at least $(r-1)^{2k_r}/5$ pairs. But now we must subtract pairs that correspond to leaves of $H_u,H_v$. By construction there are at most $4(r-1)^{2k_r-5}\leq (r-1)^{2k_r}/10$. So, at least $(r-1)^{2k_r}/10$ pairs can be used to complete the rest of the proof as before. We finally deal with those $T_x$ containing a cycle of length 10 or less, no edge of which is further than distance 10 from $x$. Now the expected number of vertices on cycles of length $k\leq 10$ is given by $$k\binom{n}{k}\frac{(k-1)!}{2}\binom{r}{2}^k2^k\frac{\Psi(rn-2k)}{\Psi(rn)} \sim \frac{(r-1)^k}{2k},$$ where $\Psi(m)=m!/(2^{m/2}(m/2)!)$. It follows that the expected number of edges $\mu$ that are within 10 or less from a cycle of length 10 or less is bounded by a constant. Hence $\mu=o(\log n)$ w.h.p.\ and we can give each of these edges a distinct new color after the first round of coloring. Any rainbow colored set of edges will remain rainbow colored after this change. Then to find a rainbow path beginning at $x$ we first take a rainbow path to some $x'$ that is distance 10 from $x$ and then seek a rainbow path from $x'$. The path from $x$ to $x'$ will not cause a problem as the edges on this path are unique to it. \section{The case $d=3$} An easy generalization of the example in Figure \ref{fig1} shows that Lemma \ref{lemcol} does not extend to binary trees. It indicates that $\kappa_{\ell,2}\leq 2^\ell$ and not $\Omega(2^{2\ell})$ as we would like. In this case we have not been able to prove Corollary \ref{cor1}. \begin{figure}[h] \includegraphics[width=\textwidth]{tree.pdf} \caption{Two rainbow trees $T_1$ and $T_2$ with $m(T_1,T_2) = 2^3$.} \label{fig1} \end{figure} Note that while the example shows that $m(T_1,T_2)=2^\ell$, it does show there is a bijection $f$ between the leaves of the two trees so that $P_{x,T_1}\cup P_{f(x),T_2}$ is rainbow. In fact, an elegant probabilistic argument due to Noga Alon \cite{noga} shows that with the hypothesis of Lemma \ref{lemcol}, there are always sets $S_i\subseteq L_i$ and a bijection $f:S_1\to S_2$ such that (i) $|S_1|=|S_2|=\Omega(2^\ell)$ and such that (ii) $x\in S_1$ implies that $P_{x,T_1}\cup P_{f(x),T_2}$ is rainbow. This is a step in the right direction and it can be used to show that $O\brac{\bfrac{\log n}{\log\log n}^2}$ colors suffice, beating the bound implied by \cite{BCRR}. \section{Conclusion} \label{sec:concl} We have shown that w.h.p.\ $rc(G(n,r))=O(\log n)$ for $r\geq 4$ and $r=O(1)$. Determining the hidden constant seems challenging. We have seen that the argument for $d\geq 4$ cannot be extended to $d=3$ and so this case represents a challenge. At a more technical level, we should also consider the case where $r\to\infty$ with~$n$. Part of this can be handled by the sandwiching results of Kim and Vu \cite{KV} (see also~\cite{DFRS}). \bigskip \noindent {\bf Acknowledgement} We are grateful to Noga Alon for help on the case $d=3$. \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\section{Introduction} Compressed Sensing (CS) \cite{CandRomTao06,Don2006} asserts a signal can be recovered from a small number of linear measurements if it can be regarded as a sparse or compressible signal in an a priori basis. An important consequence is its application in analog-to-digital conversion and spectrum estimation, where the signal of interest is spectrally sparse, composing of a small number $r$ of frequency components. It is shown that if the frequencies of a signal all lie on the DFT grid, the signal of length $n$ can then be recovered exactly from a random subset of $O(r\log n)$ samples with high probability \cite{candes2007sparsity}. However, in reality the frequencies of a signal never lie on a grid, no matter how fine the grid is; rather, they are continuous-valued and determined by the mother nature, e.g. a point spread function. Performance degeneration of CS algorithms is observed and studied in \cite{chi2011sensitivity,scharf2011sensitivity,pakroohanalysis} when a ``basis mismatch'' between the actual frequencies and the assumed grid occurs, and many algorithms have been proposed to mitigate the basis mismatch effect \cite{Yang2011offgrid,fannjiang2012coherence}. More recently, several approaches have been proposed to assume away the need for an a priori basis while maintaining the capabilities of subsampling in CS with rigorous performance guarantees. One recent approach is via atomic norm minimization \cite{chandrasekaran2012convex}, which is a general recipe for designing convex solutions to parsimonious model selection. It has been successfully applied to recover a spectrally-sparse signal from a subset of its consecutive samples \cite{CandesFernandez2012SR} or randomly selected samples \cite{TangBhaskarShahRecht2012}. In particular, Tang et. al. showed that a spectrally-sparse signal can be recovered from $O(r\log n \log r)$ random samples with high probability when the frequencies satisfy a mild separation condition \cite{TangBhaskarShahRecht2012}. This framework has also been extended to handle higher-dimensional frequencies \cite{chi2013compressive}. Another approach is proposed in \cite{chen2013spectral,chen2013robust}, where the problem is reformulated into a structured matrix completion inspired by matrix pencil \cite{Hua1992}. For this approach, it is shown that $O(r\log^2 n)$ randomly selected samples are sufficient to guarantee perfect recovery with high probability under some mild incoherence conditions and the approach is amenable to higher-dimensional frequencies. With slightly more samples, both approaches can recover off-the-grid frequencies at an arbitrary precision. We refer interested readers for respective papers for details. In this paper, we study the problem of simultaneously recovering multiple spectrally-sparse signals that are supported on the same frequencies lying arbitrarily on the unit circle. With Multiple measurement Vectors (MMV) in a CS framework, it is shown to further reduce the required number of samples \cite{tropp2006algorithms2,tropp2006algorithms,lee2012subspace,kim2012compressive,davies2012rank,mishali2008reduce} by harnessing the joint sparsity pattern of different signals. Our proposed algorithm can be regarded as a continuous counterpart of the MMV model in CS, and is based on atomic norm minimization which can be solved efficiently using semidefinite programming. We characterize a dual certificate for the optimality of the proposed algorithm. Comparisons are given for joint sparse recovery between conventional CS algorithms and the proposed algorithm through numerical experiments, which indicate that the number of samples per signal can be reduced by harnessing the joint sparsity pattern of multiple signals using atomic norm minimization. \begin{comment} We further explore the benefits of MMV models by studying the case when the number of measurement vectors goes to infinity. With an additional statistical assumption on the coefficient vectors, the problem is cast as completion of a low-rank positive semidefinite Toeplitz matrix, which exhibits superior performance. This is distinct from the correlation-aware support recovery algorithm \cite{pal2012correlation} which considers all frequencies lie on a uniform grid. \end{comment} The rest of the paper is organized as below. Section~\ref{backgrounds} describes the problem formulation. Section~\ref{proposal} formulates the atomic norm minimization problem for joint sparse recovery. Section~\ref{numerical} gives numerical experiments and we conclude in Section~\ref{concl}. Throughout the paper, we use bold letters to denote matrices and vectors, and unbolded letters to denote scalars. The transpose is denoted by $(\cdot)^T$, the complex conjugate is denoted by $(\cdot)^*$, and the trace is denoted by $\trace(\cdot)$. \section{Problem Formulation and Backgrounds} \label{backgrounds} Let ${\boldsymbol x}=[x_1,\ldots,x_n]^T\in\mathbb{C}^n$ be a spectrally-sparse signal with $r$ distinct frequency components, which can be written as \begin{equation}\label{single} {\boldsymbol x}= \sum_{k=1}^r c_{k} {\boldsymbol a}(f_k) \triangleq {\boldsymbol V} {\boldsymbol c}, \end{equation} where each atom ${\boldsymbol a}(f)$ is defined as \begin{equation} {\boldsymbol a}(f ) = \frac{1}{\sqrt{n}}\left[1, e^{j2\pi f}, \ldots, e^{j2\pi f(n-1)}\right]^T , \quad f \in [0,1). \end{equation} The Vandermonde matrix ${\boldsymbol V}$ is given as ${\boldsymbol V}= [{\boldsymbol a}(f_1),\ldots, {\boldsymbol a}(f_r)]\in\mathbb{C}^{n\times r}$, and the coefficient vector ${\boldsymbol c}=[c_1,\ldots,c_r]^T\in\mathbb{C}^r$. The set of frequencies $\mathcal{F}=\{f_k\}_{k=1}^r$ can lie anywhere on the unit circle, such that $f_k$ is continuously valued in $[0,1)$. In an MMV model, we consider $L$ signals, stacked in a matrix, ${\boldsymbol X}=[{\boldsymbol x}_1,\ldots,{\boldsymbol x}_L]$, where each ${\boldsymbol x}_l\in\mathbb{C}^{n}$, $l=1,\ldots, L$, shares the same set of frequencies and is composed of the form of \eqref{single} as \begin{equation} {\boldsymbol x}_l = \sum_{k=1}^r c_{k,l} {\boldsymbol a}(f_k) = {\boldsymbol V} {\boldsymbol c}_l, \end{equation} with ${\boldsymbol c}_l = [c_{1,l},\ldots,c_{r,l}]^T$. Hence ${\boldsymbol X}$ can be expressed as \begin{equation}\label{data_rep} {\boldsymbol X} = {\boldsymbol V}{\boldsymbol C}, \end{equation} where $ {\boldsymbol C} = \begin{bmatrix} {\boldsymbol c}_1 & \cdots & {\boldsymbol c}_L \end{bmatrix} \in\mathbb{C}^{r \times L}$. Assume we observe the \textit{same} subset of entries of each ${\boldsymbol x}_l$, and denote the location set of observed entries by $\Omega$, with $m=|\Omega|$. When there is no noise, the observations can be given as $${\boldsymbol Z}_{\Omega} = \mathcal{P}_{\Omega}({\boldsymbol X}) \in \mathbb{C}^{m\times L},$$ where $\mathcal{P}_{\Omega}$ is a projection operator that only preserves the rows of ${\boldsymbol X}$ indexed by $\Omega$.\footnote{We remark that, this restriction of fixing the observation pattern across different signals, is actually unnecessary for the proposed algorithm.} The traditional CS approach assumes that ${\boldsymbol x}_l$'s are sparse in an a priori determined DFT basis or DFT frame ${\boldsymbol F}\in\mathbb{C}^{n \times d}$ $(d\geq n)$, and they share the same sparsity pattern. Hence, the signal ${\boldsymbol X}$ is \textit{modeled} as \begin{equation} \label{model} {\boldsymbol X} = {\boldsymbol F} {\boldsymbol \Theta}, \end{equation} where ${\boldsymbol \Theta}=[{\boldsymbol \theta}_1,\ldots,{\boldsymbol \theta}_d]^*\in\mathbb{C}^{d\times L}$, and the number of nonzero rows of ${\boldsymbol \Theta}$ is small. Define the group sparsity $\ell_1$ norm of ${\boldsymbol \Theta}$ as $\|{\boldsymbol \Theta}\|_{2,1} =\sum_{i=1}^d \| {\boldsymbol \theta}_i\|_2$. A convex optimization algorithm \cite{tropp2006algorithms} that motivates group sparsity can be posed to solve the MMV model as \begin{equation} \label{cs_mmv} \hat{{\boldsymbol \Theta}}=\argmin_{{\boldsymbol \Theta}}\; \|{\boldsymbol \Theta}\|_{2,1} \quad \mbox{s.t.}\quad {\boldsymbol F}_{\Omega}{\boldsymbol \Theta} = {\boldsymbol Z}_{\Omega}, \end{equation} where ${\boldsymbol F}_{\Omega}$ is the subsampled DFT basis or DFT frame on $\Omega$. The signal then is recovered as $\hat{{\boldsymbol X}}={\boldsymbol F} \hat{{\boldsymbol \Theta}}$. However, when the frequencies $\mathcal{F}$ are off-the-grid, the model \eqref{model} becomes highly inaccurate due to spectral leakage along the Dirichlet kernel, making \eqref{cs_mmv} degrades significantly in performance. We will compare against this conventional approach with our proposed algorithm in Section~\ref{numerical}. \section{Atomic Norm Minimization For MMV Models} \label{proposal} In this section we develop the atomic norm minimization algorithm for solving the MMV model with spectrally-sparse signals. We first define an atom as \begin{equation} {\boldsymbol A}(f,{\boldsymbol b} ) = {\boldsymbol a}(f){\boldsymbol b}^* , \end{equation} where $f \in [0,1)$, ${\boldsymbol b}\in\mathbb{R}^{L}$ satisfying $\|{\boldsymbol b}\|_2 =1$, and the set of atoms as ${\mathcal A}=\{{\boldsymbol A}(f,{\boldsymbol b} ) | f \in [0,1), \|{\boldsymbol b}\|_2 =1\}$. Define \begin{align*} \| {\boldsymbol X}\|_{{\mathcal A},0} & = \inf_r \left\{ {\boldsymbol X} = \sum_{k=1}^r c_k {\boldsymbol A}(f_k,{\boldsymbol b}_k), c_k \geq 0 \right\}. \end{align*} as the smallest number of atoms to describe ${\boldsymbol X}$. Our goal is thus to minimize $\| {\boldsymbol X}\|_{{\mathcal A},0}$ that satisfies the observation, given as \begin{equation} \label{atomic_zero} \min \|{\boldsymbol X}\|_{{\mathcal A},0} \quad \mbox{s.t.} \quad {\boldsymbol Z}_\Omega = \mathcal{P}_{\Omega}({\boldsymbol X}) . \end{equation} It is easy to show that $\| {\boldsymbol X}\|_{{\mathcal A},0}$ can be represented equivalently a \begin{align*} \| {\boldsymbol X}\|_{{\mathcal A},0} & = \inf_{{\boldsymbol u} ,{\boldsymbol W} } \left\{ \rank(\toep({\boldsymbol u})) \Big| \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & {\boldsymbol W} \end{bmatrix} \succ \bf{0} \right\}, \end{align*} where $\toep({\boldsymbol u})$ is the Toeplitz matrix generated by ${\boldsymbol u}$. Hence \eqref{atomic_zero} is NP-hard. We will alternatively consider the convex relaxation of $\| {\boldsymbol X}\|_{{\mathcal A},0}$, defining the atomic norm \cite{chandrasekaran2012convex} of ${\boldsymbol X}$ as \begin{align} \| {\boldsymbol X}\|_{{\mathcal A}} &= \inf \left\{ t>0: \; {\boldsymbol X}\in t\; \mbox{conv}({\mathcal A}) \right\} \nonumber \\ &= \inf \left\{ \sum_k c_k \Big| {\boldsymbol X} = \sum_k c_k {\boldsymbol A}(f_k,{\boldsymbol b}_k), c_k \geq 0 \right\}. \label{atomic_def} \end{align} The atomic norm of a single vector ${\boldsymbol x}_l$ defined in \cite{TangBhaskarShahRecht2012} becomes a special case of \eqref{atomic_def} for $L=1$. We propose to solve the following atomic norm minimization algorithm: \begin{equation}\label{primal} \hat{{\boldsymbol X}}=\argmin_{\boldsymbol X} \|{\boldsymbol X}\|_{\mathcal A} \quad \mbox{s.t.} \quad {\boldsymbol Z}_\Omega = \mathcal{P}_{\Omega}({\boldsymbol X}). \end{equation} \subsection{Semidefinite Program Characterization} We now prove the following equivalent semidefinite program (SDP) characterization of $\|{\boldsymbol X}\|_{\mathcal A}$, showing that \eqref{primal} can be solved efficiently using off-the-shelf SDP solvers. \begin{theorem}\label{atomic-sdp} The atomic norm $\|{\boldsymbol X}\|_{\mathcal A}$ can be written equivalently as \begin{align*} \| {\boldsymbol X}\|_{\mathcal A} = \inf_{{\boldsymbol u}\in\mathbb{C}^n,{\boldsymbol W}\in\mathbb{C}^{L\times L}} & \Big\{ \frac{1}{2}\trace(\toep({\boldsymbol u})) + \frac{1}{2}\trace({\boldsymbol W}) \Big| \\ & \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & {\boldsymbol W} \end{bmatrix} \succ \bf{0} \Big\}. \end{align*} \end{theorem} \begin{proof} Denote the value of the right hand side as $\|{\boldsymbol X}\|_{\mathcal T}$. Suppose that ${\boldsymbol X} =\sum_{k=1}^r c_k {\boldsymbol a}(f_k){\boldsymbol b}_k^*$, there exists a vector ${\boldsymbol u}$ such that $$ \toep({\boldsymbol u}) = \sum_{k=1}^r c_k {\boldsymbol a}(f_k) {\boldsymbol a}(f_k)^*, $$ by the Vandermonde decomposition lemma \cite{Caratheodory}. It is obvious that \begin{align*} &\quad \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & \sum_{k=1}^r c_k {\boldsymbol b}_k {\boldsymbol b}_k^* \end{bmatrix} \\ &= \begin{bmatrix} \sum_{k=1}^r c_k {\boldsymbol a}(f_k) {\boldsymbol a}(f_k)^* & \sum_{k=1}^r c_k{\boldsymbol a}(f_k) {\boldsymbol b}_k^* \\ \sum_{k=1}^r c_k {\boldsymbol b}_k{\boldsymbol a}(f_k)^* & \sum_{k=1}^r c_k {\boldsymbol b}_k {\boldsymbol b}_k^* \end{bmatrix} \\ & = \sum_{k=1}^r c_k \begin{bmatrix} {\boldsymbol a}(f_k) \\ {\boldsymbol b}_k \end{bmatrix} \begin{bmatrix} {\boldsymbol a}(f_k)^* & {\boldsymbol b}_k^* \end{bmatrix} \succ 0, \end{align*} and $$ \frac{1}{2}\trace(\toep({\boldsymbol u})) + \frac{1}{2}\trace({\boldsymbol W}) = \sum_{k=1}^r c_k = \| {\boldsymbol X}\|_{\mathcal A},$$ therefore $\|{\boldsymbol X}\|_{{\mathcal T}}\leq \|{\boldsymbol X}\|_{\mathcal A}$. On the other hand, suppose that for any ${\boldsymbol u}$ and ${\boldsymbol W}$ that satisfy $$ \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & {\boldsymbol W} \end{bmatrix} \succ \bf{0}, $$ with $\toep({\boldsymbol u}) = {\boldsymbol V}{\boldsymbol D}{\boldsymbol V}^*$, ${\boldsymbol D}=\diag(d_i)$, $d_i> 0$. It follows that ${\boldsymbol X}$ is in the range of ${\boldsymbol V}$, hence ${\boldsymbol X}={\boldsymbol V}{\boldsymbol B}$ with the columns of ${\boldsymbol B}^T$ given by ${\boldsymbol b}_i$. Since ${\boldsymbol W}$ is also PSD, ${\boldsymbol W}$ can be written as ${\boldsymbol W} = {\boldsymbol B}^*{\boldsymbol E}{\boldsymbol B}$ where ${\boldsymbol E}$ is also PSD. We now have $$ \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & {\boldsymbol W} \end{bmatrix} =\begin{bmatrix} {\boldsymbol V} & \\ & {\boldsymbol B}^* \end{bmatrix} \begin{bmatrix} {\boldsymbol D} & {\boldsymbol I} \\ {\boldsymbol I} & {\boldsymbol E} \end{bmatrix} \begin{bmatrix} {\boldsymbol V}^* & \\ & {\boldsymbol B} \end{bmatrix} \succ \bf{0}, $$ which yields $$ \begin{bmatrix} {\boldsymbol D} & {\boldsymbol I} \\ {\boldsymbol I} & {\boldsymbol E} \end{bmatrix} \succ \bf{0} $$ and ${\boldsymbol E}\succ {\boldsymbol D}^{-1}$ by the Schur complement lemma. Now observe \begin{align*} \trace({\boldsymbol W})& = \trace({\boldsymbol B}^*{\boldsymbol E}{\boldsymbol B})\geq \trace({\boldsymbol B}^*{\boldsymbol D}^{-1}{\boldsymbol B}) \\ &= \trace({\boldsymbol D}^{-1}{\boldsymbol B}\bB^*) = \sum_i d_i^{-1} \| {\boldsymbol b}_i\|^2. \end{align*} Therefore, \begin{align*} &\quad\frac{1}{2}\trace(\toep({\boldsymbol u}))+\frac{1}{2} \trace({\boldsymbol W}) =\frac{1}{2} \trace({\boldsymbol D})+\frac{1}{2} \trace({\boldsymbol W}) \\ & \geq \sqrt{ \trace({\boldsymbol D})\cdot \trace({\boldsymbol W})} \\ & \geq \sqrt{\left(\sum_i d_i \right) \left( \sum_i d_i^{-1} \| {\boldsymbol b}_i\|^2 \right) } \\ & \geq \sum \|{\boldsymbol b}_i\| \geq \|{\boldsymbol X} \|_{\mathcal A}, \end{align*} which gives $\|{\boldsymbol X}\|_{\mathcal T} \geq \|{\boldsymbol X}\|_{\mathcal A}$. Therefore, $\|{\boldsymbol X}\|_{\mathcal T} = \|{\boldsymbol X}\|_{\mathcal A}$. \end{proof} From Theorem~\ref{atomic-sdp}, we can now equivalently write \eqref{primal} as the following SDP: \begin{align} \hat{{\boldsymbol X}}&= \argmin_{{\boldsymbol X}}\inf_{{\boldsymbol u},{\boldsymbol W}} \; \frac{1}{2}\trace(\toep({\boldsymbol u})) + \frac{1}{2}\trace({\boldsymbol W}) \label{primal-sdp}\\ & \mbox{s.t.} \; \begin{bmatrix} \toep({\boldsymbol u}) & {\boldsymbol X} \\ {\boldsymbol X}^* & {\boldsymbol W} \end{bmatrix} \succ \mathbf{0} , \; {\boldsymbol Z}_{\Omega} = \mathcal{P}_{\Omega}({\boldsymbol X}). \nonumber \end{align} \subsection{Dual Certification} An important question is when the algorithm \eqref{primal} admits perfect recovery. To this end, we study the dual problem of \eqref{primal}. Denote the optimal solution of \eqref{primal} as ${\boldsymbol X}^{\star}$. Let ${\boldsymbol Y}\in\mathbb{C}^{n\times L}$, define $\langle {\boldsymbol Y},{\boldsymbol X} \rangle = \trace({\boldsymbol X}^*{\boldsymbol Y})$, and $\langle {\boldsymbol Y},{\boldsymbol X} \rangle_{\mathbb{R}}=\mbox{Re}(\langle {\boldsymbol Y},{\boldsymbol X} \rangle)$. The dual norm of $\|{\boldsymbol X}\|_{\mathcal A}$ can be defined as \begin{align*} \|{\boldsymbol Y}\|_{{\mathcal A}}^* &=\sup_{\|{\boldsymbol X}\|_{\mathcal A}\leq 1} \langle {\boldsymbol Y},{\boldsymbol X}\rangle_{\mathbb{R}} \\ & = \sup_{f\in[0,1),\|{\boldsymbol b}\|_2=1} \langle {\boldsymbol Y}, {\boldsymbol a}(f) {\boldsymbol b}^* \rangle_{\mathbb{R}} \\ & = \sup_{f\in[0,1),\|{\boldsymbol b}\|_2=1} \left| \langle {\boldsymbol b}, {\boldsymbol Y}^*{\boldsymbol a}(f) \rangle \right| \\ & = \sup_{f\in[0,1)} \|{\boldsymbol Y}^*{\boldsymbol a}(f) \|_2 \triangleq \sup_{f\in[0,1)} \| {\boldsymbol Q}(f)\|_2, \end{align*} where ${\boldsymbol Q}(f) = {\boldsymbol Y}^*{\boldsymbol a}(f)$ is a length-$L$ vector with each entry a polynomial in $f$. The dual problem of \eqref{primal} can thus be written as \begin{equation} \label{dual} \max_{{\boldsymbol Y}} \; \langle {\boldsymbol Y}_\Omega, {\boldsymbol X}_{\Omega}^{\star} \rangle_{\mathbb{R}} \quad \mbox{s.t.} \quad \|{\boldsymbol Y}\|_{{\mathcal A}}^* \leq 1, {\boldsymbol Y}_{\Omega^c} = 0, \end{equation} where ${\boldsymbol Y}_{\Omega^c}$ denotes the projection of ${\boldsymbol Y}$ on the rows denoted by the location set $\Omega^c=\{1,\ldots, n\} \backslash \Omega$. Let $({\boldsymbol X},{\boldsymbol Y})$ be primal-dual feasible to \eqref{primal} and \eqref{dual}, we have $\langle {\boldsymbol Y},{\boldsymbol X}\rangle_{\mathbb{R}}=\langle {\boldsymbol Y},{\boldsymbol X}^{\star}\rangle_{\mathbb{R}}$. Strong duality holds since Slater's condition holds, and it implies that the solutions of \eqref{primal} and \eqref{dual} equal if and only if ${\boldsymbol Y}$ is dual optimal and ${\boldsymbol X}$ is primal optimal \cite{boyd2004convex}. Using strong duality we can obtain a dual certification to the optimality of the solution of \eqref{primal}. \begin{prop} \label{dual_certificate}The solution of \eqref{primal} $\hat{{\boldsymbol X}}={\boldsymbol X}^{\star}$ is the unique optimizer if there exists ${\boldsymbol Q}(f)= {\boldsymbol Y}^*{\boldsymbol a}(f) $ such that \begin{equation} \label{conditions} \begin{cases} (C1):\quad {\boldsymbol Q}(f_k) = {\boldsymbol b}_k, & \forall f_k \in \mathcal{F}, \\ (C2):\quad \| {\boldsymbol Q}(f) \|_2 < 1, &\forall f\notin \mathcal{F}, \\ (C3):\quad {\boldsymbol Y}_{\Omega^c} = 0.& \end{cases} \end{equation} \end{prop} \begin{proof} First, any ${\boldsymbol Y}$ satisfying \eqref{conditions} is dual feasible. We have \begin{align*} \|{\boldsymbol X}^{\star}\|_{{\mathcal A}} & \geq\|{\boldsymbol X}^{\star}\|_{{\mathcal A}}\|{\boldsymbol Y}\|_{{\mathcal A}}^* \\ &\geq \langle {\boldsymbol Y}, {\boldsymbol X}^{\star} \rangle_{\mathbb{R}} =\langle {\boldsymbol Y}, \sum_{k=1}^r c_k {\boldsymbol a}(f_k){\boldsymbol b}_k^* \rangle_{\mathbb{R}} \\ & = \sum_{k=1}^r \mbox{Re}\left( c_k \langle {\boldsymbol Y}, {\boldsymbol a}(f_k){\boldsymbol b}_k^* \rangle \right) \\ & = \sum_{k=1}^r \mbox{Re} \left( c_k \langle {\boldsymbol b}_k, {\boldsymbol Q}(f_k) \rangle \right) \\ & = \sum_{k=1}^r \mbox{Re} \left( c_k \langle {\boldsymbol b}_k, {\boldsymbol b}_k \rangle \right) = \sum_{k=1}^r c_k \geq \|{\boldsymbol X}^{\star}\|_{{\mathcal A}} . \end{align*} Hence $\langle {\boldsymbol Y}, {\boldsymbol X}^{\star} \rangle_{\mathbb{R}} =\|{\boldsymbol X}^{\star}\|_{{\mathcal A}}$. By strong duality we have ${\boldsymbol X}^{\star}$ is primal optimal and ${\boldsymbol Y}$ is dual optimal. For uniqueness, suppose $\hat{{\boldsymbol X}}$ is another optimal solution which has support outside $\mathcal{F}$. It is trivial to justify if $\hat{{\boldsymbol X}}$ and ${\boldsymbol X}^{\star}$ have the same support, they must coincide since the set of atoms with frequencies in $\mathcal{F}$ is independent. Let $\hat{{\boldsymbol X}}=\sum_k \hat{c}_k {\boldsymbol a}(\hat{f}_k)\hat{{\boldsymbol b}}_k^*$. We then have \begin{align*} \langle{\boldsymbol Y}, \hat{{\boldsymbol X}} \rangle_{\mathbb{R}} & = \sum_{\hat{f}_k\in\mathcal{F}}\mbox{Re} \left( \hat{c}_k \langle \hat{{\boldsymbol b}}_k, {\boldsymbol Q}(\hat{f}_k) \rangle \right) + \sum_{\hat{f}_l\notin\mathcal{F}}\mbox{Re} \left( \hat{c}_l \langle \hat{{\boldsymbol b}}_l, {\boldsymbol Q}(\hat{f}_l) \rangle \right) \\ & < \sum_{\hat{f}_k\in\mathcal{F}} \hat{c}_k + \sum_{\hat{f}_l \notin\mathcal{F}} \hat{c}_l = \|\hat{{\boldsymbol X}}\|_{{\mathcal A}}, \end{align*} which contradicts strong duality. Therefore the optimal solution of \eqref{primal} is unique. \end{proof} Proposition~\ref{dual_certificate} offers a way to certify the optimality of \eqref{primal} as long as we can find a dual polynomial ${\boldsymbol Q}(f)$ that satisfies \eqref{conditions}. This is left for future work. \begin{comment} \section{PSD Toeplitz Covariance Matrix Completion with Infinite Measurement Vectors}\label{infinite} If we assume the coefficient of the atoms $c_{k,l}$ are i.i.d. $\mathcal{N}(0,\sigma_k^2)$, then the covariance matrix of ${\boldsymbol x}_l$ can be written as \begin{align*} \boldsymbol{R} = \mathbb{E}[{\boldsymbol x}_l{\boldsymbol x}_l^*] = \sum_{k=1}^r \sigma_k^2 {\boldsymbol a}(f_k){\boldsymbol a}(f_k)^*, \end{align*} which is a positive semidefinite (PSD) Toeplitz matrix with $\mbox{rank}(\boldsymbol{R})=r\ll n$. Let the first column of $\boldsymbol{R}$ be denoted by $\boldsymbol{r}\in\mathbb{C}^{n}$. Assume each ${\boldsymbol x}_l$ share the same observation pattern $\Omega$, then we observe a subset of entries of $\boldsymbol{r}$, denoted by $\boldsymbol{r}_{\Omega}$. We aim to recover the matrix $\boldsymbol{R}$ by minimizing its rank given the observation as: \begin{align*} \min_{\toep({\boldsymbol u})\succeq 0}\;\mbox{rank}(\toep({\boldsymbol u})), \quad {\boldsymbol u}_{\Omega}={\boldsymbol r}_{\Omega}. \end{align*} Instead of solving this rank minimization directly, we propose to minimize the trace of $\toep({\boldsymbol u})$ instead using \begin{align} \label{toeplitz} \min_{\toep({\boldsymbol u})\succeq 0}\;\trace(\toep({\boldsymbol u})), \quad {\boldsymbol u}_{\Omega}={\boldsymbol r}_{\Omega}. \end{align} In \cite{chen2013spectral,chen2013robust}, it is shown that with $\mathcal{O}(r\log^2 n)$ measurements it is enough to recover a Toeplitz matrix under certain incoherence condition. However, when the Toeplitz matrix is further known as PSD, this problem is connected to the problem studied in \cite{donoho2005sparse,fuchs2005sparsity} on finding nonnegative sparse solutions in a Vandermonde dictionary, which shows a unique nonnegative solution exists when $\Omega=\{1,2,\ldots,m \}$ where $2m<n$ and $r\leq \lfloor\frac{m-1}{2} \rfloor$ without any incoherence condition or separation condition. However, neither of these existing results seem to predict the behavior of \eqref{toeplitz}, and we leave the theoretical performance of \eqref{toeplitz} to future work. In the numerical example, the algorithm \eqref{toeplitz} yields superior performance without imposing a separation condition. Finally, we note that \cite{pal2012correlation} proposed a correlation-aware algorithm to recover the frequencies when they are on a grid also using the second-order statistics. Here the frequencies are off the grid and therefore the nonnegative constraints appear as a PSD constraints on the Toeplitz matrix. \end{comment} \begin{figure*} \centering \begin{tabular}{cc} \includegraphics[width=0.44\textwidth]{dualpoly_single} & \includegraphics[width=0.44\textwidth]{dualpoly_multi} \\ (a) $ L=1$ & (b) $ L=3$ \end{tabular} \caption{The reconstructed dual polynomial for a randomly generated spectrally-sparse signal with $r=10$, $n=64$, and $m=32$: (a) $L=1$, (b) $L=3$.} \label{dualpoly} \end{figure*} \section{Numerical Experiments} \label{numerical} In this section, we evaluation the performance of the proposed algorithm \eqref{primal}, showing that the number of measurements per signal may be reduced as we increase the number of signals $L$ for achieving the same performance. \subsection{Phase transition when varying the number of signals} Let $n=64$ and $m=32$. In each Monte Carlo experiment, we randomly generate a spectrally-sparse signal with $r$ frequencies randomly located in $[0,1)$ that satisfies a separation condition $\Delta =\min_{k\neq l} |f_k-f_l| \geq 1/n$. This separation condition is slightly weaker than the condition asserted in \cite{TangBhaskarShahRecht2012} to guarantee the success of \eqref{primal} with high probability for $L=1$. For each frequency component, we randomly generate the amplitude for each signal from the standard complex Gaussian distribution $\mathcal{CN}(0,1)$. We run \eqref{primal-sdp} using CVX \cite{grant2008cvx} and calculate the reconstruction normalized mean squared error (NMSE) as $\|\hat{{\boldsymbol X}}-{\boldsymbol X}^{\star} \|_F/\|{\boldsymbol X}^{\star}\|_F$ where ${\boldsymbol X}^{\star}$ is the ground truth. The experiment is claimed successful if $\mbox{NMSE}\leq 10^{-5}$. For each pair of $r$ and $L$, we run a total of $50$ Monte Carlo experiments and output the average success rate. Fig.~\ref{fixed_m} shows the success rate of reconstruction versus the sparsity level $r$ for $L=1$, $2$, and $3$ respectively. As we increase $L$, the success rate becomes higher for the same sparsity level. \begin{figure}[h] \centering \includegraphics[width=0.45\textwidth]{scs_fixed_m} \caption{Success rate of reconstruction versus the sparsity level $r$ for $L=1,2,3$ when $n=64$, $m=32$ and the frequencies are generated satisfying a separation condition $\Delta\geq1/n$.} \label{fixed_m} \end{figure} Fig.~\ref{dualpoly} shows the reconstructed dual polynomial for a randomly generated spectrally-sparse signal with $r=10$ when $L=1$ and $L=3$ respectively. It can be seen that although the algorithm achieves perfect recovery for both cases, the reconstructed dual polynomial has a much better localization property when multiple signals are present. \begin{figure}[htp] \centering \begin{tabular}{cc} \hspace{-0.1in}\includegraphics[height=1.8in, width=0.25\textwidth]{cmp_offgrid_single} & \hspace{-0.15in} \includegraphics[height=1.8in,width=0.25\textwidth]{cmp_offgrid} \\ (a) $L=1$ & (b) $L=3$ \end{tabular} \caption{The NMSE of reconstruction versus the number of samples per signal $m$ for a randomly generated spectrally-sparse signal satisfying a separation condition $\Delta\geq 1/n$ with $r=8$, $n=64$: (a) $L=1$, (b) $L=3$.} \label{cmp_offgrid} \end{figure} \subsection{Comparison of CS and the proposed algorithm} Let $n=64$. We randomly generate $r=10$ frequencies on the unit circle $[0,1)$ satisfying a separation condition $\Delta\geq 1/n$ as in the previous setup. Fig.~\ref{cmp_offgrid} shows the NMSE of the reconstructed signal using the CS-MMV algorithm \eqref{cs_mmv} with a DFT basis, a DFT frame with an oversampling factor $c=2$, a DFT frame with an oversampling factor $c=4$, and the proposed atomic norm minimization algorithm \eqref{primal} for (a) $L=1$ and (b) $L=3$. The atomic norm minimization algorithm outperforms \eqref{cs_mmv} even when the reconstruction is not exact at smaller values of $m$. The CS algorithm \eqref{cs_mmv} can never achieve exact recover since a randomly generated frequency is always off the grid, while the recovery of the atomic norm algorithm \eqref{primal} is exact after $m$ exceeds a certain threshold for success. As we increase the number of signals $L$, both CS algorithms and the atomic norm algorithm improve their performance. \begin{comment} \subsection{Infinite Number of Measurement Vectors} In this example, we assume the number of measurement vectors is infinite and perfect correlation information is available. Using the same experiment setup as Fig.~\ref{fixed_m}, except that no separation conditions are imposed on the locations of frequencies, Fig.~\ref{random_cov} shows the success rate of reconstruction versus the sparsity level $r$ using the algorithm \eqref{toeplitz} by completing a PSD Toeplitz matrix. The performance is significantly better than that of Fig.~\ref{fixed_m} with a fixed number of signals. It also enjoys a smaller computational cost. \begin{figure}[htp] \centering \includegraphics[width=0.42\textwidth]{random_cov} \caption{Success rate of reconstruction versus the sparsity level $r$ using \eqref{toeplitz} when $n=64$, $m=32$ and the frequencies are generated {\em randomly} for the same observation patterns across signals.} \label{random_cov} \end{figure} \end{comment} \section{Conclusions} \label{concl} In this paper we study the problem of simultaneously recovering multiple spectrally-sparse signals that are supported on the same frequencies lying arbitrarily on the unit circle. We propose an atomic norm minimization problem, and solve it efficiently via semidefinite programming. Through numerical experiments, we show that the number of samples per signal may be further reduced by harnessing the joint sparsity pattern of multiple signals. The proposed atomic norm minimization algorithm can also be applied when the observation patterns are different across different signals. Future work is to develop theoretical guarantees of the proposed algorithm and examine its performance in the presence of noise. Another interesting direction is to study MMV extensions of the off-the-grid approach in \cite{chen2013spectral,chen2013robust}. \bibliographystyle{IEEEtran}
\section{Supplementary Material} \subsection*{1. Synthesis and Measurement Techniques} Our two series of heterostructures A and B were both grown using a standard pulsed laser deposition system manufactured by Twente Solid State Technology B.V., equipped with a reflection high-energy electron diffraction (RHEED) facility. 10 unit cells of {\La} were deposited on TiO$_2$-terminated 0.5mm thick commercial {\Sr} (001) substrates from Shinkosha. The total incident laser energy was 9~mJ, focussed onto a 6~mm$^2$ rectangular spot. For both samples, the oxygen pressure was maintained at 10$^{-3}$~mbar during growth at 800$^\circ$C. However, series A underwent a subsequent annealing stage: after cooling to 500$^\circ$C at 10$^{-3}$~mbar, the oxygen pressure was increased to 0.1~bar. The temperature was held at 500$^\circ$C for 30 minutes before natural cooling to room temperature, still in 0.1~bar oxygen. In contrast, series B was allowed to cool naturally to room temperature in 10$^{-3}$~mbar oxygen. Hall bars with Au-Ti contact pads were patterned onto the top surface of the {\La} using electron-beam lithography, while sample B also had an Au-Ti back gate sputter-deposited across the entire base of the {\Sr} substrate prior to patterning. The Hall bar width was 80~$\mu$m and the voltage contact separation 660~$\mu$m. Patterning onto the {\La} surface rather than directly contacting the interface allows us to remain sensitive to AlO$_2$ surface transport in parallel with interfacial states, even when the interface is superconducting (see section 2). Transport measurements were performed in a Janis cryogen-free dilution refrigerator, using an ac technique with two SRS 830 lock-in amplifiers and a Keithley 6221 AC current source. All data were acquired using an excitation current of 500~nA at 19~Hz; this value was chosen to maximise the signal-to-noise ratio whilst eliminating any sample heating effects at milliKelvin temperatures. Our noise threshold is approximately 2~nV. The evolution of the capacitance ($\sim$~1~nF) of the {\Sr} substrate with gate voltage $V_g$ was verified using a General Radio 1621 manual capacitance bridge: no leakage or breakdown occurred even at $V_g$~=~500~V. \subsection*{2. Electrical Resistivity and Superconducting Transitions} Upon cooling from room temperature, the electrical resistance of both heterostructures passes through a minimum at low temperature (fig.~S\ref{FigS0}), then rises logarithmically prior to the onset of superconductivity at $\sim$~0.3~K. The minima are located at $T_m$~=~10~K and 25~K for samples A and B respectively. A logarithmic divergence in the low-temperature resistance is a signature of the Kondo effect, i.e. scattering off dilute magnetic impurities. Within the Kondo scenario, sample B (which has a high oxygen vacancy concentration) must contain a higher density of magnetic scattering centres due to its higher $T_m$. \begin{figure}[htbp] \centering \includegraphics*[width=8cm,clip]{FigS1} \caption[]{\label{FigS0} Variation of the sheet resistance $R_{xx}(T)$ upon cooling from room temperature for sample A (left) and B (right). The resistance below 35~K is magnified in the insets to each graph, illustrating the Kondo minima. } \end{figure} The resistive transitions to the superconducting state used to determine the critical fields in fig.~1 of our manuscript are shown in fig.~S\ref{FigS1}a. Although the transitions qualitatively follow the behaviour expected for a quasi-2D superconducting film, the resistance does not fall to zero even at the lowest temperatures measured (0.035~K). This does not necessarily imply that our interface is inhomogeneous; in fact, non-zero resistance is a natural consequence of the patterning technique which we have used. A schematic diagram for our pattern is shown in fig.~S\ref{FigS1}c: since the contacts for our Hall bars are only deposited onto the top surface of the {\La}, there is no direct contact with the conducting interface. Instead, contact is made vertically through the 10 unit cells of {\La}, which exhibits a weak conductivity dependent on the oxygen vacancy concentration. This provides a small resistive component in series with the interface, leading to a measured non-zero resistance even with a homogeneous superconducting interface. Any conducting AlO$_2$ surface states will generate a parallel contribution to the measured resistance; the advantage of this patterning technique is that it enables these surface states to be probed without being ``shorted out'' by the superconducting interface. Conversely, this does not easily permit us to differentiate between homogeneous and inhomogeneous superconducting layers (or even field-induced inhomogeneous nucleation of superconductivity). However, this does not affect the conclusions of our work in any manner: firstly, inhomogeneities at the interface are entirely expected due to the tendency of the $d_{xy}$ electrons to localise and form ferromagnetic zones where superconductivity is suppressed. For sufficiently high densities of ferromagnetic inclusions above a shallow superconducting channel, the percolative zero-resistance current path may vanish. Secondly, the observed quantum oscillations originate from carriers deeper below the interface and only emerge at high magnetic fields when the superconductivity has been entirely quenched. Thirdly, if we consider the oxygen-deficient high carrier density heterostructure B, it is plausible that clusters of oxygen vacancies at the AlO$_2$ surface could locally overdope the interface beyond its maximum superconducting carrier density within certain discrete zones (also enhancing the ferromagnetism, as suggested by our Kondo effect and theory~\cite{SPavlenko-2012,*SPavlenko-2012-2}). This is a natural consequence of vacancy-doping the {\La}/{\Sr} interface and in no way reflects negatively on our results. \begin{figure}[htbp] \centering \includegraphics*[width=7.5cm,clip]{FigS2} \caption[]{\label{FigS1} (a) Temperature dependence of the longitudinal resistance $R_{xx}(T)$ in sample A (left) and B (right) within the superconducting (SC) phase, for a range of magnetic fields applied perpendicular to the interface ($H{\perp}(001)$, above) and parallel to the interface ($H//[110]$, below). All data are normalised to the resistance at $T$~=~0.5~K. (b) Temperature dependence of the parallel and perpendicular upper critical fields $H_{c2//,\perp}(T)$ for each heterostructure. These data are extracted from our $R_{xx}(T)$ curves: we define $T_c(H){\equiv}H_{c2}(T)$ as the temperatures at which $R_{xx}$ falls by 20\% of the difference between its values at 0.5~K and 0.04~K in zero field, indicated by the dashed lines intersecting the two $R_{xx}(T)$ plots in (a). Note that this 20\% criterion is arbitrary, since although changing the percentage will lead to small variations in the calculated coherence length $\xi$, the anisotropy and hence our determination of the superconducting channel thickness $d$ will remain unchanged. A scaling analysis for 2D SC~\cite{SSchneider-2008,SReyren-2009} is also shown, which provides an alternative means to determine $d$ using $H^{2}_{c2//}(T)$~=~$\frac{\pi\Phi_0}{2d^2}$$H_{c2\perp}(T)$. This yields SC layer thicknesses $d=16\!\pm\!1$~nm and $9\!\pm\!1$~nm for samples A and B respectively, in excellent agreement with the values from Ginzburg-Landau theory ($18\!\pm\!1$~nm and $9\!\pm\!1$~nm). (c) Schematic of the experimental setup and Hall bar patterning, indicating the resistive component from the 10 unit cells of {\La} which we always measure in series with the superconducting interface, together with the AlO$_2$ surface states in parallel with the interface. } \end{figure} Furthermore, using data for the critical current density from the literature~\cite{SStornaiuolo-2012} ($\sim$~40~nA per micron channel width), we estimate that the critical current in our Hall bars is of the order of several microAmps at zero field. Our measurement current (500~nA) is only one order of magnitude smaller than this value, thus contributing to the broadening of the transitions which we see in a magnetic field. We stress that a current of this magnitude is essential to achieve an adequate signal-to-noise ratio in highly conductive materials such as sample B, especially for Shubnikov-de Haas measurements. It also facilitates depinning by exciting a lateral ``shaking'' force on vortices, thus influencing (though not causing) the hysteresis in our in-plane magnetoresistance data. \subsection*{3. Hall Effect Data} All stated sheet carrier densities in our work have been obtained by linear fits to the high-field Hall resistance $R_{xy}(H)$ ($H>$~5~T), where $n_{2D}\frac{dR_{xy}}{dH}=-\frac{1}{e}$. However, both our Shubnikov-de Haas effect and various magnetotransport studies in the literature~\cite{SBell-2009,SShalom-2010,SJoshua-2011} have revealed evidence for multiple conduction bands at the interface, which should lead to Lorentzian forms for both the perpendicular magnetoresistance (MR) $R_{xx}(H_{\perp})$ and the Hall coefficient $R_H$. We plot $R_{xx}(H_{\perp})$ and $R_H$ in fig.~S\ref{FigS2}. \begin{figure}[htbp] \centering \includegraphics*[width=8cm,clip]{FigS3} \caption[]{\label{FigS2} (a) Perpendicular magnetoresistance $R_{xx}(H_\perp)$ for both heterostructures. (b) Above: Hall resistance $R_{xy}$ for both samples, including gate voltage dependence for sample B. Below: Field-dependent Hall coefficient $R_H$ for each sample, highlighting the Lorentzian form for sample B characteristic of a multi-band conductor. Inset: variation of the gradient of the Hall coefficient $dR_H/dH$ with gate voltage at $H$~=~9~T. } \end{figure} The simple message which we wish to convey here is the dramatically different behaviour of both $R_{xx}$ (fig.~S\ref{FigS2}a) and $R_H$ (fig.~S\ref{FigS2}b) for the two heterostructures. Examining the MR first, we observe that the curvature of $R_{xx}$ is positive for sample A, compared with negative curvature and a Lorentzian form in sample B. The Hall effect is even more revealing, with $R_H$ in sample A displaying approximately linear behaviour, which for {\La}/{\Sr} implies $d_{xy}$ single-band occupancy (slight deviations from linearity may be due to limited hole-like contributions from carriers at the AlO$_2$ top surface). In contrast, $R_H$ in sample B again shows the Lorentzian shape expected for a two-band system (i.e. $d_{xy}$ and $d_{xz,yz}$ band occupancy). Beyond the ability to extract the total carrier density, the failings of simple two-band Hall coefficient models are well-known~\cite{SJoshua-2011} due to the significant differences in the field dependence of the mobilities of carriers in each band; we therefore do not attempt any more detailed quantitative analysis of our data. One further unusual feature in our Hall data is worthy of mention: a crossover in the gradient of $R_H$ from negative to positive at high gate voltages (fig.~S\ref{FigS2}b, inset). This cannot be explained by a simple interfacial two-band model and is a consequence of the gradual population of states deeper within the substrate. We note that the total carrier density which we measure for sample A is 2.3$\times$10$^{13}$~cm$^{-2}$, slightly larger than the critical density $1.68\!\pm\!0.18\!\times\!10^{13}$~cm$^{-2}$ recently obtained for the Lifshitz transition~\cite{SJoshua-2011}. However, we see no evidence for two-band behaviour in the Hall coefficient of sample A and we conclude that the extra carriers which we detect most probably originate from deeper within the {\Sr} substrate (since any AlO$_2$ surface states should still be hole-like within this doping range). This is an important point, since it absolves the $d_{xz,yz}$ bands of responsibility for generating our observed ferromagnetic domains. \subsection*{4. Coexistent Ferromagnetism and Superconductivity, and their Evolution with Gate Voltage} \begin{figure}[htbp] \centering \includegraphics*[width=8cm,clip]{FigS4} \caption[]{\label{FigS3} (a) Comparison of the hysteretic parallel magnetoresistance $R_{xx}(H_{//}$ at $V_g$~=~0~V and 350~V. (b) Parallel and perpendicular upper critical fields $H_{c2//,\perp}$ at $V_g$~=~350~V, together with Ginzburg-Landau fits (pink and grey lines) indicating a coherence length $\xi$~=~82~$\pm$~2~nm and SC channel thickness $d~$~=~19~$\pm$~2~nm. A 2D scaling analysis which yields $d$~=~16~$\pm$~3~nm is also shown. (c) Hysteresis in the perpendicular magnetoresistance $R_{xx}(H_\perp)$. The minima in $R_{xx}$ occur at $\pm$~0.005~T, which is of the same order of magnitude as the remanent field in most large superconducting coils. } \end{figure} Let us consider the effects of applying a gate voltage to sample B on the ferromagnetic and superconducting phases. SQUID microscopy studies have indicated that gating the {\La}/{\Sr} interface to modulate its carrier density does not have any effect on the density of ferromagnetic inclusions~\cite{SKalisky-2012,SBert-2012}. In contrast, applying a positive gate voltage to increase $n_{2D}$ in sample B does change the shape of the hysteretic peaks, which are broader and clearer at $V_g$~=~350~V (fig.~S\ref{FigS3}a). Numerous potential explanations exist for this effect. Firstly, we suggest that the electric field across the {\Sr} may lead to a further increase in the vortex mobility once depinning has occurred, thus increasing the measured resistance even for small applied fields. We must also consider the expansion of the superconducting channel upon gating: the channel roughly doubles in thickness between $V_g$~=~0~V and 350~V (fig.~S\ref{FigS3}b). A thicker channel will increase the probability of pinning any given vortex during the rotation of its respective ferromagnetic dipole, hence broadening the hysteretic peak. However, the ``hardest'' pins (from the largest defects) will be located closer to the interface and hence the maximum field at which hysteresis is observed should remain similar: from fig.~S\ref{FigS3}a, this is indeed the case. Another relevant factor in modifying the hysteretic peak shape may be a partial suppression of superconductivity very close to the interface due to the high carrier density in sample B. This would also explain its narrower as-grown superconducting channel compared with sample A, although it is important to remember that the as-grown vertical charge distribution profile is a crucial factor in determining the absolute superconducting layer thickness and this may vary significantly between heterostructures. A very small hysteresis is visible in the out-of-plane MR $R_{xx}(H_\perp)$, although this occurs at fields close to zero, comparable to the typical remanence in superconducting magnets. For completeness, we plot this in fig.~S\ref{FigS3}c. We stress that the absence of any large hysteretic peaks in $R_{xx}(H_\perp)$ is entirely expected, since in this configuration there is no rotation of the in-plane moments; instead, the vortex density is much higher and we enter a liquid phase at very low applied fields. \subsection*{5. Data Reproducibility} \begin{figure*}[htbp] \centering \includegraphics*[width=17.5cm,clip]{FigS5} \caption[]{\label{FigS4} (a) Temperature-dependent resistance R(T) of another A-type {\La}/{\Sr} heterostructure, grown by pulsed-laser deposition (PLD) at 10$^{-3}$~mbar and post-annealed in 0.1~mbar O$_2$. Data from sample A studied in our manuscript is included for comparison. Inset: Kondo effect in this A-type heterostructure. \newline(b) R(T) in two B-type {\La}/{\Sr} heterostructures (grown by PLD at 10$^{-3}$~mbar with no post-annealing), compared with sample B from our manuscript. The low temperature resistance is less than 1$\Omega/\square$ in all three heterostructures. Inset: zoom on R(T) at low temperature for B-type heterostructure 1, with data from sample B for comparison. Although the Kondo effect is not as clear in this film as in sample B (due to a large parallel conductance from mobile electrons deeper within the non-magnetic bulk {\Sr}), a distinct kink is still present below 25~K and is indicative of a high density of magnetic scattering centres. \newline(c) Hall effect at 1.5~K for the A-type heterostructure shown in (a), together with data from sample A in our manuscript. Both heterostructures exhibit single-band transport, with 2D carrier densities $n_{2D}$ = 2.3$\times$10$^{13}$~cm$^{-2}$ for sample A and 3.0$\times$10$^{13}$~cm$^{-2}$ for the second A-type heterostructure. It should be noted that the low-temperature values of R(T) (roughly 1k$\Omega/\square$) and $n_{2D}$ in our A-type heterostructures are typical of annealed {\La}/{\Sr} films in the literature. \newline(d) Hall effect at 1.5~K for the B-type heterostructure 1 shown in (b), together with data from sample B in our manuscript. The Hall resistance exhibits a similar non-linear trend in both heterostructures, which is characteristic of multi-band transport. We measure $n_{2D}$~=~7.6$\times$10$^{15}$~cm$^{-2}$ for the B-type heterostructure 1, compared with 6.4$\times$10$^{14}$~cm$^{-2}$ for sample B. \newline(e) Interfacial ferromagnetism and superconductivity in the 3$\times$10$^{13}$~cm$^{-2}$ A-type heterostructure from figs.~(a,c). Hysteretic behaviour for H~$<$~0.16~T is observed in the in-plane magnetoresistance below the superconducting transition, similar to that seen in sample A (Fig. 2a in our manuscript). \newline(f) Interfacial ferromagnetism and superconductivity in the 7.6$\times$10$^{15}$~cm$^{-2}$ B-type heterostructure from figs.~(b,d). The in-plane magnetoresistance is hysteretic for H~$<$~0.15T, although the amplitude of the hysteresis is slightly smaller than that in sample B (fig.~2b in our manuscript) due to the larger parallel conductance from the high-mobility electron gas deeper within the {\Sr} substrate. Shubnikov-de Haas oscillations are visible at in-plane fields above 2.5~T (top inset), with a characteristic frequency of 23.8~T, similar to that reported in sample B. } \end{figure*} Over an 18-month period spanning the duration of this project, we synthesized numerous ``A-type'' and ``B-type'' heterostructures, all of which exhibited qualitatively similar behaviour. Series A have $n_{2D}\sim10^{13}$~cm$^{-2}$ and display single-band transport, while $n_{2D}>10^{14}$~cm$^{-2}$ and the Hall coefficient is non-linear in series B. Superconductivity is observed to coexist with ferromagnetism regardless of the carrier density, while Shubnikov-de Haas oscillations emerge in parallel fields below series B interfaces. Data-sets from several other heterostructures may be found in fig.~S\ref{FigS4}.
\section{Introduction} The Conformal Field Theory (CFT) approach aims at constructing two-dimensional quantum field theories whose correlation functions satisfy the infinite set of constraints imposed by conformal invariance \cite{DiFra,ItSaZu}. A CFT is characterized by a set of states and operators which are determined by the representations of the Virasoro algebra, or of other chiral algebras if additional symmetries are present. A consistent CFT has to satisfy general requirements of a quantum field theory, in particular the conformal bootstrap equations expressing the associativity of the operator algebra \cite{BPZ}. The CFT approach provides access to the study of critical phases of statistical systems in two dimensions. Certain properties of critical statistical models, such as the spin or the energy correlation functions in the Ising and in the Potts model, are given by a particular family of CFTs, the minimal models. These are exactly solvable CFTs which are constructed from a finite number of degenerate Virasoro representations, the correlation functions of which satisfy certain differential equations \cite{BPZ}. The two-dimensional critical phases can also be characterized by geometrical objects which take the form of conformally invariant random fractals \cite{Durev}. Examples of conformally invariant random fractals are the Potts spin domain walls or the Fortuin and Kasteleyn \cite{FK1, FK2} clusters. The question is then if one can use CFT methods as a tool to capture the geometric properties of critical phases, or more in general, to describe the conformal random geometry. In the last twenty years, there has been an intense effort in this direction, and a series of remarkable results followed: the fractal dimension of many random paths as well as different geometric exponents controlling, for instance, the reunion probability of an ensemble of self-avoiding walks, have been obtained \cite{Durev}. These results have been possible also thanks to the development of a palette of advanced techniques, such as Coulomb Gas (CG) methods \cite{Ni} or Stochastic Loewner Evolutions\cite{BaBe}. However, despite many successes, the CFT approach to the study of conformally invariant fractals is in many respects, unsatisfactory. On the one hand, many of the results found so far are for quantities related to two-point correlation functions, while the fine structure of CFT fully manifests itself only at the level of three- and four-points correlation functions. The only exceptions are the observables that satisfy some known differential equation, such as the probability measure associated to an SLE interface \cite{BaBe}. On the other hand, besides notable exceptions\cite{GaKa,GaKao,Ka,GaRuWo}, the only models which are known to satisfy the conformal bootstrap and which describe statistical models are the minimal models. The minimal models are too simple to capture the geometrical properties of critical phases whose description requires instead representations of the Virasoro algebra with more complicated features than those of the minimal models. In particular, differently from the case of the minimal models, the theory may contain indecomposable representations with a non-diagonalizable dilatation operator: this leads to the appearance of logarithms in the correlation functions \cite{Gu}. These last years have seen extensive investigations on logarithmic CFTs (LCFT)\cite{Ca_rev_log,CrRi,Gu_rev, GJRSV}. A better comprehension of LCFT representation modules and their fusion rules has been possible from the study of non semi-simple representations of the Virasoro algebra \cite{FuHwSeTi,FeGaSeTio,FeGaSeTi,PeRa} and from lattice approaches \cite{PeRaZu,ReSa,GaJaVaSa}. However, the construction of a consistent CFT, obtained by combining the chiral and anti-chiral spaces of representation with indecomposable Virasoro modules, is an hard problem and is much less under control. Recent progress in this direction come from the study of the $c=0$ bulk theories \cite{VaGaJaSa,Ri} and by the use of the category theory \cite{FuScSt}. Other insights into the CFTs describing conformal invariant fractals came from a completely new perspective. It was conjectured in \cite{DeVi} that the probability that a percolation cluster, or more in general an FK cluster, connects three given points is given at criticality by the structure constants computed in \cite{Za,KoPe}, see eq. (\ref{Zam_structure}). This conjecture found a strong numerical verification for the percolation clusters \cite{Ziff} and the FK clusters \cite{DPSV}. In the following we will refer to these structure constants as the time-like Liouville structure constants since they are obtained within the time-like Liouville theory \cite{HaMaWi}, corresponding at the classical level to the analytic continuation of the usual Liouville theory \cite{TeLiouville, Zam_lectures}. Their relevance in the percolation properties of Potts models, together with the fact that they represent the only consistent \cite{DPSV} analytic continuation of the minimal model structure constants \cite{DoFa,DoFa2,DoFa3}, strongly suggests the time-like Liouville structure constants are the basic building blocks for constructing new non-minimal Virasoro CFTs, at least when additional symmetries do not play any non-trivial role \cite{DPSV2}. A natural question then is whether and how the logarithmic behavior of the correlation functions can be spotted directly from the time-like Liouville structure constants. An analogous question has been answered in \cite{Zamhem} for the Liouville (DOZZ) structure constants \cite{DoOt,ZaZa}. In this paper we consider this problem analyzing the four-point correlation functions which posses an integral representation, the so-called CG integral representations \cite{DoFa,DoFa2}. The CG integrals provide correlation functions in which the chiral anti-chiral sectors of the CFT are glued symmetrically. This is in general not the case in LCFTs. There are nevertheless notable examples in which logarithmic CG correlation functions and thus diagonal LCFTs have been applied in the computation of bulk critical geometrical properties \cite{Salog,HaYa,Ya,JDPSW, DeVi,DPSV} and their validity checked numerically \cite{JDPSW, Ziff, DPSV}. Our analysis extends in particular these results to the multi-screening case and can be potentially relevant for the study of critical bulk systems. This paper is organized as follows. In Section~2 we introduce the basic CFT tools, needed for our analysis: the recursive formula for the conformal blocks \cite{Zam_rec1, Zam_rec2} and the time-like Liouville structure constants \cite{Za, KoPe}. In Section~3, we examine in details the one-screening CG integral and show that the mechanism of cancellation of the leading singularity and the generation of logarithmic terms is related to a precise relation between the residue of the conformal block and the structure constants, see eq. \eqref{simp_rel_log}. Such considerations are generalized in Sec.~4 to the multi-screening case for irrational values of the central charge, see in particular eqs. (\ref{canc_sing},~\ref{canc_sing0}), whereas in Sec.~5 the special cases $c=-2$ and $c=0$ are analyzed. After summarizing our findings in the Conclusions in Sec.~6, two appendices complete the paper. \section{Conformal blocks and structure constants in Virasoro CFT} In the first part of this introductory section we will review the definition of conformal block in CFT. In the second part we will discuss, within the Coulomb gas framework, the structure constants for a Virasoro CFT. \subsection{Algebric data: conformal blocks} Conformal symmetry implies the existence of an holomorphic $T(z)$ and an antiholomorphic $\bar{T}(\bar{z})$ stress-energy tensor, whose coefficients $L_n$ and $\bar{L}_n$ in the Laurent expansion \begin{equation} \label{Laurent} T(z)=\sum_{n=-\infty}^{\infty}\frac{L_n}{z^{n+2}},\quad \bar{T}(\bar{z})=\sum_{n=-\infty}^{\infty}\frac{\bar{L}_n}{\bar{z}^{n+2}} \end{equation} are the generators of the conformal algebra $\mathcal{A}$ acting on the field $\Phi(z,\bar{z})$. Coordinate transformations which involve only analytic or antianalytic functions lead to independent field variations and, as a consequence, $\mathcal{A}$ is the tensor product of the two commuting Virasoro algebra of central charge $c$ \begin{equation} \left[L_n,L_m\right]=(n-m)L_{n+m}+\frac{c}{12} n(n^2-1)\delta_{n+m,0},~\left[\bar{L}_n,\bar{L}_m\right]=(n-m)\bar{L}_{n+m}+ \frac{c}{12} n(n^2-1)\delta_{n+m,0}. \label{vir} \end{equation} The action on the Hilbert space of the primary field \cite{BPZ} $\Phi(z,\bar{z})$ of conformal dimension $\Delta+\bar{\Delta}$ is equivalent to that of the tensor product $\phi_{\Delta}(z)\otimes\bar{\phi}_{\bar{\Delta}}(\bar{z})$, with $\phi_{\Delta}$ (resp. $\bar{\phi}_{\bar{\Delta}}$) primary fields of conformal dimension $\Delta$ (resp. $\bar{\Delta}$). In the following we will always assume $\Delta=\bar{\Delta}$, which in particular corresponds to the case of scalar operators. The Hilbert space of a CFT based on a semi-simple Virasoro representation is a direct sum of Verma modules $\mathcal{V}_{\Delta}\otimes\mathcal{\bar{V}}_{\bar{\Delta}}$ and each of them, for example $\mathcal{V}_{\Delta}$, is built from the highest weight state $|\Delta\rangle\equiv\phi_\Delta(0)|0\rangle$, upon applications of the generators $L_{n}$ ($n<0$). A standard basis for the Verma module $\mathcal{V}_{\Delta}$ is the basis of the descendant states $L_{-n_1}\dots L_{-n_k}|\Delta\rangle$, with $n_{i+1}\geq n_{i}>0$. A descendant at level $D$, with $D=\sum_{i=1}^k n_i$, is indexed by the partition ${N}=\{n_1,\dots,n_k\}$ of $D$ and indicated with the notation $|\{N\}\rangle$. An inner product can be defined in the Verma module $\mathcal{V}_{\Delta}$ through \begin{equation} (L_n)^{\dagger}\equiv L_{-n},\quad \langle\Delta|\equiv\lim_{z\rightarrow\infty}z^{2\Delta}\langle 0|\phi_{\Delta}(z). \end{equation} and CFT correlation functions can be regarded as bilinear forms in the Hilbert space. The two-point function corresponds to the normalization choice of the inner product \begin{equation} \label{2point} \langle\Delta|\Delta'\rangle=\delta_{\Delta,\Delta'} \end{equation} and the three-point function of scalar primary fields follows from the knowledge of the structure constants \begin{equation} \label{3point} C(\Delta_1,\Delta_2,\Delta_3)=\lim_{z\rightarrow\infty}|z|^{4\Delta_3}\langle\Phi_{\Delta_3}(z,\bar{z})\Phi_{\Delta_2}(1) \Phi_{\Delta_1}(0)\rangle. \end{equation} Structure constants are not determined algebraically but are rather free parameters of the theory that must be fixed imposing additional constraints. Once they are known all the correlation functions can be in principle computed, starting from the four-point function which will be at length discuss in this paper. Introducing the short-hand notation $\{\Delta\}$ for the set of external dimensions and recalling \cite{BPZ} one has \begin{eqnarray} \label{4point} \lim_{z_4\rightarrow\infty}&& |z_4|^{4\Delta_4}\langle \Phi_{\Delta_4}(z_4,\bar{z_4})\Phi_{\Delta_3}(1) \Phi_{\Delta_2}(x,\bar{x})\Phi_{\Delta_1}(0)\rangle=\nonumber \\ && \sum_p |x|^{-2(\Delta_1-\Delta_2+\Delta_p)} C(\Delta_1,\Delta_2,\Delta_p)C(\Delta_3,\Delta_4,\Delta_p)~|F(x|c,\Delta_p,\{\Delta\})|^2, \end{eqnarray} The summation in \eqref{4point} is over all the possible internal fields with conformal dimensions $\Delta_p$ and those can constitute a continuous set. The CFT is well-defined when \eqref{4point} produces crossing symmetric correlation functions. $F(x|c,\Delta_p,\{\Delta\})$ is the holomorphic conformal block, usually represented by the diagram \begin{equation} \begin{tikzpicture} \draw (-3,0.5) node[left]{$F(x|c,\Delta_p,\{\Delta\})=$}--(-3.0001,0.5); \draw (-1,0.0) node[left]{$\Delta_1,0$}-- (0,0); \draw (0,1) node[above]{$\Delta_2,x$}--(0,0); \draw (0,0)--(1,0) node[above]{${\small \Delta_p}$}--(2,0); \draw (2,0)-- (2,1) node[above]{$\Delta_3, 1$}; \draw (2,0)--(3,0) node[right]{$\Delta_4, \infty$}; \end{tikzpicture} \label{diag_cb} \end{equation} The holomorphic conformal block $F(x|c,\Delta_p,\{\Delta\})$ can be written in term of the following expansion \begin{equation} \label{conformal_block} F(x|c,\Delta_p,\{\Delta\})=\sum_{D=0}^{\infty}x^D\sum_{\{N\},\{M\}}\Gamma_N H^{-1}_{NM} \tilde{\Gamma}_M, \end{equation} where \begin{equation} H_{NM}=\langle\{N\}|\{M\}\rangle, \end{equation} is the Shapovalov matrix and the coefficients $\Gamma_N$ are given by $\Gamma_N=\prod_{j=1}^k(\Delta_p+\sum_{i=1}^{j-1} n_i+n_j\Delta_3-\Delta_4)$, $\tilde{\Gamma}_M=\prod_{j=1}^l(\Delta_p+\sum_{i=1}^{j-1} m_i+m_j\Delta_2-\Delta_1)$. Let us introduce the standard parameterizations for the central charge \begin{equation} \label{central} c=1-24\alpha_0^2,\quad 2\alpha_0=\beta-\beta^{-1} \end{equation} and for the conformal dimensions \begin{equation} \label{dimension} \Delta(\alpha)=\alpha(\alpha-2\alpha_0), \end{equation} where $\beta$ and $\alpha$ (the charge) are reals. The equation \eqref{dimension} shows that the charges $\alpha$ and $2\alpha_0-\alpha$ lead to the same conformal dimensions; since we are assuming a non-degenerate spectrum, through all the paper the primary fields with dimension $\Delta(\alpha)$ and $\Delta(2\alpha_0-\alpha)$ will be identified. From the study of the representation theory of the Virasoro algebra it follows that when the charges $\alpha$ belong to the lattice\footnote{Notice that the symmetry $\alpha\rightarrow 2\alpha_0-\alpha$ of the conformal dimensions is equivalent to identify field with charges $\alpha_{r,s}$ and $\alpha_{-r,-s}$.} \begin{equation} \label{kac_lattice} \alpha_{r,s}=\frac{1-r}{2}\beta+\frac{s-1}{2\beta},\quad r,s\in\mathbb N \end{equation} the Verma module $\mathcal{V}_{rs}$ of the field $\phi_{r,s}$ with conformal dimension $\Delta_{r,s}\equiv\Delta(\alpha_{r,s})$, contains a vector $|\chi_{r,s}\rangle$ at level $rs$ with vanishing norm. In general, the vector $|\chi_{r,s}\rangle$ has the form \begin{equation} |\chi_{r,s}\rangle=\bigl[L_{-1}^{rs}+d_1L_{-2}L_{-1}^{rs-2}+\dots]|\phi_{r,s}\rangle, \end{equation} where a particular normalization has been chosen. The null vector $|\chi_{r,s}\rangle$ renders singular the Shapovalov matrix $H$ and ill-defined the expansion \eqref{conformal_block}. The fields $\phi_{r,s}$ with $r,s>0$, are termed degenerate. The existence of null vectors in Verma modules led \cite{Zam_rec1, Zam_rec2} to the recursive formula for the conformal block \begin{equation} \label{rec_conf} F(x|c,\Delta_p,\{\Delta\})=g(x|c,\Delta_p,\{\Delta\})+ \sum_{r,s > 0}\frac{S_{rs}x^{rs}}{\Delta_p-\Delta_{r,s}}F(x|c,\Delta_p+rs,\{\Delta\}). \end{equation} The residue $S_{rs}$ is a polynomial of degree $rs$ in each external charge that vanishes when the fusion rules \cite{BPZ} $\Phi_{\Delta_1}\cdot\Phi_{\Delta_p}$ (resp. $\Phi_{\Delta_3}\cdot\Phi_{\Delta_p}$) allow the dimension $\Delta_2$ (resp. $\Delta_4$). Namely \begin{equation} \label{residue} S_{rs}=\frac{R_{rs}(\alpha_1,\alpha_2)R_{rs}(\alpha_4,\alpha_3)}{B_{rs}}, \end{equation} where one has\cite{Zam_rec1, Zam_rec2} \begin{equation} R_{rs}(\alpha_1,\alpha_2)=\prod_{\substack{p=-r+1,-r+3,\dots,r-1\\ q=-s+1,-s+3,\dots,s-1}} \left(-\alpha_1-\alpha_2+\alpha_{p-1,q-1}\right) \left(-\alpha_1+\alpha_2+\alpha_{p+1,q+1}\right) \label{deg_con} \end{equation} and, after defining $\lambda_{rs}=-r\beta+s\beta^{-1}$, \cite{Zam_rec2} \begin{equation} B_{rs}=-2\prod_{k=1-r}^{r}\prod_{l=1-s}^s \lambda_{kl},\quad\{k,l\}\not=\{0,0\},\{r,s\}. \label{deg_norm} \end{equation} The $B_{rs}$ \cite{Zamhem, Yan} are related to the vanishing norm of the null vector $|\chi_{rs}\rangle$ in the following way \begin{equation} \lim_{\Delta\rightarrow\Delta_{r,s}}\langle\chi_{rs}|\chi_{rs}\rangle=B_{rs}(\Delta-\Delta_{r,s})+O\bigl((\Delta-\Delta_{r,s})^2\bigr). \end{equation} Finally, the null field $\chi_{r,s}(x)$ associated to the null vector $|\chi_{r,s}\rangle$ is said to decouple from the theory when all of its correlation functions vanish. Such condition requires in particular \begin{equation} \label{null_vector_cond} \lim_{x\rightarrow\infty}x^{2rs}\frac{\langle\chi_{r,s}(x)\phi_{\alpha_1}(1)\phi_{\alpha_2}(0)\rangle}{\langle\phi_{r,s}(x)\phi_{\alpha_1}(1)\phi_{\alpha_2}(0)\rangle}=R_{rs}(\alpha_1,\alpha_2)=0, \end{equation} i.e. $R_{r,s}(\alpha_1,\alpha_2)=0$. \subsection{Structure constants and Coulomb gas representation} As we discussed in the previous section, the structure constants \eqref{3point} are not fixed by conformal invariance and additional dynamical constraints are needed. The request of associativity of the operator product expansion (OPE) was shown in \cite{BPZ} to produce a set of functional equations for the structure constants $C(\Delta_1,\Delta_2,\Delta_3)$, the so-called conformal bootstrap, whose solution would have completely solved the field theory. Under the assumptions of a non-degenerate spectrum and of the normalization choice \eqref{2point}, \cite{Za} solved the conformal bootstrap, proving the existence of a unique solution for $\beta^2\not \in\mathbb Q$. Such solution has the following form \begin{equation} \label{Zam_structure} C(\alpha_1,\alpha_2,\alpha_3)=A_{\beta}\frac{\Upsilon_{\beta}(\beta-\alpha_{13}^2)\Upsilon_{\beta}(\beta-\alpha_{23}^1)\Upsilon_{\beta}(\beta-\alpha_{12}^3)\Upsilon_{\beta}(2\beta-\beta^{-1}-\alpha_{123})} {\Bigl[\prod_{i=1}^3\Upsilon_{\beta}(\beta-2\alpha_i)\Upsilon_{\beta}(2\beta-\beta^{-1}-2\alpha_i)\Bigr]^{1/2}}, \end{equation} where we chose to write the structure constants as a function of the charges $\alpha_i$ ($i=1,2,3$) related to the conformal dimensions by \eqref{dimension} for the value of the central charge \eqref{central} and $\alpha_{ijk}=\alpha_i+\alpha_j+\alpha_k$, $\alpha_{ij}^k=\alpha_i+\alpha_j-\alpha_k$. The special function $\Upsilon_{\beta}(x)$, first introduce in \cite{ZaZa}, is briefly considered in the Appendix B; here we remind that it has integral representation \begin{equation} \label{Zam_int} \log\Upsilon_{\beta}(x)=\int_{0}^{\infty}\frac{dt}{t}\left[\frac{(Q/2-x)^2}{e^t}-\frac{\sinh^2\frac{t}{2}(Q/2-x)}{\sinh\frac{\beta t}{2} \sinh\frac{t}{2\beta}}\right], \end{equation} convergent for $0<x<Q$ ($Q=\beta+\beta^{-1}$) and that satisfies the shift relations \begin{equation} \label{rec_Zam} \frac{\Upsilon_{\beta}(x+\beta)}{\Upsilon_{\beta}(x)}=\gamma(\beta x)\beta^{1-2\beta x} \quad \frac{\Upsilon_{\beta}(x+\beta^{-1})}{\Upsilon_{\beta}(x)}=\gamma(\beta^{-1} x)\beta^{-1+2\beta^{-1} x}, \end{equation} with $\gamma(x)=\Gamma(x)/\Gamma(1-x)$; notice $\gamma(x)\gamma(-x)=-x^{-2}$ and $\gamma(x+1)=-x^2\gamma(x)$. The constant $A_{\beta}$ is given by \begin{equation} \label{A_const} A_{\beta}=\frac{\beta^{\beta^{-2}-\beta^2-1}\gamma(\beta^2)\gamma(\beta^{-2}-1)}{\Upsilon_{\beta}(\beta)}. \end{equation} In \cite{DPSV}, it has then been shown that the solution \eqref{Zam_structure} can actually be recovered from analytic continuation of the three-point correlation functions of scalar vertex operator in the CG formalism \cite{DoFa}. Also for $\beta^2$ rational, \eqref{Zam_structure} has intriguing physical applications in statistical mechanics \cite{DPSV} which are well beyond the predictions of the minimal conformal models. Let us briefly outline the CG approach to CFT \cite{DoFa}. Coulomb gas is a free boson theory with an additional background charge $2\alpha_0$ at infinity which produces the total central charge \eqref{central}. Primary fields are vertex operator $V_{\alpha}=e^{i\alpha\varphi(z,\bar z)}$ with conformal dimension \eqref{dimension}, $\varphi(z,\bar z)$ is the free bosonic field. The double degeneracy of the scaling dimensions \eqref{dimension} is solved \cite{Doco} by assuming that the vertex operators $V_{\alpha}$ and $V_{2\alpha_0-\alpha}$ represent the same primary field $\Phi_{\Delta(\alpha)}\equiv \Phi_{\alpha}=\Phi_{2\alpha_0-\alpha}$ but acquire non trivial normalizations. Correlation functions in the CG approach are non-zero if the charges of the vertex operators $\alpha_i$ sum to the background charge $2\alpha_0$. If such condition cannot be satisfied one is free to add vertex operators in the correlation function of charges $\beta$ and $-\frac{1}{\beta}$ and integrate over them. This peculiar vertex operators (screening operators) do not spoil conformal invariance and do not modify the value of the central charge. The correlation function of four scalar primary fields $\Phi_{\alpha_i}$, $i=1,\dots,4$: \begin{equation} \langle \Phi_{\alpha_4}(\infty)\Phi_{\alpha_3}(1) \Phi_{\alpha_2}(x,\bar{x})\Phi_{\alpha_1}(0)\rangle, \end{equation} which satisfy the neutrality condition \begin{equation} \sum_{i=1}^4 \alpha_i +n \beta-m \beta^{-1}= 2\alpha_0 \quad n,m=0,1,2\dots \label{ch_neu} \end{equation} coincides, up to a normalization constant, to the following integral \begin{align} \label{4vertex} I_{nm}=|1-x|^{4\alpha_2\alpha_3}|x|^{4\alpha_1\alpha_2}&\int\prod_{i=1}^nd\vec{w}_i~\prod_{i=1}^n|w_i-1|^{4\beta\alpha_3}|w_i-x|^{4\beta\alpha_2}|w_i|^{4\beta\alpha_1} \prod_{l<k}|w_l-w_k|^{4\beta^2}\nonumber\\ &\hspace*{-3.5cm}\times\int\prod_{j=1}^m d\vec{u}_j~\prod_{j=1}^m|u_j-1|^{-4\alpha_3/\beta}|u_j-x|^{-4\alpha_2/\beta}|u_j|^{-4\alpha_1/\beta} \prod_{l<k}|u_l-u_k|^{4/\beta^2} \prod_{i,j}|w_i-u_j|^{-4}. \end{align} A very remarkable property of the integral \eqref{4vertex} is that it can be decomposed \cite{DoFa,DoFa2,Doco} into the sum of $(n+1)(m+1)$ terms \begin{equation} \label{exp_int} I_{nm}= \sum_{i=0}^{n}\sum_{j=0}^{m} |x|^{-2(\Delta_1+\Delta_2-\hat{\Delta}_{i,j})} X_{i,j} |F(x|c(\beta),\hat{\Delta}_{i,j},\{\Delta\})|^2, \end{equation} where the constants $X_{i,j}$ are proportional to the structure constants \begin{equation} \label{struc_X} \frac{X_{i,j}}{X_{i',j'}}=\frac{C(\alpha_1,\alpha_2,\hat{\alpha}_{i,j})C(\alpha_4,\alpha_3,\hat{\alpha}_{i,j})} {C(\alpha_1,\alpha_2,\hat{\alpha}_{i',j'})C(\alpha_4,\alpha_3,\hat{\alpha}_{i',j'})} \end{equation} and the function $F(x|c(\beta),\hat{\Delta}_{i,j},\{\Delta\})$ is the conformal block defined in \eqref{conformal_block}. For a correlation function that is represented by the CG integral (\ref{exp_int}) the summation (\ref{4point}) is restricted to the discrete set of internal channels with the following dimensions and charges \begin{equation} \hat{\Delta}_{i,j}=\hat{\alpha}_{i,j}(\hat{\alpha}_{i,j}-2\alpha_0), \quad \hat{\alpha}_{i,j}=\alpha_1+\alpha_2+i\beta-j\beta^{-1}. \end{equation} We observe that the radius of converge of the series expansions (\ref{conformal_block}) is one. Conformal blocks can be analytically continued in the region $|x|>1$ implementing the condition of associativity of the OPE \cite{BPZ}. In our study of logarithmic singularities of the integrals \eqref{4vertex} we will however restrict to the domain $|x|<1$, even if a characterization of such singularities in the limit $x\rightarrow 1$, could be carried out analogously exchanging the role of $\alpha_1$ and $\alpha_3$. \section{Logarithmic singularities in one-screening Coulomb gas integrals} In this section we will discuss the appearance of logarithmic singularities in the integral \eqref{4vertex}, focusing on the simplest example: the case where the neutrality condition \eqref{ch_neu} is satisfied with $n=1$ and $m=0$; the case $n=0$ and $m=1$ follows through the replacement $\beta\rightarrow-1/\beta$. For simplicity we will take $\beta^2\not\in\mathbb Q$ and $\alpha_4$ fixed by \eqref{ch_neu} as $\alpha_4=-(\beta+\alpha_1+\alpha_2+\alpha_3)+2\alpha_0$. We will also assume $\alpha_1+\alpha_2=\alpha_{r,s}$, with $r,s>0$ ensuring that the internal channel with charge $\hat{\alpha}_{00}$ is a degenerate field. For $n=1$ and $m=0$, the integral \eqref{4vertex} has the following form \cite{Doco} \begin{equation} \label{onescreanig} I_{10}=|x|^{-2(\Delta_1+\Delta_2-\hat{\Delta}_{0,0})} \Bigl[X_{0,0}~|F(x|c,\hat{\Delta}_{0,0},\{\Delta\})|^2+X_{1,0}~|x|^{2(\hat{\Delta}_{1,0}-\hat{\Delta}_{0,0})} |F(x|c,\hat{\Delta}_{1,0},\{\Delta\})|^2\Bigr], \end{equation} where $\hat{\alpha}_{0,0}=\alpha_{r,s}$, $\hat{\alpha}_{1,0}=\alpha_{r,s}+\beta$. The corresponding conformal blocks are expressed through hypergeometric functions as \begin{align} \label{F00} &F(x|c,\hat{\Delta}_{0,0},\{\Delta\})=(1-x)^{-2\alpha_2\alpha_3}{}_2F_1\bigl(-2\beta\alpha_2, -1-2\beta\alpha_{123}, -2\beta(\alpha_1+\alpha_2),x\bigr),\\ \label{F10} &F(x|c,\hat{\Delta}_{1,0},\{\Delta\})=(1-x)^{-2\alpha_2\alpha_3}{}_2F_1\bigl(-2\beta\alpha_3, 1+2\beta\alpha_1, 2+2\beta(\alpha_1+\alpha_2),x\bigr); \end{align} after defining $s(x)=\sin(\pi x)$, the coefficients $X$ read \begin{align} \label{X00} &X_{0,0}=\frac{s(2\beta\alpha_3)s(2\beta\alpha_{123})}{s(2\beta(\alpha_1+\alpha_2))}\left[\frac{\Gamma(1+2\beta\alpha_3) \Gamma(-1-2\beta\alpha_{123})}{\Gamma(-2\beta(\alpha_1+\alpha_2))}\right]^2,\\ \label{X10} &X_{1,0}=\frac{s(2\beta\alpha_1)s(2\beta\alpha_{2})}{s(2\beta(\alpha_1+\alpha_2))}\left[\frac{\Gamma(1+2\beta\alpha_1) \Gamma(1+2\beta\alpha_{2})}{\Gamma(2+2\beta(\alpha_1+\alpha_2))}\right]^2. \end{align} Inside the circle $|x|<1$, the hypergeometric function ${}_2F_1(\alpha,\beta,\gamma,x)$ is represented by the power series \begin{equation} \label{power_series} {}_2F_1(\alpha,\beta,\gamma,x)=\sum_{k=0}^{\infty}\frac{[\alpha]_k[\beta]_k}{[\gamma]_k}\frac{x^k}{k!}, \end{equation} where $[q]_k=\Gamma(q+k)/\Gamma(q)$ is the Pochhammer symbol, see for example \cite{Lebedev}. As a function of the variable $\gamma$, \eqref{power_series} has simple poles for $\gamma=(1-s)$ with $s$ a positive integer. The residue is proportional to the hypergeometric function ${}_2F_1(\alpha+s,\beta+s,s+1,x)$ \begin{equation} \label{hyper_exp} \text{Res}_{\gamma=1-s}~{}_2F_1(\alpha,\beta,\gamma,x)= \frac{(-1)^{s-1}[\alpha]_s[\beta]_s}{\Gamma(s)\Gamma(1+s)}x^{s}~{}_2F_1(\alpha+s,\beta+s,s+1,x). \end{equation} If the sum $\alpha_1+\alpha_2=\alpha_{1,s}=\frac{s-1}{2\beta}$ the conformal block $F(x|c,\hat{\Delta}_{00},\{\Delta\})$ in \eqref{F00} is then singular\footnote{If $\alpha_1+\alpha_2+\beta=\alpha_{-1,-s}$, the conformal block $F(x|c,\hat{\Delta}_{1,0},\{\Delta\})$ is singular and $\hat{\Delta}_{1,0}=\Delta_{1,s}$ now.} with residue on the pole at $\hat\Delta_{0,0}=\Delta_{1,s}$ proportional to the conformal block $F(x|c,\hat{\Delta}_{10},\{\Delta\})$, see eq. \eqref{F10}. The singularity of the conformal block $F(x|\hat{\Delta}_{00},\{\Delta\})$ and its residue could have been computed directly form the recursive formula \eqref{rec_conf}. Indeed when $\alpha_1+\alpha_2=\alpha_{1,s}$ and, by the neutrality condition, $\alpha_3+\alpha_4=\alpha_{1,-s}$, eq. \eqref{residue} predicts \begin{equation} \label{residue_one_s} S_{1s}=\frac{(-1)^{s-1}\lambda_{1,s}[-2\beta\alpha_2]_s[-s-2\beta\alpha_3]_s}{2\beta~\Gamma(s)\Gamma(1+s)}, \end{equation} which is non-vanishing\footnote{The eq. \eqref{residue_one_s} shows that the analysis we carried out in this section extends to the case $\alpha_2\not=\alpha_{1,j}$ and $\alpha_3\not=\alpha_{1,j-s}$, $j=1,\dots,s$.}. Expanding the sum of the charges $\alpha_1$ and $\alpha_2$ near the singular value $\alpha_{1,s}$ as \begin{equation} \label{reg} \alpha_1+\alpha_2=\alpha_{1,s}+\delta, \end{equation} and observing that $\Delta-\Delta_{r,s}=\lambda_{r,s}(\alpha-\alpha_{r,s})+O\bigl((\alpha-\alpha_{r,s}\bigr)^2)$, one concludes that \eqref{hyper_exp} is consistent with \eqref{rec_conf}. This singularity is associated to the presence of a null vector at level $s$ in the Verma module of the degenerate primary $\phi_{\hat{\Delta}_{0,0}}=\phi_{1,s}$ which cannot be decoupled since $S_{1,s}\not=0$. Following a general scheme in LCFT, the divergence is cured by the mixing of the null-vector in the Verma module $\mathcal{V}_{\hat\Delta_{00}}$ with the highest weight of the Verma module $\mathcal{V}_{\hat{\Delta}_{1,0}}$ corresponding to the other internal channel in \eqref{onescreanig}. When $\alpha_{1}+\alpha_2=\alpha_{1,s}$, the two states have indeed the same conformal dimension $\Delta_{1,s}+s=\Delta_{-1,s}$. A schematic picture is presented in Fig. \ref{log_pic}.\\ In the remaining part of the section, we point out that the existence of logarithmic singularities in the four-point function requires moreover a precise relation between the values of the structure constants \eqref{Zam_structure} and the residue \eqref{residue}. This mechanism has been outlined for some special cases, see for instance \cite{HaYa,Ya,CrRi}. \begin{figure}[t] \begin{center} \begin{tikzpicture} \draw (-1,1)node[left]{$\alpha_1$}-- (1,-1); \draw (-1,-1)node[left]{$\alpha_2$}-- (1,1); \draw (1,1)node[above]{$\hspace*{3.25cm}\hat{\alpha}_{0,0}=\alpha_1+\alpha_2$}--(4.5,1); \draw (1,-1)node[above]{$\hspace*{3.25cm}\hat{\alpha}_{1,0}=\alpha_1+\alpha_2+\beta$}--(4.5,-1); \draw(4.5,-1)--(6.5,1)node[right]{$\alpha_3$}; \draw(4.5,1)--(6.5,-1)node[right]{$\alpha_4$}; \end{tikzpicture} \end{center} \caption{Diagram representing the integral $I_{10}$ in \eqref{onescreanig}. The internal charges are $\hat{\alpha}_{0,0}=\alpha_{r,s}$ and $\hat{\alpha}_{1,0}=\alpha_{r-2,s}$. For $r=1$, the corresponding conformal dimensions \eqref{dimension} differ by a positive integer $s$ and the conformal block $F(x|c,\hat{\Delta}_{0,0},\{\Delta\})$ is singular. The singularity is due to the null vector at level $s$ which does not decouple and is cured by the contribution of the conformal block $F(x|c,\hat{\Delta}_{1,0},\{\Delta\})$. The integral $I_{1,0}$ has in this case a logarithmic singularity as $x\rightarrow 0$.} \label{log_pic} \end{figure} In order to show this we regularize the CG integral by introducing a parameter $\delta$ as in \eqref{reg} and to ensure that the neutrality condition \eqref{ch_neu} will be always satisfied we also take \begin{equation} \alpha_3+\alpha_4=\alpha_{1,-s}-\delta. \label{reg2} \end{equation} In the limiting procedure the charges $\alpha_1$ and $\alpha_4$ will depend on the $\delta$, through the (\ref{reg}, \ref{reg2}) whereas the charges $\alpha_2$ and $\alpha_3$ are fixed; we will take $\delta\to 0$ at the end. Note that the regularization prescriptions (\ref{reg}, \ref{reg2}) do not modify the central charge, i.e. $\beta$ is independent from $\delta$. The conformal blocks (\ref{F00}, \ref{F10}) and the coefficients (\ref{X00}, \ref{X10}) have the series expansion in $\delta$ \begin{eqnarray} F(x|c,\hat{\Delta}_{0,0},\{\Delta\})&=& \frac{1}{\delta} G^{(-1)}_{00}(x)+ G^{(0)}_{00}(x) +O(\delta), \\ F(x|c,\hat{\Delta}_{0,0},\{\Delta\})&=& G^{(0)}_{10}(x)+\delta\; G^{(1)}_{10}(x)+O(\delta^2), \label{exp_cb}\\ X_{0,0}=X^{(1)}_{00} \delta + X_{0,0}^{(2)} \delta^2+O(\delta^3),& & X_{1,0}=\frac{1}{\delta}X^{(-1)}_{1,0}+ X_{1,0}^{(0)} +O(\delta), \end{eqnarray} explicitly computed in Appendix A. The leading singularity of order $\delta^{-1}$ in the correlation functions $I_{10}$ disappears due to the identity $X^{(1)}_{0,0}|G^{(-1)}_{0,0}(x)|^2+X_{1,0}^{(-1)}|x|^{2s}|G^{0}_{(1,0)}(x)|^2=0$, which can be recast in the more expressive form \begin{equation} \label{rel_log} \frac{C(\alpha_1,\alpha_2,\alpha_{1,s}+\delta)C(\alpha_3,\alpha_4,\alpha_{1,s}+\delta)} {C(\alpha_1,\alpha_2,\alpha_{-1,s}+\delta)C(\alpha_3,\alpha_4,\alpha_{-1,s}+\delta)}= -\delta^2\left[\frac{\lambda_{1,s}}{S_{1,s}}\right]^2+O(\delta^3). \end{equation} Notice that due to the form of the structure constants \eqref{Zam_structure}, eq. (\ref{rel_log}) can be simplified into \begin{equation} \label{simp_rel_log} \frac{C(\alpha_1,\alpha_2,\alpha_{1,s}+\delta)}{C(\alpha_1,\alpha_2,\alpha_{-1,s}+\delta)}=i\delta\frac{\lambda_{1,s}B_{1,s}}{ R^2_{1,s}(\alpha_1,\alpha_2)}+O(\delta^2), \end{equation} valid for any value of the $\alpha_i$; the formula \eqref{simp_rel_log} can be also verified by using the shift relation \eqref{rec_Zam} and its generalization in the multi-screening case will be the object of the next section. It is interesting to observe that the vanishing in the $\delta\rightarrow 0$ limit of the ratio of the structure constants in \eqref{simp_rel_log} is due, see Appendix B, to zeros in the function $\Upsilon_{\beta}(x)$ at $x=\beta-2\alpha_{-1,s}$ and $x=2\beta-\beta^{-1}-2\alpha_{-1,s}$. Such zeros are responsible \cite{DPSV} for the singular behavior of the normalization factors of the field $\phi_{-1,s}$ in the CG representation. As it is well known, the cancellation of the leading $\delta^{-1}$ singularity in the integral \eqref{onescreanig} leads to a logarithmic divergence and one has \begin{align} I_{1,0}=|x|^{-2(\Delta_1+\Delta_2-\hat{\Delta}_{00})}&\left[a\log|x|^2|x|^{2s} X_{1,0}^{-1}|G_{10}^{(0)}(x)|^2+ X_{0,0}^{(1)}|G_{00}^{(-1)}(x)|^2+X_{1,0}^{(0)}|x|^{2s}|G_{10}^{(0)}(x)|^2\right.\nonumber\\ &\hspace*{-1cm}\Bigl.+X_{0,0}^{(1)}\bigr(G_{00}^{(-1)}(x)G_{00}^{(0)}(\bar{x})+c.c.\bigl)+ X_{1,0}^{(-1)}|x|^{2s}\bigl(G_{10}^{(0)}(x)G_{10}^{(1)}(\bar{x})+c.c.\bigr)\Bigr], \label{cg_1s} \end{align} where the logarithmic prefactor is \begin{equation} a=\lim_{\delta\to 0}\left(s-\hat{\Delta}_{1,0}(\delta)+\hat{\Delta}_{0,0}(\delta)\right)=-2\beta. \end{equation} Let us mention that some examples of bulk correlation functions of the kind (\ref{cg_1s}) appeared in the study of classical statistical systems at criticality. For instance, the function (\ref{cg_1s}) with $\alpha_{1}=2\alpha_{0}-\alpha_4=\alpha_{1,0} \quad \alpha_2=\alpha_{3}=\alpha_{1,2}$ and $1/2\leq \beta^2 \leq 1$, was argued to be relevant in the study of interfaces in the random bond Potts model \cite{JDPSW}. It determines, in particular, the effects of quenched bond disorder on the fractal dimension of FK clusters boundaries. These functions are, at our knowledge, the only bulk logarithmic correlation functions to have found a numerical test\cite{JDPSW}. Similarly, the bulk correlation function (\ref{cg_1s}) with $\alpha_{1}=2\alpha_0-\alpha_4=\alpha_{2,0}$, determines the disorder effects on the fractal dimensions of the pivotal bonds and has found numerical verification \cite{JDPSW_un}. Finally, the correlation function (\ref{cg_1s}) with $\alpha_{1}=2\alpha_{0}-\alpha_4=\alpha_{0,1} \quad \alpha_2=\alpha_{3}=\alpha_{2,1}$ has been argued in \cite{CaSi} to describe the propagation of a loop in the dilute phase of the $O(n)$ model in the presence of two twist operators. \section{Logarithmic singularities in multi-screening Coulomb gas integrals} We now extend the results obtained in previous section to the case where four-point correlation functions of vertex operators are computed from \eqref{4vertex} with an arbitrary number $n$ and $m$ of screening operators. We assume $\beta^2\not\in \mathbb Q$ and defer the discussion of the peculiarity of $\beta^2$ rational to the next section. Analogously to the one-screening case we consider \begin{equation} \alpha_1+\alpha_2=\alpha_{r,s} \quad \alpha_3+\alpha_4=\alpha_{2n-r,2m-s} \quad r,s\in\mathbb N. \label{sum_12} \end{equation} Under this assumption the internal charges take the form $\hat{\alpha}_{ij}=\alpha_{r-2i,s-2j}$. The condition (\ref{sum_12}) satisfies the neutrality condition (\ref{ch_neu}) and assures us that the integral (\ref{exp_int}) contains conformal block of degenerate fields. Indeed,among the $(n+1)(m+1)$ internal charges $\hat{\alpha}_{ij}$, $i=0,..,n$ and $j=0,..,m$, the ones with $(r-2i\geq 1 \wedge s-2j \geq 1)$ or $(r-2i\leq -1 \wedge s-2j \leq -1)$ belongs to the lattice \eqref{kac_lattice}. It is also useful to observe that the channels $\hat{\alpha}_{ij}$ and $\hat{\alpha}_{r-i,s-j}$ correspond to the same conformal block and to the same structure constants \begin{equation} F(x|c,\hat{\Delta}_{ij},\{\Delta\})=F(x|c,\hat{\Delta}_{r-i,s-j},\{\Delta\});\quad X_{i,j}=X_{r-i,s-j}, \label{tr_sym} \end{equation} due to the identity $\hat{\Delta}_{i,j}=\hat{\Delta}_{r-i,s-j}$. \subsection{Singularities of the internal channels} Logarithmic terms in the four-point correlation function may arise when some of the conformal blocks are singular, i.e. when the dimension $\hat{\Delta}_{i,j}$ in \eqref{exp_int} corresponds to that of a degenerate field and the condition \begin{equation} R_{r-2i,s-2j}(\alpha_{r,s}-\alpha_2,\alpha_2)R_{r-2i,s-2j}(\alpha_{2n-r,2m-s}-\alpha_3,\alpha_3) \neq 0, \label{cond_sing2} \end{equation} is satisfied. From (\ref{rec_conf}) it follows that the conformal block has in this case a singularity at the level $(r-2i)(s-2j)$ and, as we have previously noticed, the condition (\ref{cond_sing2}) is equivalent to consider a CFT where the null-vector $|\chi_{r-2i,s-2j}\rangle$ cannot be set to zero. Using (\ref{deg_con}), one can check that the requirement of the degeneracy of the internal channel and the condition (\ref{cond_sing2}) are met by the charges $\hat{\alpha}_{i,j}$ such that \begin{align} &(r-2i\geq 1) \;\wedge\; (s-2j\geq 1)\;\wedge \;(i\geq r-n \vee j\geq s-m),\\ &\hspace*{5cm}\text{or}\nonumber\\ &(r-2i\leq -1) \;\wedge\; (s-2j\leq -1)\;\wedge \;(i<n+1 \vee j< m+1\;). \label{ij_deg_sing} \end{align} Note that we are always assuming $\alpha_2$ and $\alpha_3$ to take general real values. In particular, the condition (\ref{cond_sing2}) also requires \begin{eqnarray} \alpha_{2}&\neq& \alpha_{i+k,j+l}\quad \mbox{and} \nonumber \\ \alpha_3&\neq& \alpha_{n-r+i+k,m-s+j+l}, \quad \mbox{for} \; k=1,\dots,r-2i,\;l=1,\dots,s-2j. \end{eqnarray} Now that we have identified the possible sources of singularities of the conformal blocks, we pass to analyze those of the coefficients $X_{i,j}$ defined in (\ref{struc_X}). Taking into account (\ref{sum_12}) one has \begin{equation} \label{X_ij_dege} X_{i,j}= \mathcal{N}(c,\{\alpha\})~C(\alpha_{r,s}-\alpha_2,\alpha_2,\hat{\alpha}_{i,j})C(\alpha_{2n-r,2m-s}-\alpha_3,\alpha_2,\hat{\alpha}_{i,j}). \end{equation} where $\mathcal{N}$ is a normalization constant which depends only on the external charges and the central charge. For arbitrary real values of $\alpha_2$ and $\alpha_3$, we can single out the factors $X_{i,j}|_{\text{sing}}$ present in $X_{i,j}$ which are zero or diverge \begin{align} X_{i,j}|_{\text{sing}}&= \frac{1} {\underbrace{\Upsilon_{\beta}(\beta-2\alpha_{r-2 i,s-2 j})}_{\equiv A}\underbrace{\Upsilon_{\beta}(2\beta-\beta^{-1}-2\alpha_{r-2 i,s-2 j})}_{\equiv B}}\times \nonumber \\ &\times\underbrace{\Upsilon_{\beta}(\beta-\alpha_{r,s}+\alpha_{r-2i,s-2j})\Upsilon_{\beta}(2\beta-\beta^{-1}-\alpha_{r,s}- \alpha_{r-2i,s-2j})}_{\equiv C}\times \nonumber \\ &\times \underbrace{\Upsilon_{\beta}(\beta-\alpha_{2n-r,2m-s}+\alpha_{r-2i,s-2j})\Upsilon_{\beta}(2\beta-\beta^{-1}-\alpha_{2n-r,2m-s}- \alpha_{r-2i,s-2j})}_{\equiv D}. \label{exp_sing} \end{align} From the location of the zeros of the function $\Upsilon_{\beta}(x)$ defined in \eqref{Zam_int}, see also Appendix B, one finds \begin{align} &A=0:~ (r-2 i\geq 0 \;\wedge\; s-2j\leq -1) \quad \text{or} \quad (r-2 i\leq -1\; \wedge\; s-2j\geq 0), \label{infies1} \\ &B=0:~ (r-2 i\geq 1 \;\wedge\; s-2j\leq 0) \quad \text{or}\quad (r-2 i\leq 0 \;\wedge\;s-2j\geq 1), \label{infies2} \\ &C=0:~ (i\leq r \;\wedge\; j> s)\quad \text{or}\quad ( i> r\; \wedge\; j\leq s), \label{zeros1} \\ &D=0:~ (i\geq r-n \;\wedge\; j< s-m)\quad \text{or}\quad (i< r-n \;\wedge\; j\geq s-m). \label{zeros2} \end{align} We need now to regularize the sum (\ref{exp_int}). This can be done by shifting the internal channels by \begin{equation} \hat{\alpha}_{i,j}\to \hat{\alpha}_{i,j}(\delta)=\alpha_{r,s}+i\beta -j \beta^{-1}+\delta. \label{regul} \end{equation} Note that on may use different regularization procedures. We can regularize the sum (\ref{exp_int}) by keeping the neutrality condition satisfied. This is obtained by shifting $\alpha_1,\alpha_2\to\alpha_1+\delta/2,\alpha_2+\delta/2$; $\alpha_3,\alpha_4\to\alpha_3-\delta/2,\alpha_4-\delta/2$, as we have done in the one-screening case. In principle, one can also study the integral (\ref{exp_int}) with the internal charges shifted as in (\ref{regul}) by keeping the external charges fixed: the sum (\ref{exp_int}) remains well defined also in this case even if it does not correspond to a CG integral, which is only recovered in the limit $\delta\to 0$. If $i,j$ satisfies the condition (\ref{ij_deg_sing}), the corresponding term $|F(x|c,\hat{\Delta}_{ij},\{\Delta\})|^2$ has a singularity $ O(\delta^{-2})$. The coefficient $X_{i,j}$ is $O(\delta)$ (resp. $O(\delta^{-1})$) anytime one of the conditions (\ref{zeros1}) and (\ref{zeros2}) (resp. (\ref{infies1}) and (\ref{infies2})) is satisfied. We find useful to consider in the following, the points $(x,y)$ with integer coordinates, where $(x,y)$ label the charges $\alpha_{x,y}=1/2(1-x)\beta+1/2(y-1)\beta^{-1}$. We can then visualize the set of channels entering in (\ref{exp_int}) as a grid of points of size $2n \times 2m$ whose upper right corner has coordinates $(r,s)$. The symmetry (\ref{tr_sym}) relates the pair of points symmetric with respect to the origin. Points in the lattice representing the internal channels with charge $\hat{\alpha}_{ij}$ will have coordinates \begin{equation} (r-2i,s-2j)\equiv[i,j], \end{equation} and to each of them we associate the order in the expansion in powers of $\delta$ of the corresponding coefficient $X_{i,j}$ as well as of the conformal block $|F(x|\hat{\Delta}_{ij},\{\Delta\})|^2$. Let us consider first the behavior of the conformal blocks. The points $[i,j]$ with positive coordinates, i.e. $r-2 i\geq 1$ and $s-2j\geq 1$ (belonging to the first quadrant of the lattice), and the ones with negative coordinates, $r-2 i\leq -1$ and $s-2j\leq -1$ (in the third quadrant), represent degenerate operators. If the point $[i,j]$ and at least one of the two points $[r-i,j]$, $[i,s-j]$, obtained by reflection with respect to the horizontal and vertical axis, belong to the grid of internal channels, then the pair $i,j$ satisfies the condition (\ref{ij_deg_sing}) and is associated to a singular conformal block. On the other hand, given a point $[i,j]$, for what concerns the behavior of the coefficients $X_{i,j}$ we should distinguish three cases. \begin{itemize} \item[\textit{(i)}] The point $[r-i,s-j]$ symmetric with respect to the origin belongs to the grid. In this case, the (\ref{zeros1})-(\ref{zeros2}) are not satisfied. If $[i,j]$ or $[r-i,s-j]$ belong to the first or third quadrant, $X_{i,j}=O(\delta^{0})$ otherwise $X_{i,j}=O(\delta^{-2})$ (we remind the identification (\ref{tr_sym})). For those points sitting on the two axis but excluding the origin, i.e. $2i=r$ and $2j\neq s$ or $2i\neq r$ and $2j= s$ , $X_{i,j}= O(\delta^{-1})$ while for the origin, we have $X_{r/2,s/2}=O(\delta^0)$; \item[\textit{(ii)}] One among the axis-symmetric points $[r-i,j]$ or $[r,s-j]$ belong to the grid. This implies that one of the two conditions (\ref{zeros1}) and (\ref{zeros2}) is verified. If $[i,j]$ belong to the first or third quadrant, $X_{i,j}= O(\delta)$, if they are sitting on one of two axis, $X_{i,j}=O(\delta^0)$, otherwise $X_{i,j}=O(\delta^{-1})$; \item [\textit{(iii)}] The two-axis symmetric points $[r-i,j]$ or $[r,s-j]$ do not belong to the grid: in this case $X_{i,j}=O(\delta^0)$. \end{itemize} The terms in (\ref{exp_int}) can have amplitudes of different order in $\delta$, in particular singularities of order $\delta^{-2}$ or $\delta^{-1}$. The important observation is that, in any case, to any singular term $X_{i,j}F(x|\hat{\Delta}_{i,j},\{\Delta\})|^2$ of $O(\delta^{-2})$ or $O( \delta^{-1})$, corresponds an other term $X_{r-i,j}|F(x|c,\hat{\Delta}_{r-i,j},x)|^2$ or $X_{i,s-j}F(x|x,\hat{\Delta}_{i,s-j},\{\Delta\})|^2$ of the same order. This pair of singular channels is represented by points which can (if $i=r/2$ or $s=j/2$) or not belong to one of the two axis. In this last case, the singular vector $|\chi_{r-2i,s-2j}\rangle$ of the degenerate primary $\phi_{r-2i,s-2j}$ enters in collision with the highest weight $|\phi_{2i-r,s-2j}\rangle$ (or $|\phi_{r-2i,2j-s}\rangle$) while in the first case one has a collision between the two primaries with the same dimension $\Phi_{0,s-2j}$ and $\Phi_{0,2j-s}$ (or $\Phi_{r-2i,0}$ and $\Phi_{2i -r,0}$). In both situations, the leading order singularity in the integral cancels and the subleading order in the power series in $\delta$ contains logarithms. Before showing such cancellation explicitly, let us discuss some examples. \begin{itemize} \item $r>2n$ and $s>2 m$ All the points $[i,j]$ are in the first quadrant. Since (\ref{ij_deg_sing}) cannot be satisfied, the ratio of the all coefficients $X_{i,j}$ is finite and there are not singular conformal blocks. In the geometrical interpretation we proposed, the grid does not contain points obtained by reflection with respect to the horizontal or vertical axis. In Fig. \ref{grid_ch} we show as an example the case $n=2,m=3$ and $r=5,s=7$. \begin{figure}[t] \begin{tikzpicture}[scale=0.4] \draw (-2,10) node[right]{$X_{i,j}$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,7) node[below]{{\tiny $[0,0]$}}; \draw (5,7) node[above]{{\tiny $\delta^0$}}; \draw (5,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,5) node[below]{{\tiny $[0,1]$}}; \draw (5,5) node[above]{{\tiny $\delta^0$}}; \draw (5,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,3) node[below]{{\tiny $[0,2]$}}; \draw (5,3) node[above]{{\tiny $\delta^0$}}; \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[below]{{\tiny $[0,3]$}}; \draw (5,1) node[above]{{\tiny $\delta^0$}}; \draw (3,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,7) node[below]{{\tiny $[1,0]$}}; \draw (3,7) node[above]{{\tiny $\delta^0$}}; \draw (3,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,5) node[below]{{\tiny $[1,1]$}}; \draw (3,5) node[above]{{\tiny $\delta^0$}}; \draw (3,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,3) node[below]{{\tiny $[1,2]$}}; \draw (3,3) node[above]{{\tiny $\delta^0$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[below]{{\tiny $[1,3]$}}; \draw (3,1) node[above]{{\tiny $\delta^0$}}; \draw (1,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,7) node[below]{{\tiny $[2,0]$}}; \draw (1,7) node[above]{{\tiny $\delta^0$}}; \draw (1,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,5) node[below]{{\tiny $[2,1]$}}; \draw (1,5) node[above]{{\tiny $\delta^0$}}; \draw (1,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,3) node[below]{{\tiny $[2,2]$}}; \draw (1,3) node[above]{{\tiny $\delta^0$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[below]{{\tiny $[2,3]$}}; \draw (1,1) node[above]{{\tiny $\delta^0$}}; \end{tikzpicture} \begin{tikzpicture}[scale=0.4] \draw (-5,10) node[right]{$|F(x|c,\hat{\Delta}_{ij},\{\Delta\})|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,7) node[below]{{\tiny $[0,0]$}}; \draw (5,7) node[above]{{\tiny $\delta^0$}}; \draw (5,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,5) node[below]{{\tiny $[0,1]$}}; \draw (5,5) node[above]{{\tiny $\delta^0$}}; \draw (5,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,3) node[below]{{\tiny $[0,2]$}}; \draw (5,3) node[above]{{\tiny $\delta^0$}}; \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[below]{{\tiny $[0,3]$}}; \draw (5,1) node[above]{{\tiny $\delta^0$}}; \draw (3,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,7) node[below]{{\tiny $[1,0]$}}; \draw (3,7) node[above]{{\tiny $\delta^0$}}; \draw (3,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,5) node[below]{{\tiny $[1,1]$}}; \draw (3,5) node[above]{{\tiny $\delta^0$}}; \draw (3,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,3) node[below]{{\tiny $[1,2]$}}; \draw (3,3) node[above]{{\tiny $\delta^0$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[below]{{\tiny $[1,3]$}}; \draw (3,1) node[above]{{\tiny $\delta^0$}}; \draw (1,7) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,7) node[below]{{\tiny $[2,0]$}}; \draw (1,7) node[above]{{\tiny $\delta^0$}}; \draw (1,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,5) node[below]{{\tiny $[2,1]$}}; \draw (1,5) node[above]{{\tiny $\delta^0$}}; \draw (1,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,3) node[below]{{\tiny $[2,2]$}}; \draw (1,3) node[above]{{\tiny $\delta^0$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[below]{{\tiny $[2,3]$}}; \draw (1,1) node[above]{{\tiny $\delta^0$}}; \end{tikzpicture} \caption{Example of a grid with $r>2n$ and $s>2m$. According to our analysis all the internal channels have finite coefficients $X_{i,j}$ and all the conformal blocks $F(x|c,\hat{\Delta}_{ij},\{\Delta\})$ are non-singular.} \label{grid_ch} \end{figure} \item $r>2n$ and $s\leq2m$ In this case the grid contains points which are obtained by reflection with respect to the $x$ axis. Correspondingly, we expect singularities $O(\delta^{-1})$ coming from the pair $X_{i,j}|F(x|c,\hat{\Delta}_{ij},\{\Delta\})|^2$ and $X_{i,s-j}|F(x|c,\hat{\Delta}_{i,s-j},\{\Delta\})|^2$. As we previously remarked, the sum of these two terms is $O(\delta^0)$ and produces a logarithmic term which adds with the other (non-logarithmic) contributions when considering the $\delta\to 0$ limit. Note that there are no singularities $O(\delta^{-2})$, because there are no points of the grid which are symmetric with respect to the origin. Fig. \ref{grid_ch2} and Fig \ref{grid_ch3} show the examples $n=2,m=3$ with $r=5,s=4$ and $n=2,m=3$ with $r=5,s=2$ \begin{figure}[t] \begin{tikzpicture}[scale=0.4] \draw (-2,10) node[right]{$X_{i,j}$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,4) node[below]{{\tiny $[0,0]$}}; \draw (5,4) node[above]{{\tiny $\delta^0$}}; \draw (5,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,2) node[below]{{\tiny $[0,1]$}}; \draw (5,2) node[above]{{\tiny ${\color{blue} \delta}$}}; \draw (5,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,0) node[below]{{\tiny $[0,2]$}}; \draw (5,0) node[above]{{\tiny $\delta^0$}}; \draw (5,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,-2) node[below]{{\tiny $[0,3]$}}; \draw (5,-2) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (3,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,4) node[below]{{\tiny $[1,0]$}}; \draw (3,4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[1,1]$}}; \draw (3,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[1,2]$}}; \draw (3,0) node[above]{{\tiny $\delta^0$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[1,3]$}}; \draw (3,-2) node[above]{{\tiny ${\color{red}\delta^{-1}}$}}; \draw (1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,4) node[below]{{\tiny $[2,0]$}}; \draw (1,4) node[above]{{\tiny $\delta^0$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[2,1]$}}; \draw (1,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[below]{{\tiny $[2,2]$}}; \draw (1,0) node[above]{{\tiny $\delta^0$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[2,3]$}}; \draw (1,-2) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \end{tikzpicture} \begin{tikzpicture}[scale=0.4] \draw (-5,10) node[right]{$|F(x|c,\hat{\Delta}_{ij},\{\Delta\})|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,4) node[below]{{\tiny $[0,0]$}}; \draw (5,4) node[above]{{\tiny $\delta^0$}}; \draw (5,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,2) node[below]{{\tiny $[0,1]$}}; \draw (5,2) node[above]{{\tiny ${\color{red}\delta^{-2}}$}}; \draw (5,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,0) node[below]{{\tiny $[0,2]$}}; \draw (5,0) node[above]{{\tiny $\delta^0$}}; \draw (5,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,-2) node[below]{{\tiny $[0,3]$}}; \draw (5,-2) node[above]{{\tiny $\delta^0$}}; \draw (3,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,4) node[below]{{\tiny $[1,0]$}}; \draw (3,4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[1,1]$}}; \draw (3,2) node[above]{{\tiny ${\color{red}\delta^{-2}}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[1,2]$}}; \draw (3,0) node[above]{{\tiny $\delta^0$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[1,3]$}}; \draw (3,-2) node[above]{{\tiny $\delta^0$}}; \draw (1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,4) node[below]{{\tiny $[2,0]$}}; \draw (1,4) node[above]{{\tiny $\delta^0$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[2,1]$}}; \draw (1,2) node[above]{{\tiny ${\color{red}\delta^{-2}}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[below]{{\tiny $[2,2]$}}; \draw (1,0) node[above]{{\tiny $\delta^0$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[2,3]$}}; \draw (1,-2) node[above]{{\tiny $\delta^0$}}; \end{tikzpicture} \caption{In this figure we show the correlation function with $r=5,~s=4$ and $n=2,~m=3$. Notice that the degenerate primaries represented by the points $[2,1],~[1,1]$ and $[0,1]$ have a singular conformal block but their coefficient $X_{i,j}$ is $O(\delta)$. On the other hand the reflected points $[2,3],~[1,3]$,and $[0,3]$ corresponds to non-degenerate primaries with a finite conformal block but with a coefficient $X_{i,j}$ which is now $O(\delta^{-1})$. This situation is completely analogous to the one we encountered in the one-screening case of Sect. 3.} \label{grid_ch2} \end{figure} \begin{figure}[t] \begin{tikzpicture}[scale=0.4] \draw (-2,10) node[right]{$X_{i,j}$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,2) node[below]{{\tiny $[0,0]$}}; \draw (5,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (5,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,0) node[below]{{\tiny $[0,1]$}}; \draw (5,0) node[above]{{\tiny $\delta^0$}}; \draw (5,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,-2) node[below]{{\tiny $[0,2]$}}; \draw (5,-2) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (5,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {};bookos \draw (5,-4) node[below]{{\tiny $[0,3]$}}; \draw (5,-4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[1,0]$}}; \draw (3,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[1,1]$}}; \draw (3,0) node[above]{{\tiny $\delta^0$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[1,2]$}}; \draw (3,-2) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (3,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-4) node[below]{{\tiny $[1,3]$}}; \draw (3,-4) node[above]{{\tiny $\delta^0$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[2,0]$}}; \draw (1,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {};bookos \draw (1,0) node[below]{{\tiny $[2,1]$}}; \draw (1,0) node[above]{{\tiny $\delta^0$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[2,2]$}}; \draw (1,-2) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (1,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-4) node[below]{{\tiny $[2,3]$}}; \draw (1,-4) node[above]{{\tiny $\delta^{0}$}}; \end{tikzpicture} \begin{tikzpicture}[scale=0.4] \draw (-5,10) node[right]{$|F(x|c,\hat{\Delta}_{ij},\{\Delta\})|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,2) node[below]{{\tiny $[0,0]$}}; \draw (5,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (5,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,0) node[below]{{\tiny $[0,1]$}}; \draw (5,0) node[above]{{\tiny $\delta^0$}}; \draw (5,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,-2) node[below]{{\tiny $[0,2]$}}; \draw (5,-2) node[above]{{\tiny $\delta^0$}}; \draw (5,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,-4) node[below]{{\tiny $[0,3]$}}; \draw (5,-4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[1,0]$}}; \draw (3,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[1,1]$}}; \draw (3,0) node[above]{{\tiny $\delta^0$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[1,2]$}}; \draw (3,-2) node[above]{{\tiny $\delta^0$}}; \draw (3,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-4) node[below]{{\tiny $[1,3]$}}; \draw (3,-4) node[above]{{\tiny $\delta^0$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[2,0]$}}; \draw (1,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[below]{{\tiny $[2,1]$}}; \draw (1,0) node[above]{{\tiny $\delta^0$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[2,2]$}}; \draw (1,-2) node[above]{{\tiny $\delta^0$}}; \draw (1,-4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-4) node[below]{{\tiny $[2,3]$}}; \draw (1,-4) node[above]{{\tiny $\delta^0$}}; \end{tikzpicture} \caption{Figure representing the case $r=5~,s=2$ and $n=2,~m=3$. Analogous considerations to the ones of Fig. \ref{grid_ch2} apply.} \label{grid_ch3} \end{figure} \item $r \leq 2n$ and $m\leq s \leq 2m$ In this case there are also points of the grid which are symmetric with respect to the origin. These points are associated to contributions $O(\delta^{-2})$. In Fig. \ref{grid_ch4}, we show the example $n=2$, $m=3$ and $r=3$, $s=4$. In this case we expect that the correlation function gets dominant contributions of order $\delta^{-1}$. To get a finite correlation function, one has to choose a vanishing normalization constant $\mathcal{N}(c,\{\Delta\})$ in \eqref{X_ij_dege}, of $O(\delta)$. Note also that in this case there are terms in the (\ref{exp_int}) whose contribution is $0(\delta^0)$ and therefore disappear in the limit $\delta\to 0$. \begin{figure}[t] \begin{tikzpicture}[scale=0.4] \draw (-2,10) node[right]{$X_{i,j}$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (3,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,4) node[below]{{\tiny $[0,0]$}}; \draw (3,4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[0,1]$}}; \draw (3,2) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[0,2]$}}; \draw (3,0) node[above]{{\tiny $\delta^{0}$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[0,3]$}}; \draw (3,-2) node[above]{ {\tiny ${\color{red} \delta^{-1}}$}}; \draw (1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,4) node[below]{{\tiny $[1,0]$}}; \draw (1,4) node[above]{{\tiny ${\color{blue}\delta}$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[1,1]$}}; \draw (1,2) node[above]{{\tiny $\delta^0$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[below]{{\tiny $[1,2]$}}; \draw (1,0) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[1,3]$}}; \draw (1,-2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (-1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,4) node[below]{{\tiny $[2,0]$}}; \draw (-1,4) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (-1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,2) node[below]{{\tiny $[2,1]$}}; \draw (-1,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (-1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,0) node[below]{{\tiny $[2,2]$}}; \draw (-1,0) node[above]{{\tiny ${\color{red} \delta^{-1}}$}}; \draw (-1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-2) node[below]{{\tiny $[2,3]$}}; \draw (-1,-2) node[above]{{\tiny $\delta^{0}$}}; \end{tikzpicture} \begin{tikzpicture}[scale=0.4] \draw (-5,10) node[right]{$|F(x|c,\hat{\Delta}_{ij},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (3,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,4) node[below]{{\tiny $[0,0]$}}; \draw (3,4) node[above]{{\tiny $\delta^0$}}; \draw (3,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,2) node[below]{{\tiny $[0,1]$}}; \draw (3,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[below]{{\tiny $[0,2]$}}; \draw (3,0) node[above]{{\tiny $\delta^0$}}; \draw (3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-2) node[below]{{\tiny $[0,3]$}}; \draw (3,-2) node[above]{{\tiny $\delta^0$}}; \draw (1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,4) node[below]{{\tiny $[1,0]$}}; \draw (1,4) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,2) node[below]{{\tiny $[1,1]$}}; \draw (1,2) node[above]{{\tiny ${\color{red} \delta^{-2}}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[below]{{\tiny $[1,2]$}}; \draw (1,0) node[above]{{\tiny $\delta^0$}}; \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[below]{{\tiny $[1,3]$}}; \draw (1,-2) node[above]{{\tiny $\delta^0$}}; \draw (-1,4) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,4) node[below]{{\tiny $[2,0]$}}; \draw (-1,4) node[above]{{\tiny $\delta^0$}}; \draw (-1,2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,2) node[below]{{\tiny $[2,1]$}}; \draw (-1,2) node[above]{{\tiny $\delta^0$}}; \draw (-1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,0) node[below]{{\tiny $[2,2]$}}; \draw (-1,0) node[above]{{\tiny $\delta^0$}}; \draw (-1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-2) node[below]{{\tiny $[2,3]$}}; \draw (-1,-2) node[above]{{\tiny $\delta^{0}$}}; \end{tikzpicture} \caption{Example of a grid with $r=3,~s=4$ and $n=2,~m=3$. The associated correlation function will be $O(\delta^{-1})$, due to the cancellation of leading $O(\delta^{-2})$ singularity. Such kind of correlation function can be render finite by the choice $\mathcal{N}(c,\{\Delta\})=O(\delta)$.} \label{grid_ch4} \end{figure} \end{itemize} \clearpage \subsection{Origin of logarithmic singularities in the four-point function} We have seen that the correlation function (\ref{exp_int}) has singularities $O(\delta^{-1})$ when only one pair of colliding channels $\hat{\alpha}_{ij}$, $\hat{\alpha}_{r-i,j}$ or $\hat{\alpha}_{i,j}$, $\hat{\alpha}_{i,s-j}$ appears in its decomposition. The singularity is $O(\delta^{-2})$ when all the four channels $\hat{\alpha}_{i,j}$, $\hat{\alpha}_{r-i,j}$, $\hat{\alpha}_{i,s-j}$ and $\hat{\alpha}_{r-i,s-j}$ appear in its decomposition. Here we prove a generalization of \eqref{simp_rel_log} and show that the leading order in $\delta$ to the integral cancels, leaving us with a logarithmic singularity. We consider first the leading order of the ratios $X_{i,j}/X_{r-i,j}$ and $X_{i,j}/X_{i,s-j}$ in the $\delta\to 0$ limit. By using the recursion relation (\ref{rec_Zam}), we obtain for general values of $\alpha_a$ and $\alpha_b$, with $\alpha_{a}^{b}\equiv \alpha_a-\alpha_b$ and $\alpha_{ab}\equiv \alpha_a+\alpha_b$ the following relations \begin{align} &\hspace*{-2cm}\frac{\Upsilon_{\beta}(\beta-\alpha_{a}^b-\alpha_{P,Q})\Upsilon_{\beta}(\beta-\alpha_{b}^a-\alpha_{P,Q}) \Upsilon_{\beta}(\beta-\alpha_{ab}+\alpha_{P,Q})} {\Upsilon_{\beta}(\beta-\alpha_{a}^b-\alpha_{-P,Q})\Upsilon_{\beta}(\beta-\alpha_{b}^a-\alpha_{-P,Q}) \Upsilon_{\beta}(\beta-\alpha_{ab}+\alpha_{-P,Q})}\times \nonumber \\ &\hspace*{2.5cm}\times\frac{\Upsilon_{\beta}(2\beta-\beta^{-1}-\alpha_{ab}-\alpha_{P,Q})} {\Upsilon_{\beta}(2\beta-\beta^{-1}-\alpha_{ab}-\alpha_{-P,Q})}= \frac{1}{R_{P,Q}(\alpha_a,\alpha_b)^2}, \end{align} as well as \begin{equation} \sqrt{\frac{\Upsilon_{\beta}(\beta-2\alpha_{-P,Q}-2\delta)\Upsilon_{\beta}(2\beta-\beta^{-1}-2\alpha_{-P,Q}-2\delta)} {\Upsilon_{\beta}(\beta-2\alpha_{P,Q}) \Upsilon_{\beta}(2\beta-\beta^{-1}-2\alpha_{P,Q})}}= i \;\delta \; B_{P,Q}\lambda_{P,Q}+O(\delta^{2}) \end{equation} where $P$ and $Q$ are two positive integers. The ratios of the structure constant $C(\alpha_a,\alpha_b,\alpha_{P,Q})$, $C(\alpha_a,\alpha_b,\alpha_{-P,Q}+\delta)$ and $C(\alpha_a,\alpha_b,\alpha_{P,-Q}+\delta)$ take then the form \begin{eqnarray} \frac{C(\alpha_a,\alpha_b,\alpha_{P,Q}+\delta)} {C(\alpha_a,\alpha_b,\alpha_{-P,Q}+\delta)} &=&i \;\delta\; \frac{B_{P,Q}\lambda_{P,Q}}{R_{P,Q}(\alpha_a,\alpha_b)^2}+O(\delta^2) \nonumber \\ \frac{C(\alpha_a,\alpha_b,\alpha_{P,Q}+\delta)} {C(\alpha_a,\alpha_b,\alpha_{P,-Q}+\delta)} &=&i \;\delta\; \frac{B_{P,Q}\lambda_{P,Q}}{R_{P,Q}(\alpha_a,\alpha_b)^2}+O(\delta^2). \label{logza1} \end{eqnarray} Notice that the leading order in $\delta$ of the ratios \eqref{logza1} are the same. This is expected since $\Delta_{-P,Q}=\Delta_{P,-Q}$. We stress that a very similar equation relating the Liouville structure constants with the so-called logarithmic primaries has been found in \cite{Zamhem}. Using the property of the function $\Upsilon_{\beta}(x)$ and in particular its invariance with respect to the transformation $x\to\beta+\beta^{-1}-x$, we computed \begin{equation} \frac{C(\alpha_a,\alpha_b,\alpha_{0,Q}+\delta)} {C(\alpha_a,\alpha_b,\alpha_{0,-Q}+\delta)}=i;\quad \quad \frac{C(\alpha_a,\alpha_b,\alpha_{P,0}+\delta)} {C(\alpha_a,\alpha_b,\alpha_{-P,0}+\delta)}=i. \end{equation} From these results, we can finally write the relation between the coefficient $X_{i,j}$ associated to the point $[i,j]$ and to its symmetric under reflection $[r-i,j]$, $[i.s-j]$ \begin{eqnarray} \label{canc_sing} \frac{X_{i,j}(\delta)}{X_{r-i,j}(\delta)}&=&-\delta^{-2}\left(\frac{B_{r-2i,s-2j}\lambda_{r-2i,s-2j}}{R_{r-2i,s-2j} (\{\alpha_{i}\})}\right)^2 +O\left(\delta^{-1}\right) \\ \label{canc_sing0} \frac{X_{i,s/2}(\delta)}{X_{r-i,s/2}(\delta)}&=&-1 +O\left(\delta\right);\quad \frac{X_{r/2,j}(\delta)}{X_{r/2,s-j}(\delta)} =-1 +O\left(\delta\right) \end{eqnarray} Clearly, one can verify that the results (\ref{canc_sing}) and (\ref{canc_sing0}) are consistent with the previous findings based on the analysis of the poles and zeros of the structure constants. Let us focus on the situation in which one pair of colliding channels is present, for example the operators with charge $\hat{\alpha}_{i,j}(\delta)$ and $\hat{\alpha}_{r-i,j}(\delta)$. In the case ($i\neq r/2 \;\wedge\; j\neq s/2$), we have seen that the primary operator $\phi_{2i-r,s-2j}$ interferes with the null field $\chi_{r-2i,s-2j}$ responsible for the pole in the conformal block of $\phi_{r-2i,s-2j}$. We can fix the global normalization in (\ref{X_ij_dege}) such that \begin{equation} X_{i,j}(\delta) = X_{i,j}^{(1)}\delta+X^{(2)}_{ij}\delta^2+O(\delta^3) \quad X_{r-i~j}(\delta) = X_{r-i,j}^{(-1)}\delta^{-1}+X^{(0)}_{r-i,j}+O(\delta). \end{equation} On the other hand, the expansions in powers of $\delta$ of the singular conformal block $F(x|c,\hat{\Delta}_{ij},\{\Delta\})$ and of the non-singular conformal block $F(x|c,\hat{\Delta}_{r-i,j},\{\Delta\})$ can be written as \begin{eqnarray} F(x|c,\hat{\Delta}_{ij},\{\Delta\})&=&\delta^{-1}G^{(-1)}_{i,j}(z) +G^{(0)}_{i,j}(x)+O(\delta)\\ F(x|c,\hat{\Delta}_{r-i,j},\{\Delta\})&=&G^{(0)}_{r-i,j}+\delta G^{(1)}_{r-i,j}(z)+O(\delta^2), \end{eqnarray} where as it follow from \eqref{rec_conf} \begin{equation} \label{delta_conf} G^{(0)}_{r-i,j}(x)=F(x|c,\hat{\Delta}_{r-i,j},\{\Delta\}), \quad G^{(-1)}_{i,j}(x)=\frac{x^{P Q}S_{P,Q}}{\lambda_{P,Q}}F(x|c,\hat{\Delta}_{ij}+PQ,\{\Delta\}). \end{equation} In the equations \eqref{delta_conf}, $P= r-2i \geq 1$, and $Q=s-2j \geq 1$ and we recall that $\hat{\Delta}_{i,j}(\delta)=\hat{\Delta}_{i,j}+\lambda_{P,Q}\delta +O(\delta^2)$. The functions \begin{equation} G^{(0)}_{i,j}(x)=1+\sum_{k=1}^{\infty}a_k~x^k;\quad G^{(1)}_{r-i,j}(x)=\sum_{k=1}^{\infty}b_k~x^k \end{equation} are obtained by differentiating the conformal blocks with respect to $\delta$. In general, compact analytic expressions for the expansion coefficients $a_k$ and $b_k$ are not available. Nevertheless one could derived them from the recursion formula \cite{Zam_rec1,Zam_rec2} or applying the AGT correspondence \cite{AGT} which relates a conformal block to the Nekrasov instantons partition function of a given $\mathcal N=2$ supersymmetric four-dimensional gauge theory, see for instance \cite{SaTa10}. Applying the identity (\ref{canc_sing}), we can finally write the finite contribution $I^{PQ}_{nm}$ to the integral \eqref{exp_int} of the two colliding channels $\hat{\alpha}_{i,j}\equiv\alpha_{P,Q}$ and $\hat{\alpha}_{r-i,j}\equiv\alpha_{-P,Q}$ \begin{align} &I_{nm}^{PQ}\equiv|x|^{2\Delta_{PQ}(\delta)}~X_{i,j}|F(x|c,\Delta_{PQ}(\delta), \{\Delta\})|^2+ |x|^{2\Delta_{-P,Q}(\delta)}~X_{r-i,j}|F(x|c,\Delta_{-P,Q}(\delta),\{\Delta\})|^2\nonumber\\ &=|x|^{2\Delta_{PQ}}\left[a^{(1)}_{P,Q}\log|x|^2|x|^{2PQ} X_{r-i,j}^{(-1)}|G_{r-i,j}^{(0)}(x)|^2+ X_{i,j}^{(1)}|G_{i,j}^{(-1)}(x)|^2+X_{r-i,j}^{(0)}|x|^{2PQ}|G_{r-i,j}^{(0)}(x)|^2\right.\nonumber\\ &\Bigl.\hspace*{0.5cm}+X_{i,j}^{(1)}\bigr(G_{i,j}^{(-1)}(x)G_{i,j}^{(0)}(\bar{x})+c.c.\bigl)+ X_{r-i,j}^{(-1)}|x|^{2PQ}\bigl(G_{r-i,j}^{(0)}(x)G_{r-i,j}^{(1)}(\bar{x})+c.c.\bigr)\Bigr], \label{cg_ms} \end{align} where the factor $a_{P,Q}^{(1)}$ is defined by \begin{equation} a_{P,Q}^{(1)}\equiv\lim_{\delta \to 0}\frac{1}{\delta}\left(PQ-\Delta_{-P,Q}(\delta)+ \Delta_{P,Q}\right)=2P\beta. \label{blog} \end{equation} If the colliding channels are $\hat{\alpha}_{i,j}$ and $\hat{\alpha}_{i,s-j}$, a formula identical to \eqref{cg_1s} will hold, but with the substitutions $r-i\to i $, $j\to s-j$ and of course the parameter $a_{P,Q}^{(1)}$ will have to be replaced by \begin{equation} \label{bdef} a_{P,Q}^{(2)} =\lim_{\delta \to 0}\frac{1}{\delta}\left(PQ-\Delta_{P,-Q}(\delta)+ \Delta_{PQ}(\delta)\right)=-2Q\beta^{-1}. \end{equation} Notice that the two coefficients $a_{P,Q}^{(1)}$ and $a_{P,Q}^{(2)}$ have always different signs. We briefly outlined now the case where the channel with charge $\hat{\alpha}_{i,s/2}$ collides with $\hat{\alpha}_{r-i,s/2}$. Using now the relation (\ref{canc_sing0}), we can write \begin{equation} X_{i,j}(\delta) = X_{i,s/2}^{(-1)}\delta^{-1}+X^{(1)}_{(i,s/2)}+O(\delta) \quad X_{r-i,s/2}(\delta) = -X_{i,s/2}^{(-1)}\delta^{-1}+X^{(1)}_{(r-i,s/2)}+O(\delta) \label{canc0} \end{equation} while the expansions of the corresponding conformal blocks, both non-singular, are \begin{align} &F(x|c,\hat{\Delta}_{i,s/2},\{\Delta\})=G^{(0)}_{i,s/2}(x)+\delta G^{(1)}_{i,s/2}(x)+O(\delta^2)\\ &F(x|c,\hat{\Delta}_{i,s/2},\{\Delta\})=G^{(0)}_{(i,s/2)}(x)+\delta G^{(1)}_{(r-i,s/2)}(x)+O(\delta^2). \end{align} The finite contribution in the correlation function can be then easily derived and the coefficient of the logarithm is $a_{P,0}^{(1)}$, see \eqref{bdef}. From the result (\ref{cg_1s}) one can infer an Operator Product Expansion (OPE), compatible with the cancellation of the leading singularities in the four point function. Consider two scalar operators $\Phi_{\alpha_a}$ and $\Phi_{\alpha_b}$ which fuse into a third operator $\Phi_{P,Q}^{\delta}(x,\bar{x})=\phi_{P,Q}^{\delta}(x) \otimes \bar{\phi}_{P,Q}^{\delta}(\bar{x})$, having charge $\alpha_{P,Q}(\delta)=\alpha_{P,Q}+\delta$ and dimension $\Delta_{P,Q}(\delta)=\Delta_{P,Q}+\lambda_{P,Q}\delta+O(\delta^2)$ \begin{align} \Phi_{\alpha_a}(x,\bar{x})\Phi_{\alpha_b}(0)=\frac{C(\alpha_a,\alpha_b,\alpha_{P,Q}(\delta))} {|x|^{2\Delta_{\alpha_a}+2\Delta_{\alpha_b}- 2\Delta_{P,Q}(\delta)} }&\left(\phi^{\delta}_{P,Q}(0)+\dots+\eta_{P,Q} x^{PQ}\chi^{\delta}_{P,Q}(0)+\dots\right) \times \nonumber \\ \times &\left(\bar{\phi}^{\delta}_{P,Q}(0)+\dots+\eta_{P,Q} \bar{x}^{PQ} \bar{\chi}^{\delta}_{P,Q}(0)+\dots\right), \label{ope_s} \end{align} where we have considered both the holomorphic and anti-holomorphic parts of the OPE. In the limit $\delta\to 0$ the descendant $\chi_{P,Q}^{\delta}$ is the null-field $\chi_{P,Q}$ of the representation $\mathcal{V}_{\Delta_{P,Q}}$ of the Virasoro algebra. From (\ref{deg_con}), the coefficient $\eta_{P,Q}$ is given by \begin{equation} \eta_{P,Q}=\frac{\langle \chi^{\delta}_{P,Q}(\infty) \phi_{\alpha_a}(1) \phi_{\alpha_b}(0)\rangle} {\langle \chi_{P,Q}(\infty) \chi_{P,Q}(0) \rangle} =\frac{R_{P,Q}(\alpha_a,\alpha_b)}{\delta \lambda_{P,Q} B_{P,Q}}+O(\delta^0), \end{equation} and is singular in the limit $\delta\to 0$ if $R_{P,Q}(\alpha_a,\alpha_b)\neq 0$. As we have seen before local correlation functions may have singularities $O(\delta^{-1})$ and $O(\delta^{-2})$. The first one is always eliminated by the vanishing of the structure constant $C(\alpha_a,\alpha_b,\alpha_{P,Q}(\delta))$ which is $0(\delta)$. The collision between $\Lambda_{P,Q}^{\delta}\equiv \chi^{\delta}_{P,Q}\otimes\bar{\chi}^{\delta}_{P,Q}$ and the operator $\Phi^{\delta}_{-P,Q}$ (or $\Phi^{\delta}_{P,-Q}$) which enters in the correlation function and is produced in the OPE \begin{equation} \Phi_{\alpha_a}(x,\bar{x})\Phi_{\alpha_b}(0)=\frac{C(\alpha_a,\alpha_b,\alpha_{-P,Q}(\delta))} {|x|^{2\Delta_{\alpha_a}+2\Delta_{\alpha_b}- 2\Delta_{-P,Q}(\delta)}}\left(\Phi_{-P,Q}^{\delta}+\dots\right), \label{reg_ter} \end{equation} cancels instead the remaining $\delta^{-1}$ singularity. Consider for instance the mixing between the fields $ \Lambda_{P,Q}^{\delta}$ and $\Phi^{\delta}_{-P,Q}$. The formula \eqref{logza1} suggests to define a pair of fields $(\mathcal{C},\mathcal{D})$ \begin{eqnarray} \label{log_field1} \mathcal{D}&\equiv&\frac{1}{\sqrt{\delta}}\left(\Phi_{-P,Q}^{\delta}+\frac{i\Lambda_{P,Q}^{\delta}}{\delta \lambda_{P,Q}B_{P,Q}}\right) \\ \label{log_field2} \mathcal{C}&\equiv&-\left(PQ -\Delta_{-P,Q}(\delta)+\Delta_{P,Q}(\delta) \right)\frac{\Phi^{\delta}_{-P,Q}}{\sqrt{\delta}} \end{eqnarray} such that in the limit $\delta\to 0$ one has the finite correlation functions \begin{eqnarray} \label{log_pair1} \langle \mathcal{D}(x,\bar{x})\mathcal{D}(0)\rangle &=&\frac{-2a^{(1)}_{P,Q}(\ln{|x|^2}+O(1))}{|x|^{4(\Delta_{PQ}+ PQ)}} \\ \label{log_pair2} \langle \mathcal{C}(x,\bar{x})\mathcal{D}(0)\rangle &=&\frac{a^{(1)}_{P,Q}}{|x|^{4(\Delta_{PQ}+ PQ)}} \\ \label{log_pair3} \langle \mathcal{C}(x,\bar{x})\mathcal{C}(0)\rangle&=&0 \end{eqnarray} The fields $\mathcal{C},\mathcal{D}$ are called a logarithmic pair, see for example \cite{Ca_rev}. Similar conclusions in the chiral case have been reached in \cite{VaJaSa}, see in particular eq. (1.22). \section{Appearance of logarithms for rational central charges} In the previous section we have shown that logarithms in the CG integrals (\ref{exp_int}) are necessarily generated by singular conformal blocks. The results we have presented so far are based on the assumption that $\beta^2\not\in\mathbb Q$. Indeed, when $\beta$ (and therefore the central charge $c$) takes the value \begin{equation} \beta=\beta_c(p,p')\equiv\sqrt{\frac{p'}{p}};\quad c(p',p)=1-6\frac{(p-p')^2}{p p'}, \label{scc} \end{equation} with $p$ and $p'$ two coprime positive integers, one has to take into account that \begin{equation} p \beta_c(p,p')-p' \beta_c(p,p')^{-1}=0. \label{spcco} \end{equation} Accordingly, the charges $\alpha_{r,s}$ satisfy the additional symmetry \begin{equation} \alpha_{r,s}=\alpha_{r\pm p,s\pm p'} \label{nsym} \end{equation} which implies new identifications between operators. The minimal models $\mathcal{M}_{p,p'}$ are CFT with central charge $c(p,p')$ which are constructed from the finite set of $\frac{1}{2}(p-1)(p'-1)$ irreducible representations $\Phi_{r,s}$, $1\leq r\leq p-1$ and $1\leq s\leq p'-1$. The algebraic structure of extensions of minimal models, which contain in general indecomposable representations has been intensively studied, see for instance \cite{PeRaZu,FeGaSeTi} and references therein. Here we are interested in obtaining local correlation functions for theories with central charge (\ref{scc}) by taking the limits $\beta\to \beta_c$ of integrals of type (\ref{exp_int}), associated to a set of external charges $\alpha_i$ obeying the condition (\ref{sum_12}). As we previously showed, the presence of zeros and poles in the structures constants, as well as the singularities in the conformal block, depend upon the position of the points $[i,j]$, representing the internal charges $\hat{\alpha}_{ij}=\alpha_1+\alpha_2+i\beta-j\beta^{-1}$ on the grid of Sec.4. In order to classify every type of singularity entering in the correlation function \eqref{exp_int}, we can use the expansion \begin{equation} p\beta-p'\beta^{-1}=2 p(\beta-\beta_c)+O\left((\beta-\beta_c)^2\right), \end{equation} to define new charges $\tilde{\alpha}_{i}$ such as \begin{eqnarray} \alpha_{i}&=& \underbrace{\alpha_i+\frac{k p}{4}\beta-\frac{k p'}{4}\beta^{-1}-\frac{k p}{2} (\beta-\beta_c)}_{\tilde{\alpha}_{i}}+O\left((\beta-\beta_c)^2\right)\quad i=1,2\nonumber \\ \alpha_{i}&=& \underbrace{\alpha_i-\frac{k p}{4}\beta+\frac{k p'}{4}\beta^{-1} +\frac{k p}{2}(\beta-\beta_c)}_{\tilde{\alpha}_{i}}+O\left((\beta-\beta_c)^2\right) \quad i=3,4 \label{ncr} \end{eqnarray} with $k$ a positive integer. In this way the correlation function (\ref{exp_int}) with given initial external charges $\alpha_i$ can be also written in terms of an integral with the same number of screening $n$ and $m$ but with new external charges $\tilde{\alpha_{i}}=\alpha_i+O\left((\beta-\beta_c)^2\right)$ and with new values of $r,s$: $r\to r- k p, s\to s-k p'$. The grid of points representing the internal channels is therefore obtained by translating the initial grid by a multiple $k$ of $(p,p')$ \begin{equation} \hat{\alpha}_{i,j}=\alpha_{r-kp -2i,s-k p'-2j}+k p(\beta-\beta_c). \end{equation} Note that the distance from the central charge (\ref{scc}) play the role of a small regularization parameter $\delta$ \begin{equation} \delta=-k p(\beta-\beta_c). \label{delta_c} \end{equation} In the case of special values of $\beta$, we can compare the order of the different contribution of the integral using the \eqref{logza1}, (\ref{canc_sing}) and (\ref{canc_sing0}) applied to any translated grid, in particular when the translated grid intersects different quadrants. It is important to stress that, in the limit $\beta\to \beta_c(p,p')$, the new symmetry (\ref{spcco}) affect the analysis of the zeros of the function (\ref{deg_con}) and of the function $\Upsilon_{\beta}(x)$. Therefore the conditions (\ref{ij_deg_sing}) and the (\ref{infies1})-(\ref{zeros2}) fail to take into account new zeros. Moreover, in the limit $\beta\to \beta_c(p,p')$, the norm $B_{PQ}$, see (\ref{deg_con}), can vanish due to the (\ref{spcco}). For these reasons, differently from the case $\beta^2 \notin \mathbb{Q}$, we did not find a simple graphical rule to determine the order of contribution of each channel. Nevertheless, the analysis can be done case by case and we will show some concrete examples below. Finally, note that for $\beta^2\not\in\mathbb Q$, we have regularized the expressions by introducing a parameter $\delta$ which does not depend on $\beta$. Here, instead, we consider $\delta=O(\beta-\beta_c)$ and we study the limit $\beta\to \beta_c$. The parameters $a_{P,Q}^{(1)}$ and $a_{P,Q}^{(2)}$, which are the prefactors of term $\log{|x|^2}$, are given by \begin{equation} a_{P,Q}^{(1)}=2P \beta_c\quad a_{P,Q}^{(2)}=-2 Q \beta_c. \label{bspc} \end{equation} \subsection{$c=-2$, $\beta_c^2=1/2$} The theories with $c=-2$ occupy a very particular place as paradigms of LCFTs. Indeed, local $c=-2$ theories can be constructed as models of free simplectic fermions and correlation functions can be computed exactly, see for instance \cite{Salog,GaKa,Ka}. The $c=-2$ models have been intensively studied in polymer theory as they describe the dense phase of the $n\rightarrow 0$ of the $O(n)$ model known to capture the critical properties of dense polymers and spanning trees\cite{DuLERW,DuSa3,Maj}. In \cite{Salog}, the CG formalism has been used to determine local four-point correlation functions of leg operators. In particular the correlation function of four $\Phi_{2,1}$ operators with conformal weight $\Delta_{2,1}=-1/8$ has been shown to contain logarithmic terms. As a physically relevant example we discuss the behavior of correlation functions \begin{equation} F^{l}(x)\equiv\langle \Phi_{2+2l,1}(0)\Phi_{2+2l,1}(1)\Phi_{2+2l,1}(x)\Phi_{2+2l,1}(\infty) \rangle. \end{equation} which for $l\in\mathbb N$ is the four-point function of the $4l$ leg operator for dense polymers \cite{Salog}. By choosing \begin{equation} \alpha_1=\alpha_2=\alpha_{3}=\alpha_{2+2l,1}\quad \alpha_4=\alpha_{-2-2l,-1} \label{extchargecm2} \end{equation} the function $F^{l}$ can be expressed by the CG integral with $n=2 l+1$ and $m=0$ screenings. In the following we define: \begin{equation} \delta=-2 \left(\beta-\sqrt{\frac{1}{2}}\right) \end{equation} \underline{Four-point function with l=0} The case $l=0$ is particularly simple as it corresponds to the one-screening case, i.e. it has only two internal fusion channels, see Fig. \ref{l0cm2}. As can be seen by translating the grid, these two fusion channels can be related to the fields $\Phi_{1,0}$ and $\Phi_{-1,0}$ with the same conformal dimension. The (\ref{canc_sing}) tells us that the coefficients $X_{1,0}$ and $X_{-1,0}$ are of the same order in $\delta$ and in particular that $X_{1,0}=-X_{-1,0}$. The function $F^{0}(x)$ takes therefore the form (\ref{cg_ms}). This was pointed out in detail in \cite{Gu}, where the conformal block corresponding to the case $l=0$ was shown to satisfy a degenerate hypergeometric differential equation. Note that the divergence $\delta^{-1}$ of the coefficients $X_{i,j}$ associate to points on the axis is here canceled by the vanishing of the global normalization $A_\beta$ in \eqref{Zam_structure} for $\beta=\sqrt{1/2}$. \begin{figure}[t] \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (5,2)--(3,1); \draw[->,dashed] (3,2)--(1,1); \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[above]{{\tiny $\delta^0$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^0$}}; \draw[->] (3,1)--(1,0); \draw[->] (1,1)--(-1,0); \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,0) node[above]{{\tiny $\delta^{0}$}}; \draw[->] (-1,0)--(-3,-1); \draw[->] (1,0)--(-1,-1); \draw (-1,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (-3,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-1) node[above]{{\tiny $\delta^{0}$}}; \draw[->,dashed] (-3,-1)--(-5,-2); \draw[->,dashed] (-1,-1)--(-3,-2); \end{tikzpicture} \end{center} \caption{The two internal channels $[0,0]$ and $[1,0]$ and the points obtained by translation of the vector $(2,1)$. The analysis of Sec.~4 can be carried out analogously with the \textit{caveat} that new sources of zeros and poles may arise at $\beta^2$ rational in the structure constants.} \label{l0cm2} \end{figure} By choosing the normalization factor $\mathcal{N}=O(\delta^{-1})$, the CG integral gives a finite $F^{0}$. In order to familiarize with our notations, it may be useful to write explicitly the result for $l=0$ by setting the (\ref{extchargecm2}) into the equations (\ref{F00},\ref{F10}) and (\ref{X00},\ref{X10}) and take the limit $\beta \to \sqrt{1/2}$. One obtains \begin{align} &G^{(0)}_{0,0}(x)=G^{(0)}_{1,0}(x)=\; _2F_1(\frac{1}{2},\frac{1}{2},1,x) \nonumber \\ &G^{(1)}_{0,0}=\sqrt{2}(\beta-\sqrt{1/2})\sum_{k=1}^{\infty} \frac{[1/2]_k^2}{(k!)^2} \left(4\psi(k+1/2)-4\psi(1/2)-2\psi(k+1)+2\psi(1)\right)\nonumber \\ &G^{(1)}_{1,0}=\sqrt{2}(\beta-\sqrt{1/2})\sum_{k=1}^{\infty} \frac{[1/2]_k^2}{(k!)^2} \left(2\psi(k+1)-2\psi(1)\right) \nonumber \\ &X^{(0)}_{0,0}=-X^{(0)}_{(1,0)}=-\frac{\pi}{2\sqrt{2} (\beta-\sqrt{1/2})}+\frac{\pi}{2}\left(1-4\gamma_{E}-4\psi(1/2)\right)+O((\beta-\sqrt{1/2})^2)\nonumber\\ &X_{1,0}=\frac{\pi}{2\sqrt{2} (\beta-\sqrt{1/2})}-\frac{\pi}{2}\left(1-4\gamma_{E}-4\psi(1/2)\right)+O((\beta-\sqrt{1/2})^2), \end{align} where $\gamma_{E}$ is the Euler constant and $\psi(z)$ the digamma function. Using the above results in the (\ref{cg_1s}) one obtains \begin{equation} F^0(x)=|x|^{1/2}|x-1|^{1/2}\left[\sqrt{2}\ln{|x|^2} |_2F_1(1/2,1/2,1,x)|^2 +\left(_2F_1(1/2,1/2,1,x) M(\bar{x})+c.c.\right)\right], \label{cm2l0} \end{equation} with \begin{equation} M(x)=\sum_{k=1}^{\infty} \frac{[1/2]_k^2}{(k!)^2} \left(\psi(k+1)-\psi(1)-\psi(k+1/2)-\psi(1/2)\right). \end{equation} Note that the factor of the logarithmic term corresponds to $a^{(1)}_{1,0}$ of the equation (\ref{bspc}). The correlation function (\ref{cm2l0}) was also studied in \cite{Gu} and discussed in \cite{GaKa,HaYa}. \underline{Four-point function with l=1/2} \begin{figure} \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (7,2)--(5,1); \draw[->,dashed] (5,2)--(3,1); \draw[->,dashed] (3,2)--(1,1); \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[above]{{\tiny $\delta^0$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[above]{{\tiny $\delta^0$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^0$}}; \draw[->] (5,1)--(3,0); \draw[->] (3,1)--(1,0); \draw[->] (1,1)--(-1,0); \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[above]{{\tiny $\delta^{0}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,0) node[above]{{\tiny $\delta^{0}$}}; \draw[->] (3,0)--(1,-1); \draw[->] (1,0)--(-1,-1); \draw[->] (-1,0)--(-3,-1); \draw (-3,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (1,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-1) node[above]{{\tiny $\delta^{0}$}}; \draw[->,dashed] (1,-1)--(-1,-2); \draw[->,dashed] (-1,-1)--(-3,-2); \draw[->,dashed] (-3,-1)--(-5,-2); \end{tikzpicture} \end{center} \caption{This figure shows the points associated to the three internal channels in the case $l=1/2$ and the corresponding translated grids.} \label{l1cm2} \end{figure} Analogously to the case $l=0$, by fixing $\mathcal{N}=O(\delta^{-1})$, the contributions of the channels $\Phi_{3,1}$ and $\Phi_{1,1}$ sum to produce a finite result. As for the remaining third channel $\Phi_{5,1}=\Phi_{1,-1}$, we can determine its order in $\delta$, using (\ref{logza1}) and compare it with the one of the field $\Phi_{3,1}=\Phi_{-1,-1}$. For external charges as in (\ref{extchargecm2}) with $l=1$, one obtains \begin{equation} S_{1,1}(\alpha_1,\alpha_2,\alpha_3,\alpha_4)= O(\delta^2). \end{equation} The ratio of the coefficients $X_{1,0}$ and $X_{0,0}$ is therefore given by \begin{equation} \frac{X_{1,0}}{X_{0,0}}= \frac{C(\alpha_{1},\alpha_2,\alpha_{1,1}+\delta)C(\alpha_{1},\alpha_2,\alpha_{1,1}+\delta)} {C(\alpha_{1},\alpha_2,\alpha_{-1,1}+\delta)C(\alpha_{1},\alpha_2,\alpha_{-1,1}+\delta)}\sim \delta^2\frac{\lambda_{1,1}^2}{S_{1,1}(\{\alpha_i\})^2}=O(\delta^{-2}). \end{equation} Therefore we can conclude that in the limit $\delta\to 0$, i.e. $c\to -2$, $F^{1}(x)$ is given by the contributions of the two channels $\Phi_{3,1}$ and $\Phi_{1,1}$, which produce a logarithmic term $\propto \frac{1}{\sqrt{2}} \ln|x|^2$. Notice that we could have considered the correlation function involving more general external charges $\alpha_i$ (for instance $\alpha_1\neq \alpha_2\neq \alpha_3$) satisfying the (\ref{sum_12}) with $n=2,m=0$ and $(r=5,s=1)$. In this case the coefficient $S_{1,1}(\{\alpha_i \})$ vanishes linearly in $\delta$, $S_{1,1}(\{\alpha_i \})\sim \delta$. This would imply that the contribution of the channel $\Phi_{5,1}$ would be of the same order of the other two. As the the dominant term of the channels $\Phi_{3,1}$ and $\Phi_{1,1}$ cancels, this would imply that, in the limit $\beta\to 1/\sqrt{2}$, the correlation $F^{1}(x)$ is given by the channel $\Phi_{5,1}$ alone, without logarithmic terms. \underline{Four-point function with l=1} \\ \begin{figure}[t] \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (9,2)--(7,1); \draw[->,dashed] (7,2)--(5,1); \draw[->,dashed] (5,2)--(3,1); \draw[->,dashed] (3,2)--(1,1); \draw (7,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (7,1) node[above]{{\tiny $\delta^0$}}; \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[above]{{\tiny $\delta^0$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[above]{{\tiny $\delta^0$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^0$}}; \draw[->] (7,1)--(5,0); \draw[->] (5,1)--(3,0); \draw[->] (3,1)--(1,0); \draw[->] (1,1)--(-1,0); \draw (5,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,0) node[above]{{\tiny $\delta^0$}}; \draw (3,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,0) node[above]{{\tiny $\delta^{0}$}}; \draw (1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,0) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,0) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,0) node[above]{{\tiny $\delta^{0}$}}; \draw[->] (5,0)--(3,-1); \draw[->] (3,0)--(1,-1); \draw[->] (1,0)--(-1,-1); \draw[->] (-1,0)--(-3,-1); \draw (3,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (-3,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-1) node[above]{{\tiny $\delta^{0}$}}; \draw (1,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-1) node[above]{{\tiny $\delta^{0}$}}; \draw[->] (3,-1)--(1,-2); \draw[->] (1,-1)--(-1,-2); \draw[->] (-1,-1)--(-3,-2); \draw[->] (-3,-1)--(-5,-2); \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[above]{{\tiny $\delta^{0}$}}; \draw (-5,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-5,-2) node[above]{{\tiny $\delta^{0}$}}; \draw (-3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-2) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-2) node[above]{{\tiny $\delta^{0}$}}; \draw[->,dashed] (1,-2)--(-1,-3); \draw[->,dashed] (-1,-2)--(-3,-3); \draw[->,dashed] (-3,-2)--(-5,-3); \draw[->,dashed] (-5,-2)--(-7,-3); \end{tikzpicture} \end{center} \caption{Grids of internal channels for the correlation function $F^{2}(x)$.} \label{l2cm2} \end{figure} We consider finally the case $l=2$, see Fig. \ref{l2cm2} In order to evaluate the order of its contribution, we compare the channel $\Phi_{7,1}\equiv\Phi_{1,-2}$ with the channel $\Phi_{5,1}\equiv\Phi_{-1,-2}$ by using the (\ref{logza1}). Assuming the values of the external charges as in (\ref{extchargecm2}) with $l=2$,one can verify that \begin{equation} S_{12}(\alpha_1,\alpha_2,\alpha_3,\alpha_4)\neq 0. \end{equation} This means that the conformal block associated to $\Phi_{5,1}$ has a singularity $O(\delta^{-1})$ at level 2. As we have said before, this singularity is canceled by the collision with the field $\Phi_{7,1}$. Taking into account that $S_{1,1}(\{\alpha\})$ is of order $\delta^{-1}$ we can conclude that $F^{2}(x)$ has two logarithmic terms coming from the pairs $\Phi_{1,1}$, $\Phi_{3,1}$ and $\Phi_{5,1}$, $\Phi_{7,1}$ with prefactors $a_{1,0}^{(1)}=\sqrt{2}$ and $a_{2,1}^{(1)}=2\sqrt{2}$. The correlation function $F^{l}(x)$ for general $l$, as well as other correlation functions, such as for instance the four point correlation function $\langle \Phi_{1, l}(0)\Phi_{1,l}(1)\Phi_{1,l}(x)\Phi_{1,l}(\infty) \rangle$ of $L=4l+2$ operators \cite{Salog} can be studied case by case with the same techniques without any additional technical complication. \subsection{$c=0$, $\beta^2=2/3$} CFTs at central charge $c=0$ are of particular interest as they play a crucial role in the study of critical systems with quenched disorder or in the description of dilute self-avoiding walk and critical percolation theory. For $\beta=\sqrt{2/3}$, the vanishing of the dimension $\Delta_{2,1}$ is at the origin of the so-called $c=0$ catastrophe \cite{GuLu1, GuLu2}. Such terminology refers to the fact that OPE \begin{equation} \Phi_{\alpha}(x,\bar{x}) \Phi_{\alpha}(0)=\frac{C_{\Delta_{\alpha}}}{|z|^{4\Delta_{\alpha}}} \left(1+x^2\frac{2\Delta_{\alpha}}{c}T(0)\right)\left(1+\bar{x}^2\frac{2\Delta_{\alpha}}{c}\bar{T}(0)\right), \label{id_fusion} \end{equation} is singular in the $c\rightarrow 0$ limit. The $c=0$ catastrophe can be rephrased in the following terms: if we consider the identity as internal channel with dimension $\Delta_p=0$ in \eqref{rec_conf} then the conformal block $F(x|c=0,\Delta_p=0,\{\Delta\})$, with $\Delta_1=\Delta_2=\Delta_{\alpha}$ does not have a singularity at level 1 since $R_{1,1}(\alpha,\alpha)=0$ but is divergent at the level two because $R_{2,1}(\alpha,\alpha)\neq 0$ for $\Delta_{\alpha}\neq 0$. As we have already seen in the general case, the singularities are $O(c^{-2})$ and $O(c^{-1})$, and they are canceled inside correlation functions by the vanishing of the structure constant $C_{\Delta_{\alpha}}$ which is $O(c)$ and by the collision with one of the operators $\Phi_{-2,1}$ and $\Phi_{2,-1}$. It may be useful to compare the expression (\ref{id_fusion}) with the expression (\ref{ope_s}). Let us define $\chi^{\delta}_{2,1}=(L_{-1}^2-\beta^2 L_{-2}) V_{1,1}^{\delta}$, where $V_{1,1}^{\delta}$ is the primary with charge $\alpha_{-1,-1}(\delta)=\alpha_{-1,-1}+\delta$ and \begin{equation} \delta=-3(\beta-\sqrt{2/3}). \label{delta_c_0} \end{equation} From the (\ref{nsym}) we have $\alpha_{2,1}=\alpha_{-1,-1}(\delta)+O(\delta^2)$. In the limit $\beta\to \sqrt{2/3}$, i.e. $c\to0$, the field $\chi_{2,1}^{\delta}(x)$ is therefore proportional to the stress-energy tensor, \begin{equation} \lim_{\delta\to 0} \chi_{2,1}^{\delta}=-2/3\;T \end{equation} which becomes a null-vector at level $2$ in the Verma module of the identity. Setting $T=0$ would imply to consider a trivial theory that contains in its spectrum only fields invariant under all conformal transformations, i.e. only the identity field. The coefficient \begin{equation} \frac{3}{2}\eta_{2,1}=\frac{3}{2}\frac{\langle \chi_{2,1}^{\delta}(\infty) V_{\alpha}(1) V_{\alpha}(0)\rangle}{\langle \chi_{2,1}^{\delta}(\infty) \chi_{2,1}^{\delta}(0) \rangle} \sim \frac{1}{2}\frac{R_{2,1}(\alpha,\alpha)}{(\beta-\sqrt{2/3})(2\beta-\beta^{-1})B_{2,1}}\sim\frac{2\Delta_{\alpha}}{c} \end{equation} is consistent with the expansion (\ref{ope_s}) where the regularization parameter is played by $\delta$. In order to make connection with previous results, let us consider for instance the correlation function \begin{equation} \langle \Phi_{n+1,m+1}(0)\Phi_{n+1,m+1}(x) \Phi_{n+1,m+1}(1)\Phi_{-n-1,-m-1}(\infty) \rangle \label{cfco} \end{equation} of four operators with the same dimension $\Delta=\Delta_{n+1,m+1}$. The neutrality condition (\ref{ch_neu}) is satisfied with $n$ and $m$ screenings respectively of type $\beta$ and $\beta^{-1}$. Let us consider in the following some illustrative cases \underline{$n=2, m=0$} \begin{figure} \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (8,3)--(5,1); \draw[->,dashed] (6,3)--(3,1); \draw[->,dashed] (4,3)--(1,1); \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[above]{{\tiny $\delta^{-2}$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[above]{{\tiny $\delta^{-1}$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^{-2}$}}; \draw[->] (5,1)--(2,-1); \draw[->] (3,1)--(0,-1); \draw[->] (1,1)--(-2,-1); \draw (2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw (0,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (0,-1) node[above]{{\tiny $\delta^{-1}$}}; \draw (-2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw[->] (5,1)--(2,-1); \draw[->] (3,1)--(0,-1); \draw[->] (1,1)--(-2,-1); \draw (2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw (0,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (0,-1) node[above]{{\tiny $\delta^{-1}$}}; \draw (-2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw[->] (2,-1)--(-1,-3); \draw[->] (0,-1)--(-3,-3); \draw[->] (-2,-1)--(-5,-3); \draw (-1,-3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-3) node[above]{{\tiny $\delta^{-2}$}}; \draw (-3,-3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-3) node[above]{{\tiny $\delta^{-1}$}}; \draw (-5,-3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-5,-3) node[above]{{\tiny $\delta^{-2}$}}; \draw[->,dashed] (-1,-3)--(-4,-5); \draw[->,dashed] (-3,-3)--(-6,-5); \draw[->,dashed] (-5,-3)--(-8,-5); \end{tikzpicture} \end{center} \caption{We show an example of a correlation functions at $c=0$, with two screening charges. Notice that the case $n=1$ does not contain any logarithm.} \end{figure} Using the analysis of the previous sections, one observes that the most singular terms are of order $\delta^{-2}$ and correspond to the charges $\alpha_{-2,-1}$ and $\alpha_{2,-1}$, associated respectively to the identity $\Phi_{1,1}$ and to the $\Phi_{5,1}$ fusion channels. The dominant contribution $\delta^{-2}$ is canceled in the sum, providing an $O(\delta^{-1})$ contribution with logarithms to the correlation function. Choosing a global normalization constant of $O(\delta)$, the correlation function (\ref{cfco}) with $n=2$ and $m=0$ shows a a logarithmic contribution generated by the colliding channels $\Phi_{1,1}$ and $\Phi_{5,1}$ and a non-logarithmic contribution from the channel $\Phi_{3,1}$. The prefactor of the logarithmic terms is \begin{equation} a_{2,1}^{(1)}=4\sqrt{6}. \end{equation} As we previously noticed, see (\ref{log_field1}, \ref{log_field2}), the value $a_{2,1}^{(1)}$ can be seen as the mixing parameter between the \textit{scalar} field $\Phi_{5,1}(x,\bar{x})$ and the field $T(x) \bar{T}(\bar{x})$. Let us comment on the fact that in the chiral $c=0$ CFT \cite{GuLu1,GuLu2} which describes boundary percolation, one is general interested in the mixing parameter $b_{\tiny\mbox{perco}}$, see \cite{Ri} for a recent survey, between the \textit{chiral} operator $T(x)$ and $\phi_{5,1}(x)$. It is therefore not surprising that such parameter $b_{\tiny\mbox{perco}}=-5/8$, numerically coincides with \begin{equation} -\left(\left.\frac{d \beta_c}{d c}\right|_{c=0}~ 2 a_{2,1}^{(1)}\right)^{-1}=-\frac{5}{8}. \end{equation} The factor $-\frac{d \beta_c}{d c}|_{c=0}$ takes into account that, in the definition of the parameter $b_{{\tiny\mbox{perco}}}$ one uses the limit $c\to0$ instead of $\beta\to \sqrt{2/3}$, while the factor $2$ comes from the fact that we are considering a non-chiral theory where the chiral and anti-chiral degrees of freedom are coupled symmetrically. \underline{$n=0,m=2$} \begin{figure}[t] \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (4,7)--(1,5); \draw[->,dashed] (4,5)--(1,3); \draw[->,dashed] (4,3)--(1,1); \draw (1,5) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,5) node[above]{{\tiny $\delta^{0}$}}; \draw (1,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,3) node[above]{{\tiny $\delta^{-2}$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^{-2}$}}; \draw[->] (1,5)--(-2,3); \draw[->] (1,3)--(-2,1); \draw[->] (1,1)--(-2,-1); \draw (-2,3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,3) node[above]{{\tiny $\delta^{0}$}}; \draw (-2,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,1) node[above]{{\tiny $\delta^{-2}$}}; \draw (-2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw[->] (-2,3)--(-5,1); \draw[->] (-2,1)--(-5,-1); \draw[->] (-2,-1)--(-5,-3); \draw (-5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-5,1) node[above]{{\tiny $\delta^{0}$}}; \draw (-5,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-5,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw (-5,-3) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-5,-3) node[above]{{\tiny $\delta^{-2}$}}; \draw[->,dashed] (-5,1)--(-8,-1); \draw[->,dashed] (-5,-1)--(-8,-3); \draw[->,dashed] (-5,-3)--(-8,-5); \end{tikzpicture} \end{center} \caption{The figure shows a correlation function at $c=0$ with $m=2$ and $n=0$.} \label{c0_m2} \end{figure} Using $S_{2,1}(\{\alpha_i\})\neq 0$ and $S_{5,1}(\{\alpha_i\})=O(\delta^2)$ together with (\ref{logza1}), one obtains the contributions shown in Fig. \ref{c0_m2}. Setting the global normalization constant $\mathcal{N}$ of order $\delta$, the function (\ref{cfco}) for $n=0,m=2$ is given by the \eqref{cg_ms} with colliding channels $\Phi_{1,1}=\Phi_{-2,-1}$ and $\Phi_{1,3}=\Phi_{-2,1}$ with \begin{equation} a_{2,1}^{(2)}=-3\sqrt{6}. \end{equation} The above factor $a_{2,1}^{(2)}$ is formally related to the parameter $b_{{\tiny\mbox{poly}}}=5/6$ \cite{GuLu1,GuLu2} which is defined as the mixing factor between the chiral fields $\phi_{1,3}(x)$ and $T(x)$ in the $c=0$ theory describing dilute polymers with a boundary. Analogously to the previous case, we have \begin{equation} -\left(\left.\frac{d \beta_c}{d c}\right|_{c=0}~ 2 a_{2,1}^{(2)}\right)^{-1}=\frac{5}{6}. \end{equation} Although their numerical values are simply related, we stress again that the two mixing factors $b_{{\tiny\mbox{poly}}}$ and $a_{2,1}^{(2)}$ corresponds to different physical situations: in the first case the LCFT is chiral and the chiral field $\phi_{1,3}$ is the logarithmic partner of the stress-energy tensor while in the second case, the scalar field $\Phi_{3,1}(x,\bar{x})$ is the logarithmic partner of the operator $T\bar{T}$. \underline{$n=2, m=1$} Until now we have seen correlation functions of the $c=0$ theory given by the collision of the channel $\Phi_{1,1}$ with $\Phi_{5,1}$ or with $\Phi_{1,3}$. The case $n=2,m=1$ is particularly interesting because, fixing $\mathcal{N}=O(\delta)$ both the channels $\Phi_{5,1}$ and $\Phi_{1,3}$ provide a finite contribution to the integral, as shown in Fig. \ref{c0_n3}. \begin{figure} \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw (5,3) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw (3,3) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw (1,3) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[circle,fill=red,minimum size=1pt,scale=0.5] {}; \draw[->] (5,3)--(2,1); \draw[->] (5,1)--(2,-1); \draw[->] (3,3)--(0,1); \draw[->] (3,1)--(0,-1); \draw[->] (1,3)--(-2,1); \draw[->] (1,1)--(-2,-1); \draw (2,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (2,1) node[above]{{\tiny $\delta^{-2}$}}; \draw (2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (2,-1) node[above]{{\tiny $\delta^{-2}$}}; \draw (0,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (0,1) node[above]{{\tiny $\delta^{-1}$}}; \draw (0,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (0,-1) node[above]{{\tiny $\delta^{-1}$}}; \draw (-2,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,1) node[above]{{\tiny $\delta^{-2}$}}; \draw (-2,-1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-2,-1) node[above]{{\tiny $\delta^{-2}$}}; \end{tikzpicture} \end{center} \caption{The figure shows a correlation function computed at $c=0$ in which both the operators $\Phi_{1,3}$ and $\Phi_{5,1}$ enters in collision with the null-field $T\bar{T}$. The others two channels can be shown to give a subleading contribution.} \label{c0_n3} \end{figure} In this case, the logarithmic prefactor $a$ turns out to be \begin{equation} a=a_{2,1}^{(1)}+a_{2,1}^{(2)}=\sqrt{6}, \label{sum_a} \end{equation} This value is again formally related to the mixing parameter $b_{{\tiny\mbox{bulk}}}=-5$, recently conjectured \cite{VaGaJaSa, Ri} for the bulk percolation and dilute polymer theory indeed \begin{equation} \label{bulk} -\left(\left.\frac{d \beta_c}{d c}\right|_{c=0} a\right)^{-1}=-5. \end{equation} We emphasize that our result is obtained in a very different setting respect to the one in \cite{VaGaJaSa}. We recall that CG integrals are necessarily associated to a diagonal CFT whereas in \cite{VaGaJaSa}, the $b_{{\tiny\mbox{bulk}}}$ emerged as the mixing parameter between the non-diagonal field $\phi_{-2,1}(x)\phi_{2,1}(\bar{x})$ and the stress energy tensor $T(x)$. Moreover, since the structure constant $C_{\Delta_{\alpha}}$ in the OPE \eqref{id_fusion} is $O(c)$ the introduction of a logarithmic partner for $T(x)$, usually called $t(x)$, does not appear necessary in our treatment; the numerical coincidence \eqref{bulk} is however quite interesting. Summarizing, we have shown three different possible situations occurring for the $c=0$ logarithmic CG integral, which correspond respectively to the mixing of $T\bar{T}$ with the scalar fields $\Phi_{5,1}$, $\Phi_{1,3}$ or with both. In this latter case, the operator $T\bar{T}$ has two logarithmic partners which get identified, see eq. \eqref{sum_a}, in the $\beta^2\rightarrow 2/3$ limit, a circumstance reminiscent of the findings of \cite{Ri}. We point out that the existence of three different scenarios is not in contradiction with the unicity of the $b$ parameter. Indeed, in order to construct a consistent theory, one has to fix the normalization of the two point functions for all the operators entering in the spectrum. For instance, one could imagine that a consistent choice of these normalization factors would force all the correlation functions of the first two kinds to vanish, as it has also been argued in \cite{VaGaJaSa}. \subsection{$\mathcal{M}_{p,p'}$ minimal models} It is well known that the minimal models $\mathcal{M}_{p,p'}$ do not show logarithmic behavior. Indeed the conformal blocks are not singular as the Kac operators $\Phi_{r,s}$ satisfy two differential equations, one of order $rs$ and the other of order $(p-r)(p'-s)$. The divergences associated to the corresponding two null-vector states are therefore compensated by the vanishing of the associated matrix elements (\ref{deg_con}). Maybe less known is the fact that the property of $\mathcal{M}_{p,p'}$ to be a rational theory, does not imply the vanishing of the CG structure constants $C(\alpha_{r_1,s_1},\alpha_{r_2,s_2},\alpha_{r_3,s_3})$ \cite{DoFa,DoFa2,DoFa3} when they contain operators inside and outside the Kac table. This is well explained in \cite{Doco}, where the $\mathcal{M}_{5,4}$ minimal model is taken as an example. We review the argument given in \cite{Doco} following our scheme. Let us consider for instance the correlation function \begin{equation} \langle \Phi_{3,1}(0)\Phi_{3,1}(x) \Phi_{3,1}(1)\Phi_{-3,-1}(\infty) \rangle, \label{cfci} \end{equation} for $\beta=\sqrt{3/4}$ ($c=1/2$), i.e. the four-point energy correlation function in the Ising model. The computation of \eqref{cfci} in the CG approach has been also proposed by \cite{DiFra} as an exercise. This example shows that, even if the structure constant $C(\alpha_{31},\alpha_{31},\alpha_{51})\neq 0$ does not vanish, the field $\Phi_{5,1}$, which is out side the Kac table, do not enter in the computation of the four point correlation function. From $R_{12}(\alpha_{31},\alpha_{31})\neq 0$ one sees that the conformal block with $\Phi_{3,1}=\Phi_{1,2}$ has a singularity at level two. On the other hand, using the equations (\ref{logza1}), one obtains $C(\alpha_{31},\alpha_{31},\alpha_{31})=O(\delta)$ and $C(\alpha_{31},\alpha_{31},\alpha_{51})=O(\delta^0)$ where $\delta\propto (\beta-\sqrt{3/4})$. The internal translated channels of the correlator (\ref{cfci}) as well as their contributions in $\delta\propto (\beta-\sqrt{3/4})$ are shown in Fig. \ref{Ising}. \begin{figure} \begin{center} \begin{tikzpicture}[scale=0.6] \draw (-5,10) node[right]{$X_{i,j}|F(x|c,\hat{\Delta}_{i,j},\{\Delta\}|^2$}; \draw[->,thick] (-9,0) -- (9,0) node[below]{$x$}; \draw[->,thick] (0,-9) --(0,9)node[left]{$y$}; \foreach \x in {-8,...,8} \draw[very thin, gray] (\x,-8)--(\x,8); \foreach \y in {-8,...,8} \draw[very thin, gray] (-8,\y)--(8,\y); \draw[->,dashed] (9,4)--(5,1); \draw[->,dashed] (7,4)--(3,1); \draw[->,dashed] (5,4)--(1,1); \draw (5,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (5,1) node[above]{{\tiny $\delta^{0}$}}; \draw (3,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (3,1) node[above]{{\tiny $\delta^{0}$}}; \draw (1,1) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,1) node[above]{{\tiny $\delta^{0}$}}; \draw[->] (5,1)--(1,-2); \draw[->] (3,1)--(-1,-2); \draw[->] (1,1)--(-3,-2); \draw (1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (1,-2) node[above]{{\tiny $\delta^{0}$}}; \draw (-1,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-1,-2) node[above]{{\tiny $\delta^{0}$}}; \draw (-3,-2) node[circle,fill=black,minimum size=1pt,scale=0.5] {}; \draw (-3,-2) node[above]{{\tiny $\delta^{0}$}}; \draw[->,dashed] (1,-2)--(-3,-5); \draw[->,dashed] (-1,-2)--(-5,-5); \draw[->,dashed] (-3,-2)--(-7,-5); \end{tikzpicture} \end{center} \caption{The grid of internal channel for the four-point correlation function of the energy operator $\Phi_{3,1}$ in the Ising model.} \label{Ising} \end{figure} The dominant term of the colliding channel $\Phi_{3,1}$ and $\Phi_{5,1}$ cancels. The only contribution to the correlation function (\ref{cfci}) is therefore given by the identity field $\Phi_{1,1}=\Phi_{-3,-2}$. One can therefore verify that only operators inside the Kac table contribute to the correlation functions. \section{Conclusions} We considered local monodromy-invariant four-point correlation functions which are built from Virasoro algebra conformal blocks and which posses a CG integral representation (\ref{4vertex}). In particular we focused our attention on the singularities of the integrals (\ref{exp_int}) which generate, with an appropriate limit procedure, logarithmic terms. We have shown that the analysis of these singularities can be efficiently carried out by considering the analytic behavior of the time-like Liouville structure constants (\ref{Zam_structure}) and of Virasoro conformal blocks (\ref{rec_conf}). In particular, we found the key equations (\ref{logza1}, \ref{canc_sing}, \ref{canc_sing0}) which relate the structure constants (\ref{Zam_structure}) to combinatorial factor coming from the Virasoro algebra representation theory. We remark that in the cases under consideration, the structure constants (\ref{Zam_structure}) correspond to the Dotsenko-Fateev structure constants \cite{DoFa,DoFa2,DoFa3}. Nevertheless, the use of the properties of the function $\Upsilon_{\beta}(x)$, the building block for the formula (\ref{Zam_structure}), greatly simplifies the analysis, especially when multi-screening integrals are involved. The relations (\ref{logza1}),(\ref{canc_sing}) and (\ref{canc_sing0}) assure us that the dominant contributions coming from singular conformal block cancel and that the remaining finite terms contain logarithmic singularities. In particular they express in a compact form the fact that, for a general central charge, the logarithms are generated by the collision of a singular null field $\Lambda_{P,Q}(x,\bar{x})=\chi_{P,Q}(x)\otimes\chi_{P,Q}(\bar{x})$ with one of the non-degenerate primaries $\Phi_{-P,Q}(x,\bar{x})$ or $\Phi_{P,-Q}(x,\bar{x})$. Using our approach, we considered in detail the CG correlation functions when $\beta^2 \notin \mathbb {Q}$, see (\ref{central}). When the external charges take general values, we provided a direct graphical method to predict the contributions of the expansion (\ref{exp_int}). We adapted our approach to the case of $\beta^2=\sqrt{p'/p}$, with $p'$ and $p$ positive and coprime. We applied our method to the study of correlation functions of leg operators in dense polymers and spanning trees ($c=-2$) and to correlation functions of $c=0$ theories. We showed in particular that the situation in which the CG integral gets contributions from both the colliding fields $\Phi_{-P,Q}(x,\bar{x})$ and $\Phi_{P,-Q}(x,\bar{x})$ can occur. In the $c=0$ theory, this suggests the possibility that the operator $T(x)\bar{T}(\bar{x})$ has two logarithmic partner which get identified in the limit $c\to0$. In this particular case the prefactor of the logarithmic term is simply related to the value $b_{{\tiny\text{bulk}}}=-5$, determined in \cite{Ri, VaGaJaSa} for bulk $c=0$ LCFTs. The motivation behind this work is the construction of consistent CFTs to study geometric properties of critical phases. Although the correlation functions, for which the CG integral furnishes an integral representation necessarily come from a diagonal CFT, theoretical arguments and numerical investigations on Potts models \cite{DPSV} recently suggested that the diagonal theories based on the time-like Liouville structure constants are physically relevant to construct new CFTs capable of describing random conformal fractals. The results presented here show also that the analytic properties of time-like Liouville structure constants encode logarithmic features of Virasoro CFTs, and turn out to be very useful to determine the singularities of CG integrals as well as to compute local logarithmic correlation functions. This was shown by evaluating the logarithmic behavior of physically relevant correlation functions which satisfy differential equations of order higher than two, i,e. which are given by multi-screening CG integrals. \vspace{1cm} \noindent\textbf{ACKNOWLEDGMENTS.} We thank R. Vasseur for having share with us his knowledge of LCFTs and for showing us some of his results before publication. We are grateful to V. Dotsenko and H. Saleur for suggestions and comments and we also benefit from discussions with J. Jacobsen, Y. Ikhlef, V. Schomerus and J. Teschner. \begin{appendices} \section{Explicit expressions in the one-screening case} From the identities \begin{equation} \frac{d}{d x} [x]_s=[x]_s \sum_{j=0}^{s-1}\frac{1}{x+j},\quad \Gamma(x+n)=[x]_n \Gamma(x), \quad \sin{\pi x}=\frac{1}{\Gamma(1-x)\Gamma(x)} \end{equation} with $[x]_n=\Gamma(x+n)/\Gamma(x)$ being the Pochhammer symbol, the functions $G^{(a)}_{i,j}$ and $X_{i,j}^{(a)}$ can be easily determined \begin{align} &G^{(-1)}_{0,0}(x)=\frac{(-1)^{s-1}[-2\beta \alpha_2]_{s}[-s-2\beta \alpha_3]_{s}x^{s} }{2\beta\Gamma(s)\Gamma(1+s)} \; G^{(0)}_{1,0}(x), \label{G000}\\ &G^{(0)}_{1,0}(x)= (1-x)^{2\alpha_2\alpha_3}_2F_1(-2\beta \alpha_3,s-2\beta \alpha_2;1+s;x), \label{G011}\\ &G^{(0)}_{0,0}(x)=(1-x)^{2\alpha_2\alpha_3}\left[\sum_{l=0}^{s-1}\frac{[-2\beta \alpha_2]_l [-s-2\beta \alpha_3]_l}{l![1-s]_l} x^l+\frac{[-2\beta \alpha_2]_{s}[s-2\beta \alpha_3]_{s}x^{s} }{s![1-s]_{s-1}}H(x)\right] \label{G11} \\ &G^{(1)}_{1,0}(x)=-(1-x)^{2\alpha_2\alpha_3}\left[\sum_{l=1}^{\infty}\frac{[-2\beta \alpha_3]_{l} [1+s-2\beta \alpha_2]_l}{l![2+s]_l} x^l \sum_{j=0}^{l-1} \frac{1}{j+2+s}\right], \\ &H(x)=\left[\frac{1}{1+2\beta \alpha_3}-\sum_{l=0}^{s-2}\left(\frac{1}{l+1-s}- \frac{1}{l-s-2\beta \alpha_3}\right)\right] _2F_1(-2\beta \alpha_3,s-2\beta \alpha_2;1+s;x)\nonumber \\ &+\sum_{l=1}^{\infty}\frac{[-2\beta \alpha_3]_{l} [1+2\beta \alpha_2]_l}{l![1+s]_l} x^l \sum_{k=0}^{l-1} \left(\frac{1}{k+1}-\frac{1}{k-2\beta \alpha_3}\right),\label{HH}\\ &X^{(-1)}_{1,0}=-\frac{1}{2\beta}\left[\frac{\pi[-2\beta\alpha_2]_s}{\Gamma(1+s)}\right]^2,\\ &X^{(1)}_{0,0}=2\beta\left[\frac{\pi\Gamma(s)}{[-s-2\beta\alpha_3]_{s}}\right]^2. \end{align} \section{The special function $\Upsilon_{\beta}(x)$} This appendix contains some properties of the function $\Upsilon_{\beta}(x)$, for more details we remand to the comprehensive review \cite{Barnes}. The function $\Upsilon_{\beta}(x)$ is related to the Barnes double Gamma function $\Gamma_2(x|\omega_1,\omega_2)$ defined for $\text{Re}(\omega_1)>0$ and $\text{Re}(\omega_2)>0$ by \begin{equation} \log\Gamma_2(x|\omega_1,\omega_2)\equiv\left.\frac{d}{ds}\right|_{s=0}\sum_{n,m=0}^{\infty}\frac{1}{(x+n\omega_1+m\omega_2)^s},\quad \text{Re}(s)\geq2. \end{equation} The function $\Gamma_2(x|\omega_1,\omega_2)$ satisfies the shift relations \begin{align} \label{shift1} &\Gamma_2(x+\omega_1|\omega_1,\omega_2)=\frac{\sqrt{2\pi}\omega_2^{1/2-x/\omega_2}}{\Gamma(x/\omega_2)} \Gamma_2(x|\omega_1,\omega_2),\\ \label{shift2} &\Gamma_2(x+\omega_2|\omega_1,\omega_2)=\frac{\sqrt{2\pi}\omega_1^{1/2-x/\omega_1}}{\Gamma\bigl(x/\omega_1)} \Gamma_2(x|\omega_1,\omega_2), \end{align} which show that it has simple poles on the lattice points $x=-n\omega_1-m\omega_2$, for $n$ and $m$ non-negative. The proof of (\ref{shift1}-\ref{shift2}) is sketched below. Observe that \begin{equation} \label{app1} \frac{1}{x^s}=-\frac{1}{2\pi i}\Gamma(1-s)\int_C dt (-t)^{s-1}e^{-xt}, \end{equation} where the contour $C$ starts end at $\infty$ and encircles the origin counterclokwise. The branch cut in \eqref{app1} is chosen on the positive real axis and arguments in the complex $t$ plane are bounded in $[-\pi,\pi]$. Using \eqref{app1} and performing explicitly the contour integral\footnote{One must regularize the divergent integrals when the contour has been split.}, the relation \begin{equation} \label{int_rep} \left.\frac{d}{ds}\right|_{s=0}\sum_{n=0}^{\infty}\frac{1}{(x+n\omega)^s}=\frac{1}{\omega} \int_{0}^{\infty}\frac{dt}{t}\left[-\frac{1}{t}-\frac{e^{-t}(\omega-2x)}{2}+\frac{\omega e^{-xt}}{1-e^{-\omega t}}\right] \end{equation} can be proved. Then recalling Binet formula for the logarithm of the Gamma function \begin{equation} \log\Gamma(x)=(x-1/2)\log x+\frac{1}{2}\log 2\pi-x+\int_{0}^{\infty}\frac{dt}{t}e^{-tx}\left(\frac{1}{2}-\frac{1}{t}+\frac{1}{e^t-1}\right), \end{equation} we obtain \begin{equation} \left.\frac{d}{ds}\right|_{s=0}\sum_{n=0}^{\infty}\frac{1}{(x+n\omega)^s}=\frac{1}{2}\log 2\pi+\left(\frac{1}{2}-\frac{x}{\omega}\right)\log\omega-\log\Gamma(x/\omega) \end{equation} and the shift relations are now easily verified. Analogously to \eqref{int_rep} one derives the integral representation for the Barnes double Gamma function, valid for $\text{Re}(x)>0$ \begin{align} \label{doubleGamma} \log\Gamma_2(x|\omega_1,\omega_2)&=\frac{1}{\omega_1\omega_2}\int_{0}^{\infty}\frac{dt}{t}\left[-\frac{1}{t^2}-\frac{Q-2x} {2 t}-\frac{e^{-t}[(2x-Q)^2+2x(x-Q)+\omega_1\omega_2]}{12}+\right.\nonumber\\ &\hspace*{3cm}\left.+\frac{\omega_1\omega_2 e^{-xt}}{(1-e^{-\omega_1 t})(1-e^{-\omega_2t})}\right], \end{align} where we defined $Q\equiv\omega_1+\omega_2$. Following \cite{Zam_lectures}, we introduce the function $\Upsilon(x|\omega_1,\omega_2)$ as \begin{equation} \Upsilon(x|\omega_1,\omega_2)=\frac{\Gamma_2^2(Q/2|\omega_1,\omega_2)}{\Gamma_2(x|\omega_1,\omega_2)\Gamma_2(Q-x|\omega_1,\omega_2)}. \end{equation} It is then straightforward to check that for $\omega_1=\omega_2^{-1}=\beta$, $\Upsilon(x|\beta,\beta^{-1})\equiv\Upsilon_\beta(x)$ has the integral representation \eqref{Zam_int} and that satisfies the shift relations \eqref{rec_Zam}. Notice that the function $\Upsilon_{\beta}(x)$ has zeros at $x=-n\beta-m\beta^{-1}$ or $x=(n+1)\beta+(m+1)\beta^{-1}$ with $n,m$ non-negative. \end{appendices}
\section{Introduction} Unusual physical properties of magnetite (Fe$_3$O$_4$), the oldest known magnetic material, gave birth, among others, to Mott's model of metal-insulator transition (MIT) and to the variable range hopping mechanism of electrical conductivity \cite{Mott1968, Mott1956}. Many physical properties of the material were apparently explained by a relative simple model of charge ordering of Fe$^{2+}$ and Fe$^{3+}$ cations below so-called Verwey temperature $T_{V}\approx 123$~K at which the MIT together with a structural transition occur (the Verwey model of charge ordering \cite{Verwey1939, Wright2002, Jeng2004}). And although in recent studies \cite{Subias2004, Shchennikov2009, Senn2012, Senn2012b} the question of the strict meaning of Verwey-like charge ordering and the low-T atomic structure is solved, the exact mechanism of the Verwey transition is still in dispute and the phenomena below $T=50$~K are far from being well understood. These phenomena are treated in our paper. Magnetite is ferrimagnetic with a N\'eel temperature $T_{N} \approx 860$~K; it forms a spinel-like cubic lattice at high temperatures ($\approx 1400$~K) which gradually transforms from random spinel to inverse-spinel cubic at room temperature \cite{Mack2000} and discontinuously transforms to monoclinic Cc at $T_V$ , as was determined by x-ray diffraction \cite{Zuo1990,Senn2012} and NMR \cite{Novak2000b}. Other measurements, however, suggest even lower crystal symmetry (magnetoelectric measurements \cite{Rado1975}, appearance of ferroelectricity \cite{Kato1981} and x-ray topography measurements \cite{Medrano1999}). In this paper we focus on the behavior of magnetite well below $T_{V}$, especially its magnetic properties in weak ac and dc fields. We have noticed a striking difference between the ac susceptibilities accompanying magnetic after-effect studies (MAE, disaccommodation effect) of stoichiometric magnetite \cite{Walz2005} and the susceptibilities measured by other techniques (e.g. an inductive bridge method \cite{Simsa1985}) below 35~K. More specifically, the notion of the anomaly stems from the bifurcation of the $\chi'(T)$ dependence at low temperatures (see inset of FIG.~\ref{fig_anomaly}), where the upper curve represents a MAE measurement and the lower represents a classical result (e.g. a SQUID measurement; a notable exception is the result \cite{Balanda2005} of ac susceptibility studies). Bearing in mind the specifics of MAE measurements we devised a series of experiments with the aim of explaining these discrepancies. \section{Samples} The samples used in our experiments were high quality single crystals of synthetic magnetite grown in the laboratory of J.M. Honig of Purdue University (samples denoted H) and by V.A.M. Brabers of Eindhoven University of Technology (samples denoted B). The samples B were grown by a floating zone technique \cite{Brabers1971}. The samples H were prepared by a skull melter technique \cite{Harrison1978}. Namely, the sample B1 was annealed to a stoichiometric composition and cut to the shape of a cylinder 5~mm in diameter and 4~mm high. The cylinder axis is parallel to the crystallographic $c$ axis, direction $\langle 001 \rangle$. This sample was also used for NMR measurements \cite{Novak2000}. The sample H1 was a prism with dimensions $2.1\times2.3\times3.1\ \rm{mm}^3$ (the longest edge along $\langle 001 \rangle$) annealed for stoichiometry \cite{Aragon1983}. Since the annealing procedure was different from that for B1, it resulted in different defect pattern, as measured by MAE technique \cite{Walz2002}. This sample was also used for ac susceptibility measurements \cite{Balanda2005}. The sample H2 - a platelet of irregular shape roughly $8\times13\ \rm{mm}^2$ and thickness of 1.8~mm - although not annealed, was apparently close to stoichiometry due to preparation conditions \cite{Harrison1978}. Thus, three samples, very close to stoichiometry, but with possibly different defect pattern were measured. For comparison, the measurements were also performed on the deliberately nonstoichiometric sample, H3 a roughly pyramidal single crystal of Fe$_{3(1-\delta)}$O$_4$ with $\delta=0.002$ and dimensions approximately $1.5\times1.5\times2.0$~mm$^3$. \section{Technique} The magnetic properties were measured using a non-commercial, continuously operating SQUID magnetometer with an immobile sample, where the magnetic field is generated by a superconducting solenoid operating in non-persistent mode, and the variations of the magnetic moment of the sample is read continuously using a SQUID \cite{Youssef2009}. The workspace, including the superconducting solenoid, superconducting gradiometer, SQUID, thermometers, heater, and the sample was magnetically shielded, using both a superconducting (lead) can and soft magnetic material surrounding the cryostat. The achieved spectral density of magnetic moment was $\approx$17~pA~m$^2$~Hz$^{-1/2}$ and residual field was less than 2 $\mu$T inside the sample chamber. The sample, mounted on a sapphire holder, was in $^{4}$He gas environment at atmospheric pressure and was subjected to time-varying magnetic field of maximum intensity of about 20~mT. The spectral response of the magnetometer was flat from dc up to about 150~Hz. Both the applied field generation and the sample response acquisition were carried out by a dedicated computer equipped with appropriate analog converters. In a typical susceptibility measurements reported here the applied field was varied sinusoidally, the SQUID signal (sampled at 6.4~kHz) was processed by fast Fourier transform and the results were expressed as a set of complex harmonic susceptibilities. With weak driving ac fields the sample response was linear and only the fundamental susceptibility was considered, the higher harmonics being negligible. \section{Results} When a single crystal of stoichiometric magnetite is cooled from room temperature down to a few K in the absence of a dc magnetic field and in a weak ac field, the thermal dependence of its complex ac susceptibility $\chi_{ac} = \chi' - i\chi''$ is remarkably simple (see inset of FIG. \ref{fig_anomaly}). \begin{figure} \includegraphics [clip=true,scale=1.0]{anomaly_inset_symb.eps} \caption{\label{fig_anomaly} Real part (top symbols) and imaginary part (bottom symbols) of complex ac susceptibility, H1 sample, as a function of temperature, ac field amplitude 5~$\mu$T, ac field frequency 1.6~Hz ($\circ$), 3.1~Hz ($\bullet$,$\blacktriangle$), 6.3~Hz ($\triangle$) and 12.5~Hz ($\times$).The anomaly ($\bullet$) was excited by increasing the dc field by 2~mT at 10~K. The inset: the overall temperature dependence of real susceptibility in weak magnetic field, $\mu_0 H_{ac} = 10\ \mu$T at 6.3~Hz.} \end{figure} Above the Verwey transition the value of $\chi'$ may reach as high \cite{Simsa1985} as $10^3$ (SI units), especially near the easy axis reorientation temperature \cite{Reznicek2012}, but due to finite sample size the measured value is limited to $1/N$, $N$ being the demagnetising factor. Below $T_V$ the crystal symmetry changes and new (and large) components of magnetocrystalline anisotropy energy appear \cite{Chiba1983}, $|\chi_{ac}|$ reduces to the order of unity \cite{Simsa1985}, and demagnetising effects can be neglected in a qualitative approach. Further below $T_V$, in the range 60~K to 35~K, a frequency-dependent thermally activated process, denoted here as "glass-like transition" after \cite{Janu2007}, occurs. Both the strong frequency dependence of this transition and the accompanying peak in absorption (the imaginary part, $\chi''$) agree well with Debye's relaxation model of thermally activated magnetic moments. Below about 30~K, when these moments are frozen in for all practical frequencies, we measured and described a novel manifestation of relaxation processes in magnetite, which we call the \emph{low temperature anomaly} (FIG. \ref{fig_anomaly}). The anomaly is described as follows: During cooling of a stoichiometric magnetite in a "normal" way (i.e. in relatively weak magnetic field without abrupt changes - ac field up to 100~$\mu$T and dc field up to several mT), $\chi'$ reaches its lowest value below the glass-like transition (further denoted as the "frozen" state) and retains this value down to the lowest temperatures. Also, $\chi''$ is nearly zero in this region. Upon subsequent warming at the same conditions, the same $\chi(T)$ is observed (symbols "$\blacktriangle$" in FIG. \ref{fig_anomaly}). However, when an abrupt disturbance is applied at low temperature (e.g. step change of dc field by several mT), $\chi'$ jumps to a new higher steady value, the anomaly is "excited". Upon subsequent warming of such excited sample, $\chi(T)$ relaxes back to the "frozen" state value before the glass-like transition takes place (symbols "$\bullet$" in FIG. \ref{fig_anomaly}). This universal behaviour of stoichiometric magnetite was found to be insensitive to the ac field frequency and amplitude (provided it is relatively weak, up to about 100~$\mu$T) and the absolute intensity of dc magnetic field (within our accessible range of +/-~20~mT). Only the magnitude and character of excitation event (even a short pulse of several mT amplitude may excite the anomaly) determines the extent to which the anomaly is excited. To further explore this behavior, we have performed an experiment in which the dc field was swept in a triangular fashion ( 0 $\rightarrow$ +1mT $\rightarrow$ 0 $\rightarrow$ -1mT $\rightarrow$ 0 with a period of 120~s) with superimposed weak ac field at several fixed temperatures (FIG. \ref{fig_sweep}). \begin{figure} \center \includegraphics [clip=true,scale=1]{sweep_small.eps} \caption{\label{fig_sweep} The DC field dependence of ac susceptibility at various temperatures: (a) 10~K, (b) 28~K, (c) 35~K - the "frozen" state, (d) 43~K and (e) 50~K - above the glass-like transition. The field was swept continuously in the range of +/-1~mT in a triangular fashion, the whole cycle taking 120~s. The susceptibility was measured simultaneously with an ac field of 20~$\mu$T at 6~Hz, H1 sample.} \end{figure} Apparently, at 35~K the sample maintained the "frozen" state regardless of the changing dc field (curve (c)). At lower temperatures the change of dc field causes $\chi'$ to increase, but in this case the process is largely reversible, and the sample returns close to the "frozen" state after removal of the dc field (curves (a), (b)). With other samples (H2, B1) or with bigger changes of the dc field the process is irreversible. \begin{figure} \includegraphics [clip=true,scale=1]{sdc_symb.eps} \caption{\label{fig_sdc} Real part of ac susceptibility (full symbols) and accompanying changes of spontaneous dc magnetisation (open symbols), H1 sample, as a function of temperature, ac field amplitude 5~$\mu$T at 6.3~Hz. The anomaly was excited by decreasing the dc field by 2~mT at 10~K. (The units of magnetisation are essentially A/m except for an additive constant - an offset - which is unknown due to working principle of a SQUID. This offset is also different for cooling ($\triangle$) and warming ($\circ$).)} \end{figure} Changes of spontaneous magnetisation of magnetite sample during the transition are depicted in FIG. \ref{fig_sdc}. Although more dependent on the thermal and magnetic history of the sample, the magnetisation shows several general features: (i) no dramatic changes of magnetisation occur in the region of glass-like transition, neither on cooling nor on warming the sample, (ii) if the anomaly was excited at low temperature the sample fully accommodates to the new conditions by a change of spontaneous magnetisation that takes place at higher temperature than the relaxation of ac susceptibility (to the "frozen" state), and (iii) the magnitude of the change of spontaneous magnetisation is rather low, in this case only about 10\% of the total response of the sample to the step change of the applied dc field. The relaxation of the anomaly was also found to depend on the rate of sample warming: the lower the rate the lower the temperature at which the "frozen" state is reached. This feature is further explored in relaxation measurements where a step change of magnetic field is applied at a defined temperature in the range of 10~K to 35~K and the time dependence of the ac susceptibility is recorded (FIG. \ref{fig_relaxation}). \begin{figure} \includegraphics [clip=true,scale=1]{relaxace_small.eps} \caption{\label{fig_relaxation} Relaxation of the real part of ac susceptibility after a 2~mT step of applied magnetic field at 25~K (top symbols), 30~K, and 33~K (bottom symbols). H1 sample, ac field amplitude 5~$\mu$T at 6.3~Hz.} \end{figure} The properties of several samples (including a nonstoichiometric single crystal) were measured and the results summarised in Table \ref{table_temperatures}. \begin{table} \caption{\label{table_temperatures}A summary of transition temperatures (the Verwey transition, the glass-like transition $T_g$ at 6~Hz, and the temperature of relaxation of the anomaly $T_a$, warming rate 1~K/m) of measured samples. Ac amplitude was 10~$\mu$T, the anomaly was excited by a dc field step of 2~mT at 10~K. $T_g$ and $T_a$ are defined as inflection points of the $\chi'(T)$ curve. The differences in the magnitude of the anomaly (in percent of the magnitude of the glass-like transition) is mainly caused by different demagnetising factors of the samples. H3 is nonstoichiometric with $\delta = 0.002$.} \begin{ruledtabular} \begin {tabular}{lcccc} sample & $T_V$ & $T_g$ at 6 Hz & $T_a$ & anomaly\\ & (K) & (K) & (K) & magnitude \\\hline H1 & 122.2 & 41.7 & 27.2 & 20 \% \\ H2 & 122.0 & 40.3 & 23.0 & 80 \% \\ B1 & 121.5 & 36.4 & 18.0 & 50 \% \\ H3 & 117.5 & 21.0 & - & - \\ \end{tabular} \end{ruledtabular} \end{table} Note, that in the samples departing from perfect stoichiometry the glass-like transition shifts quickly to lower temperatures and the anomaly cannot be observed even at lowest attainable temperature (about 6~K in the current setup). \section{Discussion} The microscopic origin of the observed phenomena cannot be determined unambiguously from the experimental findings. However, we present several arguments suggesting that the measured changes of susceptibility of magnetite below $T_V$ arise solely from the changes of the mobility of \emph{domain walls}: (i) both the saturation magnetisation ($4\mu_B$ per Fe$_3$O$_4$ formula agrees well with the experimental value of 0.5~MA/m \cite{Ozdemir1999}) and the magnetocrystalline anisotropy (the largest component \cite{Chiba1983} is about 0.25~MJ/m$^3$) are approximately constant below $T_V$. Thus the susceptibility arising from the rotation of domain magnetisation should not depend on temperature. Moreover (ii) this susceptibility is about an order of magnitude lower \cite{Lewis1958} than the measured susceptibility. (iii) A notably similar effect was observed seven decades ago in iron with carbon or nitrogen impurities at 250~K \cite{Snoek1939}, attributed to diffusion of the impurities to the domain walls in order to accommodate magnetostrictive strain. These impurities then act as potential minima that hinder the domain wall motion in a certain temperature range. Similar measurements of high purity single crystal magnetite \cite{Balanda2005} at low temperatures also suggest that magnetic domain walls are responsible for the observed effects. The authors also eliminated the effect of structural domain walls by detwinning the sample in strong magnetic field. However, in those measurements the anomaly always appears in the "excited" state, probably due to the high amplitude of ac magnetic field used in their experiments ($\geq$~100~$\mu$T). To explain the low-temperature anomaly we hypothesise the existence of pinning centres that create a net of potential minima for the domain walls. Starting from the "frozen" state at 35~K (see FIG. \ref{fig_anomaly}) all domain walls lie in these minima, and a weak applied ac field induces only a small response of the sample magnetisation, $\chi'$ is small. According to the assumptions \cite{Snoek1938} about domain wall parameters we may assume that an ac field with amplitude $\mu_0 H_{ac} \leq 100\ \rm{\mu T}$ is weak enough to cause domain wall displacement that is much less than the wall thickness. This process is reversible and, if the pinning centres are stationary, no energy is dissipated ($\chi''$ is also small). But as the temperature is elevated the pinning centres are no longer fixed, thermal activation allows them to follow the displacing walls, and energy is dissipated. This process gives rise to the frequency-dependent Debye-like relaxation at 40~-~50~K as described previously \cite{Janu2007}. Above 60~K the pinning centres move freely with the domain walls, the potential minima are not effective, and $\chi'$ is large. To complete the explanation of the anomaly let us again start from the "frozen" state at 35~K, where the domain walls reside in their potential minima. This state is preserved when the sample is further cooled to the lowest temperature (6~K in this experiment). Now the anomaly can be "excited", e.g., by changing the dc component of the applied magnetic field by a sizable amount (several mT). The domain walls are pulled from their potential minima and $\chi'$ rises (see curve (a) in FIG. \ref{fig_sweep}). When this "excited" sample is warmed gradually, the hypothetical pinning centres start to move to their new equilibrium positions, i.e. into the domain walls, new potential minima appear and the susceptibility approaches the "frozen" state. This gives rise to strongly temperature-dependent relaxation of the ac susceptibility (see FIG. \ref{fig_relaxation}), and the very same relaxation is responsible for the large peak in MAE spectra at about 30~K \cite{Walz2005}. The connection between our results and MAE spectra is very intimate for two reasons: (i) both methods are based on ac magnetic susceptibility measurements after some excitation with a strong magnetic field (we excite the sample only once at the lowest temperature, whereas during a MAE measurement the magnetic pulse is applied periodically), and (ii) the characteristic times of both methods are similar. It is important to note that there are in fact two characteristic times incorporated in these techniques. One is determined by the frequency of probing magnetic field (1~kHz in MAE measurements and 1~Hz to 100~Hz in our measurements, corresponding to 1~ms and 10~ms to 1~s, respectively). The other characteristic time is the isochronal time of MAE experiment (1~s to 180~s); in our experiments it is determined by the rate of sample warming, and although it is not strictly defined, it is of the order of seconds to hundreds of seconds. The above mentioned peak in MAE spectra around 30~K is attributed by the authors \cite{Walz2005} to "\emph{intra-ionic} thermal inter-level excitations", i.e. an excitation of the "free" electron of an individual Fe$^{2+}$ ion from the ground sate across a gap that appears due to the energy level splitting in a trigonal crystal field. Our explanation, on the other hand, takes into account diffusion phenomena. The microscopic nature of the hypothesised pinning centres remains unclear. However, some arguments can be summarised to elucidate the underlying mechanism: (i) a similar effect in iron \cite{Snoek1939} is ascribed to the diffusion of impurities (at concentration of 0.01\%). This most definitely will \emph{not} be the case of magnetite, as even minute concentration of impurities or structural defects shifts the glass-like transition to lower temperatures (in accordance with MAE results \cite{Walz1982}) and destroys the low-temperature anomaly (see Table \ref{table_temperatures}). Also, relaxation effects of vacancies and impurities in magnetite are expected well above $T_V$. (ii) The crystallographic domains that appear when cooling magnetite single crystal through $T_V$ have a substantial impact on the magnetic susceptibility \citep{Balanda2005}, but their structure is fixed at low temperature (easy axis switching can be induced by much larger fields close below $T_V$ \cite{Krol2007}) and the structural domains themselves cannot account for the observed phenomena. (iii) One plausible explanation is the interplay between the structural, magnetic and ferroelectric domain structure, as spontaneous polarisation was identified in magnetite thin films below 40~K \cite{Alexe2009} and the coupling between magnetic and dielectric properties of magnetite was identified long ago \cite{Rado1975}. However, a comparison of relative sizes of structural, magnetic end ferroelectric domains, that might support this explanation, cannot be easily done, as the properties of ferroelectric domains in bulk magnetite are currently unknown and the size of magnetic domains was reported in the range 30~$\mu$m to 100~$\mu$m \cite{Blackman1963,Medrano1999}. In thin samples \cite{Kasama2010} the size of both structural and magnetic domains is in the range 1~$\mu$m to 10~$\mu$m with possibly even smaller twin domains. Other experimental findings that may be important with respect to the low-temperature anomaly are anomalous behaviour of dielectric susceptibility \citep{Akishige1985} and significantly shortened NMR spin-spin relaxation times \cite{Novak2000}, both observed in magnetite at similar temperatures. \section{Conclusion} We have measured and described a novel manifestation of relaxation phenomena in high quality single crystals of stoichiometric synthetic magnetite in weak low-frequency magnetic fields. The largely unnoticed discrepancy between ac susceptibilities accompanying magnetic after-effect measurements and the results of conventional experiments (performed on top quality synthetic samples) was consistently explained as resulting from the interaction between magnetic domain walls and pinning centres that can diffuse towards the domain walls at elevated temperatures (above about 25~K). A single diffusion process thus manifests itself both as a frequency-dependent glass-like transition above 35~K and time dependent susceptibility relaxation below 35~K, the characteristic experiment timescales being the frequency of the ac magnetic field and the rate of sample warming, respectively. However, the microscopic origin of the pinning centres remains unclear. Impurities and structural defects that can give rise to similar efects in pure iron \cite{Snoek1939} are ruled out, as even minute amounts of defects shift the glass-like transition quickly to lower temperatures and the anomaly disappears completely. The rigidity of structural domain walls at these temperatures and the results in strong fields \cite{Balanda2005} make it evident that the interaction between structural and magnetic domains cannot account for the observed effects. A process encompassing ferroelectricity is proposed, whose elucidation will require more profound experiments to be carried out, such as combined magnetic and transport measurements in strong electric field, or nonlinear susceptibility measurements by means of higher order harmonic susceptibilities. \begin{acknowledgments} We would like to thank to Dr. V. A. M. Brabers of Eindhoven University of Technology, The Netherlands, for supplying us with one of the samples used in the study. This work was supported by Institutional Research Plan Contract No. AVOZ10100520, Research Project MSM Contract No. 0021620834 and grant ESO MNT-ERA (ME10069 of MEYS CR) and SVV-2012-265303. \end{acknowledgments}
\section{Introduction} We consider the problem of recovering an unknown matrix $\mathbf X$ from the following observation model \begin{eqnarray} \mathbf Y = \mathbf A \mathbf X \mathbf B^T\label{obs_1} \end{eqnarray} where $\mathbf X\in \mathbb R^{N\times N}$, $\mathbf A \in \mathbb R^{M\times N}$, $\mathbf B \in \mathbb R^{L\times N}$ and $\mathbf A^T$ denotes the transpose of the matrix $\mathbf A$. This problem has been studied by many researchers in different contexts for arbitrary matrices $\mathbf X$ \cite{Penrose1,Dai2,Peng1}. In many applications dealing with high dimensional data, \emph{sparsity} is one of the low dimensional structures widely observed. Most popular transforms applied to 2-dimensional signals are in the form of (\ref{obs_1}) where compression is obtained by a transformation of rows followed by a transformation of columns of the data matrix \cite{Caiafa1,Fang2,Rivenson1,Dasarathy1}. With an arbitrarily distributed sparse matrix $\mathbf X$ in which each column/row has only a few non zeros, a natural question to ask is whether it is possible to design sensing matrices in the form of (\ref{obs_1}) so that $\mathbf X$ can be uniquely recovered from $\mathbf Y$ when $M,L < N$. Sparse signal recovery has attracted much attention in the recent literature in the context of \emph{compressive sensing (CS)} \cite{candes1,donoho1,Eldar_B1}. In the standard CS framework, a commonly used mechanism is to stack the high dimensional data into vector form to recover the sparse vector uniquely from an underdetermined linear system \cite{candes1,donoho1}. The observation model (\ref{obs_1}) can be equivalently written in vector form using Kronecker products as: \begin{eqnarray} \mathbf y = \mathbf C \mathbf x\label{obs_2} \end{eqnarray} where $\mathbf y = \mathrm{vec}(\mathbf Y) \in \mathbb R^{ML}$, $\mathbf C = \mathbf B \otimes \mathbf A \in \mathbb R^{ML\times N^2} $, $\mathbf x = \mathrm{vec}(\mathbf X) \in \mathbb R^{N^2}$, $\otimes$ denotes the Kronecker operator and $\mathrm{vec}(\mathbf X)$ is a column vector that vectorizes the matrix $\mathbf X$ (i.e. columns of $\mathbf X$ are stacked one after the other). The sensing matrix in (\ref{obs_2}) has a special structure, i.e., it can be represented as a Kronecker product of two matrices $\mathbf A$ and $\mathbf B$. It has been shown \cite{Duarte4,Jokar1,Duarte5,Jokar2} that the sparse signal $\mathbf x$ from (\ref{obs_2}) can be recovered by solving the following $l_1$ norm minimization problem \begin{eqnarray} \min ||\mathbf x||_1~ s.t.~ \mathbf C \mathbf x = \mathbf y \label{l_1_norm_min} \end{eqnarray} under certain conditions on the matrices $\mathbf A$ and $\mathbf B$ where $||\mathbf x||_p$ denotes the $l_p$ norm of $\mathbf x$. In particular, these results imply that the capability of recovering $\mathbf x$ based on (\ref{obs_2}) is ultimately determined by the worst behavior of $\mathbf A$ or $\mathbf B$. Also, this approach is computationally complex especially when the matrix dimension $N$ increases\cite{Rivenson1,Dasarathy1}. Several recent papers addressed the problem of recovering a sparse $\mathbf X$ from (\ref{obs_1}) without employing the Kronecker product. In \cite{Dasarathy1}, it was shown that a unique solution for $\mathbf X$ can be found when $\mathbf X$ is distributed sparse under certain conditions on $\mathbf A$ and $\mathbf B$ by solving the following optimization problem: \begin{eqnarray} \min ||\mathbf X||_1 ~ \mathrm{s}. ~\mathrm{t}. ~ \mathbf A \mathbf X \mathbf B^T = \mathbf Y\label{matrix_l1} \end{eqnarray} where $||\mathbf X||_1$ is the $l_1$ norm of $\mathrm{vec}(\mathbf X)$. The authors derive recovery conditions when the matrices $\mathbf A$ and $\mathbf B$ contain binary elements which are better than that obtained via the Kronecker product approach. In \cite{Rivenson1}, the authors discuss advantages in terms of computational, storage, calibration and implementation while solving (\ref{matrix_l1}) in matrix form compared to that with vector form. However, no specific algorithm was developed to solve for $\mathbf X$. In \cite{Fang2}, a version of orthogonal matching pursuit (OMP) (dubbed 2D OMP) is presented to find a sparse $\mathbf X$ in the matrix form (\ref{obs_1}) when $\mathbf A= \mathbf B$. Our goal in this paper is to develop algorithms to solve for sparse $\mathbf X$ from (\ref{obs_1}) without the employment of Kronecker products. We extend fast iterative shrinkage threshold algorithm (FISTA) \cite{Beck1,Yang1} developed for the vector case to sparse matrix recovery with matrix inputs. We further consider a greedy based approach via OMP to find the sparse solution. We show that both algorithms with matrix inputs are equivalent to their vector counterparts obtained via Kronecker products in terms of performance. However, the computational complexity of the matrix approach is shown to be much less, especially with FISTA, compared to solving the problem in vector form. \section{Sparse Matrix Recovery via $\l_1$ Norm Minimization}\label{matrix_l1norm} \subsection{Vector formulation} While numerous algorithms have been proposed in the literature to solve (\ref{l_1_norm_min}), in this paper we consider FISTA as discussed in \cite{Beck1,Yang1}. We consider the noisy observation model so that FISTA with vector inputs as given in Algorithm \ref{algo_FISTA_vec} \cite{Yang1}, is the solution of \begin{eqnarray} \underset {\mathbf x}{\min} \left\{\frac{1}{2} ||\mathbf y - \mathbf C \mathbf x||_2^2 + \lambda ||\mathbf x||_1\right\}\label{l1_norm} \end{eqnarray} where $\lambda$ is a regularization parameter. In Algorithm \ref{algo_FISTA_vec}, $L_f=||\mathbf C||_2$ is the Lipschitz constant of $\nabla f(\mathbf x)$ where $||\mathbf C||_2$ denotes the spectral norm of $\mathbf C$, $\nabla$ denotes the gradient operator, and $f(\mathbf x) = \frac{1}{2}||\mathbf y - \mathbf C \mathbf x||_2^2$, and \begin{eqnarray} \mathrm{soft}(\mathbf u, a) = \begin{array}{ccc} \mathrm{sgn}(\mathbf u_i)(|\mathbf u_i| - a)_+ \end{array} \end{eqnarray} for $i=1,\cdots,N^2$ where $\mathbf u_i$ is the $i$-th element of $\mathbf u$, $x_+$ equals $x$ if $x>0$ and equals $0$ otherwise. \begin{algorithm} \textbf{Input:} observation vector $\mathbf y$, measurement matrix $\mathbf C$\\ \textbf{output:} estimate for signal, $\hat{\mathbf x}$ \begin{enumerate} \item Initialization: $\mathbf x^{0} =\mathbf 0$, $\mathbf x^{1}=\mathbf 0$, $t_0=1$, $t_1=1$, $k=1$\\ Initialize: $\lambda_1$, $\beta\in (0,1)$, $\bar\lambda > 0$ \item \textbf{while} not converged \textbf{do} \item $\mathbf z^{k} = \mathbf x^k + \frac{t_{k-1}-1}{t_k} (\mathbf x^k - \mathbf x^{k-1})$ \item $ \mathbf u^k = \mathbf z^k - \frac{1}{L_f}\mathbf C^T (\mathbf C \mathbf z^k - \mathbf y) $ \item $\mathbf x^{k+1} = \mathrm{soft} \left(\mathbf u^k, \frac{\lambda_k}{L_f}\right)$ \item $t_{k+1} = \frac{1+\sqrt{4t_k^2 + 1}}{2}$ \item $\lambda_{k+1} = \max(\beta \lambda_k, \bar\lambda)$ \item $k=k+1$ \item \textbf{end while} \end{enumerate} $\hat{\mathbf x} = \mathbf x^{k} $ \caption{FISTA for sparse signal recovery with vector inputs}\label{algo_FISTA_vec} \end{algorithm} The computational complexity of FISTA is dominated by step 4 in Algorithm \ref{algo_FISTA_vec}. The matrix-vector multiplications require $\mathcal O(N^4ML +N^4+N^2ML)$ computations. Since $M,L \leq N$, the complexity is in the order of $\mathcal O(N^4ML)$. Thus, FISTA is feasible only when $N, M, L$ are fairly small. \subsection{Matrix formulation} With the noisy version of (\ref{obs_1}), we aim to solve the following $l_1$ norm minimization problem: \begin{eqnarray} \underset{\mathbf X}{\min} ~ \left\{\frac{1}{2} ||\mathbf Y - \mathbf A \mathbf X \mathbf B^T||_F^2 + \lambda ||\mathbf X||_1 \right\} \label{FISTA_matrix} \end{eqnarray} where $\lambda$ is a regularization parameter and $||\mathbf A||_F$ is the Frobenius norm of $\mathbf A$. In generalizing FISTA to solve (\ref{FISTA_matrix}), we follow a similar approach as discussed in \cite{Toh1}. Consider the more general unconstrained optimization problem: \begin{eqnarray} \underset{\mathbf X\in \mathbb R^{N\times N}}{\min} ~ F(\mathbf X) + \lambda G(\mathbf X) \end{eqnarray} where $G(\cdot)$ is a proper, convex, lower semicontinuous function, and $F(\cdot)$ is a convex smooth (continuously differentiable) function on an open subset of $\mathbb R^{N\times N}$ containing $\mathrm{dom} G = \{\mathbf X | G(\mathbf X) < \infty\}$. We assume that $\mathrm{dom} G$ is closed and $\nabla F$ is Lipschitz continuous on $\mathrm{dom} G$ with Lipschitz constant $L_f$; i.e. \begin{eqnarray} ||\nabla F(\mathbf X) - \nabla F(\mathbf Z) ||_F \leq L_f ||\mathbf X - \mathbf Z||_F, ~X, Z\in \mathrm{dom}~G. \label{lips_mat} \end{eqnarray} When $F(\mathbf X) = \frac{1}{2} ||\mathbf Y - \mathbf A \mathbf X \mathbf B^T||_F^2$, it can be shown that $||\nabla F(\mathbf X) - \nabla F(\mathbf Z) ||_F = ||\nabla f(\mathbf x) - \nabla f(\mathbf z) ||_2$ and $||\mathbf X - \mathbf Z||_F = ||\mathbf x - \mathbf z||_2$ where $\mathbf z = \mathrm{vec}(\mathbf Z)$, $f(\mathbf x) = \frac{1}{2}||\mathbf y - \mathbf C \mathbf x||_2^2$ and $\mathbf C=\mathbf B \otimes \mathbf A$ are as defined before. Thus, the Lipschitz constant of $\nabla F(\mathbf X)$ is the same as $\nabla f(\mathbf x)$, and we use the same notation $L_f$ as used in Algorithm \ref{algo_FISTA_vec}. Consider the following quadratic approximation of $F(\cdot)$ at $\mathbf Z$ for any $\mathbf Z \in \mathrm{dom} G$: \begin{eqnarray} Q_L(\mathbf X, \mathbf Z) &:=& F(\mathbf Z) + \mathrm{tr}(\nabla F(\mathbf Z)^T (\mathbf X - \mathbf Z))\nonumber\\ & +& \frac{L_f}{2} ||\mathbf X - \mathbf Z||_F^2 + \lambda G (\mathbf X)\label{Q_L} \end{eqnarray} where $\mathrm{tr}(\cdot)$ denotes the trace of a matrix. We can rewrite (\ref{Q_L}) as, \begin{eqnarray} Q_L(\mathbf X, \mathbf Z) = \frac{L_f}{2} ||\mathbf X - \mathbf U(\mathbf Z)||_F^2 + \lambda ||\mathbf X||_1 + \tilde F(\mathbf Z)\label{obj_matrix_Fista_mod} \end{eqnarray} where $\tilde F(\mathbf Z)$ is a function of only $\mathbf Z$ and \begin{eqnarray} \mathbf U(\mathbf Z) = \mathbf Z - \frac{1}{L_f}\nabla F(\mathbf Z)=\mathbf Z - \frac{1}{L_f} \mathbf A^T (\mathbf A \mathbf X \mathbf B^T - \mathbf Y)\mathbf B. \end{eqnarray} Thus, \begin{eqnarray*} \underset{\mathbf X}{\arg\min}~ Q_L(\mathbf X, \mathbf Z) = \underset{\mathbf X}{\arg\min} \left\{ \frac{L_f}{2} ||\mathbf X - \mathbf U(\mathbf Z)||_F^2 + \lambda ||\mathbf X||_1 \right\}. \end{eqnarray*} Since both terms are element wise separable, we have \begin{eqnarray} \underset{\mathbf X\in \mathrm{dom} G}{\arg\min} ~ Q(\mathbf X, \mathbf Z)= \mathrm{soft}\left(\mathbf U(\mathbf Z), \frac{\lambda}{L_f}\right)\label{P_l_z} \end{eqnarray} where $\mathrm{soft}\left(\mathbf U(\mathbf Z), \frac{\lambda}{L_f}\right)$ denotes an element wise operation with \begin{eqnarray*} \mathrm{soft}\left(\mathbf W, L_0\right)= \begin{array}{ccc} \mathrm{sgn}(\mathbf W_{ij})(|\mathbf W_{ij}| - L_0)_+ \end{array} \end{eqnarray*} for all indices $i,j$ of the $N\times N$ matrix $\mathbf W$. These steps lead to a generalization of FISTA with matrix inputs, as given in Algorithm \ref{algo_FISTA_matrix}. \begin{algorithm} \textbf{Input:} observation matrix $\mathbf Y$, measurement matrices $\mathbf A$ and $\mathbf B$ \\ \textbf{Output:} estimate for sparse signal matrix, $\hat{\mathbf X}$ \begin{enumerate} \item Initialization: $\mathbf X^{0} =\mathbf 0$, $\mathbf X^{1}=\mathbf 0$, $t_0=1$, $t_1=1$, $k=1$\\ Initialize: $\lambda_1$, $\beta\in (0,1)$, $\bar\lambda > 0$ \item \textbf{while} not converged \textbf{do} \item $\mathbf Z^{k} = \mathbf X^k + \frac{t_{k-1}-1}{t_k} (\mathbf X^k - \mathbf X^{k-1})$ \item $ \mathbf U^k = \mathbf Z^k - \frac{1}{L}\mathbf A^T (\mathbf A \mathbf Z^k\mathbf B^T - \mathbf Y) \mathbf B $ \item $\mathbf X^{k+1} = \mathrm{soft} \left(\mathbf U^k, \frac{\lambda_k}{L_f}\right)$ \item $t_{k+1} = \frac{1+\sqrt{4t_k^2 + 1}}{2}$ \item $\lambda_{k+1} = \max(\beta \lambda_k, \bar\lambda)$ \item $k=k+1$ \item \textbf{end while} \end{enumerate} $\hat{\mathbf X} = \mathbf X^{k} $ \caption{FISTA for sparse matrix recovery with matrix inputs}\label{algo_FISTA_matrix} \end{algorithm} As in Algorithm \ref{algo_FISTA_vec}, the computational complexity is dominated by step 4. The matrix-matrix multiplication at step 4 in Algorithm \ref{algo_FISTA_matrix} is performed with $\mathcal O(N^2(N+M+3L) + NML)$ computations. Since $M,L \leq N$, the worst case complexity is $\mathcal O(N^3)$. Recall, that FISTA in vector form has worst case complexity of $\mathcal O(N^4ML)$. Thus, there is a $\mathcal O(NML)$ gain in the matrix version compared to the vector approach. \subsection{Equivalence of Algorithms \ref{algo_FISTA_vec} and \ref{algo_FISTA_matrix}} It is easy to see that $\mathbf z^k$ and $\mathbf x^{k+1}$ computed in steps 3 and 5 in Algorithm \ref{algo_FISTA_vec} are the same as $\mathrm{vec}(\mathbf Z^k)$ and $\mathrm{vec } (\mathbf X^{k+1})$, respectively, if $\mathbf u^k = \mathrm{vec}(\mathbf U^k)$ where $\mathbf Z_k$, $\mathbf U_k$ and $\mathbf X^{k+1}$ are as computed at steps 3, 4 and 5 in Algorithm \ref{algo_FISTA_matrix} . Now, \begin{eqnarray*} &~&\mathrm{vec}(\mathbf A^T \mathbf A \mathbf Z^k \mathbf B^T \mathbf B - \mathbf A^T \mathbf Y \mathbf B)\nonumber\\ &=&((\mathbf B^T \mathbf B)\otimes (\mathbf A^T \mathbf A)) \mathrm{vec}(\mathbf Z^k) - (\mathbf B^T\otimes \mathbf A^T )\mathrm{vec}(\mathbf Y)\nonumber\\ &=&(\mathbf B^T \otimes \mathbf A^T)(\mathbf B\otimes \mathbf A) \mathrm{vec}(\mathbf Z^k) - (\mathbf B^T\otimes \mathbf A^T )\mathrm{vec}(\mathbf Y)\nonumber\\ &=&(\mathbf C^T \mathbf C)\mathbf z^k -\mathbf C^T \mathbf y, \end{eqnarray*} where $\mathbf C= \mathbf B \otimes \mathbf A$. Thus, $\mathbf u_k$ computed at step 4 in Algorithm \ref{algo_FISTA_vec} is the vectorized version of $\mathbf U^k$ of step 4 in Algorithm \ref{algo_FISTA_matrix}. We conclude that Algorithms \ref{algo_FISTA_vec} and \ref{algo_FISTA_matrix} provide the same output, however, Algorithm \ref{algo_FISTA_matrix} is more efficient. \begin{algorithm} \textbf{Input:} observation matrix $\mathbf Y$, measurement matrices $\mathbf A$ and $\mathbf B$ \\ \textbf{Output:} index set $\Lambda$ containing locations of the non zero indices of the matrix $\mathbf X$, estimate for signal matrix $\hat{\mathbf X}$ \begin{enumerate} \item Initialization: residual $\mathbf R_0 =\mathbf Y$, index set $\Lambda_0=\emptyset$, $t=1$ \item Find the two indices $\lambda_t=[\lambda_t(1)~ \lambda_t(2)]$ such that \begin{eqnarray} [\lambda_t(1)~\lambda_t(2)]& = &\underset{i,j}{\arg\max}~ |\mathbf b_j^T\mathbf R_{t-1}^T \mathbf a_i|\label{eq_OMP_matx} \end{eqnarray} \item Augment index set $\Lambda_t = \Lambda_t \cup \{\lambda_t\}$ \item Find the new signal estimate \begin{eqnarray} \mathbf x_t = \mathbf D_t^{-1} \mathbf d_t\label{x_t} \end{eqnarray} where $\mathbf D_t$ and $\mathbf d_t$ are as in (\ref{hat_t}) \item Compute new residual \begin{eqnarray} \mathbf R_t &=& \mathbf Y - \sum_{m=1}^{t} \mathbf x_t(m) \mathbf a_{\Lambda_t(m,1)} \mathbf b_{\Lambda_t(m,2)}^T\label{residual_M} \end{eqnarray} \item Increment $t$ and return to step $2$ if $t \leq d $, otherwise stop \item Estimated support set $\hat\Lambda=\Lambda_d$ \\ Estimated signal matrix $\hat{\mathbf X}$: $(\Lambda_{d}(m,1), \Lambda_{d}(m,2))$-th component of $\hat{\mathbf X}$ is given by $\mathbf x_{d}(m)$ for $m=1=,\cdots, d$ while rest of the elements are zeros. \end{enumerate} \caption{OMP with matrix inputs}\label{algo_OMP_matrix} \end{algorithm} \section{Sparse Matrix Recovery via OMP} Next, we consider the extension of standard OMP to the matrix form (\ref{obs_1}). We can write the observation $\mathbf Y$ in (\ref{obs_1}) as a summation of $N^2$ matrices as given below: \begin{eqnarray} \mathbf Y =\underset{i,j} {\sum} X_{ij} \mathbf a_i \mathbf b_j^T.\label{obs_sum} \end{eqnarray} When $\mathbf X$ is sparse with $d$ nonzeros, the summation in (\ref{obs_sum}) has only $d$ terms. Let $\Sigma_d$ denote the support of $\mathbf X$ so that $\mathbf X_{ij}$ is non zero for $i,j=1,\cdots, N$, and let $\bar\Sigma_d$ be its complement. We can write (\ref{obs_sum}) as $ \mathbf Y =\underset{(i,j) \in \Sigma_d} {\sum} X_{ij} \mathbf a_i \mathbf b_j^T. $ Our goal is to recover $\mathbf X_{ij}$ for $(i,j)\in \Sigma_d$ in a greedy manner. The proposed OMP version with matrix inputs is given in Algorithm \ref{algo_OMP_matrix}. In Algorithm \ref{algo_OMP_matrix}, $\Lambda_t$ contains estimated $(i,j)$ pairs up to $t$-th iteration in which the $m$-th pair is denoted by $(\Lambda_t(m,1), \Lambda_t(m,2))$ for $m=1,\cdots, t$. Once $\Lambda_t$ is updated as in step 3, the signal is estimated solving the following optimization problem: \begin{eqnarray} \mathbf x_t = \underset{\mathbf x} {\arg\min}\parallel \mathbf Y - \sum_{m=1}^{t} x_m \mathbf a_{\Lambda_t(m,1)} \mathbf b_{\Lambda_t(m,2)}^T ) \parallel_F. \label{eq_frob_norm} \end{eqnarray} The solution of (\ref{eq_frob_norm}) is given by \begin{eqnarray} \mathbf x_t = \mathbf D_t^{-1} \mathbf d_t\label{hat_t} \end{eqnarray} where $\mathbf D_t$ is a $t\times t$ matrix in which the $(m,r)$-th element is given by \begin{eqnarray} (\mathbf D_t)_{m,r} = \mathbf b_{\Lambda_t(r,2)}^T \mathbf b_{\Lambda_t(m,2)} \mathbf a_{\Lambda_t(m,1)}^T \mathbf a_{\Lambda_t(r,1)} \end{eqnarray} for $m,r=1,\cdots, t$ and \begin{eqnarray}\mathbf d_t = [\mathbf b_{\Lambda_t(1,2)}^T \mathbf Y^T \mathbf a_{\Lambda_t(1,1)}~\cdots ~\mathbf b_{\Lambda_t(t,2)}^T \mathbf Y^T \mathbf a_{\Lambda_t(t,1)}]^T \end{eqnarray} is a $t\times 1$ vector. Then the new approximation at the $t$-th iteration is given by \begin{eqnarray} \mathbf Q_t &=& \sum_{m=1}^{t} \mathbf x_t(m) \mathbf a_{\Lambda_t(m,1)} \mathbf b_{\Lambda_t(m,2)}^T\label{Q_t} \end{eqnarray} where $\mathbf x_t(m)$ denotes the $m$-th element of $\mathbf x_t$. Algorithm \ref{algo_OMP_matrix} is a trivial extension of the standard OMP (and was also considered in \cite{Fang2} for $\mathbf A=\mathbf B$). \subsection{Computational complexity } As shown in \cite{Fang2} for $\mathbf A= \mathbf B$, it can be easily verified that Algorithm \ref{algo_OMP_matrix} and the standard OMP \cite{tropp1} with vector inputs (\ref{obs_2}) provide the same performance at each iteration. However, the computational complexity of Algorithm \ref{algo_OMP_matrix} is less than that of its vector counterpart. Step 2 in Algorithm \ref{algo_OMP_matrix} can be implemented as a matrix multiplication $\mathbf A^T \mathbf R_{t-1} \mathbf B$. Thus, the computational complexity of this step is in the order of $\mathcal O(NML+N^2L)$. It is noted that, when implementing the standard OMP as in \cite{tropp1} with vector form (\ref{obs_2}), the equivalent step is computed with complexity of $\mathcal O(N^2ML)$. With respect to step 4 in Algorithm \ref{algo_OMP_matrix}, the matrix $\mathbf D_t$ requires $\mathcal O(t^2(M+L) )$ computations at the $t$-the iteration. The vector $\mathbf d_t$ requires $\mathcal O(t(ML+M))$ computations. Worst case complexity of the inverse operation is $\mathcal O(t^3)$. Matrix-vector multiplication in (\ref{x_t}) requires $\mathcal O(t^2)$ computations. Thus, at a given iteration, worst case complexity of step 4 in Algorithm \ref{algo_OMP_matrix} is in the order of $\mathcal O(tML)$. It can be shown that the worst case computational complexity of the equivalent step of standard OMP with Kronecker products to estimate the signal at $t$-th iteration, is in the order of $\mathcal O(t^2 ML)$. Thus, steps 2 and 4 in Algorithm \ref{algo_OMP_matrix} provide us with a computational gain over the equivalent steps of the standard OMP with Kronecker products. Therefore, we conclude that Algorithm \ref{algo_OMP_matrix} is an efficient way to find sparse $\mathbf X$ from (\ref{obs_1}) compared to its vector counterpart (\ref{obs_2}) although both provide the same performance. It is further observed that this computational gain is not as significant as with FISTA. \section{Numerical Results} In this section, we demonstrate the capability of recovering sparse $\mathbf X$ from observation model (\ref{obs_1}) via different algorithms and provide insights into the computational gains achievable with the matrix version. First, we illustrate the performance of FISTA with different choices for $\mathbf A$ and $\mathbf B$. For numerical results, we assume that $\mathbf X$ is a distributed sparse matrix in which each column has a maximum of $K$ nonzeros and the locations are generated uniformly. The values of nonzero entries are drawn from a uniform distribution in the range $ [-250, ~-200] \cup[200, ~250]$. We consider that the observation matrix $\mathbf Y$ in (\ref{obs_1}) is corrupted by additive noise and the elements of the noise matrix are assumed to be independently and identically distributed Gaussian random variables with mean zero and variance $\sigma_v^2$ \begin{figure}[htb] \centering \centerline{\includegraphics[width=9.0cm]{K2_DCT_Matrix_error}} \caption{Normalized reconstruction error vs $M=L$ with FISTA and different projection matrices, $N=40$, $\sigma_v^2=0.01$} \label{fig_1} \end{figure} In Fig. \ref{fig_1}, we plot the normalized reconstruction error $\frac{||\mathbf X - \hat{\mathbf X}||_F}{||\mathbf X||_F}$ vs $M$ when $M=L$ averaging over $500$ trials. We let $N=40$, and $\sigma_v^2=0.01$. In Fig. \ref{fig_1}, we illustrate two aspects. First, for given $K$, different types of matrices $\mathbf A$ and $\mathbf B$ are examined. We consider independent random rows of the $N\times N$ DCT matrix, Gaussian, and binary matrices. In the case of a Gaussian, elements are drawn from a normal ensemble and then orthogonalized. By binary matrix, we mean that the elements of the matrix can take values $1/N$ or $0$ with equal probability. Note that, random rows of DCT matrix and Gaussian matrix obey uniform uncertainty principle with good isometry constants in contrast to a binary matrix. When both matrices $\mathbf A$ and $\mathbf B$ are either random rows of DCT matrix or zero mean Gaussian, the recovery of the sparse matrix is guaranteed with less measurements compared to $N$. When $\mathbf A$ and $\mathbf B$ are binary, the recovery is not so good, which is intuitive since binary matrices are not "good" compressive sensing matrices. However, when $\mathbf A$ is binary and $\mathbf B$ is Gaussian, we see an improved performance compared to the case where both are binary. This provides an insight that even when one matrix does not obey uniform uncertainty principle with good isometry constant, still the sparse matrix can be recovered reliably when the other matrix is a "good compressive sensing" matrix. We will further investigate this observation in our future work. Second, with a given type of matrices (in the case of $\mathbf A$, $\mathbf B$ are Gaussian) we vary $K$. It is seen that, recovery capability of FISTA does not degrade significantly as $K$ increases. \begin{figure}[htb] \centering \centerline{\includegraphics[width=9.0cm]{DCT_Matrix_error_ML_new}} \caption{Normalized reconstruction error vs $L$ for given $M$ with FISTA, $\mathbf A$ and $\mathbf B$ contain independent random rows of the $N\times N$ DCT matrix, $N=40$, $\sigma_v^2=0.01$} \label{fig_01} \end{figure} In Fig. \ref{fig_01}, we plot the reconstruction error vs $L$ with FISTA keeping $M$ fixed. As a benchmark, the curve corresponding to $M=L$ is also plotted. It can be seen that, when one dimension of the observation matrix is fixed, an improved performance in terms of signal reconstruction error is observed as the other dimension increases. However, when $M$ is very small (or below a certain value), complete recovery is not guaranteed even if $L=N$. This implies that when the dimension of the matrix $\mathbf A$ is fixed, increasing the number of columns of $\mathbf B^T$ does not necessarily guarantee complete recovery when $M$ is very small. \begin{table} \caption{Runtime (in $s$) of FISTA with vector and matrix inputs} \centering \begin{small} \begin{tabular}{|l|l|l|l|} \hline $~$ & N=20 & N=40 & N=60 \\ \hline Matrix & $0.3863$ & $1.3064$ & $6.5595$ \\ Vector & $1.0987$ & $26.4162$ & $142.7584$ \\ \hline \end{tabular} \label{table_comparison} \end{small} \end{table} In Table \ref{table_comparison}, we compare the average runtime with MATLAB (in Intel(R) Core(TM) i7-3770 CPU$@$ 3.40GHzz processor with 12 GB RAM) for FISTA for matrix and vector versions as the sparse matrix dimension $N$ varies given that the number of iterations in both Algorithms \ref{algo_FISTA_vec} and \ref{algo_FISTA_matrix} is fixed at the same value ($=10000$). We let $K=N/20$ and $M=L=N/2$. Matrices $\mathbf A$ and $\mathbf B$ are assumed to be Gaussian. It reflects the computational efficiency of the matrix approach compared to the vector approach especially as $N$ increases although both algorithms provide the same performance. To illustrate the performance of OMP with matrix inputs, we plot the fraction of the support correctly recovered with Algorithm \ref{algo_OMP_matrix} for different choices for $\mathbf A$ and $\mathbf B$ with $K=1$ in Fig. \ref{fig_3}. From Fig. \ref{fig_3}, it is again observed that, although not with the same scale as with FISTA, the recovery capability can be improved when one projection matrix is binary and the other is Gaussian compared to the case where both $\mathbf A$ and $\mathbf B$ are binary. Another observation is that even with a Gaussian matrix, as $K$ increases the performance of OMP degrades significantly leaving OMP not a better choice when the sparsity level increases. \begin{figure}[htb] \centering \centerline{\includegraphics[width=9.0cm]{OMP_2D_spt_A_B}} \caption{Fraction of the support correctly recovered vs $M=L$ with OMP with different projection matrices with no noise} \label{fig_3} \end{figure} \section{Discussion} In this paper, we showed numerically that recovering $\mathbf X$ based on (\ref{obs_1}) in its matrix form is more computationally efficient than solving it after converting to vector form via Kronecker products when $\mathbf X$ is arbitrarily distributed sparse. We developed matrix versions of FISTA to solve $l_1$ norm minimization in (\ref{matrix_l1}) efficiently and OMP to solve for $\mathbf X$ in a greedy manner. It has been shown that a significant computational gain is achieved by FISTA with matrix form compared to its vector counterpart. We further illustrated the recovery capability with different choices for projection operators. The results provide insight into the following. If a linear system of the form (\ref{obs_2}) can be converted into a matrix form as in (\ref{obs_1}), the problem can be solved more efficiently without losing performance with respect to the original vector form. Thus, it is worth investigating such scenarios where the matrix approach can be efficiently used to solve linear systems which are computationally demanding otherwise. \newpage \bibliographystyle{IEEEtran}
\section{Introduction} Phase space gradient of thermodynamic quantities in a nonequilibrium process and those limitations by information theoretic ones should bring fundamental insights into the understandings of the system. Among others, dissipation and work are central quantities in thermodynamic operation between equilibrium states. The connecting path between the two distinct equilibrium states in phase space has a diversified range depending on the steps of the procedure and it determines the amount of the mechanical work needed to perform the process. It is well recognized that as long as the process is not quasistatic, the free energy at the end of the process becomes less than that of the initial state plus the work invested from outside \cite{Landau}. In other words, there is dissipative loss of work into the surroundings of the system -- the second law of thermodynamics.\\ The dissipated work is thus an indicator of the excess of the injected work into the system, and it is defined in terms of the difference in the equilibrium free energy $\Delta F$ as $W_{diss}(\Gamma,\lambda)= W(\Gamma,\lambda)-\Delta F$, where $\lambda$ is a protocol parameter. $W(\Gamma,\lambda)$ denotes the average work done on the system by the external operator as perturbation and it is a function of the position in phase space. The dissipated work also represents the total change in entropy of the system as a result of the operated transition. The free energy difference between terminal states is also called the reversible work. Instead of taking one instance in the definition, we consider a statistical ensemble of realizations of the process as the consequence of infinitely many repetitions of this process. Therefore, we refer to the dissipated work in the sense of mean $\langle W_{diss}(\Gamma,\lambda)\rangle$ throughout this paper, but for simplicity we omit the angular brackets in the following.\\ Any displacement in phase space from a specific equilibrium state to another equilibrium phase point induced by a dynamics has a counterpart process by reversing time. The to-and-fro movement makes us to expect finding a universal relation in the dissipated work in an averaged way. One such intriguing example was discovered as a relation \cite{JE2,Kawai} \begin{eqnarray} \frac{\langle W_{diss} \rangle}{k T}= D_{KL}(\mathcal{P}_F\|\mathcal{P}_B)\label{eqn:Kawai}, \end{eqnarray} under a sequence of procedure where the system is initially at canonical equilibrium with temperature $T$, and then it is detached from the heat reservoir to let the system evolve according to the Liouville equation and lastly equilibrate again to a new canonical state determined by the protocol parameter by attaching the same bath at the end of the process. The equality is replaced by inequality in case the system is kept contacting with the bath \cite{Kawai}. $\mathcal{P}_F$ and $\mathcal{P}_B$ are phase densities of the forward and the time-reversed processes at a particular time, respectively. Whether or not the prompt thermalization for each instance at any attained phase space point is realized remains a matter of careful thought, however, it is a standard device for consideration commonly used in the literature. The relation gives a meaning of a degree of time-reversal asymmetry measured by the Kullback-Leibler (KL) relative entropy $D_{KL}(\mathcal{P}_F\|\mathcal{P}_B)$ from $\mathcal{P}_F$ and $\mathcal{P}_B$ \cite{KL,KL2}. The left-hand side of Eq.(\ref{eqn:Kawai}) is a physical content in units of thermal energy with the Boltzmann constant $k$ and, on the other hand, the right-hand side represents purely information theoretic quantity.\\ In this paper, based on the above consideration, we see how the gradient of the dissipated work in phase space is inextricably linked to an information theoretic distance between phase density functions of forward and backward processes. Specifically, we show that a novel constraint on the averaged square of the gradient of the dissipated work $\langle |\nabla W_{diss}|^2\rangle$ can be obtained via the relative Fisher information and the logarithmic Sobolev inequality. We further present a possible inequality expression indicating a deep connection in terms of two kinds of information quantities. These studies are in conformity with a modern approach that bases the construction of statistical thermodynamics upon the concept of information (e.g., \cite{Ben}).\\ In Sec.II, we first present a general relation corresponding to Eq.(\ref{eqn:Kawai}) when we employ an alternative to the KL relative entropy in order to recognize a distance notion in more general context in terms of information measure. The main ingredient of our consideration is introduced in Sec.III. We then show a constraint of the obtained relation with the inequality in Sec.IV. We give a concluding summary in Sec. V. \section{A general relation for dissipated work} Let $M^n$ be the $n$-dimensional configuration space of the present system. The two participating distributions $\mathcal{P}_F$ and $\mathcal{P}_B$ are compared at equal-time when calculating the general relative entropy, which is specified by a convex function $\chi(\mathcal{P}_F/\mathcal{P}_B)$ with $\chi(1)=0$, \begin{eqnarray} D_{G}(\mathcal{P}_F\|\mathcal{P}_B):=\int \mathcal{P}_F \chi\left(\frac{\mathcal{P}_F}{\mathcal{P}_B}\right)d\Gamma.\label{eqn:DG} \end{eqnarray} This form of the general relative entropy is first introduced in \cite{Csi,Morimoto} and it is now well recognized as the Csisz{\'a}r-Morimoto divergence (less known in physics, though). It was employed in the proof of the H-theorem for Markov processes \cite{Morimoto}. In general, we do not require the asymmetry in the arguments but assume positivity. Let $T^*M^n$ be the cotangent bundle to $M^n$, i.e., phase space. The system whose Hamiltonian is parameter dependent such as $H_{\lambda}(\Gamma_\lambda)$ starts its evolution from a phase space point $\Gamma_0\in T^*M^n$ at time $t_0$ and reaches another point $\Gamma_1\in T^*M^n$ at $t_1$. During this interval of time, the protocol parameter changes from $\lambda(t_0)=\lambda_0$ to $\lambda(t_1)=\lambda_1$. Initially, the system is put in the canonical equilibrium form $\mathcal{P}_F(\Gamma_0)=e^{-\beta H_{\lambda_0}(\Gamma_0)}/Z_0$ with $\beta=(k T)^{-1}$. Similarly, prior to the backward process the system is supposed to have a density function $\mathcal{P}_B(\Gamma_1)=e^{-\beta H_{\lambda_1}(\Gamma_1)}/Z_1$ by preparing the canonical form at the start of the return. The reverse protocol that goes from $\lambda_0$ to $\lambda_1$ corresponds to $\Gamma_1$ and $\Gamma_0$ in the phase space. The free energy difference between the initial and the final states is given by $\Delta F=-\beta^{-1}(\ln Z_1-\ln Z_0)$ with the partition functions $Z_0$ and $Z_1$ for each state. We note that the protocol path in $T^*M^n$ joining $\Gamma_0$ to $\Gamma_1$ that represents the operation does not necessarily proceed along the geodesic associated with it. The geodesic on the surface of the phase space manifold is a significant locus, however, the protocol given by an external agent or by an arbitrary schedule does not always go through the shortest path. Moreover, even if an operation for one realization takes the geodesic, other realizations do not trace exactly the same route at every time. They fluctuate around the geodesic path. That is why the work performed on the system is statistically distributed representing the ensemble of realizations and it is averaged over the ensemble. \\ Let $C$ be a curve (path) on a surface of $M^n$ connecting $\Gamma_0$ and $\Gamma_1$. Then, the amount of work the system receives on completion of the process depends not only on the location of the two terminal points $\Gamma_0$ and $\Gamma_1$, but also on the route $C$. It should be precisely denoted as $W^C(\Gamma_0,\Gamma_1)$ and equals to the difference in the Hamiltonian $H_{\lambda_1}(\Gamma_1) - H_{\lambda_0}(\Gamma_0)=\int^{\lambda_1}_{\lambda_0}d\lambda \partial H_\lambda(\Gamma_\lambda)/\partial \lambda$ (the first law of thermodynamics). For simplicity, we omit both the subscript and arguments, and we denote $W$ in the following.\\ We note that for the well-defined distance $D_G$, the relative density $\mathcal{P}_F/\mathcal{P}_B$ must have a finite value on a support (that is, $\mathcal{P}_F$ is absolutely continuous with respect to $\mathcal{P}_B$), otherwise it is defined as infinity. Then, the relative phase density becomes \begin{eqnarray} \frac{\mathcal{P}_F}{\mathcal{P}_B}=\frac{Z_1}{Z_0} e^{-\beta\left(H_{\lambda_0}(\Gamma_0) - H_{\lambda_1}(\Gamma_1)\right)} = e^{\beta (W-\Delta F)}.\label{eqn:reld} \end{eqnarray} Inserting this into Eq.(\ref{eqn:DG}), we obtain the general reformulation of Eq.(\ref{eqn:Kawai}) as the following form \begin{eqnarray} D_{G}(\mathcal{P}_F\|\mathcal{P}_B)=\Big\langle\chi(e^{\beta(W-\Delta F)}) \Big\rangle_{\mathcal{P}_F}, \end{eqnarray} where $\langle \cdot \rangle_{\mathcal{P}_F}$ denotes the average with respect to the distribution $\mathcal{P}_F$. This is the most general expression relating the dissipated work and the distance measure in this setting. We remark that the choice $\chi(x)=\ln x$ in Eq.(\ref{eqn:DG}) with $x=\mathcal{P}_F/\mathcal{P}_B$ defines the KL distance and recovers exactly the result Eq.(\ref{eqn:Kawai}). It is also worth mentioning another example. If we choose $\chi(x)=1/\sqrt{x}$, we have the overlap distance between $\mathcal{P}_F$ and $\mathcal{P}_B$, i.e., $D_O(\mathcal{P}_F\|\mathcal{P}_B):=\int \sqrt{\mathcal{P}_F\mathcal{P}_B} d\Gamma = \langle \sqrt{\mathcal{P}_F/\mathcal{P}_B}\rangle_{\mathcal{P}_F}$. If two phase densities are identical, it gives unity. However, $\mathcal{P}_F$ and $\mathcal{P}_B$ can never be identical ($\mathcal{P}_F\neq \mathcal{P}_B$) in our consideration. Substituting Eq.(\ref{eqn:reld}) stemming from the canonical forms into the definition of the overlap distance, we have \begin{eqnarray} D_O(\mathcal{P}_F\|\mathcal{P}_B)&=& \Big\langle e^{-\frac{\beta}{2}(W-\Delta F)} \Big\rangle_{\mathcal{P}_F}\nonumber\\ &=& e^{\frac{\beta}{2}\Delta F}\Big\langle e^{-\frac{\beta}{2}W}\Big\rangle_{\mathcal{P}_F}\nonumber\\ &\stackrel{\text{J.E.}}{=}& 0. \label{eqn:DO} \end{eqnarray} The second line follows from the fact that the $\Delta F$ is not a statistically distributed quantity. The last line is due to the nonequilibrium work relation called the Jarzynski equality (J.E.) $\langle e^{-\beta W}\rangle=e^{-\beta \Delta F}$ \cite{JE1,Crooks}, where the angular brackets denotes an average over a statistical ensemble of realizations of a single process through phase space. It is known that for general circumstances including the Hamilton's evolution, this average is equivalent to taking an average using the initial equilibrium distribution designated by $\lambda_0$ \cite{JE3}. Furthermore, the J.E. can be potentially useful in various equilibrium statistical physics, in which the estimation of the free energy differences is difficult to obtain. In these cases, such an equilibrium property is obtained from observations (experiments) of work distributions performed on the system. Note that distinct distributions in phase space do not share the common domain, so that Eq.(\ref{eqn:DO}) means that the vanishing overlap between the two equilibrium states is consistent with the fact that the J.E. holds. On the other hand, the J.E. is irrelevant to the derivation of Eq.(\ref{eqn:Kawai}), i.e. the case of the KL distance. We also remark that when the system realizes $W=\Delta F$ (reversible process), the overlap $D_o$ vanishes, meaning that it can be a measure of how much the nonequilibrium behavior deviates from the reversibility. \section{Relative Fisher information and dissipated work} In Section II, we recognized that Eq.(\ref{eqn:Kawai}) is just one realization of the information theoretic expression of the averaged dissipated work among other possibilities. The choice of $\chi$ is basically at our disposal and there is no legitimate criterion in terms of information theory. For operational purposes such as statistical inferences and estimation of measurement values, relative entropy with parameters can be considered advantageous over the one without them. Such a generalized distance measure with keeping the original properties of the Kullback-Leibler has been proposed \cite{TY2} (see also e.g., \cite{Special}). But we do not delve into such possibilities including the overlap distance in this paper. Instead, to advance the understanding in line with previous works, we hereafter set the gauge of the distance measure as KL in our consideration. As we pronounced in Introduction, we are concerned with the quantity $\nabla_{\Gamma} W_{diss}$, where $\nabla_{\Gamma}$ indicates the gradient in phase space, because the dissipated work depends on the location in phase space and its gradient reflects a local structure (geometry) of the dissipation occurred as an excess of work performed on the system. \\ The relative Fisher information from $f$ to $g$ is defined as (e.g. \cite{Villani}, p.278) \begin{eqnarray} D_{\rm RFI}(f\|g):=\int f \Big| \nabla_\Gamma \left(\ln \frac{f}{g}\right)\Big|^2 d\Gamma. \label{eqn:rFi} \end{eqnarray} It is non-negative and achieves zero iff $f=g$. Further, it is asymmetric, i.e., $D_{\rm RFI}(f\|g)\neq D_{\rm RFI}(g\|f)$, meaning that it is {\it directed} as is the case of the KL distance. This form can be defined independently of whether the distributions possess estimation parameters or not. We can also employ it when the two distributions are parametrized by the same family (say, $\boldsymbol{\theta}$). We can say that this metric reflects a local comparison of the two distributions in that we take the derivative of them, whereas the KL distance returns a coarse-grained quantity by gathering each contribution of the displacement (or difference) between distributions. The same remark mentioned before Eq.(\ref{eqn:reld}) also applies to this relative distance: the relative phase density must have finite value on the domain of phase space. To the best of our knowledge, there is no study to give a physical connection to that measure. Now, we show one of the applications below. That is, substituting Eq.(\ref{eqn:reld}) into the above definition, we have a relation (see Appendix \ref{app:rfi}) \begin{eqnarray} D_{\rm RFI}(\mathcal{P}_F\|\mathcal{P}_B)=\beta^2\Big\langle \Big| \nabla_\Gamma W_{diss}(\Gamma)\Big|^2 \Big\rangle_{\mathcal{P}_F}\label{eqn:reF1}. \end{eqnarray} This identity articulates that an average of the square of the gradient of the dissipated work generated by the transient process, taken over a forward equilibrium state, can be equated with the distance between the forward and backward distributions in the phase space measured by the relative Fisher information.\\ Our setting of formulation represented by an operational parameter is in a position of Hamilton equations (differential equations on symplectic manifolds). Then, in order to consider nonequilibrium processes in phase space perspective, it is principally necessary to deal with the trajectory on a manifold (either on symplectic or on Riemannian). It is known that paracompact $C^\infty$-manifolds always have their Riemannian metrics. Since the usual manifold is paracompact, the Riemannian metric is introduced if we regard the phase space as a manifold. Then, the metric expression of the above identity is of concern. To this end, we remark that a gradient vector of a differentiable function $f$ on the Riemannian manifold associated with the covector $df$ can be defined as the contravariant vector, which is given in coordinate as $(\nabla f)^i=\sum_j g^{ij}\partial f/\partial x^j$, where $x^i$ represents the generalized coordinates and the contravariant metric tensor $g^{ij}$ is the inverse of the metric tensor $g_{ij}$ ($(g^{ij})=(g_{ij})^{-1}$). We further recall that, for any vector ${\bf a}$, the differential of a function $df$ is defined by the derivative of $f$ along ${\bf a}$. Therefore, the gradient vector $\nabla f$ satisfies $df({\bf a})=\langle \nabla f, {\bf a} \rangle=\sum_i (\partial f/\partial x^i) a^i$. Then, substituting $(\nabla f)^i$ into this, we find that the square of the modulus of the gradient $f$ can be written as $|\nabla f|^2=df(\nabla f)=\sum_{ij}(\partial_{x^i}f)g^{ij}(\partial_{x^j}f)$. Since we now take $f=W_{diss}(\Gamma)$ in the present consideration, therefore, the right-hand side of Eq.(\ref{eqn:reF1}) can be expressed as \begin{eqnarray} \beta^2 \Big\langle \sum_{i j}\frac{\partial W_{diss}(\Gamma)}{\partial x^i}g^{ij} \frac{\partial W_{diss}(\Gamma)}{\partial x^j} \Big\rangle_{\mathcal{P}_F}. \end{eqnarray} We can obtain a more direct physical meaning for the quantity of the Dirichlet form $\langle | \nabla_\Gamma W_{diss}(\Gamma)|^2 \rangle_{\mathcal{P}_F}$ appearing in Eq.(\ref{eqn:reF1}) via the gradient flow interpretation as we will see shortly. To this end, we recall again what $W$ exactly expresses. It means that the work done on the system when it starts from $\Gamma_0$ and reaches $\Gamma_1$ by changing the control parameter $\lambda\in [\lambda_0,\lambda_1]$ along a path $C$ on the surface of the phase space manifold. Since $\lambda$ designates the instance of the evolution from any starting point $\Gamma_0$, we rewrite it as $W_\lambda(\Gamma)$ by denoting $\Gamma_0$ as $\Gamma$ on a specific $C$. Accordingly, the dissipated work up to the intermediate value of $\lambda$ can be denoted as $W_{diss}(\Gamma_\lambda)$, where the phase point $\Gamma_\lambda\in T^*M^n$ should be read as $\hat{T}_\lambda\Gamma$ with an evolution operator $\hat{T}_\lambda$ that represents the protocol. Recall now that the gradient flow associated with the velocity vector field $-\nabla(\delta G/\delta \rho):=v(\Gamma_\lambda)$ is defined (Appendix \ref{app:gradf}) as \begin{eqnarray} \frac{\partial \rho}{\partial t}={\rm div} \left(\rho \nabla\frac{\delta G}{\delta \rho}\right),\label{eqn:gradf} \end{eqnarray} where $G(\rho)$ is an energy functional. In our case, we can take it as \begin{eqnarray} G(\rho)=\int W_{diss}(\Gamma_\lambda) \rho(\Gamma)d\Gamma. \end{eqnarray} With abuse of notation above, we have used $\rho(\Gamma)$ as the probability density function in phase space at each time instant $t$. This is the phase space average of the dissipated work in our consideration and the first variation of it is $\delta G/\delta \rho= W_{diss}(\Gamma_\lambda)$. Therefore, the average kinetic energy $K_\lambda$ accompanied to the dissipation up to the protocol $\lambda$ is found to be just the squared mean of the gradient of the dissipated work evaluated by the initial equilibrium state $\mathcal{P}_F$ that is prepared for the start of the process: \begin{eqnarray} K_\lambda=\int |v(\Gamma_\lambda)|^2\mathcal{P}_F(\Gamma)d\Gamma = \Big\langle | \nabla_\Gamma W_{diss}(\Gamma)|^2 \Big\rangle_{\mathcal{P}_F}. \end{eqnarray} In the whole process, the total kinetic energy becomes $\int^{\lambda_1}_{\lambda_0}K_\lambda d\lambda$. \section{A lower bound for average dissipated-work gradient} In this section, we show that our primarily focused quantity can be bounded by the KL distance via the logarithmic Sobolev inequality (LSI) \cite{Gross}. In this sense, the identity that connects the relative Fisher information and the quantity $\langle | \nabla_\Gamma W_{diss}(\Gamma)|^2 \rangle_{\mathcal{P}_F}$ derived in the previous section can be regarded as an intermediate result. The LSI has wide range of applications and it takes several mathematically equivalent forms \cite{Villani}. In this paper, we employ the form (Lemma 6.1 in \cite{Gross}) \begin{eqnarray} \int |f|^2 \ln |f|d\mu \leqslant c \int |\nabla f|^2d\mu + \|f\|_2^2\ln \|f\|_2^2,\label{eqn:LSI} \end{eqnarray} which holds for all functions $f$, whose gradient $\nabla f$ and $f$ itself are square integrable in the domain. $\|f\|_2$ denotes $(\int f^2 d\mu )^{1/2}$ and the constant $c>0$ independent of $f$. The physical interpretation of this constant is not obvious in general circumstances and would depend on the physical model. However, when an equilibrium state satisfies the LSI and the probability density function has the corresponding gradient flow, the constant has a clear meaning of how fast the system equilibrates with it. The corresponding relaxation rate takes the exponential form and it appears as a pre-factor to the KL distance between the initial and the equilibrium density functions. This feature is a direct consequence of the result provided in \cite{Barron} (see also e.g., \cite{Villani}, p.288 and Appendix \ref{app:LSI}). If we substitute $d\mu=\mathcal{P}_B(\Gamma)d\Gamma$ for the probability measure and choosing $f=\sqrt{\mathcal{P}_F/\mathcal{P}_B}$ then multiplying both sides by $2$, we find \begin{eqnarray} \int \mathcal{P}_F \ln \frac{\mathcal{P}_F}{\mathcal{P}_B}d\Gamma & \leqslant & 2c\int \Big| \nabla \sqrt{\frac{\mathcal{P}_F}{\mathcal{P}_B}}\Big|^2 \mathcal{P}_B d\Gamma \nonumber\\ & = & \frac{c}{2} \int \mathcal{P}_F \Big| \nabla \left(\ln \frac{\mathcal{P}_F} {\mathcal{P}_B}\right)\Big|^2d\Gamma.\label{eqn:LSI2} \end{eqnarray} In what follows, we set $c=1$ without impeding our consideration. Combining Eq.(\ref{eqn:reF1}) and Eq.(\ref{eqn:LSI2}), we find that the mean of the gradient dissipated work is lower bounded by the information theoretic distance between forward and backward phase densities : \begin{eqnarray} \langle | \nabla_\Gamma W_{diss}(\Gamma)|^2 \rangle_{\mathcal{P}_F} \geqslant 2(kT)^2 D_{KL}(\mathcal{P}_F \| \mathcal{P}_B).\label{eqn:NW} \end{eqnarray} We further pursue a reformulation of thus obtained constraint from an information point of view, that is, in terms of Fisher information in statistics. \\ We begin with recalling the followings. For a family of probability distributions $\rho_{\boldsymbol{\theta}}(\boldsymbol{x})$ parametrized by $\boldsymbol{\theta}$, the Fisher information in estimation theory is defined as $I(\boldsymbol{\theta}):=\langle (\nabla_{\boldsymbol{\theta}} \rho/\rho)^2\rangle_{{\rho}_ {\boldsymbol{\theta}}(\boldsymbol{x})}$. In the same way, it is defined also as $I(\rho(\boldsymbol{x})):=\langle (\nabla_{\boldsymbol{x}} \rho/\rho)^2\rangle_{{\rho}(\boldsymbol{x})}$ for a differentiable distribution $\rho(\boldsymbol{x})$, and it measures how much two neighboring density functions are statistically distinguishable \cite{Cover}. The former form is invariant against any shift $\boldsymbol{\theta}$, which signifies the diagonal entries of the Fisher information matrix, thereby implying i.i.d. data. Indeed, when we choose as $\rho_{\boldsymbol{\theta}}(\boldsymbol{x})=\rho(\boldsymbol{x}-\boldsymbol{\theta})$, we easily find that due to $\nabla_{\boldsymbol{x}}\rho(\boldsymbol{x}-\boldsymbol{\theta})=-\nabla_{\boldsymbol{\theta}} \rho(\boldsymbol{x}-\boldsymbol{\theta})$, the Fisher information does not depend on $\boldsymbol{\theta}$ and $I(\boldsymbol{\theta})=I(\rho(\boldsymbol{x}))$ follows.\\ There have been comprehensive efforts to understand various physical laws in terms of this information \cite{Frieden}. It also plays a crucial role to upper bound the entropy production (e.g. \cite{TY1} and references therein). Since the present canonical equilibrium distribution specified by a protocol parameter satisfies $\nabla\mathcal{P}(\Gamma)=-\beta H(\Gamma)\mathcal{P}(\Gamma)$, it becomes \begin{eqnarray} I(\mathcal{P}) = \Big\langle \left(\frac{\nabla \mathcal{P}}{\mathcal{P}}\right)^2\Big\rangle_{\mathcal{P}} = \int \frac{\nabla \mathcal{P}}{\mathcal{P}} (-\beta \nabla H(\Gamma)) \mathcal{P}d\Gamma.\nonumber\\ \end{eqnarray} Integrating by parts, the right-hand side reduces further to \begin{eqnarray} -\beta \int \nabla H(\Gamma) \nabla \mathcal{P}(\Gamma)d\Gamma =\beta \langle \Delta H(\Gamma) \rangle_{\mathcal{P}},\nonumber \end{eqnarray} where we have assumed that the phase density vanishes at the boundary ($\partial S$) of the system, so that $[\mathcal{P}\nabla H(\Gamma)]_{\partial S}=0$, and $\Delta$ denotes the Laplacian in phase space. Therefore, we obtain a striking relation \begin{eqnarray} I(\mathcal{P}) &=& \beta \langle \Delta H(\Gamma) \rangle_{\mathcal{P}},\label{eqn:Ftmp} \end{eqnarray} which tells that the inverse temperature of the system can be intimately linked with the information quantity (Upon revision of this manuscript, the author became aware that this relation has derived also in Ref. \cite{Narayanan}.). In general coordinates, we note that $\Delta H(\Gamma)$ can be expressed as $(\sqrt{g})^{-1}\partial_i(\sqrt{g}g^{ij}\partial_jH(\Gamma))$, where $g$ is the determinant of the metric $g=det(g_{ij})$. A way to interpret this relation is that the temperature of the system can be defined by the Fisher information of the system and by the averaged second order differential (curvature) of energy \cite{Ffl}. As we shall use below, this relation is critical for the present study. Since the system contacts with the same heat bath (with common $\beta$) both at the start and the end of the process as described in Sec. II (or we could also restate it as follows; the heat reservoir is so large compared with the system, so that temperature of the reservoir is not disturbed by the heat discarded by the system), we have readily a relation from Eq.(\ref{eqn:Ftmp}) \begin{eqnarray} \frac{I(\mathcal{P}_F)}{\langle \Delta H(\Gamma) \rangle_{\mathcal{P}_F}}= \frac{I(\mathcal{P}_B)}{\langle \Delta H(\Gamma) \rangle_{\mathcal{P}_B}}= \beta.\label{eqn:Iofb} \end{eqnarray} An immediate but profound implication of this consequence is that the ratio of the Fisher information associated with probability distributions of the forward and backward processes is equivalent to the ratio of the averaged curvature of the Hamiltonian. This is true if the system takes the canonical form in distribution. The former ratio has the origin of the information quantity and the latter has the physical one. Next, substituting Eq.(\ref{eqn:Iofb}) into Eq.(\ref{eqn:NW}), we readily have an inequality \begin{eqnarray} \frac{\langle | \nabla_\Gamma W_{diss}(\Gamma)|^2 \rangle_{\mathcal{P}_F}} {\langle \Delta H(\Gamma) \rangle_{\mathcal{P}_{\gamma}}^2} \geqslant 2\frac{D_{KL}(\mathcal{P}_F \| \mathcal{P}_B)}{[I(\mathcal{P}_\gamma)]^2}.\label{eqn:last} \end{eqnarray} where the symbol $\gamma$ denotes either $F$ or $B$ representing the forward and the backward equilibrium states. The lower bound on the ratio relevant to the averaged physical quantities (the left-hand side) is nicely bounded from below in terms of the ratio of information-associated quantities only. A further interesting observation can be derived for this relation Eq.(\ref{eqn:last}) from the well-known Cramer-Rao inequality in statistical estimation theory \cite{Cover}. For simplicity's sake, we consider it for one-dimensional case. The Cramer-Rao inequality tells a tradeoff relation between Fisher information of a distribution $P(X)$ and the variance of the distribution $\sigma^2_X$, i.e., $I(P(X))\geqslant 1/\sigma^2_X$, where $X$ is a random variable. Then, we find that the lower bound in Eq.(\ref{eqn:last}) is upper bounded by $2D_{KL}(\mathcal{P}_F \| \mathcal{P}_B) \sigma^4_X$. Now that $P(X)$ is of the canonical form in our setting, the equality can be achieved when the Hamiltonian $H(X)$ is of quadratic form. \section{Summary} To deepen the understandings of a profound information theoretic relation between work and dissipation in nonequilibrium systems, we have derived a universal relation that connect the gradient of the dissipated work and the relative Fisher information within a framework of the setup repeatedly employed in the previous studies. Considering the gradient of the dissipated work at each point in phase space enables us to get geometric information that cannot be obtained from KL entropy only. The relative Fisher information plays the role. By way of this, we have established the information based lower bound for the quantity relevant to the dissipation in phase space. The instantaneous equilibration was a premise to assure the well-posedness of the free energy difference between the two canonical equilibrium states associated to the forward and backward processes. However, the notion of the nonequilibrium free energy change $\Delta F_{neq}$ has recently considered to refine the dissipation occurring in far from thermodynamic equilibrium, where the dissipated work is defined by average work minus $\Delta F_{neq}$ instead of the equilibrium free energy change $\Delta F$ \cite{Fneq}. An extension of the present result to such a case surely needed if one is to understand and to gain deeper insights into biological systems. \begin{acknowledgments} The author wishes to thank Hiroaki Yoshida for a valuable discussion on the relative Fisher information at the Ochanomizu University in August 2012. \end{acknowledgments}
\section{INTRODUCTION} The question of whether and how the morphology of galaxies evolves over time is one of the most intensely debated questions in astrophysics. Understanding this secular evolution requires a detailed theory for the origin and fate of the various components found in the central regions of galaxies, such as a super-massive black hole (SMBH), a compact nuclear star cluster, stellar bar, star-forming rings, and a (pseudo)bulge. Despite numerous theoretical studies \citep[see][and references therein]{hopkins06} and large-scale observational programs \citep[e.g.][]{sings,dale04,sauron,dr7,atlas3d}, a complete understanding of how these various features form, evolve, or influence each other, remains elusive. The problem is exacerbated by the fact that some (if not many) of these features are likely transient, and thus may no longer be obvious in observations, even though the consequences of their past existence still are. Circumnuclear star formation in disk galaxies is a good example of a highly time-variable phenomenon which is both an agent for and an indicator of secular evolution. The high gas densities required to initiate and maintain star formation in the central few hundred pc are the result of inward radial transport of large amounts of gas \citep[e.g.][]{simkin80,combes85,athanassoula94,knapen95}. Over time, the newly formed stars can alter the appearance of the galaxy, in that they contribute to the prominence of a \hbox{(pseudo-)bulge} \citep[e.g.][]{kor04}.\looseness-2 The star formation history of the nuclear region of galaxies seems to be also tightly linked to the properties of their central engines, i.e. active galactic nuclei (AGN). This has been widely investigated over the last decade from a number of observational studies which link dynamical or structural properties of the galaxy as a whole to those of the SMBH \citep[e.g.][]{ferrarese00,gebhardt00,graham+01,novak+06, shapiro+06,bandara+09,kor09,atlasxxiii}. Taken together, these results make it clear that the growth of SMBHs and the evolution of the central few $100\>{\rm pc}$ of the galaxy influence each other and thus cannot be interpreted independently. While some qualitative theoretical explanations have been proposed \citep[e.g.][]{mcLaughlin+06}, a detailed understanding of how star formation and AGN activity depend on each other is still lacking. This is hardly surprising, given that both star formation and AGN activity are highly time-dependent phenomena, and that any one galaxy can offer only a snapshot view of this time dependence. It is therefore desirable to observe as many galactic nuclei as possible with sufficient spatial resolution to separate the circumnuclear star formation activity from the central engine. Such observations are challenging because very often, the complex dust structures and resulting high extinction values make it necessary to use infrared or radio wavelengths in order to reveal the sites of active star formation (a.k.a. ``hot spots''), and their properties. In addition, it is desirable to not only derive the morphology of the various tracers of star formation, but also their kinematics in order to get a sense for the timescales involved. \begin{figure} \centering \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig1.eps} \caption{\textit{Hubble} Space Telescope colour image of the nuclear region of NGC\,613. Physical scale of the image is indicated in the lower right corner. The regions under study in this work are marked with white circles.} \label{fig:hst} \end{figure} Near-infrared integral field spectroscopy is the ideal tool for such studies. When used with modern 8m-class telescopes, it offers a combination of high sensitivity, high spatial resolution, and full spectroscopic information over the field of view. In this paper, we present VLT/{\tt SINFONI} observations of NGC\,613, a barred spiral galaxy with Hubble type Sbc at a distance of $17.5\>{\rm Mpc}$. The projected spatial scale at that distance is 84.8 parsec per arcsec. NGC\,613 was classified as a composite object by \citet{Veron86} based on its low-resolution optical spectrum (Seyfert/H{\sc ii}), and it was confirmed as an AGN using mid-infrared spectroscopy \citep{Goulding09}. NGC\,613 is part of a small sample of spirals with kpc-scale star-forming rings discussed in \citet[hereafter paper I]{boker08}, and some aspects of the data have already been presented there. Here, we use the data to give a detailed account of the physical conditions of stars and gas (both molecular and ionized) in the central $700\>{\rm pc}$, with the aim of illuminating the mutual feedback between circumnuclear star formation and nuclear activity. The paper is structured as follows. In section~\ref{sec:data} we present a brief outline of the observations and instrumental setup. Section~\ref{sec:morph} introduces the morphology of the stellar continuum and some emission-lines, while \S\ref{sec:spec} describes the location of the aperture spectra extracted for our analysis. We focus on the properties of the stellar component in the nucleus and along the ring in \S\ref{sec:stellar}. The details on the derived extinction for each aperture are presented in \S\ref{sec:extinc}. We discuss the physical state of the gas in the circumnuclear region in \S\ref{sec:ratios} and investigate the possible connection between the nucleus and the star-forming ring in \S\ref{sec:discussion}. Finally, we summarise our results in \S\ref{sec:summary}. \begin{figure*} \centering \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig2.eps} \caption{Observed flux maps for NGC\,613 ($\sim$700\,pc wide). From left to right:, Br$\gamma$\ (marking the locations and sizes of the apertures used in this paper), $\rm [Fe\,II]$, and $\rm H_2$\ maps. The right most panel displays a composite colour image using $\rm [Fe\,II]$\ (red), $\rm He\,I$\ (blue) and Br$\gamma$\ (green), see Section~\ref{sec:morph} for details. Overlaid on all the maps are the Br$\gamma$\ isophotes. In all the panels North is up and East to the left. Displayed emission-line fluxes range between 1.0$\times$10$^{-16}$ and 5$\times$10$^{-14}$\,erg\,cm$^{-2}$\,s$^{-1}$.} \label{fig:maps} \end{figure*} \section{OBSERVATIONS \& DATA REDUCTION}\label{sec:data} \subsection{Integral-field data} The near-infrared data analysed in this paper are part of the data set described in paper I. Briefly, we used the {\tt SINFONI} integral-field spectrograph at the European Southern Observatory Very Large Telescope (Program ID: 076.B-0646A) to observe a sample of five spiral galaxies with compact, star-forming circumnuclear rings selected from the imaging survey of \citet{knapen06}. In its seeing-limited mode (i.e. without the aid of adaptive optics), {\tt SINFONI} enables two-dimensional spectroscopy over a field-of-view of $8^{\prime\prime}\times8^{\prime\prime}$, and a spatial sampling of $\rm 0.125^{\prime\prime} /pixel$. We used the {\tt SINFONI} configuration for simultaneous $H+K$ spectra which yields a spectral resolution of $R\approx2000$. The total on-source integration time for NGC\,613 was 2.5 hours, divided in five identical observing blocks that were obtained over the course of five nights in Oct./Nov. 2005. Average seeing during our observations was approximately 0\farcs5. The raw data were corrected for sky background, detector dark current, and cosmic rays using standard methods. Flux calibration was performed using standard stars obtained routinely during each observing night. A detailed description of the data reduction and flux calibration procedures can be found in paper I. \subsection{Radio data} In addition to the integral-field data, and in order to confirm the presence of a jet (based on our $\rm [Fe\,II]$\ maps, see \S\ref{sec:morph}), we have searched the Very Large Array (VLA) archives for radio observations. The archival C band (4.86\,GHz) observations (project AH231) first published by \citet{hummel92} were done in A and B configuration in 1986. We reduced the two datasets using the standard routines in AIPS \citep{greisen90}. The quasars 3C84 and 0142-278 served as flux and phase calibrator, respectively. The combined datasets have a resolution of 0\farcs88$\times$0\farcs46 with a PA of 0.99$^\circ$ using robust weighting and a pixel scale of 0\farcs1 per pixel. We CLEANed the data using a single CLEANing box and a flux limit of 1.5$\sigma$. The resulting map has an rms of about 17 $\mu$Jy/beam. \subsection{Hubble Space Telescope imaging data} \label{sec:hst} In order to get the highest spatial resolution view of the inner regions of NGC\,613, we retrieved from the Hubble Legacy Archive (http://hla.stsci.edu) WFPC2 data for the F450W, F606W, and F814W filters. The dataset is part of the proposal number 9042 (PI: Stephen Smartt) aimed at detecting the progenitors of massive, core-collapse supernovae. Figure~\ref{fig:hst} shows a colour composite image (F450W, F606W, F814W) of the nuclear region of NGC\,613. For reference we indicate the main apertures used for the analysis in this paper. \section{MORPHOLOGY OF THE CENTRAL REGION}\label{sec:morph} In order to set the stage for the following discussion, we present in Fig.~\ref{fig:maps} a number of maps that define the morphology in the central $8^{\prime\prime}$ ($\approx~700\,\>{\rm pc}$) of NGC\,613. They are used here to identify the individual "hot spots" in the NGC\,613 ring that will be analysed in more detail using aperture spectra presented in \S\ref{sec:spec}.\looseness-2 Maps for the three emission lines $\rm [Fe\,II]$\ ($\lambda 1.64\,\mu$m), $\rm H_2$\ ($\lambda 2.12\,\mu$m), and Br$\gamma$\ ($\lambda 2.16\,\mu$m) have already been presented in paper I. As described there, the line maps were generated by summing all spectral channels over the width of the respective emission line, and subtracting a continuum image obtained by averaging neighbouring channels on either side. This method is equivalent to obtaining narrow band images centred on the line and the blue and red continuum, respectively. A number of grey circles in the Br$\gamma$\ map denote $1^{\prime\prime}$ diameter apertures centred on the various ``hot spots'' in the ring, one centred on the nucleus, and another one on a ``empty'' region. The three emission-line maps clearly reveal a ring-like morphology composed by seven stellar clusters. The nucleus presents high $\rm [Fe\,II]$\ and $\rm H_2$\ flux levels, compared to Br$\gamma$\ which is significantly weaker. The $\rm [Fe\,II]$\ flux distribution in the centre, however, seems somewhat more elongated than the fairly spherical $\rm H_2$\ morphology. As already noted in paper I, there appears to be a gap in the NW of the ring. This is conspicuous in all the three near-IR (hereafter NIR) line maps, as well as the radio maps shown in Fig.~\ref{fig:radio}. Interestingly, two plumes of material seem to extend from the nucleus all the way out to the edges of the gap. We believe these features are related to the presence of a radio outflow already noted by \citet{hummel87} and \citet{hummel92} (see \S\ref{sec:nuc}). The last panel in Fig.~\ref{fig:maps} displays a composite colour image produced by combining the $\rm He\,I$\ (blue), Br$\gamma$\ (green) and $\rm [Fe\,II]$\ (red) emission lines. This image illustrates the {\it pearls on a string} scenario proposed in paper I for the evolution of star formation of the hot spots in NGC\,613 ring, in which the hot spots age as they move along the ring away from the over-density region. \section{NEAR-INFRARED SPECTRA}\label{sec:spec} In order to perform a more detailed analysis of the central region of NGC\,613, we have extracted $H$- and $K$-band spectra for the apertures indicated in Fig.~\ref{fig:hst} and the Br$\gamma$\ map in Fig.~\ref{fig:maps}. The spectra are presented in Fig.~\ref{fig:spectra}. Their flux calibration is accurate to within 15\% as discussed in paper I. We have performed a quantitative analysis of the emission line fluxes and kinematics, and made an effort to decompose the spectra into the stellar continuum and the ``pure'' emission line spectrum. For this, we made use of the {\sc GANDALF} package \citep{sarzi06}. The software performs a simultaneous least-squares fit to both the stellar continuum and emission lines. The stellar continuum is described using theoretical spectra of red giants and supergiants which dominate the NIR emission of evolved stellar populations and star-forming regions, respectively. We made use of the library of theoretical model spectra by \citet{lancon07}. The spectra used for our fits have solar abundances, and cover a range of effective temperatures (T$_{\rm eff}$=2900--5900\,K) and gravity (log(g)=0--2). The models were re-binned to match the resolution of the data. The emission lines are treated simply as Gaussian templates. Their fluxes were left unconstrained during the fitting process, so that the amount of extinction could be estimated. The exact peak position and the width of the $\rm [Fe\,II]$\ line were determined independently from those of the Hydrogen Brackett-series and the different transitions of the H$_2$ molecule. For the latter, however, the different lines in each series were forced to share the same kinematics, i.e., line-of-sight velocity and velocity dispersion, as they did not show any substantial differences. Given the instrumental resolution of our data ($R\approx2000$), most of the emission lines are unresolved, with the exception of the $\rm [Fe\,II]$\ line. Uncertainties in the different line fluxes have been determined by generating 100 Monte Carlo realisations of the input spectra. We achieved this by perturbing our input spectra with white noise using the amplitude of the residuals from our spectral fits with GANDALF as an estimate of their variance. Tables~\ref{tab:fluxes} and~\ref{tab:h2fluxes} summarize the results of the emission-line analysis for all nine apertures. We list all lines that are detected with a minimum signal-to-noise ratio of 3. \begin{figure*} \begin{center} \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig3.eps} \caption{Observed aperture spectra extracted from the regions indicated in Fig.~\ref{fig:maps}. Black, the original spectra. Red, the best stellar fit to the spectra (see text for details). Green the best Gaussian fit to the emission lines present in the spectra. Main spectral features are also indicated.} \label{fig:spectra} \end{center} \end{figure*} \section{THE STELLAR COMPONENT}\label{sec:stellar} To our knowledge the only work measuring CO line-strengths in detail in NGC\,613 is that of \citet{js99}. In their study they placed several slits across different locations of the galaxy: the nuclear region and one of the spiral arms. One of the main findings in their paper was the fact that in the spiral arm the measured CO line-strength (using the \citealt{doyon94} definition of the index) was significantly lower than that expected from old stellar populations. Ruling out metallicity and extremely young stellar populations as the causes for this CO depletion, dilution by host dust was their main argument to explain the observed CO values. The results on the circumnuclear region presented in their paper confirm the presence of a young population (up to 24\% of the light in the $K$-band) in the innermost 4$^{\prime\prime}$, a fraction that decreases significantly as we move into the bulge dominated region (which is consistent with an old stellar population). \begin{figure} \begin{center} \includegraphics[angle=0,width=\linewidth]{NGC0613_fig4.eps} \caption{Two dimensional distribution of the D$_{\rm CO}$ line index \citep{marmol08} in the centre of NGC\,613} \label{fig:DCOmap} \end{center} \end{figure} Here we make use of the CO line-strength index (D$_{\rm CO}$) developed by \citet{marmol08} to interpret our observations. This definition is better suited for the study of stellar populations than previous ones, as it is based on observations of a new intermediate resolution $K$-band stellar library that exceeds the coverage of the stellar atmospheric parameters of previous works. It also has the advantage of being fairly independent on the broadening of the CO line, and thus avoiding often uncertain corrections. The index was measured over the aperture spectra using the {\hbox {\small INDEXF}} code ({http://www.ucm.es/info/Astrof/software/indexf/}) by \citet{cardiel07}.\looseness-1 In Fig.~\ref{fig:DCOmap} we present the two dimensional distribution of the D$_{\rm CO}$ index in the centre of NGC\,613. The map reveals some level of structure. The North side of the ring has slightly enhanced CO absorption compared to the southern part, where the most star-forming hot spots are located. Spot~6 displays the lowest D$_{\rm CO}$ value, as also evident from its spectrum in Fig.~\ref{fig:spectra}. Spot~6 is also the most heavily obscured region observed in the \textit{HST} image presented in Fig.~\ref{fig:hst}. The nucleus also displays a slightly lower D$_{\rm CO}$ content compared to its surroundings.\looseness-1 Figure~\ref{fig:HK_CO} shows the D$_{\rm CO}$ measurements for the different apertures defined in Fig.~\ref{fig:maps}. Given the high signal-to-noise of our aperture spectra, uncertainties in the D$_{\rm CO}$ index are negligible and well below 0.01. Except for Spot~6 all the other regions have D$_{\rm CO}$ values similar to those observed in the sample of field and cluster early-type galaxies of \citet{mm09}. This result may be somewhat unexpected, at least for the star-forming regions, given the marked difference between their integrated stellar populations and those of early-type galaxies. As illustrated in \citet{mayya97} as well as by the Starburst99 predictions \citep{starburst99} the evolution with time of the CO equivalent width for instantaneous bursts of star formation exhibits several peaks at different phases (most prominently the red supergiant phase around 10\,Myr). This behaviour makes it possible for such extremely different stellar populations to display similar D$_{\rm CO}$ values. At the same time it highlights the complexity of modelling the stellar content of galaxies in this wavelength range alone. On a side note, interestingly, Fig.~\ref{fig:HK_CO} shows a fairly marked difference, well beyond the reported measurement errors, in the D$_{\rm CO}$ content of field and cluster early-type galaxies. \citet{mm09} interpret this difference as a signature of an excess of intermediate age, asymptotic giant branch (AGB) stars in the stellar light of field galaxies. An alternative explanation, also considered by these authors, is the environmental dependence of the carbon abundance. The similarity in the D$_{\rm CO}$ values between the most actively star-forming regions in NGC\,613 and those of field galaxies, with significantly older stellar populations, strongly supports the latter view that environment rather than AGB stars, plays an important role in the chemical enrichment of early-type galaxies \citep[see also][]{carretero04}. One of the discoveries in paper I was the age sequence (see the right-most panel of Fig.~\ref{fig:maps}), formed by Spots~5, 6 and 7, along the southern side of the ring. As shown in this section, the complexity in understanding stellar populations at near-IR wavelengths makes it difficult to check whether the age gradient inferred from emission line ratios in paper I can be confirmed by the mean stellar ages of the hot spots. Nevertheless we refer the interested reader to \citet{allard06}, \citet{sarzi07} and \citet{vdl13} for examples of this kind of study in nearby galaxies with star-forming nuclear rings. \begin{figure} \begin{center} \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig5.eps} \caption{$H-K$ colour vs D$_{\rm CO}$ index for the aperture spectra in Fig.~\ref{fig:maps} (filled circles). Shaded grey region marks the colour and D$_{\rm CO}$ measurement for the stellar templates of \citet{lancon07}. $H-K$ colours have been computed from the spectra themselves. For comparison, red and blue squares are measurements for the sample of (cluster and field, respectively) early-type galaxies in \citet{mm09}.} \label{fig:HK_CO} \end{center} \end{figure} \section{EXTINCTION CORRECTION}\label{sec:extinc} In this section we employ several methods to determine the amount of extinction in the inner regions of NGC\,613. The list of emission-line fluxes for the different apertures, corrected for extinction, along with the computed A$_{\rm V}$ values, are presented in Tables~\ref{tab:fluxes} and \ref{tab:h2fluxes}. \subsection{The Brackett decrement: Br$\gamma$/Br10} \label{sec:Br_ratios} The flux ratio $\psi$\,=\,Br$\gamma$ /Br10 can be used to estimate the extinction within the {\tt SINFONI} field of view, because the relative intensities of the hydrogen recombination lines for various gas densities and temperatures are well understood (e.g. \citealt{hs87}). Any deviation from the theoretical line ratio $\psi_{0} = 3.07$ for a ``standard'' environment (case B, $\rho = 10^4\,e^-\,\rm{cm}^{-3}$, and $T_e = 7500\>{\rm K}$) can then be used to infer the amount of visual extinction, using the \citet{rl85} extinction law, along the line-of-sight as:\looseness-2 \begin{equation} \rm A_V = 50.5 \cdot log(\psi/\psi_{0}) \end{equation} The measurement of A$_{\rm V}$ using this method is only possible along the ring (i.e. Spots~1 to 7). Outside this area (Spot~8 and the nucleus), neither Br$\gamma$\ nor Br10 can be measured accurately enough to determine a reliable value. Extinction in the latter apertures was determined using the method outlined in the following section. \subsection{$H-K$ vs CO index} \label{sec:HK_CO} While the use of line ratios is a relatively straightforward method to estimate extinction, it is highly sensitive to the signal-to-noise ratio of the data. An alternative approach for those cases where the line ratios are not accessible is the use of colours and line-strength indices to obtain extinction values. The combination of a colour image, which is affected by dust, with a line index, largely insensitive to extinction \citep{macArthur05} is a powerful method that has not been applied much in the literature, because of the lack of well-calibrated line strength data \citep[see][]{ganda09}. By choosing the appropriate colour, this technique gives the colour excess in an almost model-independent way, which can easily be converted to A$_{\rm V}$.\looseness-2 In our case, we can measure the $H-K$ colour by integrating the flux in the spectra of each aperture within the given bandpass. From our $H-K$ colour map alone it is very difficult to determine the amount of extinction as it is not obvious how much of the colour is due to intrinsic stellar populations and how much comes from dust. Having two colours does not help a lot, since the effects of reddening in colour-colour diagrams is almost parallel to the effect of changing metallicity or age \citep[e.g.][]{kuchinski98}.The most prominent absorption features in the $H$- and $K$-band spectra presented in Fig.~\ref{fig:spectra} are the $^{12,13}$CO bandheads. Stellar population modelling has shown that the strength of these features reaches a maximum when the spectrum is dominated by red supergiants, typically 15 to 40\,Myr after the birth of the population \citep{persson83,doyon94,rhoads98}. Several authors defined quantitative diagnostics in the NIR spectrum of evolved stars to discriminate between the different luminosity classes \citep[e.g.][]{baldwin73,frogel78,kh86,omo93,doyon94,puxley97,fs00,ryder01,riffel07}. Figure~\ref{fig:HK_CO} illustrates the method to estimate the extinction. We assume that in the absence of dust the regions should have similar colours to those predicted by the stellar templates of \citet[][shaded area in Fig.~\ref{fig:HK_CO}]{lancon07}. The colour excess E($H-K$), the difference between data and the stellar spectra, is used in conjunction with the \citet{rl85} extinction law to compute A$_{\rm V}$ values. Since both the colour and D$_{\rm CO}$ index are fairly insensitive to instrumental broadening, there is no need to degrade our data's spectral resolution to match that of the theoretical stellar spectra. For reference we draw in the same figure the sample of field and cluster early-type galaxies of \citet{mm09}, which are expected to have very small colour excess. In general the measurements obtained with the line ratios in \S\ref{sec:Br_ratios} are a factor 2--5 larger than those derived here\footnote{Typical formal uncertainties in the extinction measured from the Br$\gamma$/Br10 ratio are around 2\,mag whereas those associated with the method presented in this section are on the order of 0.5\,mag.}. The observed differences could be associated to the assumptions made by each of the methods, i.e. the colour-based extinction relies on theoretical stars that might not mimic the intrinsically dusty conditions of star-forming regions \cite[e.g.][]{calzetti94}. This behaviour is actually not unusual and it is often found in star-forming galaxies \citep[e.g.][]{hao11,kreckel13}. Following \citet{kreckel13}, who concluded that the line ratio decrement method provides a more reliable extinction measurement for dusty regions, we use the line ratio based A$_{\rm V}$ values to correct our measured fluxes in the hot spots, and the colour-based method presented in this section for Spot~8 and the nucleus only.\looseness-1 \section{EMISSION-LINE ANALYSIS}\label{sec:ratios} Emission-line ratios of different elements constitute a powerful tool to disentangle the dominant excitation mechanisms in different regions in galaxies. Here we use the ratios of the $\rm [Fe\,II]$, $\rm H_2$\ and Br$\gamma$\ lines for this purpose. At the same time we make use of the Br$\gamma$\ line to provide star formation rate estimates, as well as the amount of H{\sc ii} present in the star-forming ring. \subsection{Star formation rates}\label{sec:sfr} Now that we have an estimate for the extinction in the different regions along the ring, we can correct the observed flux in the Br$\gamma$\ line and thus estimate the intrinsic flux. Knowing the intrinsic ratio between the Br$\gamma$\ and $\rm H\alpha$\ lines for case B recombination ($\rm H\alpha$/Br$\gamma$$\approx$104, \citealt{hs87}), we can convert those luminosities into star formation rates (SFRs) following the prescription by \citet{kennicutt98}: \begin{equation} {\rm SFR}(M_{\odot}~{\rm yr}^{-1}) = \frac{1}{1.26\times10^{41}~{\rm erg~s^{-1}}}\cdot L({\rm H\alpha}). \end{equation} \noindent In addition, we derived the mass of the ionised gas in each aperture assuming the same case B recombination scenario as above.\looseness-2 The star formation rates measured in the ring spots range from 0.03 to 0.1\,$\>{M_{\odot}}$/yr. The value for the nucleus, however, drops down to 0.015\,$\>{M_{\odot}}$/yr, which is lower than the SFRs reported by \citet{Kewley02} for a sample of 81 nearby non-active galaxies. It is therefore likely that the recombination lines observed in the nucleus are not due to star formation but the AGN. The SFRs measured in apertures 5, 6 and 7 show a steady decline along the ring, which is consistent with the {\it pearls on a string} scenario presented in paper I (i.e. the hot spots passively evolve as they move along the ring and away from the over-density regions). The masses of the ionised-gas in the H{\sc ii} regions probed by the apertures vary from $\approx$\,1.2\,$\times$\,10$^{3}\>{M_{\odot}}$ in the nucleus to 11.2\,$\times$\,10$^{3}\>{M_{\odot}}$ for the most massive hot spot. The SFR densities in the hot spots are in good agreement with those of circumnuclear star-forming regions in disk galaxies \citep{kennicutt98}.\looseness-2 \subsection{The $\rm [Fe\,II]$\ \& Br$\gamma$\ ratio} Strong $\rm [Fe\,II]$\ emission is indicative of shock-excited gas in the filaments of supernova remnants, in contrast to the weak $\rm [Fe\,II]$\ emission characteristic of photoionized gas in H{\sc ii} regions. In AGN, strong $\rm [Fe\,II]$\ emission is also common, although several processes may contribute to its production: (1) photoionization by extreme UV to soft X-ray radiation from the central source, producing large partially ionized regions in the NLR clouds; (2) interaction of radio jets with the surrounding medium, inducing shocks and hence partially ionized cooling tails; and (3) fast shocks associated with supernova remnants present in starburst regions. The $\rm [Fe\,II]$/Br$\gamma$\ ratio has proved to be very useful for distinguishing between a stellar or non-stellar origin of the $\rm [Fe\,II]$\ emission. This line ratio increases from H{\sc ii} regions (photoionization by hot stars) to supernova remnants (shock excitation), passing through starburst and active galaxies \citep[e.g.][]{alonso97,ra04,ra05,ramos06,Ramos09,Riffel13}. Figure~\ref{fig:BrgFeIIratio} shows the variation of the $\rm [Fe\,II]$/Br$\gamma$\ ratio for our apertures. The ratio is relatively low ($\rm [Fe\,II]$/Br$\gamma$=[0.9,2.5]) in the star-forming hot spots, consistent with the typical values of H{\sc ii} regions. On the other hand, the ratio in the nucleus is high ($\rm [Fe\,II]$/Br$\gamma$=17.7), in the domain populated by supernova remnants and radio-loud Seyferts as e.g. NGC\,1275 and NGC\,2110 ($\rm [Fe\,II]$/Br$\gamma$=37.3 and 28.1 from \citealt{Kawara93} and \citealt{ra05} respectively). As found by \citet{Forbes93}, there is a tight correlation between $\rm [Fe\,II]$\ and radio emission in both Seyfert and star forming galaxies. In fact, as we can see in Sec.~\ref{sec:nuc}, the radio contours perfectly trace the $\rm [Fe\,II]$\ emission. We thus conclude from Fig.~\ref{fig:BrgFeIIratio} that the excitation mechanisms in the nucleus of NGC\,613 (i.e. photoionization and shocks) differ from that in the hot spots (stellar photoionization).\looseness-2 \begin{figure} \begin{center} \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig6.eps} \caption{Variation of the $\rm [Fe\,II]$\ over Br$\gamma$\ ratio for the different apertures considered here. We indicate typical ratios for different types of astrophysical objects from the compilations of \citet{mo88,mkt93,dale04,ra04,ra05}. Horizontal dashed lines mark the values observed for the two radio-loud galaxies NGC\,1275 and NGC\,2110 from \citet{Kawara93} and \citet{ra05} respectively.} \label{fig:BrgFeIIratio} \end{center} \end{figure} \subsection{Molecular hydrogen} All aperture spectra shown in Fig.~\ref{fig:spectra} show a number of emission lines from molecular hydrogen. This section investigates what can be learned about the physical state of the $\rm H_2$\ gas. In this effort, we follow a number of studies, both theoretical \citep[e.g.][]{mouri94}, and observational \citep[e.g.][]{vei97,krabbe00,davies05,ra04,ra05,ramos06,Ramos09,Riffel13, Mazzalay13}. All these studies take advantage of the fact that in most astronomical environments, the lowest vibrational levels ($v=1$) of $\rm H_2$\ tend to be well thermalised, while higher level transitions are predominantly populated by processes such as non-thermal UV fluorescence. Generally speaking, the $\rm H_2$\ molecule can be excited via three distinct mechanisms: (i) UV fluorescence, where photons with $\lambda> 912$\AA\ are absorbed by the $\rm H_2$\ molecule and then re-emitted \citep[e.g.][]{bv87}; (ii) shocks, where high-velocity gas motions heat and accelerate the ambient gas \citep[e.g.][]{holkee89}; and (iii) X-ray illumination, where hard X-ray photons penetrate deep and heat large amounts of molecular gas \citep[e.g.][]{maloney96}. Shocks and X-ray illumination are normally referred to as thermal processes, and UV fluorescence as non-thermal. Each of these three mechanisms produces a distinct $\rm H_2$\ spectrum.\looseness-2 The 1$-$0S(1)/2$-$1S(1) line ratio is an excellent discriminator between thermal and non-thermal processes. According to the models of \cite{mouri94}, this ratio has much lower values ($\leq2$) in regions that are dominated by UV fluorescence than in thermal-dominated gas ($\geq5$). This diagnostic has the advantages of being fairly insensitive to extinction because both lines have similar wavelengths and are independent of the ortho/para ratio. At the same time, the 1$-$0S(2)/1$-$0S(0) line ratio is sensitive to the strength of the incident radiation, and it can be used to discriminate the dominant excitation process. \begin{figure} \begin{center} \includegraphics[angle=0,width=0.99\linewidth]{NGC0613_fig7.eps} \caption{$\rm H_2$\ 1$-$0S(2)/1$-$0S(0) versus $\rm H_2$\ 1$-$0S(1)/2$-$1S(1) line ratios of the different selected regions in NGC\,613. In addition, literature values for other galaxies are presented (see text for details). The dashed line indicates the locus of equal T$_{\rm vib}$ and T$_{\rm rot}$. Vertical dotted lines delimit the regions of thermal and UV fluorescence excitation from \citet{mouri94}.} \label{fig:H2ratio} \end{center} \end{figure} Figure~\ref{fig:H2ratio} shows - for the apertures defined in Fig.~\ref{fig:maps} - the value of the flux ratio of the 1$-$0S(1) and 2$-$1S(1) versus the ratio of the 1$-$0S(2) and 1$-$0S(0) lines. For comparison, we also plot the ratios of five nearby Seyfert 2 galaxies from \citet{Ramos09} and of the starburst galaxy NGC\,253 from \citet{Riffel13}.\looseness-2 All the hot spots in the ring are, to varying degrees, affected by a mixture of thermal and non-thermal (i.e. fluorescent) excitation. The molecular gas in the nucleus of NGC\,613, however, falls in the thermal excitation domain, indicating a lack of strong non-thermal excitation mechanisms. The $\rm H_2$\ ratios measured for the nucleus are also far from those predicted by thermal X-ray models (e.g. \citealt{Lepp83,maloney96}), which would lie outside the upper right boundaries of Fig.~\ref{fig:H2ratio}. To confirm these results, we can compare the rotational and vibrational temperatures (T$_{\rm vib}$ and T$_{\rm rot}$) of the $\rm H_2$\ gas, which we calculated using the expressions in \citet{Reunanen02}. In the case of thermal excitation, both temperatures should be similar (i.e. close to the dashed line in Fig.~\ref{fig:H2ratio}), whereas in the case of non-thermal excitation, T$_{\rm vib}>>T_{rot}$. The latter is the case of the hot spots, which have T$_{\rm vib}$=3000--5000 K and T$_{rot}<$1200 K. On the other hand, the $\rm H_2$\ gas in the nucleus has T$_{\rm vib}\sim$2500$\pm$500 K and T$_{\rm rot}\sim$1200$\pm$40 K, characteristic of thermally excited gas. At face value, these results are consistent with the absence of strong star formation and of a luminous X-ray source in the nucleus. Its relatively high 1$-$0S(2)/1$-$0S(0) value falls in the region of shock heated (thermalised) gas, and points to mechanical energy from the radio outflow as the source of $\rm H_2$\ heating. All seven ring apertures, in contrast, show a much more pronounced contribution from UV fluorescence. This is fully consistent with them being young star-forming regions. In addition to elucidating the excitation mechanisms at play in the inner regions of NGC\,613, it is interesting to estimate the amount of molecular gas present. This information is very valuable as it can be used to estimate the level of star formation expected in the near future in these regions. Using the $1-0$S(1) line fluxes reported in Table~\ref{tab:h2fluxes}, we estimate the mass of hot $\rm H_2$\ following the equation \begin{equation} m_{\rm H_2}^{hot}(M_{\odot}) \simeq 5.0875\times10^{13}d^2I_{1-0{\rm S}(1)}, \end{equation} \noindent from \citet{Reunanen02}, as well as the assumptions for gas temperature, transition probability and population fractions therein. $I_{1-0S(1)}$ is the extinction-corrected line flux, and $d$ the distance to the galaxy in Mpc. The resulting $\rm H_2$\ masses are reported in Table~\ref{tab:h2fluxes}. We derive values between 9.4 and 38.5\,$\>{M_{\odot}}$ for the ring spots and 112\,$\>{M_{\odot}}$ for the nucleus. In the ring spots, where star formation is taking place, the mass of hot gas is low, whereas in the nucleus the mass of hot $\rm H_2$\ is an order of magnitude larger, although this measurement is likely affected by the interaction with the radio jet. Nevertheless, our masses of hot $\rm H_2$\ are consistent with those derived for a sample of Seyferts and low-luminosity AGN using {\tt SINFONI} data as those used here \citep{Mazzalay13}. The same authors also estimated the the mass of cold $\rm H_2$\ from the integrated 1$-$0S(1) luminosity ($L_{1-0{\rm S}(1)}$) in erg~s$^{-1}\>{\mu {\rm m}}^{-1}$ using the relation: \begin{equation} \label{eq:cold_mass} m_{\rm H_2}^{cold}(M_{\odot}) \approx 1174\times L_{1-0{\rm S}(1)}(L_{\odot}), \end{equation} derived by comparing a large number of $m_{\rm H_2}^{cold}$ values from the literature, calculated from CO observations, and integrated $\rm H_2$\ 1$-$0S(1) luminosities\footnote{The reported uncertainty of this calibration is about a factor 2 in mass.}. Thus, we can roughly estimate the masses of cold $\rm H_2$\ gas for the nucleus and the hot spots in NGC\,613 (also listed in Table~\ref{tab:h2fluxes}). Those masses range between $\sim$7 and 28$\times10^6\>{M_{\odot}}$ in the ring spots, and $\sim$8$\times10^7\>{M_{\odot}}$ in the nucleus. An important word of caution is necessary on the interpretation of $\rm H_2$\ masses having an associated amount of cold molecular gas. While relations like the one presented in Eq.~\ref{eq:cold_mass} are very useful as proxies for the presence of cold gas, there are known cases of galaxies with estimates of molecular mass content, based on the $\rm H_2$\ line, two orders of magnitude above the one provided by the direct measurement of CO emission \cite[e.g.~NGC\,4151,][]{dumas10}. \section{DISCUSSION}\label{sec:discussion} In light of the analysis and results described in the previous sections, we now discuss some of the most interesting unsolved questions in the circumnuclear environment of NGC\,613. These are mainly related to the interactions between the nuclear ring and the radio outflow, and to the source of excitation of the $\rm H_2$\ present in the nucleus. \subsection{The radio outflow: an active nucleus in NGC\,613?} \label{sec:nuc} The classification of NGC\,613 in terms of nuclear activity is not completely clear by looking at its optical spectrum alone. \citet{Veron86}, as part of a spectroscopic study of the complete Revised Shapley-Ames Catalogue, classified it as composite object (Seyfert-like component co-existing with an H{\sc ii} region). This classification is not surprising considering that the spectrum was extracted with an aperture of 4\arcsec, which corresponds to 328\,pc in the case of NGC\,613. The AGN nature of NGC\,613 has, however, been recently confirmed from mid-infrared spectroscopy. \citet{Goulding09} reported Spitzer/IRS observations showing high-excitation lines such as [Ne{\sc III}], [Ne{\sc V}] and [Si{\sc II}], characteristic of active galaxies. These authors claimed that the optical signatures of nuclear activity in this galaxy are likely diluted by strong star formation. In the X-rays, the galaxy was observed with XMM-Newton on December 2010 and, although the data have not been published yet, they appear to confirm the presence of an active nucleus in NGC\,613 (obsID 0654800501). The high resolution radio continuum images presented in \citet{hummel92}, and re-analysed here (see \S\ref{sec:data}), show evidence for an energetic outflow emanating from the nucleus. The radio map (see Fig.~\ref{fig:radio}) reveals a linear feature of about 300\,pc which consists of three discrete components. These radio blobs are perpendicularly oriented to the projected major axis of the star-forming ring observed in our {\tt SINFONI} maps. Whether the radio blobs are truly separate entities aligned in one direction, or bubbles of hot plasma originating from the incident radiation has not yet been determined. However, both the radio morphology and the coincidence of the central blob with the optical position (radio-optical offset of 0\farcs1) indicate that the central component of the linear feature is indeed the nucleus. This constitutes a first indication that the radio jet orientation might be close to the plane of the sky, and consequently, relatively close to the plane of the galaxy as well, which has an inclination angle of $\sim$35$^\circ$ (as listed in Hyperleda\footnote{http://leda.univ-lyon1.fr/}). Interestingly, \citet{laine06} found a tendency towards perpendicular alignments in Seyferts, but more parallel in starbursts. \citet{Kondratko06} and \citet{Castangia08} reported the detection of a H$_2$O megamaser with an isotropic luminosity of $\sim$35\,L$_{\odot}$, coincident with the position of the nucleus in the optical (with an uncertainty of 1.3\arcsec). This coincidence supports the link between the AGN and maser emission. In fact, all masers with isotropic luminosities $>$10\,L$_{\odot}$ are associated with AGN \citep{Zhang12}, as confirmed from interferometry. There is good evidence that extragalactic H$_2$O megamasers trace edge-on accretion disks in Seyfert galaxies \citep[e.g.][]{Greenhill97,Greenhill03}, since large line-of-sight column densities are required theoretically to make the masers observable \citep{Lo05,Henkel05}. However, there is a different class of H$_2$O megamasers, the so-called ``jet-masers''. In these sources, the maser emission is the result of an interaction between the radio jet and a molecular cloud on parsec scales \citep{Braatz97,Henkel98,Peck03}. Only four Seyfert galaxies have jet-masers confirmed to date: Mrk\,348, NGC\,1052, NGC\,1068 and the Circinus galaxy\footnote{NGC\,1068 and Circinus seem to have both disk- and jet-masers.} (see \citealt{Peck03} and references therein). Although it is difficult to infer the radio jet orientation from maser emission, in all the known jet-masers the radio jets are perpendicular to the line-of-sight. In the case of NGC\,613, the origin of the megamaser is uncertain. \citet{Kondratko06} reported the detection of a very broad H$_2$O emission feature (FWHM$\sim$87 km~s$^{-1}$), which was subsequently resolved into two different components of 20 and 40 km~s$^{-1}$ by \citet{Castangia08}. These two components are redshifted and blueshifted, respectively, with respect to the galaxy systemic velocity. The large linewidths and the presence of two kinematic components only point to a jet-maser, but the fact that the H$_2$O emission is centred on the systemic velocity is not typical of those. Independently of the disk- and/or jet-maser classification of the H$_2$O megamaser in NGC\,613, it is very likely that the orientation of the jet is close to the plane of the sky. This orientation favours shocks capable of yielding a strong jet-maser along our line-of-sight (see Fig.~11 in \citealt{Peck03}). If we have a disk-maser, the accretion disk and the dusty torus have to be very close to edge-on for the line-of-sight column density to make the maser detectable, and thus, the radio jet would also be orthogonally oriented to the plane of the disk (see \citealt{Drouart12} and references therein).\looseness-1 Once we know the radio jet orientation, our NIR observations provide additional information to establish the true morphology of the outflow. In paper I, we showed how the structure of the different emission lines in our spectral range were affected by the out-flowing radiation. The flux maps for all the lines (i.e. $\rm [Fe\,II]$, $\rm He\,I$, $\rm H_2$, and Br$\gamma$) display a lack of emission on the top half of the ring, coinciding with the direction of the out flowing material. This indicates that the interaction with the radio jet has probably swept out gas and dust. This is visible in Fig.~\ref{fig:hst}. \begin{figure} \begin{center} \includegraphics[angle=0,width=\linewidth]{NGC0613_fig8.eps} \caption{{\tt SINFONI} $\rm [Fe\,II]$\ flux and velocity dispersion maps for NGC\,613. Overlaid we show VLA radio contours (see \S\ref{sec:data} for details). The inset in the lower-right of each panel gives the range of values for the colour scale.} \label{fig:radio} \end{center} \end{figure} While morphologically all the lines seem altered, only the $\rm [Fe\,II]$\ line appears to be affected dynamically. In Fig.~\ref{fig:radio} we show the $\rm [Fe\,II]$\ line flux and velocity dispersion with radio flux contours overlaid. We see that the highest velocity dispersion values coincide, although not perfectly, with the linear structure resolved in three blobs observed in the radio continuum map. Although marginal, the values happen to be somewhat lower at the edges of this gap, coinciding with two plumes of $\rm [Fe\,II]$\ that extend beyond the nuclear ring. Taken at face value, this picture is consistent with a jet radiating along a cone. The high velocity dispersions can be explained by the fact that the differences in the line-of-sight velocities of the material are maximum in the central parts of the cone. At the edges of the cone, velocities are very similar, which explains the low values there. Regarding the flux, the enhancement we see at the edges would be caused by limb-brightening of the conical structure. In conclusion, our NIR data support the presence of a radio jet oriented close to the plane of the sky, which is sweeping out the gas and dust within the top half of the star-forming ring. According to \citet{hummel92}, if the ring is circular, its inclination is $\sim$55$^\circ$, larger than the 35$^\circ$ inclination of the galaxy. This larger disk inclination explains the proposed scenario: the interaction with the jet affects the material in the top half of the ring, while the bottom part remains unaltered. \subsection{The nucleus: a starburst in waiting?} It is evident from the numerous emission lines in the spectrum of the central $1^{\prime\prime}$ ($84.8\,\>{\rm pc}$) in Fig.~\ref{fig:spectra} that the nucleus of NGC\,613 harbours a substantial amount of hot $\rm H_2$\ gas. Our mass estimate based on the 1$-$0S(1) line flux is 1.12$\times10^2\>{M_{\odot}}$. This mass is an order of magnitude larger than in the star-forming knots in the ring. A large amount of cold gas is also found, which we estimated using the empirical relation derived by \citet{Mazzalay13} for Seyfert galaxies and low-luminosity AGN. We derived a cold gas mass of $\sim$8$\times10^7\>{M_{\odot}}$ for the nucleus of NGC\,613. These values are larger than the hot and cold masses reported in \citet{Mazzalay13} for a small sample of nearby AGN. Their sample includes two galaxies with circumnuclear star-forming rings, NGC\,3351 and NGC\,4536, and the masses of hot and cold gas measured by \citet{Mazzalay13} in those rings are much larger than in their nuclei, which is opposite to what we find in NGC\,613. In contrast to the presence of hot and cold molecular gas, there is no evidence for active star formation in the nucleus of NGC\,613: the recombination lines of ionized hydrogen (e.g. Br$\gamma$) are almost absent. This immediately raises a number of questions about the mechanism responsible for heating the $\rm H_2$, the amount of molecular gas within the nucleus, and the likely fate of this gas. A similarity between the two galaxies with circumnuclear rings in \citet{Mazzalay13} and NGC\,613 is the deficit of Br$\gamma$\ emission relative to the $\rm H_2$\ emission. However, the 1$-$0S(1)/Br$\gamma$\ ratio that we measure in the nucleus of NGC\,613 (14.9$\pm$3.0) is much larger than the values reported in \citet{Mazzalay13} for NGC\,3351 and NGC\,4536 on the same scales (5.9 and 1.9). This ratio is normally $\la$0.6 in starburst galaxies\footnote{In the ring hot spots we measure 1$-$0S(1)/Br$\gamma$$\leq$0.67.}, between 0.6 and 6 in Seyfert galaxies, and larger than 6 in Low lonization Nuclear Emission Regions (LINERs; \citealt{Riffel13,Mazzalay13}). The extreme value of 1$-$0S(1)/Br$\gamma$\ measured in the nucleus of NGC\,613 is thus consistent with a low-luminosity AGN (Seyfert or LINER) with a strong influence from shock heating. The interaction with the radio jet is thus likely enhancing the $\rm H_2$\ emission in the nucleus.\looseness-2 Using $K$-band {\tt SINFONI} data at the highest resolution available, \citet{Davies07} studied the AGN--star formation connection in the inner 10\,pc of nine nearby Seyferts. They found evidence for starbursts which took place in the last 10--300\,Myr but are no longer active, and for a delay of 50--100\,Myr between the triggering of the star formation in those galaxies and the onset of the nuclear activity. Considering the lack of star formation in the nucleus of NGC\,613, the large amount of cold molecular gas and the confirmed presence of nuclear activity, we speculate with a cyclical scenario, where a starburst episode was followed by another of nuclear activity. In addition, the interaction with the radio jet likely produced a substantial amount of molecular gas, creating a reservoir of cold gas that can sustain the AGN and possibly fuel the next starburst episode. We have attempted to date the last episode of star formation in the nucleus comparing the Br$\gamma$\ equivalent width with Starburst99 model predictions. The measured value of 1.89\,\AA\ implies that the latest starburst happened sometime between 8 and 13\,Myr ago (depending on the adopted metallicity)\footnote{Considering the possible effect of the radio jet, this estimate is necessarily a lower limit for the last starburst.}. Based on our current estimate of H$_2$ molecular gas available, it is interesting to determine when the next event might take place. The estimated free-fall time (t$_{\rm ff}$), i.e., the earliest possible collapse time, for a cloud of 0\farcs5 radius (our aperture size) and $\sim$8$\times10^7\>{M_{\odot}}$ is t$_{\rm ff}$\,$\sim$0.5\,Myr. The next starburst episode can thus be imminent considering the dynamical timescales ($\sim$\,Gyr) in this type of galaxies. Compared to the measurements presented by \citet{Mazzalay13} for a set of five star-forming galaxies, the amount of H$_2$ present in NGC\,613 is at least a factor of two, and up to a factor forty, larger for similar aperture measurements. This comparison emphasises the rather special situation we are observing in NGC\,613, where a starburst episode is likely about to happen. This result, together with the lack of emission lines in other galaxies in our original sample (paper I), suggest that as long as there is gas supply towards the centre of the galaxy, recurrent star formation events can take place in very short time intervals. This scenario is consistent with the one established by \citet{Davies07}, although the time scales for NGC\,613 are somewhat shorter. The lack of apparent gas transfer from the star-forming ring to the inner regions, based on the H$_2$ emission-line map in Fig.~\ref{fig:maps}, makes it rather difficult to determine the frequency of these events in NGC\,613. While interesting, NGC\,613 is by no means a unique case. Additional examples of galaxies, with large $\rm H_2$\ mass concentrations in their nuclei, that might be at the verge of undergoing a starburst phase include NGC\,6946 \citep{schinnerer06,schinnerer07} and NGC\,7552 \citep{pan13}.\looseness-2 While the discussion about recurrent star formation episodes presented here focuses on the nucleus of NGC\,613, it is worth mentioning that a similar behaviour has been observed in H{\sc ii} regions of star-forming nuclear rings and nuclear clusters (with intense starbursts followed by long inactive periods, e.g.~\citealt{allard06,walcher06,sarzi07}). \section{SUMMARY AND CONCLUSIONS}\label{sec:summary} In this paper we have made use of NIR integral-field {\tt SINFONI} observations to study the inner regions of the nearby spiral galaxy NGC\,613. This galaxy is part of the sample of five galaxies with star-forming nuclear rings presented in \citet{boker08}. We selected this galaxy for a more detailed analysis due to the peculiar nature of its circumnuclear environment: a star-forming ring, a radio jet, and an unexpectedly high level of H$_2$ concentrated in its nucleus. In \citet{boker08} we concluded that star formation along the inner ring proceeds in a ''pearls on a string`` fashion, i.e. with star clusters in the ring getting progressively older as they move away from the over-density regions. The complexity of studying stellar content with the CO bandhead alone prevented us from checking whether the same behaviour is observed in the underlying populations. Nevertheless, the measurement of the CO equivalent width (i.e. D$_{\rm CO}$ index), together with the $H-K$ colour, now allows us to determine the amount of extinction in the innermost regions of NGC\,613. Incidentally, the comparison of our D$_{\rm CO}$ values with those of early-type galaxies reinforces previous findings in support of environmentally driven carbon abundances. The analysis of the emission lines reveals that the gas in the nucleus is not exclusively photoionized by the AGN, but also shock-heated. The hot spots along the ring however show a much more pronounced contribution from UV fluorescence. The star formation rates and masses of the ionised gas in H{\sc ii} regions are consistent with those observed in star-forming regions in other galaxies. In our discussion, we used all this information to study the origin of the nuclear activity and establish the orientation of the observed radio outflow, which is nicely mapped by the $\rm [Fe\,II]$\ distribution and velocity dispersion. Based on the existence of a maser in the nucleus, we propose that the radio jet is likely oriented very close to the plane of the sky and that it is sweeping away the gas and dust in one portion of the ring. Finally, we have investigated the fate of the unusually large amount of H$_2$ gas present in the nucleus of NGC\,613. We estimate that the last episode of star formation took place around 10\,Myr ago and establish a lower limit for the onset of the next event of 0.5\,Myr. The recurrence of these episodes, however, is rather uncertain, especially since we have not found any evidence for gas transfer from the star-forming ring to the nucleus. In addition to that, one has to consider the role the radio jet may play in inhibiting or fostering the generation of new stars. Still, datasets like the one presented here have the important value of setting constraints and adding fundamental information to the still rather poorly understood AGN--star formation connection.\looseness-2 \section*{Acknowledgments} The authors are indebted to J.~A. Acosta, A. Alonso-Herrero, P. Castangia, P. Esquej and A. Vazdekis for insightful comments and suggestions at different stages of this work. J.~F.-B. acknowledges support from the Ram\'on y Cajal Program, grants AYA2010-21322-C03-02 from the Spanish Ministry of Economy and Competitiveness (MINECO). We also acknowledge support from the FP7 Marie Curie Actions of the European Commission, via the Initial Training Network DAGAL under REA grant agreement number 289313. C.~R.~A. acknowledges grant PN AYA2010-21887-C04.04 (Estallidos) from MINECO. Based on observations collected at the European Southern Observatory, Chile, for proposal 076.B-0646(A).\looseness-2 \bibliographystyle{mn2e}
\section{Introduction} Quantum computation has great possibility to offer substantial advantages in solving some sorts of mathematical problems and also in simulating physical dynamics of quantum systems. A representative instance is Shor's factoring algorithm~\cite{Shor}, which solves integer factoring problems in polynomial time, while no polynomial-time classical algorithm has been known. Furthermore, there are some evidences that quantum computation, more precisely, {\sf BQP} (bounded-error quantum polynomial-time computation~\cite{BernsteinVazirani}), can solve problems outside the polynomial hierarchy ({\sf PH}\cite{PH1,PH2}) \cite{Aaronson09}. These results strike the extended Church-Turing thesis~\cite{Turing,Church,BernsteinVazirani}, which states that every reasonable physical computing devices can be simulated efficiently (with a polynomial overhead) on a probabilistic Turing machine. One of the most revolutionary and challenging goals of human beings is to realize universal quantum computer and verify such quantum benefits in experiments. However, experimental verification, which is the most essential part in science, is still extremely hard to achieve, requiring a huge number of qubits and extremely high accuracy in controls. Is there any possible pathway to verify computational complexity benefits of quantum systems that are realizable in the near future, say, one-hundred-qubit (or particle) systems under reasonable accuracy of controls? If there is such a subclass of quantum computation that consists of experimental procedures much simpler than universal quantum computer but is still hard to simulate efficiently in classical computers, experimental verification of complex quantum systems reaches new phase. Aaronson and Arkhipov introduced {\sf B{\footnotesize OSON}S{\footnotesize AMPLING}}~\cite{boson}, a sampling problem according to a probability distribution of $n$ bosons scattered by linear optical unitary operations. The probability distribution is given by the permanent of a complex matrix, which is determined by the linear optical unitary operations. Calculation of the permanent of complex matrices is known to be \#{\sf P}-hard~\cite{Valiant}. Since a polynomial-time machine with an oracle for \#{\sf P} can solve all problems in the {\sf PH} according to Toda's theorem~\cite{Toda}, an exact classical simulation (in the strong sense~\cite{WeakSim,JozsaNest} meaning exact calculation of the probability distribution) of {\sf B{\footnotesize OSON}S{\footnotesize AMPLING}} is highly intractable in a classical computer. They showed under assumptions of plausible conjectures that if there exists an efficient classical approximation of {\sf B{\footnotesize OSON}S{\footnotesize AMPLING}} (classical simulation in the weak sense~\cite{WeakSim,JozsaNest} meaning sampling according the a given probability distribution), the {\sf PH} collapses at the third level, which unlikely occurs. This result brings a novel perspective on linear optical quantum computation and drives many researchers into the recent proof-of-principle experiments~\cite{BSex1,BSex2,BSex3,BSex4}. Another subclass of quantum computation of this kind is instantaneous quantum polynomial-time computation ({\sf IQP}) proposed by Shepherd and Bremner~\cite{IQP0}. {\sf IQP} consists only of commuting unitary gates, such as $\exp[ i \theta \prod _{k \in S} Z_k]$. Here $\theta \in [0,2\pi )$ is a rotational angle, $Z_k$ indicates the Pauli operator on the $k$th qubit, and $S$ indicates the set of qubits on which the commuting gate acts. (A detailed definition will be provided in the next section.) The input state is given by $|+\rangle ^{\otimes n}$ with $|+\rangle \equiv (|0\rangle + |1\rangle)/\sqrt{2}$, and the output state is measured in the $X$-basis. Since all unitary operations are commutable with each other, there is no temporal structure in the circuits. (This is the reason why it is called instantaneous quantum polynomial.) The commutability implies that {\sf IQP} cannot perform an arbitrary unitary operation for the input qubits and hence seems to be less powerful than standard quantum computation, i.e., {\sf BQP}. Nevertheless, Bremner, Jozsa, and Shepherd showed that if there is an efficient classical algorithm that samples the outcomes (and hence classical simulation in the weak sense~\cite{WeakSim,JozsaNest}) according to the probability distribution of {\sf IQP} within a certain multiplicative approximation error, the {\sf PH} collapses at the third level. While this does not mean {\sf P} = {\sf NP}, this is also considered to be highly implausible. The result is obtained by introducing postselection and using the fact that {\sf post-{\sf BQP}=PP} shown by Aaronson~\cite{postBQP}. Here postselection means that an additional ability to choose, without any computational cost, arbitrary measurement outcomes of possibly exponentially decreasing probabilities. However, in comparison to {\sf B{\footnotesize OSON}S{\footnotesize AMPLING}} ~\cite{boson,Aaronson13}, the origin of the classical intractability of {\sf IQP} is still not well understood. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.8\textwidth]{fig0.eps} \end{center} \caption{The summary of results obtain in this work.} \label{fig0} \end{figure} The purpose of this paper is to further explore the origin of the classical intractability of {\sf IQP}, relating it with computational complexity of Ising partition functions, which has been well studied in both statistical physics, computer science, and quantum computation. Specifically we obtain the following results (see Fig.~\ref{fig0}): \begin{enumerate}[(i)] \item We reformulate {\sf IQP} from a viewpoint of computational complexity of calculating Ising partition functions. The probability distribution of the output distribution of {\sf IQP} is mapped into a partition function of an Ising model with imaginary coupling constants (Theorem~\ref{Main1} and Theorem~\ref{Main2}). \item By using the above relation, we specify classically simulatable classes of {\sf IQP}, which correspond to exactly solvable Ising models (Theorem~\ref{sparseIQP} and Theorem~\ref{freefermion}). For example, {\sf IQP} that consists only of nearest-neighbor two-qubit commuting gates in two dimension (2D) is classically simulatable irrespective of their rotational angles. \item We show that there exists no fully polynomial randomized approximation scheme (FPRAS) for the Ising partition functions with almost all imaginary coupling constants even on a 2D planar lattice with a bounded degree, unless the {\sf PH} collapses at the third level (Theorem~\ref{FPRASHard}). We further show, by considering strong simulation of {\sf IQP}, that a multiplicative approximation of such Ising models is at least as hard as a multiplicative approximation for the output distribution of universal quantum computation. \end{enumerate} The first result bridges {\sf IQP} and computational complexity of Ising partition functions, which tells us the origin of classical intractability of classical simulation of {\sf IQP}, since exact calculation of Ising partition functions is \#{\sf P}-hard even in the ferromagnetic case \cite{Barahona,Jerrum93}. Only restricted models are known to be exactly solvable such as Ising models on the 2D planar lattices without magnetic fields. One might expect that {\sf IQP} mapped into an exactly solvable Ising model is classically simulatable in the strong sense~\cite{WeakSim,JozsaNest}, since its joint probability distribution can be calculated efficiently. However, there are exponentially many instances of the measurement outcome, and hence an efficient calculation of the joint probability distribution of the output of {\sf IQP} does not directly mean an efficient simulation of {\sf IQP}. Furthermore, in Ref.~\cite{WeakSim}, it is pointed out that there exists the case where the joint probability distribution is easily calculated but its marginal is rather hard to calculate. In order to construct an efficient simulation of {\sf IQP}, we need the marginal distributions of the output of {\sf IQP}, which allows a recursive simulation of the sampling problem. To this end, we map not only the joint probability distribution but also the marginal distributions of {\sf IQP} into the partition functions of certain Ising models. In the proof, we virtually utilize measurement-based quantum computation (MBQC)~\cite{MBQC} on graph states~\cite{GraphState}, which are defined associated with the {\sf IQP} circuits. The established relationship is useful since computational complexity of Ising models have been well studied. We can apply preexisting knowledge to understand quantum computational complexity of {\sf IQP}. In the second result, we provide classical simulatable classes of {\sf IQP} by using exact solvability of certain types of Ising models. We provide two examples of classically simulatable classes of {\sf IQP}. One is based on the sparsity of the commuting gates. Another is a class of {\sf IQP} that consists only of two-qubit commuting gates acting on nearest-neighbor qubits in 2D, which we call planar-{\sf IQP}. Planar-{\sf IQP} is mapped into a two-body Ising model on a 2D planar lattice without magnetic fields, which is known to be solvable by using the Pfaffian method~\cite{Barahona,Kasteleyn,Fisher}. In the proof, we also utilize properties of graph states in order to renormalize random $i\pi/2$ magnetic fields into two-body interactions, which originated from the random nature of the measurements. Then arbitrary marginal distributions can be efficiently calculated irrespective of their rotational angles by using the Pfaffian method \cite{Barahona,Kasteleyn,Fisher}. On the other hand, {\sf IQP} consisting of two- and single-qubit commuting gates are known to be universal under postselection~\cite{IQP}. Here the universality of {\sf IQP} means that it can simulate arbitrary quantum circuits of a uniform family with the help of postselection. (Hereafter a uniform family of quantum circuits is simply denoted by a ``quantum circuit".) Such a class of {\sf IQP} is referred to as {\sf IQP}${}_{\sf UP}$. This fact and the above classically simulatable class imply that single-qubit rotations play a very important role for {\sf IQP} to be classically intractable. Actually single-qubit rotations make a drastic change of complexity from strongly simulatable to not simulatable even in the weak sense. A similar result is also obtained for Toffoli-Diagonal circuits, where a final round of Hadamard gates plays very important role~\cite{WeakSim}. When experimentalists utilize {\sf IQP} for the purpose of verification of quantum benefits in malicious experimental setups, they should avoid the {\sf IQP} circuits of these classically simulatable classes, since a malicious experimental device can cheat experimentalists by a classical sampling instead of implementing the {\sf IQP} circuits. In most cases, however, experimental setups are well organized and not so malicious, and hence it might be possible to use these classically simulatable classes as an efficient benchmark of {\sf IQP}. In the above classically simulatable class, the probability distribution is given by the determinant, i.e., square of the Pfaffian, of a complex matrix. This result contrasts with {\sf B{\footnotesize OSON}S{\footnotesize AMPLING}} related with the permanent of a complex matrix. The exact solvability with the determinant (Pfaffian) naturally reminds us free-fermionic models, which have been also studied in standard quantum computation as match gates~\cite{matchgates0,matchgates1,matchgates2,matchgates3,matchgates4}. Since a determinant can be mapped into a probability amplitude of a free-fermionic system, the classically simulatable class of {\sf IQP} can be regarded as {\sf F{\footnotesize ERMION}S{\footnotesize AMPLING}} discussed in Ref.~\cite{Aaronson13}. This suggests that the sampling problems in physics can be classified in a unified way as sampling problems of elementary particles. In the final result, we apply the first result in an opposite direction, from quantum complexity to classical one. We consider some instances of ${\sf IQP}_{\sf UP}$ to understand classical complexity of calculating the corresponding Ising partition functions. Unless the {\sf PH} collapses at the third level, classical simulation of {\sf IQP}${}_{\sf UP}$ is hard even in the weak sense. We show that the {\sf IQP} circuits associated with Ising models with almost all imaginary coupling constants on a 2D planar graph with a bounded degree are included in {\sf IQP}${}_{\sf UP}$. By using this fact, we show that there is no fully-polynomial randomized approximation scheme (FPRAS) for such a class of Ising models, unless the {\sf PH} collapses at the third level. Moreover, for almost all imaginary coupling constants, a multiplicative approximation of such a class of Ising partition functions is at least as hard as a multiplicative approximation for the output distribution of an arbitrary quantum circuit. By considering the fact that strong simulation of a quantum circuit is \#{\sf P}-hard \cite{WeakSim}, this results indicates that a multiplicative approximation of Ising partition functions is highly intractable for almost all imaginary coupling constants even on a 2D planar graph with a bounded degree. The rest of the paper is organized as follows. In Sec.~\ref{sec2}, we introduce the definition and useful properties of the graph states with fixing notations. We also review {\sf IQP} and its measurement-based version, {\sf MBIQP}. In Sec.~\ref{sec3}, we establish a relationship between {\sf IQP} and Ising partition functions, where we frequently use the properties of the graph states. In Sec.~\ref{sec4}, we demonstrate classically simulatable classes of {\sf IQP}. One is based on the sparsity of the {\sf IQP} circuits. Another is based on exact solvability of the corresponding Ising models. In Sec.~\ref{sec5}, we apply the relationship between {\sf IQP} and Ising partition functions in an opposite direction to investigate (im)possibility of classical efficient approximation schemes for Ising models with imaginary coupling constants. Finally we summarize our results and related problems, which have not been addressed, in Sec.~\ref{sec6}. \section{Preliminary} \label{sec2} In the proofs of the main theorems, we work with a measurement-based version of {\sf IQP}, {\sf MBIQP}, introduced by Hoban {\it et al.}~\cite{Hoban}. The reason is that transformations on the resource state for MBQC~\cite{MBQC}, so-called graph states~\cite{GraphState}, are much easier and more intuitive than those for circuit themselves. Here we introduce the definition and useful properties of graph states with fixing the notations. \subsection{Basic Notations} The Pauli matrices on the $i$th qubit is denoted by $A_i$ ($A=I,X,Y,Z$). The Hadamard gate is denoted $H$. The eigenstates of $Z$ with eigenvalues $+1$ and $-1$ are denoted by $|0\rangle$ and $|1\rangle$, respectively. The eigenstates of $X$ with eigenvalues $+1$ and $-1$ are denoted by $|+\rangle$ and $|-\rangle$, respectively. We denote the controlled-$A$ gate acting on the $i$th (control) and $j$th (target) qubits by $\Lambda _{i,j}(A) = |0\rangle \langle 0| \otimes I + |1\rangle \langle 1| \otimes A$. Specifically, $\Lambda _{i,j}(Z)=\Lambda _{j,i}(Z)$ and $H_j\Lambda _{i,j} (Z) H_j = \Lambda _{i,j}(X)$. \subsection{Graph states} \begin{defi}[Graph state] \label{graphstate} For a given graph $G(V,E)$ with sets $V$ and $E$ of vertices and edges, an operator $K_i= X_i \prod _{j \in \mathcal{N}_i}Z_j $ is defined for each vertex $i$, where $\mathcal{N}_i$ indicates a set of vertices adjacent to vertex $i$. The graph state $|G\rangle$ is defined as the simultaneous eigenstate of the operator $K_i$ with eigenvalue $+1$ for all $i$: \begin{eqnarray*} K_i |G\rangle = |G\rangle. \end{eqnarray*} \end{defi} The above relation reads that the graph state $|G\rangle$ is stabilized by the operator $K_i$ for all $i$. Such a state is called a stabilizer state. The operator $K_i$ that stabilizes the stabilizer state is called a stabilizer operator. More details of the stabilizer formalism can be found in Ref.~\cite{Gottesman,NielsenChuang}. The graph state $|G\rangle$ is generated from a tensor product state of $|+\rangle$ by performing $\Lambda _{i,j}(Z)$ for all edge $(i,j) \in E$: \begin{eqnarray*} |G\rangle = \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right) |+\rangle^{ \otimes |V|}. \end{eqnarray*} This can be confirmed as follows. The product state $|+\rangle ^{ \otimes |V|}$ is the eigenstate of $X_i$ with eigenvalue $+1$ for all $i \in V$, and hence $X_i | + \rangle ^{ \otimes |V|}= |+ \rangle ^{ \otimes |V|}$. By applying $\prod _{(i,j)\in E} \Lambda_{i,j} (Z)$ for both sides, we obtain \begin{eqnarray*} \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right) X_i | + \rangle ^{ \otimes |V|} &=& \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right)|+ \rangle ^{ \otimes |V|} \\ X_i \prod _{j \in \mathcal{N}_i} Z_j \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right)| + \rangle ^{ \otimes |V|} &=& \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right)|+ \rangle ^{ \otimes |V|}, \end{eqnarray*} where we used the fact that $\Lambda _{i,j}(Z) X_i = X_i Z_j \Lambda _{i,j}(Z)$. This is the definition of the graph state, and we conclude $|G\rangle = \left( \prod _{(i,j)\in E} \Lambda_{i,j} (Z) \right)|+ \rangle ^{ \otimes |V|}$. In the proofs of the main theorems, we repeatedly consider projective measurements on the graph state and the resultant state transformations. In the following we will see two important transformation rules for the graph states by projective measurements in certain bases. \begin{remk}[$Z$-basis measurement] \label{Zmeasurement} If the $k$th qubit of the graph state $|G\rangle$ is measured in the $Z$-basis, the resultant post-measurement state is the graph state associated with the graph $G' \equiv G \backslash k$, where the byproduct operator $B_k=\prod _{j \in \mathcal{N}_k} Z_j$ is located according to the measurement outcome $m_k \in \{ 0,1\}$, i.e., $B_k ^{m_k} |G'\rangle$. \end{remk} {\it Proof}: We observe the effect of the measurement on the stabilizer operator $K_i$. If $i\neq k$ nor $i \neq \mathcal{N}_k$, the measurement does not make any effect on a stabilizer $K_k$, and hence the post-measurement state is stabilized by such a $K_k$. If $i=k$, $K_i$ anticommutes with $Z_k$ and hence does not stabilize the post-measurement state anymore. Instead, $(-1)^{m_k}Z_k$ stabilizes the post-measurement state $|m_k\rangle _k$, where $m_k=0,1$ is the measurement outcome. If $i\in \mathcal{N}_k$, we define a new stabilizer operator $K'_{i} = Z_k K_i$ such that $K_k$ does not contain $Z_k$. The post-measurement state is stabilized by $(-1)^{m_k}K'_{i}$. Thus the graph state with the byproduct operator, $B_k^{m_k}|G' \rangle$, is the post-measurement state. (Note that $B_k^{m_k}$ anticommutes with $K'_i$s for all $i$ but commutes with $K_i$s with $i \neq k$ and $i \notin \mathcal{N}$.) \hfill $\square$ Intuitively, the $Z$-basis measurement on the $k$th qubit remove the $k$th qubit from the graph state, and then the byproduct operator $B_k$ is located according the measurement outcome $m_k$. Next we consider a projective measurement on the $k$th qubit in the $\{|\theta _{k,m_k} \rangle \equiv X^{m_k} (e^{ - i \theta _k} |+\rangle + e^{ i \theta _k } |- \rangle )/\sqrt{2} \}$ basis, where $m_k \in \{0,1\}$ is the measurement outcome. \begin{remk}[Remote $Z$-rotation] \label{remote-Z} The projective measurement of the $k$th qubit on the graph state $|G\rangle$ in the $\{ | \theta _{k,m_k} \rangle \}$ basis results in \begin{eqnarray*} \exp\left[ i (\theta _k+ m_k\pi /2) \left( \prod _{j \in \mathcal{N}_k} Z_j \right) \right] | G\backslash k \rangle/\sqrt{2}. \end{eqnarray*} \end{remk} {\it Proof}: By using the fact that \begin{eqnarray*} |G\rangle = \left(\prod _{j \in \mathcal{N}_k} \Lambda _{k,j} (Z) \right) |+\rangle _k |G \backslash k\rangle , \end{eqnarray*} we can calculate the projection as follows: \begin{eqnarray*} \langle \theta _{k,m_k } |_k| G\rangle &=& \langle \theta _{k,m_k } |_k\left(\prod _{j \in \mathcal{N}_k} \Lambda (Z)_{kj} \right) |+\rangle _k| G \backslash k \rangle \\ &=&\langle +|_k e^{ i( \theta _k + m_k \pi/2) Z_k } H_k \left(\prod _{j \in \mathcal{N}_k} \Lambda (Z)_{kj} \right) |+\rangle _k| G \backslash k \rangle \\ &=& \left[\cos(\theta _k + m_k \pi/2) I + i \sin (\theta _k + m_k\pi/2)\left(\prod _{j \in \mathcal{N}_k} Z_j \right) \right] | G \backslash k \rangle /\sqrt{2} \\ &=& \exp\left[ i (\theta _k+m_k \pi /2) \left( \prod _{j \in \mathcal{N}_k} Z_j \right) \right] | G\backslash k \rangle/\sqrt{2}. \end{eqnarray*} \\ \hfill $\square$. The measurement in the $\{ | \theta _{k,m_k} \rangle \}$ basis induces a multi-body $Z$ rotation on the qubits adjacent to the $k$th qubit. The norms of the post-measurement states are both $1/2$, which indicates that the outcomes $m_k=0,1$ appear randomly. Another class of measurements, which is frequently used in MBQC, is the measurement in $\{ e^{i \theta Z} |\pm \rangle \}$ basis. It is known that adaptive measurements in this basis on a certain graph state is enough to perform universal quantum computation, i.e., {\sf BQP}~\cite{MBQC}. Here the adaptive measurement means changing the following measurement angles according to the previous measurement outcomes to cancel the effect of randomness originated from the projective measurements. This process is often called feedforward. There have been known a wide variety of graph states that are universal resource for MBQC~\cite{GraphState}. \subsection{Instantaneous quantum polynomial-time computation} Here we introduce {\sf IQP} and its measurement-based version. We first define {\sf IQP}: \begin{defi}[{\sf IQP} by Bremner et al.~\cite{IQP0,IQP}] Let $n$ be a number of qubits. We refer to a poly($n$) number of commuting gates \begin{eqnarray*} D(\theta _j,S_j)\equiv \exp\Big[i\theta_j\prod_{k \in S_j }Z_k\Big], \end{eqnarray*} including the input state $|+\rangle ^{\otimes n}$ and the $X$-basis measurements as an {\sf IQP} circuit. Here $\theta_j \in [0, 2 \pi)$ is a real number meaning the rotational angle, and $\{ S_j\}$ is a set of subsets of $\{ 1,2,...n\}$ on which the commuting gates act on. {\sf IQP} is defined as a sampling problem from the {\sf IQP} circuit, whose probability distribution is given by \begin{eqnarray*} P_{IQP} ( \{ s_i \} | \{ \theta _j\}, \{ S_j \} ) \equiv \left| \bigotimes _{i=1}^{n} \langle +_{s_i} | \prod_{j} D(\theta _j,S_j) |+\rangle ^{\otimes n} \right|^2, \end{eqnarray*} where $s_i \in \{0, 1\}$ is the measurement outcome and $|+_{s_i} \rangle= Z^{s_i} |+\rangle$. \end{defi} For each commuting circuit, we can naturally define a bipartite graph $G(V_A \cup U_B, E)$, where $V_A$ and $U_B$ are disjoint sets of vertices, and every edge $\in E$ connects a vertex in $V_A$ to one in $U_B$. Each vertex $v_i \in V_A$ is associated with the $i$th qubit, and hence $|V_A|=n$. Each vertex $u_j \in U_B$ is associated with the $j$th commuting gate $D(\theta _j,S_j)$, and hence $|U_B|= \textrm{poly($n$)}$. The set of edge $E$ is defined as $E:=\{ (u_j, v_i) | u_j \in U_B, i \in S_j \}$, that is, the set $S_j$ specifies the vertices $v_i$ that are connected with the vertex $u_j$. For a given weighted bipartite graph $G(V_A \cup U_B, E)$, where weights $\{\theta _j\}$ is defined on the vertices in $U_B$, we can define an {\sf IQP} circuit. By using Definition~\ref{graphstate} and Remark~\ref{remote-Z}, {\sf IQP} can be rewritten as MBQC on a graph state $|G\rangle$ associated with the graph $G(V_A \cup U_B, E)$. In this case, the set $\mathcal{N}_{u_j}$ of vertices corresponds to $S_j$. More precisely, for a given bipartite graph state $G(V_A\cup U_B,E)$ and weights $\{ \theta _j\}$, measurement-based {\sf IQP} ({\sf MBIQP}) is defined as follows: \begin{defi}[MBIQP by Hoban et al.~\cite{Hoban}] {\sf MBIQP} is defined as a sampling problem according to the probability distribution \begin{eqnarray*} P_{MBIQP}(\{m_{v_i}\}, \{ m_{u_j} \} | \{ \theta _{j} \}, G) \equiv \left|\bigotimes _{v_i \in V_A} \langle + _{m_{v_i}}| \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | | G\rangle \right|^2, \end{eqnarray*} where $m_{v_i} \in \{0,1\}$, $m_{u_j} \in \{0,1\}$ and $| \theta _{j,m_{u_j}} \rangle\equiv X^{m_{u_j} } (e^{-i \theta _{j}} |+ _0\rangle + e^{i \theta _{j}}|+_1\rangle)/\sqrt{2}$. \end{defi} The bit strings $\{ m_{v_i}\}$ and $\{ m_{u_j}\}$ correspond to the measurement outcomes on the qubits belonging to $V_A$ and $U_B$, respectively. We should note that there is no temporal order in the measurements since there is no feedforward of the measurement angles in {\sf MBIQP}. Then we can prove {\sf MBIQP}={\sf IPQ}. \begin{remk}[{\sf MBIQP} = {\sf IQP} by Hoban et al.~\cite{Hoban}] {\sf MBIQP} and {\sf IQP} are equivalent in the sense that if one sampler exists, another sampler can be simulated. \end{remk} {\it Proof:} Since a stabilizer operator of the graph state is given by $K_{u_j} = X_{u_j} \prod _{v_i \in \mathcal{N}_{u_j}} Z_{v_i} $, $K_{u_j} | G\rangle =| G\rangle$ for each vertex $u_j \in U_B$. By using this equality, we obtain \begin{eqnarray} P_{MBIQP}(\{m_{v_i}\}, \{ m_{u_j} \} | \{ \theta _{j} \},G) &=& \left| \bigotimes _{v_i \in V_A} \langle +_{m_{v_i}} | \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | \left( \prod _{u_j \in U_B} K_{u_j}^{m_{u_j}}\right) | G\rangle \right|^2 \nonumber \\ &=& \left| \bigotimes _{v_i \in V_A} \langle + _{m_{v_i}} | \bigotimes _{u_j \in U_B} \langle \theta _{{j},0} | \left[ \prod _{u_j \in U_B} \left (\prod _{v_i \in \mathcal{N}_{u_j}} Z_{v_i} \right)^{m_{u_j}} \right] | G\rangle \right|^2 \nonumber \\ &=& 2^{- |U_B|}P_{IQP} (\{ s_{i} \} | \{ \theta _{j} \} , \{S_j \}) \label{MBIQP} \end{eqnarray} where $m_{v_i}$ and $s _{i}$ are related via \begin{eqnarray*} s_{i} \equiv m_{v_i} \oplus \left( \bigoplus _{u_j \in \mathcal{N}_{v_i}} m_{u_j} \right). \end{eqnarray*} We used the facts that each measurement outcome $\{ m_{u_j} \}$ is randomly distributed with probability $1/2$, and the projection $\langle \theta _{j,0}|$ results in $D(\theta _j,S_j)$ (see Remark~\ref{remote-Z}). The above equality means that, regardless of the measurement outcomes $\{ m_{v_i} \}$ and $\{ m_{u_j}\}$, we can simulate {\sf IQP} by using {\sf MBIQP}. On the other hand, by using a random bit string $\{m_{u_j}\}$ with an equal probability 1/2 for each bit and $\{ s_{i}\}$ sampled from the {\sf IQP} circuit, we obtain $\{ m_{v_i} \equiv s_{i} \oplus _{u_j \in \mathcal{N}_{v_i}} m_{u_j} \}$ and $\{m_{u_j}\}$, which is equivalent to the output of {\sf MBIQP}. \hfill $\square$ As mentioned previously, there is no feedforward for the measurement angles in {\sf MBIQP}, and hence the measurements can be done simultaneously. This means that we cannot perform universal quantum computation in {\sf MBQIP} unless {\sf BQP} can be simulated by a quantum circuit of a constant depth. However, if postselection is allowed, we can choose the measurement outcomes in such a way that no byproduct operator is applied. Thus, with appropriately chosen graph structure and weights, we can simulate universal quantum computation without any feedforward if postselection is allowed. This means that postselected-{\sf MBIQP} (post-{\sf MBIQP}) with an appropriate graph state and weights (measurement angles) is equivalent to post-{\sf BQP}. Thus we conclude that post-{\sf MBIQP} = post-{\sf IQP} = post-{\sf BQP}. On the other hand, Aaronson showed that post-{\sf BQP}={\sf PP}~\cite{postBQP}. Accordingly, post-{\sf IQP}=post-{\sf MBIQP} = {\sf PP}. In order to simulate universal quantum computation (i.e., {\sf BQP}), it is sufficient to consider post-{\sf IQP} or post-{\sf MBIQP} associated with planar bipartite graphs $G(V_A\cup U_B, E)$ with $|S_j|\leq 2$ and $\theta _j = \pi /8$ for all $j$~\cite{IQP}. (As shown in Sec.~\ref{sec5}, we can obtain the same result not only for $\theta _j =\pi/8$ but also for almost all angles $\theta _j$.) Such a universal class of {\sf IQP} that can simulate an arbitrary quantum circuit with postselection is referred to as ${\sf IQP}_{\sf UP}$ ({\sf UP} stands for universal with postselection). In this case, quantum algorithms are determined according to structures of the graphs. We can also fix the structure of the graph and choose the angle $\theta _j$ from $\{ \pi /4, \pi /8, 0\}$. Specifically, $\theta _j=0$ corresponds to a deletion of vertex $u_j$ from the graph (see Remark~\ref{Zmeasurement}). $\theta _j = \pi /4$ and $\pi /8$ correspond to Clifford and non-Clifford gates, respectively. Examples of graphs and weights of {\sf MBIQP} that are universal under postselection are presented in Fig.~\ref{fig0-1} (a) and (b). \begin{figure}[t] \begin{center} \includegraphics[width=0.8\textwidth]{fig0-1.eps} \end{center} \caption{(a) An example of a planar bipartite graph and weights for universal {\sf IQP}, where $\theta _j \in \{0, \pi /8, \pi/4\}$. (2) Another example of a planar bipartite graph and wights, where $\theta _j=\pi /4$ for all $j$ with $|S_j|=2$ (corresponding to two-qubit commuting gates) and $\theta _j \in \{ 0,\pi/8\}$ for all $j$ with $|S_j|=1$ (corresponding to single-qubit rotations). The associated graph state is a decollated version of the brickwork state utilized in blind quantum computation~\cite{Blind,MoriBlind}. Each dotted square indicates a unit cell of the brickwork state. The brickwork state allows us to perform universal quantum computation with measurements only in $\{ |\pm \rangle\}$ and $\{e^{ i (\pi /8) Z}|\pm \rangle\}$ bases. } \label{fig0-1} \end{figure} Bremner, Jozsa, and Shepherd showed that if {\sf IQP} can be simulated efficiently by using a classical randomized algorithm within multiplicative approximation error, we can solve all problems in {\sf PP} by post-{\sf BPP}. Here the multiplicative error means that the approximated distribution, say $P^{ap}_{IQP}$, lies within \begin{eqnarray*} \frac{1}{c} P_{IQP} \leq P^{ap}_{IQP} \leq c P_{IQP}, \end{eqnarray*} with a certain constant $c$. \begin{remk}[Hardness of {\sf IQP} by Bremner et al.~\cite{IQP}] \label{IQP} If {\sf IQP} can be simulated by a classical polynomial randomized algorithm within multiplicative error $1\leq c \leq \sqrt{2}$, {\sf PP} $\subseteq$ post-{\sf BPP}. \end{remk} {\it Proof}: (See also Ref.~\cite{IQP}.) Let us utilize post-{\sf IQP}, sampling problems with postselection, to solve a decision problem with bounded error. Let $L$ be an arbitrary language decided by post-{\sf IQP} with bounded error, that is, \begin{eqnarray} \textrm{when the instance is in $L$: } P_{IQP}[ \textrm{yes} | \textrm{postselection}] \geq 1/2 +\delta \label{decision1} \\ \textrm{when the instance is not in $L$: } P_{IQP}[ \textrm{yes} | \textrm{postselection}] \leq 1/2 -\delta \label{decision2} \end{eqnarray} for some $0 < \delta < 1/2$. Here $P_{IQP}[x|y]$ indicates that the conditional probability distribution of the decision $x \in \{ \textrm{yes, no}\}$ conditioned on the register $y$ for the postselection. Suppose we have a polynomial-time approximation $P^{ap}_{IQP}$ of $P_{IQP}$ with a multiplicative error $c$. The conditional distribution with postselection can be calculated as \begin{eqnarray*} P^{ap}_{IQP}[ x | \textrm{postselection}]=\frac{P^{ap}_{IQP} [x, \textrm{postselection}]}{P^{ap}_{IQP} [\textrm{postselection}]}. \end{eqnarray*} Thus the multiplicative error for $P^{ap}_{IQP}[ x | \textrm{postselection}]$ becomes $c^2$: \begin{eqnarray*} \frac{1}{c^2} P_{IQP}[x|\textrm{postselection}] \leq P^{ap}_{IQP}[x |\textrm{postselection}] \leq c^2 P_{IQP}[x |\textrm{postselection}]. \end{eqnarray*} By using this and Eqs. (\ref{decision1}) and (\ref{decision2}), we obtain \begin{eqnarray*} \textrm{when the instance is in $L$: } P^{ap}_{IQP}[\textrm{yes}|\textrm{postselection}]\geq \frac{1}{c^2} (1/2+\delta) \\ \textrm{when the instance is not in $L$: } P^{ap}_{IQP}[\textrm{yes}|\textrm{postselection}]\leq c^2 (1/2-\delta) \end{eqnarray*} Thus if both $c^{-2} (1/2+\delta ) > 1/2$ and $c^{2} (1/2-\delta ) <1/2$ are satisfied, we obtain the correct answer with a probability higher than $1/2$. Since post-{\sf IQP} does not depend on the level of error $\delta$, we can choose any value $\delta <1/2$. Then we conclude that if $c<\sqrt{2}$, $L$ is decided in post-{\sf BPP}. Since post-{\sf IQP}=post-{\sf BQP}={\sf PP}, this implies {\sf PP} $\subseteq$ post-{\sf BPP}. \hfill $\square$ Due to Toda's theorem~\cite{Toda}, {\sf P} with an oracle for {\sf PP} includes whole classes in the {\sf PH}, i.e., {\sf PH} $\subseteq$ {\sf P}${}^{{\sf PP}}$. On the other hand, {\sf P} with an oracle for post-{\sf BPP} is in the third level of the {\sf PH}, i.e, P${}^{\rm post-{\sf BPP}} \subseteq \Delta _3$. Thus {\sf PP} $\subseteq$ post-{\sf BPP} (i.e., the existence of a polynomial classical approximation scheme of {\sf IQP} with multiplicative error $c <\sqrt{2}$) implies collapse of the {\sf PH} at the third level, which is highly implausible. In other words, unless the {\sf PH} collapses at the third level, {\sf IQP} cannot be simulated efficiently in a classical computer. \subsection{Strong and weak simulations of quantum circuits} Here we provide definitions of two important notions of simulation of quantum circuits, strong and weak simulations~\cite{WeakSim,JozsaNest}. \begin{defi}[Strong and weak simulations] Let $\mathcal{U}$ and $P(\mathcal{U})$ denote a quantum circuit of a uniform family and the probability distribution by computational basis measurements, respectively. Strong simulation of a quantum circuit $\mathcal{U}$ is calculation of the probability distribution $P(\mathcal{U})$. Weak simulation of a quantum circuit $\mathcal{U}$ is to sample the output (measurement outcomes) according to the probability distribution $P(\mathcal{U})$. \end{defi} In addition to these notions of simulation, we can further consider both exact and approximated simulations. In an approximated simulation with multiplicative error $c$, we replace the probability distribution $P(\mathcal{U})$ with its approximation $P^{ap}(\mathcal{U})$ that lies within the following approximation range \begin{eqnarray*} \frac{1}{c} P(\mathcal{U}) \leq P^{ap}(\mathcal{U}) \leq c P(\mathcal{U}). \end{eqnarray*} It has been known that exact strong simulation of a uniform family of quantum circuits is \#{\sf P}-hard~\cite{WeakSim} and hence much harder than weak simulation of them or {\sf BQP}, which are solved by a universal quantum computer itself. \section{Bridging {\sf IQP} and Ising partition functions} \label{sec3} In this section, we establish a bridge between {\sf IQP} and partition functions of Ising models. We will first show that the joint probability distribution of the output of an {\sf IQP} circuit associated with a graph $G$ is given by normalized squared norm of the partition function of the Ising model defined by the graph $G$. This is shown first by mapping {\sf IQP} into {\sf MBIQP} and then by relating an inner product between a product state and the graph state $|G\rangle$ with the partition function of the Ising model following the overlapping map~\cite{VandenNest07}. However, this is not sufficient for our purpose. Since there are exponentially many instances of the measurement outcome, a straightforward sampling by using the joint probability distribution is not available. Instead, we simulate {\sf IQP} in a recursive way, where the next measurement outcome is sampled according to the conditional distribution on the previous measurement outcome. To this end, we need the marginal distributions with respect to the measured qubits. If the marginal distribution can be calculated efficiently, the recursive method succeeds to simulate a sampling according to the joint probability distribution of {\sf IQP} efficiently. In this section, we will also establish a relationship between the marginal distribution with respect to a set $M$ of the measured qubits and the partition function of the Ising model defined by another graph $\tilde G_M$, which is systematically constructed from the graph $G$ and the set $M$. \subsection{Joint probability distribution} We define an Ising model, which may include multibody interactions, according to the bipartite graph $G(V_A\cup U_B,E)$ and weights $\{\theta _j\}$. The Ising model consists of the sites associated with the vertices $v_i \in V_A$ and multibody interactions assigned for the vertices $u_j \in U_B$. The spins engaged in the $j$th interaction and its coupling constant are given by $\mathcal{N}_{u_j}$ (or equivalently $S_j$) and $\theta _j$, respectively. \begin{defi}[Multibody Ising Model with random $i\pi/2$ magnetic fields] For a given bipartite graph $G(V_A\cup U_B, E)$ and weights $\{\theta _j\}$ defined on the vertices in $U_B$, a Hamiltonian of an Ising model with random $i \pi /2$ magnetic fields is defined by \begin{eqnarray} H(\{s_{i}\}, \{ \theta _{j} \}, G) \equiv - \sum_{v_i \in V_A} i\pi s_{i} \frac{1-\sigma _{v_i}}{2} - \sum _{u_j \in U_B} i \theta _{j}\left( \prod _{v_i \in \mathcal{N}_{u_j} } \sigma _{v_i} \right), \label{Isingmodel} \end{eqnarray} where $\sigma _{v_i} \in \{ +1,-1\}$ is an Ising variable defined associated with vertex $v_i \in V_A$. The partition function of the Ising model is defined by \begin{eqnarray*} \mathcal{Z}( \{s_{v_i}\}, \{ \theta _{j} \}, G) = \sum _{\{ \sigma _{v_i} \}} e^{ - H(\{s_{i}\}, \{ \theta _{j} \}, G)}, \end{eqnarray*} where $\sum _{\{ \sigma _{v_i} \}}$ means the summation over all configuration $\{\sigma _{v_i} \}$. \end{defi} We should note that, in addition to the interactions defined by the graph and weights, random $i \pi /2$ magnetic fields are also introduced according to the bit string $\{ s_{v_i}\}$. This corresponds to the measurement outcome of ${\sf IQP}$ as seen below. Furthermore, in Sec.~\ref{sec4}, these random $i\pi/2$ magnetic fields will be removed in a certain class of Ising models by renormalizing them into the coupling constants $\{ \theta _j \}$. The probability distribution of {\sf IQP} associated with $G(V_A\cup U_B,E)$ and weights $\{ \theta _j\}$ is now shown to be equivalent to the normalized squared norm of the partition function of Ising model defined by the graph $G$ and weights $\{ \theta _j\}$ as follows: \begin{theo}[{\sf IQP} and Ising partition functions] \label{Main1} {\sf IQP} associated with the graph $G(V_A\cup U_B,E)$ and weights $\{\theta _j\}$ is equivalent to the sampling problem according to the normalized squared norm of an Ising partition function defined by the graph $G$ and weights $\{ \theta _j\}$: \begin{eqnarray*} P_{IQP}(\{s_i \} | \{ \theta _{j} \}, \{ S_j \}) &=&2^{ |U_B|} P_{MBIQP} (\{m_{v_i}\}, \{ m_{u_j}\} | \{ \theta _{j} \}, G) \\ &=&2^{-2|V_A|} \left|\mathcal{Z}(\{ s_{i}\}, \{ \theta _{j} \}, G)\right|^2. \end{eqnarray*} \end{theo} {\it Proof:} We reformulate the left hand side of Eq. (\ref{MBIQP}) using the overlap mapping developed by Van den Nest, D{\"u}r, and Briegel~\cite{completeness,FujiiStat}: \begin{eqnarray} && P_{IQP}(\{s_i\} | \{ \theta _{j} \}, \{S_j \}) \nonumber \\ &=&2^{|U_B|}P_{MBIQP}(\{m_{v_i}\}, \{ m_{u_j} \} | \{ \theta _{j} \},G) \nonumber \\ &=& 2^{|U_B|} \left| \left( \bigotimes _{v_i \in V_A} \langle + _{s_i} | \right) \left( \bigotimes _{u_j \in U_B} \langle \theta _{j,0} |H \right) \prod _{u_j \in U_B} H_{u_j}| G \rangle \right|^2 \nonumber \\ &=& 2^{|U_B|} \left| \left(\bigotimes _{v_i \in V_A} \frac{\langle 0| + e^{i s_i \pi } \langle 1| }{\sqrt{2}} \right) \left( \bigotimes _{u_j \in U_B} \frac{\langle 0 | e^{ i \theta _{j}} +\langle 1| e^{- i \theta _{j}}}{\sqrt{2}} \right) \left( 2^{-|V_A|/2}\sum _{\{ \bar \sigma _{v_i} \} } |\{ \bar \sigma _{v_i} \} \rangle \bigotimes_{u_j\in U_B}\left| \bigoplus _{ v_i \in N_{u_j} } \bar \sigma _{v_i} \right\rangle \right) \right|^2 \nonumber \\ &=& 2^{|U_B|} \left| 2^{-|U_B|/2-|V_A|} \sum _{\{ \bar \sigma _{v_i} \} } \exp\left[\sum_{v_i \in V_A} i\pi s_{i} \bar \sigma _{v_i} \right] \exp \left[ \sum _{u_j \in U_B } -i \left[2 \theta _{j} \left( \bigoplus _{ v_i \in N_{u_j} } \bar \sigma _{v_i} \right) - \theta _j \right] \right] \right|^2 \nonumber \\ &=& 2^{-2|V_A| } \left|\sum _{ \{ \sigma _i \}} e^{- H(\{s_{i}\}, \{ \theta _{j} \},G) } \right|^2 \nonumber \\ &= & 2^{-2|V_A| } \left|\mathcal{Z}(\{s_{i}\}, \{ \theta _{j} \},G)\right|^2, \label{Mapping} \end{eqnarray} where we define a binary variable $\bar \sigma _{v_i} \equiv (1-\sigma _{v_i})/2$, and $\sum _{\bar \sigma _{v_i}}$ indicates a summation over all binary strings. From the second to the third lines, we used the fact that \begin{eqnarray*} |G\rangle &=& \left( \prod _{u_j \in U_B} \prod _{v_i \in \mathcal{N}_{u_j} } \Lambda _{v_i,u_j}(Z) \right) |+\rangle ^{\otimes |V_A|} |+\rangle ^{\otimes |U_B|} \\ &=& \left( \prod _{u_j \in U_B} H_{u_j}\right) \left( \prod _{u_j \in U_B} \prod _{v_i \in \mathcal{N}_{u_j} } \Lambda _{v_i,u_j}(X) \right) \sum _{\{ \bar \sigma _{v_i}\}} | \{ \bar \sigma _{v_i}\} \rangle | 0 \rangle ^{ \otimes |U_B|} \\ &=& \left( \prod _{u_j \in U_B} H_{u_j}\right) \sum _{\{ \bar \sigma _{v_i}\}} | \{ \bar \sigma _{v_i}\} \rangle \bigotimes _{u_j \in U_B} | \oplus _{v_i \in \mathcal{N}_{u_j} } \bar \sigma _{v_i } \rangle. \end{eqnarray*} \hfill $\square$ \\ Equation (\ref{Mapping}) shows that {\sf IQP} is equivalent to the sampling problem according to the probabilities proportional to the squared norm of the partition functions of an Ising model with imaginary coupling constants. Note that the measurement outcome $\{ s_i \}$ correspond to the random $i \pi /2$ magnetic fields. The present sampling problem is not related directly to what is well studied in the fields of statistical physics, such as the Metropolis sampling according to the Boltzmann distribution. However, as we will see below, the relation between {\sf IQP} and Ising partition functions leads us to several interesting results about complexity of {\sf IQP}, since calculation of the Ising partition functions are well studied in both fields of statistical physics and computer science. It was shown in Ref.~\cite{Barahona} that exact calculation of partition functions of two-body Ising models with magnetic fields even on the planar graphs is NP-hard. Furthermore, in general, exact calculation of partition functions of two-body Ising models with magnetic fields is \#{\sf P}-hard~\cite{Jerrum93}. No polynomial approximation scheme with multiplicative error exists unless {\sf NP}={\sf RP}. While {\sf IQP} does not provide the exact values of the partition functions, it is surprizing that the sampling according to the partition functions of many-body Ising models $H(\{s_{v_b}\}, \{ \theta _{v_a} \},G)$ with imaginary coupling constants, can be done in {\sf IQP}, which consists only of commuting gates and seems much weaker than {\sf BQP}. Only in the limited cases, partition functions of Ising models can be calculated efficiently. Such an example is two-body Ising models on planar lattices without magnetic fields. In the next section, we show that certain classes of {\sf IQP} are classically simulatable, by using the fact that the associated Ising models are exactly solvable. To this end, we need not only the joint distribution of the output of {\sf IQP} circuits but also marginal distributions with respect to measured qubits, in order to simulate the sampling problem recursively. \subsection{Marginal distribution} Even if we can calculate the probability distribution $P_{IQP}(\{ s_i \} | \{ \theta _j \} ,\{S_j\})$ efficiently, it does not directly mean that the corresponding {\sf IQP} is classically simulatable, since there are exponentially many instances of the measurement outcomes $\{s_i \}$. An efficient classical simulation of {\sf IQP} requires the marginal distribution with respect to measured qubits, by which we can simulate {\sf IQP} recursively. In the following we will establish a mapping between the marginal distribution with respect to the set $M$ of measured qubits and the partition function of an Ising model defined by a merged graph $\tilde G_{M}$. The merged graph $\tilde G_{M}$ constructed by merging a subgraph $G_M$ and its copy $G'_M$ corresponding to the measured part of the graph $G$ (see Fig.~\ref{fig3}). (The detailed definition of the subgraph $G_M$ and the merged graph $\tilde G_M$ are given in the proof of the following theorem.) \begin{theo}[Marginal distribution of {\sf IQP}] \label{Main2} Let $M \subset \{ 1,2,..., n\}$ and $\bar M \subset \{ 1,2,...,n\}$ be sets of the measured and unmeasured qubits, respectively (and hence $M \cup \bar M = \{ 1,2,...,n\}$ and $M \cup \bar {M} = \emptyset$). A marginal distribution with respect to the set $M $ \begin{eqnarray*} P_{IQP}(\{ s_ i\}_{i \in M}| \{ \theta _j \} , \{S_j\}, M ) \equiv \sum _{ \{ s_i \}_{ i \in \bar M} } P_{IQP}(\{ s_i \} | \{ \theta _j \} , \{ S_j\} ) \end{eqnarray*} is related to the partition function of the Ising model defined by the merged graph ${\tilde G_M}$ and weights $\{\theta _j\} \cup \{- \theta _j \}$. \end{theo} {\it Proof:} In order to prove this, we consider the corresponding {\sf MBIQP}. However, it is just for a proof, and hence we do not need to simulate {\sf MBIQP} in classical simulation as seen later. Thus without loss of generality, we can assume that the measurement outcome is subject to $m_{u_j}=0$ for all $u_j \in U_B$. \begin{figure}[t] \begin{center} \includegraphics[width=0.9\textwidth]{fig3.eps} \end{center} \caption{(a) The graph state $|G\rangle$ associated with the graph $G$. The gray and white circles indicate qubits associated with $v_i \in V_A$ and $u_j \in U_B$, respectively. (b) The subgraph state $|G_M\rangle$ and its copy $|G'_M\rangle$ are merged via the qubits $|+\rangle ^{ |\partial M_{AB}|}$ located on the boundary. The merged graph is denoted by $\tilde G_M$.} \label{fig3} \end{figure} Following the sets $M$ and $\bar M$, the sets of measured and unmeasured qubits in $V_A$ is defined as $M_A$ and $\bar M_A$, i.e., $M_A \cup \bar M_A =V_A$. We define a subgraph $G_M (M_A \cup M_B, E_M)$, where $M_B \subset U_B$ is a set of vertices that are connected with any vertices in $M_A$, i.e., $M_B=\{ u_j\in U_B | (u_j,v_i) \in E, v_i \in M_A \}$. $E_M$ is a set of edges whose two incident vertices both belong to $M_A\cup M_B$. We denote $M_A \cup M_B$ simply by $M_{AB}$ and $(V_A \cup U_B)\backslash M_{AB}$ by $\bar M_{AB}$ (see Fig.~\ref{fig3} (a)). The marginal distribution can be written as measurements on the reduced density matrix on the qubits $M_{AB}$: \begin{eqnarray*} P_{IQP}(\{ s_{i}\}_{i \in M}| \{ \theta _{j} \} , \{ S_j \} , M ) &=& \langle \Theta | {\rm Tr}_{\bar M_{AB}}\bigl[|G\rangle \langle G| \bigr] | \Theta \rangle, \end{eqnarray*} where $| \Theta \rangle \equiv \bigotimes _{v_i \in M_A} |+_{s_{i}} \rangle \bigotimes _{u_j \in M_B} |\theta _{j,0}\rangle $, and ${\rm Tr}_{\bar M_{AB}}$ indicates the partial trace with respect to the unmeasured qubits $\bar M_{AB}$. We define a subset $\partial M_{AB} \subset \bar M_{AB}$ as a set of vertices connected with any vertices in $M_{AB}$, i.e., $\partial M_{AB} = \{ v_i \in \bar M_{A} |(v_i, u_j) \in E, u_j \in M_{B} \}$ (note that $\partial M_{AB} \subset \bar M_A$). We refer to the qubits associated with the vertices in $\partial M_{AB}$ as the boundary qubits, since they are the boundary of the measured and unmeasured qubits in the graph state as shown in Fig.~\ref{fig3} (a). For the graph state $|G\rangle$, the trance out with respect to the unmeasured qubits $\bar M_{AB}$ can be equivalently done by $Z$ basis measurements on the boundary qubits and forgetting about the measurement outcomes. This is because, $Z$-basis measurements on the boundary qubits separate the measured and unmeasured qubits (see Remark~\ref{Zmeasurement}), and hence the tracing out of the qubits in $\bar M_{AB} \backslash \partial M^{AB}$ does not have any effect on the measured qubits $M_{AB}$. From this observation we obtain \begin{eqnarray*} {\rm Tr}_{\bar M_{AB}}\bigl[|G\rangle \langle G| \bigr] =2^{ -|\partial M_{AB}|} \sum _{\{m_{v_i} \}_{\partial M_{AB} }} \left(\prod _{v_i \in \partial M_{AB} } B({v_i})^{m_{v_i}} \right) |{G_M} \rangle \langle {G_M} |\left(\prod _{v_i \in \partial M_{AB} } B({v_i})^{m_{v_i}} \right) \end{eqnarray*} where $\{ m_{v_i} \}_{\partial M_{AB} }$ is the set of the measurement outcomes on the boundary qubits, and we define a byproduct operator $B(v_i) = \prod _{u_j \in \mathcal{N}_{v_i} \cap M_{AB}} Z_{u_j}$ (see Remark~\ref{Zmeasurement}). Let us consider a merged graph $\tilde G_M$ that is constructed from the graph $G_M$ and its copy $G'_M$, and the boundary $\partial M_{AB}$. Two copies of graph states, $|G_M\rangle$ and $|G'_M\rangle$, are merged via $|+\rangle^{\otimes |\partial M_{AB}|}$ as shown in Fig.~\ref{fig3} (b). The vertices in $\partial M_{AB}$ and those in $G_M$ and $G'_M$ are connected iff there is an edge between them in the original graph $G$ and its copy $G'$. The graph state associated with the merged graph $\tilde G_M$ is written as \begin{eqnarray*} |\tilde G_M \rangle = \prod _{v_i \in \partial M_{AB} } \left( \prod_{ u_j \in \mathcal{N}_{v_i}\cup M_{B} } \Lambda _{v_i,u_j} (Z) \prod_{ u'_j \in \mathcal{N}'_{v_i}\cup M'_{B} } \Lambda _{v_i,u'_j} (Z) \right) |G_M\rangle |+\rangle ^{\otimes |\partial M_{AB} |} | G'_M \rangle. \end{eqnarray*} Let us consider a projection of $|\tilde G_M\rangle$ by $|+\rangle ^{\otimes |\partial M_{AB} |}$: \begin{eqnarray*} \langle + | ^{ \otimes | \partial M_{AB} |} |\tilde G_M \rangle = 2^{- |\partial M_{AB}|} \sum _{\{ m_{v_i} \} _{\partial M_{AB} }} \left[ \prod _{v_i \in \partial M_{AB}} \left[B(v_i) {B'}(v_i) \right]^{ m_{v_i}}\right] |G_M\rangle | G'_M \rangle, \end{eqnarray*} where $B'(v_i)$ is defined similarly to $B(v_i)$ on the graph state $|G'_M\rangle$. Let us define \begin{eqnarray*} |\Theta ' \rangle \equiv \bigotimes _{v_i \in M_A} |+_{s_{i}} \rangle \bigotimes _{u_j \in M_B} |-\theta _{j,0}\rangle , \end{eqnarray*} where we should note that the sign of the angle $\theta _{j,0}$ is flipped. Next we consider a projection by $| \Theta \rangle | \Theta ' \rangle$ as follows: \begin{eqnarray} &&\langle \Theta | \langle + |^{ \otimes | \partial M_{AB} |} \langle \Theta ' | | \tilde G_M\rangle \nonumber\\ &=& 2^{- |\partial M_{AB}|} \sum _{\{ m_{v_i} \} _{\partial M_{AB} }} \langle \Theta |\left[ \prod _{v_i \in \partial M_{AB}} \left[B(v_i) \right]^{ m_{v_i}}\right] |G_M\rangle \langle \Theta ' | \left[ \prod _{v_i \in \partial M_{AB}} \left[B'(v_i) \right]^{ m_{v_i}}\right] | G'_M \rangle \nonumber \\ &=& \langle \Theta | {\rm Tr} _{\bar M_{AB} }\bigl[|G\rangle \langle G| \bigr] |\Theta \rangle \nonumber \\ &=& P_{IQP}(\{ s_{i}\}_{i \in M}| \{ \theta _{j} \} , \{S_j \}, M ). \label{overlap1} \end{eqnarray} This indicates that the summation over exponentially many variables for the marginalization is taken simply in an overlap between the product state and the merged graph state. On the other hand, the overlap $\langle \Theta | \langle + |^{ \otimes | \partial M_{AB} |} \langle \Theta ' | | \tilde G_M\rangle$ is also reformulated by the partition function of an Ising model as done in the proof of Theorem~\ref{Main1}. Specifically, the interaction patterns are given by the merged graph $\tilde G_M$. The coupling strengths are given by two copies of $\{ \theta _{j} \}_{u_j \in M_B}$ and $\{ -\theta _{j} \}_{u'_j \in M'_B}$: \begin{eqnarray} &&\langle \Theta | \langle + |^{ \otimes | \partial M_{AB} |} \langle \Theta ' | | \tilde G_M\rangle \nonumber \\ &=& 2^{-2|M_A|-|\partial M_{AB}| - |M_B|} \left| \mathcal{Z} ( \{ s_{i}\} _{M} \cup \{ 0 \}_{ v_i \in \partial M_{AB}} \cup \{s'_i \}_{M'} , \{\theta _{j} \}_{u_j \in M_B} \cup \{ -\theta _{j} \}_{u_j \in M'_B},\tilde G_M)\right|, \nonumber \\ &\equiv& 2^{-2|M_A|-|\partial M_{AB}|-|M_B|} \left| \mathcal{Z} ( \{ s_i\}^{*}, \{ \theta _{j} \}^{*},\tilde G_M)\right| \label{overlap2} \end{eqnarray} where we defined $\{ s_i \}^{*} \equiv \{ s_{i}\} _{M} \cup \{ 0 \}_{ v_i \in \partial M_{AB}} \cup \{s'_i \}_{M'} $ and $\{ \theta _j \}^{*}\equiv \{\theta _{j} \}_{u_j \in M_B} \cup \{ -\theta _{j} \}_{u_j \in M'_B}$. We should note that $s_i$ and $s'_i$ take the same value but $\theta _j$'s sign is flipped on its copy $u'_j \in M'_B$. From Eqs. (\ref{overlap1}) and (\ref{overlap2}), \begin{eqnarray*} P_{IQP}(\{ s_{i}\}_{i \in M}| \{ \theta _{j} \} , \{S_j \}, M ) &=&2^{-2|M_A|-|\partial M_{AB}|} \left| \mathcal{Z} ( \{ s_i\}^{*}, \{ \theta _{j} \}^{*},\tilde G_M)\right| \end{eqnarray*} That is, the marginal distribution with respect to the set $M$ of the measured qubits is given by the normalized squared norm of the partition function of the Ising model defined by the merged graph $\tilde G_M$. \hfill $\square$ The above theorem also indicates that the marginal distribution is equivalent to the square root of the joint probability of the {\sf IQP} circuit associated with the merged graph $\tilde G_M$, weights $\{ \theta _j \}^{*}$ and the measurement outcomes $\{ s_i\}^{*}$: \begin{eqnarray*} P_{IQP}(\{ s_{i}\}_{i \in M}| \{ \theta _{j} \} , \{S_j \}, M ) =\left[ P_{IQP}(\{ s_i\}^*,\{\theta _j\}^*, \{ \mathcal{N}_{u_j} | u_j \in \tilde G_M\} )\right]^{1/2}. \end{eqnarray*} This indicates that if the {\sf IQP} circuits associated with a class of graphs can be strongly simulated (in a complexity class), and the class of graphs is closed under merging mentioned above, then the marginal distributions of such a class of {\sf IQP} circuits can also be strongly simulated (in the same complexity class). An example of such a class is planar graphs, where the merged graph $\tilde G_{M^{(k)}}$ is also a planar graph with an appropriately chosen measurement order such that $M^{(k)}$ is always connected. This class of {\sf IQP} includes nontrivial problems, ${\sf IQP}_{\sf UP}$, which becomes as powerful as {\sf PP} under postselection. This property is very useful, and we employ several times in the following sections in order to prove our main claims. As noted in Ref.~\cite{WeakSim}, this property does not always hold in general quantum circuits. There is a case in which the joint probability can be efficiently strongly simulatable, but its marginal (or partial measurements) is highly intractable. Conditioned on the measurement outcome $\{ s_i \}_{i \in M}$ on the set $M$, the probability of obtaining the next measurement outcome $s_k$ is calculated by using the Bayse rule as \begin{eqnarray*} p(s_k | \{ s_i \} _{i \in M})=\frac{ P_{IQP}(\{ s_ i\}_{i \in M\cup k}| \{ \theta _j \} , \{S_j\}, M \backslash k ) }{ P_{IQP}(\{ s_ i\}_{i \in M}| \{ \theta _j \} , \{S_j\}, M ) }. \label{recursive} \end{eqnarray*} By denoting the set of all measured qubits after the $k$th measurements as $M^{(k)}$ (since there is no order in the measurements in {\sf IQP}, we can choose an arbitrary order of measurements for our convenience), we can reconstruct the joint probability distribution of {\sf IQP} as follows: \begin{eqnarray*} P_{IQP}( \{ s_i \} | \{\theta _j\} , \{ S_j \}) = \prod _{k=1}^{n} p( s_{i_k} | \{s_i \}_{i \in M^{(k)}} ), \end{eqnarray*} where the $i_k$th qubit is measured at step $k$, i.e., $\{i_k \} \cup M^{(k-1)} = M^{(k)}$. If the marginal distribution, that is, partition functions of the Ising models defined by $\tilde G_{M^{(k)}}$ can be calculated efficiently for all $M^{(k)}$ for any measurement order, {\sf IQP} is classically simulatable. \section{Classical simulatable classes of {\sf IQP}} \label{sec4} In general, exact calculation of partition functions of Ising models in the presence of magnetic fields is highly intractable in classical computer even on 2D planar lattice~\cite{Barahona,Jerrum93}. The Ising models, to which we have mapped {\sf IQP} in Sec.~\ref{sec3}, includes the random $i\pi/2$ magnetic fields depending on the output $\{ s_i\}$. Thus one might think that we cannot find a nontrivial class of {\sf IQP} that is classically simulatable. This is, however, not the case. Below we will show that if the geometries of the graphs have some properties, we can safely remove the magnetic fields renormalizing it into the coupling constants $\{ \theta _j\}$. In this section, we will provide two classes of {\sf IQP} that are classically simulatable efficiently in the strong sense, that is, the probability distribution including its marginals can be calculated efficiently. One is based on the sparsity of the commuting gates. The other is based on the exact solvability of 2D Ising models on planar lattices without magnetic fields~\cite{Kasteleyn,Barahona,Fisher}. In both cases, classical simulatability can be shown under arbitrary rotational angles $\{\theta _j\}$. \subsection{Classical simulatability: sparse commuting circuits} Let us define a $|V_A| \times |U_B|$ matrix $R$, associated with the bipartite graph $G(V_A \cup U_B,E)$, such that $R_{v_i}^{u_j} = 1$ iff a vertex $v_i \in V_A$ is in $\mathcal{N}_{u_j}$, otherwise $R_{v_i}^{u_j}=0$. We consider a class of bipartite graph, for which the row vectors of $R$ are linearly independent and full rank (and therefore $|V_A|=|U_B|$) in $\mathbf{Z}_2^{|U_B|}$ (later we will weaken the latter condition). This condition implies that the column vectors of $R$ are also linearly independent and full rank. We call such a bipartite graph as independent and full rank bipartite (IFRB) graph. An example of an IFRB graph is depicted in Fig.~\ref{fig1} (a). \begin{figure}[t] \begin{center} \includegraphics[width=0.9\textwidth]{fig1.eps} \end{center} \caption{Bipartite graph states (top) and associated commuting circuits (bottom). The white and gray shaded circles indicate qubits in $U_B$ and $V_A$, respectively. (a) An independent and full rank bipartite graph. (b) A non-independent bipartite graph. (c) An independent but non-full rank bipartite graph. } \label{fig1} \end{figure} Now we consider the Ising model associated with an IFRB graph. If we consider only computational basis, we can replace the classical spin variable $\sigma$ with the Pauli $Z$ operator. Therefore, we can rewrite the Ising Hamiltonian Eq. (\ref{Isingmodel}) as \begin{eqnarray*} \hat H(\{s_{i}\}, \{ \theta _{j} \},G) \equiv - \sum_{i} \frac{i\pi}{2} s_{i} (1-Z _{v_i}) - \sum _{j} i \theta _{j}\left( \bigotimes _{ v_i \in N_{u_j} } Z _{v_i} \right). \end{eqnarray*} Then the partition function is given by \begin{eqnarray*} \mathcal{Z}(\{s_{i}\}, \{ \theta _{j} \},G) = {\rm Tr} \left[ e^{ - \hat H (\{s_{i}\}, \{ \theta _{j} \},G) } \right]. \end{eqnarray*} Our main goal here is to calculate $|\mathcal{Z}(\{s_{i}\}, \{ \theta _{j} \},G)| ^2$ exactly. To this end, let us first consider the case $s_{i}=0$ for all $v_i$. In this case, there is no magnetic field, and hence we can transform the Hamiltonian into an interaction-free Ising model by virtue of the properties of the IFRB graph. \begin{lemm}[Mapping to interaction-free Ising model] \label{renoma1} For any Ising model associated with an IFRB graph, there exists a unitary operator $W$ that transforms $\hat H(\{0\}, \{ \theta _{j} \},G)$ to interaction-free Ising Hamiltonian: \begin{eqnarray*} W\hat H (\{0\}, \{ \theta _{j} \},G)W^{\dag}= \sum _{j} i \theta _j Z_{v_j} \end{eqnarray*} \end{lemm} {\it Proof}: Since the column vectors of $R$ are independent and full rank, we can transform the matrix $R$ to the identity matrix by using the Gauss-Jordan elimination method. Since the matrix $R$ defines the graph and the Hamiltonian, the Gauss-Jordan elimination can be viewed as a transformation of the graph and the corresponding Hamiltonian. The graph associated with the identity matrix consists of pairs of vertices $(v_i,u_i)$ connected by edges. Since each vertex in $U_B$ is always connected only one vertex in $V_A$, the corresponding Ising Hamiltonian is interaction-free. Each process in the Gauss-Jordan elimination for the matrix $R$ can be implemented on the Hamiltonian by conjugations of controlled-Not (CNOT) and swapping gate operations. The CNOT gate from the $i$th to the $j$th qubits is equivalent to adding the $j$th row vector to the $i$th one on the matrix $R$. The swapping gate exchanges the labels $\{ v_i\}$ of the vertices. Thus there exists a unitary operator $W$ consisting of swapping and CNOT gates such that $W \hat H(\{ 0\} , \{ \theta _j \} , G) W^{\dag} = \sum _{j} i \theta _j Z_{v_j}$. \hfill $\square$ For example, in the case of the IFRB graph shown in Fig.~\ref{fig1} (a), the set of operators in the Hamiltonian is given by $\{ Z_{v_1}Z_{v_2}, Z_{v_1} Z_{v_2} Z_{v_3}, Z_{v_2} \}$. This can be mapped to $\{ Z_{v_1}, Z_{v_2} ,Z_{v_3}\}$ by using the unitary operator $W=S^{wap}_{v_2,v_3} \Lambda (X) _{v_1,v_3} \Lambda (X)_{v_2,v_1}$, where $S^{wap}_{v_i,v_j}$ is the swapping operation between qubits $v_i$ and $v_j$. By using such a $W$, the partition function can be calculated as \begin{eqnarray*} \mathcal{Z}(\{s_{i}\}, \{ \theta _{j} \},G) &=& {\rm Tr} \left[ e^{ - \hat{H} (\{s_{i}\}, \{ \theta _{j} \}) } \right] \\ &=& {\rm Tr} \left[W e^{ - W^{\dag}\hat H (\{s_{i}\}, \{ \theta _{j} \})W} W^{\dag} \right] \\ &=& 2^{ |U_B|} \prod _{u_j} \cos \theta _{j} . \end{eqnarray*} Thus the probability of obtaining $\{ s_{i} = 0 \}$ is computed as \begin{eqnarray*} P_{\rm IQP}(\{ s_{i}=0 \} | \{ \theta _{j} \},\{S_j\}) = \Big( \prod _{u_j} \cos \theta _{j}\Big)^2. \end{eqnarray*} Since the joint probability is factorized for each $\theta_j$, we can easily calculate its marginal distribution (without using Theorem~\ref{Main2} in this case). Next we extend the above result to the general measurement outcomes $\{ s_{i} \}$. This is done by renormalizing the random $i\pi /2$ magnetic fields into the coupling constants as follows. \begin{lemm}[Renormalization of $i \pi/2$ magnetic fields] \label{renoma2} For any {\sf IQP} associated with an IFRB graph, we can find a bit string $\{c_{u_j}\}$ such that \begin{eqnarray*} P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \}) =P_{\rm IQP}(\{ s_{i}=0 \} | \{ \tilde \theta _{j} \}), \end{eqnarray*} with $\tilde \theta _{j} \equiv \theta _{j} + c_{u_j} \pi /2$. \end{lemm} {\it Proof}: Let us consider the corresponding {\sf MBIQP}. From the definition of {\sf MBIQP}, \begin{eqnarray*} P_{MBIQP} (\{ m_{v_i} \},\{ m_{u_j}\} | \{ \theta _j \}, G) &=& \left| \bigotimes _{v_i \in V_A} \langle +_{m_{v_i}} | \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | | G\rangle \right|^2 \\ &=& \left| \langle +_{0} |^{\otimes |V_A|} F(\{ m_{v_i}\}) \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | | G\rangle \right|^2, \end{eqnarray*} where $F(\{m_{v_i} \})\equiv\bigotimes _{v_i \in V_A} Z_{v_i} ^{ m_{v_i} }$. Since the row vectors of $R$ are independent and full rank, we can find a vector $c_{u_j}$ in ${\bf Z}_2^{|U_B|}$ such that $m_{v_i} = \sum_{u_j}R_{v_i}^{u_j} c_{u_j}$ for any $\{ m_{v_i} \}$. By using this vector $c_{u_j}$, we obtain the following equality, \begin{eqnarray*} \prod _{u_j \in U_B } (X_{u_j}K_{u_j})^{c_{u_j}}= \prod _{u_j \in U_B } \left ( \prod _{v_i\in \mathcal{N}_{u_j}}Z_{v_i} \right) ^{c_{u_j}} = F(\{ m_{v_i} \}). \end{eqnarray*} By using this and the fact that $K_{u_j}$ stabilizes $|G\rangle$, we obtain \begin{eqnarray*} P_{MBIQP} (\{ m_{v_i} \},\{ m_{u_j}\} | \{ \theta _j \}, G) &=& \left| \langle +_{0} |^{\otimes |V_A|} \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | \left( \prod _{u_j \in U_B } X_{u_j}^{m_{v_i}} \right)| G\rangle \right|^2 \\ &=& \left| \langle +_{0} |^{\otimes |V_A|} \bigotimes _{u_j \in U_B} \langle \tilde \theta _{j,m_{u_j}} || G\rangle \right|^2 \\ &=&P_{MBIQP} (\{ \tilde s_{v_j}=0 \}, \{ m_{u_j}\}| \{ \tilde \theta _{j}\},G ), \end{eqnarray*} where $\tilde \theta _{j} \equiv \theta _{j} + c_{u_j} \pi /2 $. Specifically, if we consider the case $m_{u_j}=0$, we obtain that \begin{eqnarray*} P_{IQP}(\{s_i\}| \{ \theta _j \} , \{ S_j\}) &=&2 ^{ |U_B|}P_{MBIQP}(\{s_{v_i}\}, \{ m_{u_j}=0 \}| \{ \theta _j \} , G) \\ &=&2 ^{ |U_B|}P_{MBIQP}(\{s_{v_i}=0\}, \{ m_{u_j}=0 \}| \{ \tilde \theta _j \} , G) \\ &=&P_{IQP}(\{s_i=0\}| \{ \tilde \theta _j \} , \{ S_j\}). \end{eqnarray*} \hfill $\square$ Let us consider the example shown in Fig.~\ref{fig1} (a) again. For instance, if $\{s_{v_i}\}=\{0,0,1\}$, $F(\{0,0,1\})= Z_{v_3}$, and $\{c_{u_1}=1, c_{u_2}=0,c_{u_3}=1\}$. By multiplying the stabilizer operators of the graph state with respect to the $4$th and $6$th vertices, we obtain another stabilizer operator $(X_{u_1} Z_{v_1} Z_{v_2})(X_{u_3} Z_{v_1} Z_{v_2} Z_{v_3})= X_{u_1} X_{u_3} Z_{v_3}$. Thus the action of $F(\{ 0,0,1\})$ is equivalent to that of $X_4 X_6$, which rotates the angles $\theta _{u_1}$ and $\theta _{u_3}$ by $\pi /2$. By combining Lemma~\ref{renoma1} and Lemma~\ref{renoma2}, we can show classical simulatability of {\sf IQP} associated with IFRB graphs. \begin{theo}[Classical simulatability: sparse circuits] \label{sparseIQP} {\sf IQP} associated with an IFRB graph is classically simulatable. \end{theo} {\it Proof}: From Lemma~\ref{renoma1} and~\ref{renoma2}, we can calculate $P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \})$ exactly for an IFRB graph. Thus such a class of {\sf IQP} is classically simulatable for arbitrary angles $\{ \theta _{j}\}$. \hfill $\square$ Finally, we slightly weaken the condition, full rank. Even if the column vectors of $R$ is not full rank, i.e., $|U_B|< |V_A|$ [as shown in Fig.~\ref{fig1} (c)], there exist $W$ such that transforms the many-body Ising Hamiltonian to interaction-free Ising Hamiltonian as long as the column vectors of $R$ are independent. Such a class of graphs are called independent bipartite (IB) graphs. In this case, the existence of $c_{u_j}$ for all $\{m_{u_j} \}$ is not guaranteed, and hence we have to find another way to deal with this situation. To settle this, we add ancilla vertices $u_{j'} \in U_{B'}$ to the set $U_B$ in such a way that $R_{v_i}^{u_j}$ ($u_j \in U_B \cup U_{B'}$) has full rank [The 5th qubit in Fig.~\ref{fig1} (a) can be viewed as the ancilla qubit for the non-full rank graph in Fig~\ref{fig1} (c)]. Due to Theorem~\ref{sparseIQP}, we can exactly calculate the probability for the slightly enlarged problem, $P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \} \cup \{ \theta _{j'} \})$. Then, the probability $P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \} )$, with which we want to sample $\{ s_{i} \}$, can be obtained by considering a specific case $\theta _{j'}=0$ for all $u_{j'} \in U_{B'}$, i.e., \begin{eqnarray*} P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \} \cup \{ \theta _{j'}=0 \} ) = P_{\rm IQP}(\{ s_{i} \} | \{ \theta _{j} \} ). \end{eqnarray*} A representative example of classically simulatable {\sf IQP} circuits are depicted in Fig.~\ref{fig1} (a) and (c). If we restrict ourself into two-body Ising models (i.e., $|S_j|=2$), the meaning of independence becomes clear; independence means that the lattice does not contain any loop, such as Ising models on one-dimensional chain or tree graphs. Thus {\sf IQP} with two-qubit commuting gates whose interaction geometry does not contain any loop can be efficiently simulated. In order to avoid the present class of classically simulatable {\sf IQP}, the {\sf IQP} circuits that consist of at least $n$ ($=|V_A|$) commuting gates acting on different subsets $\{S_j\}$ of qubits are sufficient. \subsection{Classical simulatability: planar-{\sf IQP}} Classical simulatability in the previous case is based on the sparsity of the commuting gates, where at most only $n-1$ commuting gates are included. In such a case we can calculate the partition functions without using Theorem~\ref{Main2}. Next we will provide another classically simulatable class of {\sf IQP}, that includes commuting gates much more than $n$. Specifically, we will show below that {\sf IQP} with two-qubit commuting gates acting on nearest-neighbor two qubits in 2D, which we call planar-{\sf IQP}, is classically simulatable. To this end, we first show, by using properties of the graph states, that for two-body Ising models we can always remove the random $i \pi /2$ magnetic fields by appropriately renormalizing their effects into coupling constants $\{ \theta_j \}$. This allows us to map planar-{\sf IQP} to two-body Ising models without magnetic fields. Then we utilize Theorem~\ref{Main2} and exact solvability of two-body Ising models on planar lattices to construct an efficient classical simulation of {\sf IQP}. Consider a planar bipartite graph $G$ with $|S_j|=2$, that is, every vertex $u_j \in U_B$ are connected with just two vertices $v_i \in V_A$. The weights $\{ \theta _j\}$ are arbitrary. For simplicity, we assume that $G$ is connected. Let us consider properties of the graph state associated with such a planar bipartite graph $G$. \begin{remk}[Property of Graph states 1] \label{property1} For any connected bipartite graph $G$ with $|S_j|=2$ for all $j$, the associated graph state $|G\rangle$ is subject to the following property: \begin{eqnarray*} \left( \prod _{v_i \in V_A} \langle + _{m_{v_i}} | \right) |G\rangle = 0 \end{eqnarray*} for any $\{m_{v_i}\}$ such that $\bigoplus _{v_i \in V_A} m_{v_i} =1$. Here the addition is taken modulo two. \end{remk} {\it Proof}: The bipartite graph state is stabilized by \begin{eqnarray*} \prod_{v_i \in V_A} \left( X_{v_i} \prod_{u_j \in \mathcal{N}_{v_i}} Z_{u_j} \right) = \prod_{v_i \in V_A} X_{v_i}, \end{eqnarray*} and hence $\left(\prod_{v_i \in V_A} X_{v_i} \right)|G\rangle =|G\rangle$. By using this, we obtain \begin{eqnarray*} \left( \prod _{v_i \in V_A} \langle + _{m_{v_i}} | \right) |G\rangle = \left( \prod _{v_i \in V_A} \langle + _{m_{v_i}} | \right) \left(\prod_{v_i \in V_A} X_{v_i} \right) |G\rangle =\left( \prod _{v_i \in V_A} \langle + _{m_{v_i}} | \right) (-1)^{\bigoplus _{v_i \in V_A} m_{v_i}} |G\rangle . \end{eqnarray*} Thus if $\bigoplus _{v_i \in V_A} m_{v_i} =1$, then $\left( \prod _{v_i \in V_A} \langle + _{m_{v_i}} | \right) |G\rangle =0$. \hfill $\square$ Thus we only consider the case $\bigoplus _{v_i \in V_A} m_{v_i}=0$, that is, the number of vertices with $m_{v_i} =1$ is even. In such a case, we can show that the random $i \pi/2$ magnetic fields can be renormalized by modifying the coupling constants $\{\theta _j\}$ appropriately as follows. \begin{remk}[Property of Graph states 2] \label{property2} For any {\sf IQP} associated with a connected bipartite graph $G$ with $|S_j|=2$ for all $j$, by appropriately choosing $\{ \tilde \theta _j\}$, \begin{eqnarray*} P_{IQP}(\{s_i \} | \{ \theta _j\}, \{ S_j\} ) =P_{IQP}(\{ s_i=0 \} | \{ \tilde \theta _j\}, \{ S_j\} ), \end{eqnarray*} where $\{ s_i=0 \}$ means that $s_i=0$ for all $i$. Equivalently, for the corresponding Ising models, we have \begin{eqnarray*} H(\{s_i\}, \{ \theta _j\}, G) =H(\{s_i=0\}, \{ \tilde \theta _j\}, G) , \end{eqnarray*} that is, the random $i \pi /2$ magnetic fields can be renormalized into the coupling constants $\{\tilde \theta _j \}$. \end{remk} {\it Proof}: Consider the graph state $|G\rangle$. Due to Remark~\ref{property1}, the number of $\tilde s_i=1$ is always even. The graph is connected. Thus we can always make pairs of vertices $v_i \in V_A$ of $m_{v_i} =1$. Apparently this can be done in polynomial time, since arbitrary paring is allowed. Let us denote such a pair as $(v_k \sim v_{k'})$ and a set of vertices on a path (arbitrarily) connecting them as ${\rm path}(v_k \sim v_{k'})$. The graph state is stabilized by \begin{eqnarray*} \prod _{u_j \in {\rm path}(v_k \sim v_{k'}) \cap U_B} K_{u_j} = Z_{v_k} \left( \prod_{u_j \in {\rm path}(v_k \sim v_{k'}) \cup U_B} X_{u_j} \right) Z_{v_{k'}}, \end{eqnarray*} [see Fig.~\ref{fig2} (b) and (c)]. \begin{figure}[htbp] \begin{center} \includegraphics[width=0.9\textwidth]{fig2.eps} \end{center} \caption{(a) A planar lattice. (b) An associated bipartite graph state, where gray and white circles denote qubits in $V_A$ and $U_B$, respectively. (c) A path between a pair of qubits in $V_A$. (d) The corresponding commuting circuit. } \label{fig2} \end{figure} By using this fact, we can obtain \begin{eqnarray*} &&\left(\bigotimes _{v_i \in V_A} \langle +_{m_{v_i} } |\right) \left( \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | \right) | G\rangle \\ &=&\left(\bigotimes _{v_i \in V_A} \langle +_{m_{v_i} } |\right) \left( \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | \right) \left[ Z_{v_k} \left( \prod_{u_j \in {\rm path}(v_k \sim v_{k'}) \cup U_B} X_{u_j} \right) Z_{v_{k'}} \right] |G\rangle \\ &=& \left(\bigotimes _{v_i \in V_A} \langle +_{m_{v_i} \oplus \delta _{v_i,v_k} \oplus \delta _{v_i,v_{k'}} } |\right) \left( \bigotimes _{u_j \in U_B} \left\langle \theta _{j,m_{u_j}\bigoplus _{u_{j'} \in {\rm path}(v_k \sim v_{k'})} \delta _{u_j,u_{j'}}} \right| \right) |G\rangle. \end{eqnarray*} By doing this repeatedly for all pairs of $m_{v_i}=1$, i.e., a perfect matching of $m_{v_i}=1$ vertices, we can transform all $m_{v_i}=1$ to $m_{v_i}=0$. Let us define an arbitrary perfect matching $\mathcal{M}$ of vertices of $m_{v_i}=1$ and a set ${\rm path}(\mathcal{M})$ of paths of the matching $\mathcal{M}$. By denoting the addition modulo two over $u_{j'}$s on all these paths by $\bigoplus _{ u_{j'} \in{\rm path} (\mathcal{M})}$, the renormalized coupling constant is given by \begin{eqnarray*} \tilde \theta _j = \theta _j + \left( \bigoplus _{u_{j'} \in{\rm path}(\mathcal{M})} \delta _{ u_j, u_{j'}} \right) \pi /2. \end{eqnarray*} Then we obtain \begin{eqnarray*} \left(\bigotimes _{v_i \in V_A} \langle +_{m_{v_i} } |\right) \left( \bigotimes _{u_j \in U_B} \langle \theta _{j,m_{u_j}} | \right) | G\rangle = \langle +_{0 } |^{\otimes |V_A|} \left( \bigotimes _{u_j \in U_B} \langle \tilde \theta _{j,m_{u_j}} | \right) | G\rangle. \end{eqnarray*} This leads that \begin{eqnarray*} P_{IQP} (\{s_i\} | \{\theta _j\}, \{ S_j\})&=& 2^{|U_B|}P_{MBIQP}( \{ m_{v_i}\},\{ m_{u_j}=0 \}| \{ \theta _j \} , G) \\ &=&2^{|U_B|}P_{MBIQP}( \{\tilde s _{v_i}=0\},\{ m_{u_j}=0 \}| \{ \tilde \theta _j \} , G) \\ &=&P_{IQP} (\{s_i =0\} | \{\tilde \theta _j\}, \{ S_j\}) \end{eqnarray*} \hfill $\square$ Note that in the proofs of the properties of graph states with $|S_j|=2$, we did not use the planeness of the graph. Thus Remark~\ref{property1} and Remark~\ref{property2} hold even for nonplanar graphs as long as $|S_j |=2$ for all $j$. Accordingly, we can always remove the random $i \pi /2$ magnetic fields of arbitrary two-body Ising models by appropriately renormalizing them into the two-body coupling constants. Interestingly, these properties of the graph states are closely related to the properties of anyonic excitations on surface codes with a smooth boundary~\cite{Kitaev}. On the graph state with $|S_j|=2$ for all $j$, if one project the qubits in $V_A$ by $|+\rangle^{\otimes |V_A|}$, we obtain the surface code state defined on a lattice $\mathcal{L}$, where vertex and edge corresponds to vertices in $V_A$ and $U_B$ of $G$ respectively, and a qubit is assigned on each edge. This can be confirmed as follows. The post-measurement state is stabilized by $\prod _{u_j \in \mathcal{N}_{v_i}} Z_{u_j} \equiv A_{v_i}$ for all $v_i$. Furthermore, for all faces $f$ of the lattice $\mathcal{L}$, $ \prod _{u_j \in \partial f} K_{u_j} =\prod _{u_j \in \partial f} X_{u_j}\equiv B_f$ stabilizes the post-measurement state, where $\partial f$ is the set of the edges that are boundary of the face $f$. These two types operators are called star and plaquette operators in Ref.~\cite{Kitaev}. The post-measurement state or equivalently the surface code state is the ground state of the Hamiltonian, so-called Kitaev's toric code Hamiltonian, \begin{eqnarray*} H= -J \sum _{i} A_i - J \sum _{f} B_f. \end{eqnarray*} A projection by $|-\rangle _{v_i} $ results in the eigenvalue $-1$ of the star operator at vertex $v_i$, which corresponds to the anyonic excitation in the Kitaev model. Then Remark~\ref{property1} indicates that the parity of anyonic excitations is always even. They are created and annihilated in pairs. Remark~\ref{property2} corresponds a way to annihilate the pairs of the anyonic excitations. The trajectory of anyonic excitations in the annihilation process corresponds to ${\rm path}(\mathcal{M})$. Now we are ready to show that classical simulatability of {\sf IQP} consisting of 2D nearest-neighbor two-qubit commuting gates. \begin{theo}[Classical simulatability: planar-{\sf IQP}] \label{freefermion} {\sf IQP} with two-qubit commuting gates acting on nearest-neighbor qubits in two-dimension, which we call planar-{\sf IQP}, is classically simulatable (in the strong sense including its marginal distributions). \end{theo} {\it Proof}: According to Theorem~\ref{Main1}, the joint probability distribution of planar-{\sf IQP} can be calculated from the partition function of a two-body Ising model on a planar lattice. Since the graph $G$ is a planar bipartite graph, we can easily find an order of measurements such that $\tilde G_{M^{(k)}}$ is also planar at any measurement step $k$ (Any order of measurements such that the subgraph $G_{M^{(k)}}$ becomes a connected graph for all $k$ can be utilized). Due to Theorem~\ref{Main2}, the marginal distributions are also obtained by the partition functions of Ising models on planar lattices. Furthermore, in the merged graph, the vertices $u_j \in M_B^{(k)} \cup {M'}^{(k)}_{B}$ are connected with just two vertices, i.e., $|\mathcal{N}_{u_j}|=2$. For such Ising models, by using Remark~\ref{property1} and Remark~\ref{property2}, the random magnetic $i \pi /2$ fields can be renormalized into the coupling constants $\{ \theta \} \rightarrow \{ \tilde \theta _j \}$. Thus all marginal distributions can be calculated from the partition functions of two-body Ising models on planar lattices without magnetic fields. On the other hand, it is well known that the partition function of two-body Ising models on planar lattices without magnetic fields can be calculated efficiently by expressing them as the Pfaffians~\cite{Kasteleyn,Fisher,Barahona}. Thus we conclude that {\sf IQP} of this class can be simulated efficiently in the strong sense including its marginals. This also guarantees weak simulation by a recursive simulation. \hfill $\square$ Note that a similar argument is also made in Ref.~\cite{BravyiRaussendorf} by considering classical simulatability of MBQC on the planar surface codes~\cite{Kitaev}. Indeed, as mentioned before, if we apply the projection by $|+ \rangle ^{ \otimes |V_A|}$ on the bipartite planar graph state with $|S_j|=2$, we obtain an unnormalized planar surface code state consisting of the qubits on $U_B$. The effect of $m_{v_i}=1$ (i.e., the projection by $|+_{1}\rangle$) can be renormalized into the coupling constants $\{\theta _j \} \rightarrow \{ \tilde \theta _j\}$, where an arbitrary perfect matching is chosen as shown in Remark~\ref{property2}. Thus we may construct an alternative proof of Theorem~\ref{freefermion} without using Theorem~\ref{Main2}. However, Theorem~\ref{Main2}, employing the properties of the graph states, is much straightforward and simple for our purpose. Furthermore, Theorem~\ref{Main2} is valid not only for the case with $|S_j|=2$, but also the general cases, which cannot be regarded as MBQC on the planar surface codes. The Pfaffian is the square root of the determinant, and hence the probability distribution of planar-{\sf IQP} is given by the determinant of an appropriately defined complex matrix. The determinant appears in the probability distribution of fermions scattered by fermionic linear optical unitary operators. Thus the present classical simulatable class of {\sf IQP} is regarded as a {\sf F{\footnotesize ERMION}S{\footnotesize AMPLING}}. The sampling problems in physics might be classified in a unified way as collision problems of elementary particles. Important implications of Theorem~\ref{freefermion} are twofold. One is that entanglement is necessary ingredient (since discord-free dynamics can be classically simulatable~\cite{Eastin}) for quantum speedup, but not sufficient at all. This is also the case for the Clifford circuits and match gates, which generate genuinely entangled states but are classically simulatable ~\cite{NielsenChuang,matchgates0,matchgates1,matchgates2,matchgates3}. The 2D nearest-neighbor two-qubit gates on the input state $|+\rangle ^{\otimes n}$ can create genuinely entangled states. However, the probability distribution of the $X$-basis measurements on them can be classically simulatable. On the other hand, {\sf IQP} associated with planar graphs with $|S_j|\leq 2$ belongs to ${\sf IQP}_{\sf UP}$ and is as powerful as {\sf PP} by introducing postselection. Thus single-qubit rotations take a quite important rule for {\sf IQP} to be classically intractable. Indeed, single-qubit rotations make a drastic change of computational complexity from strongly simulatable to not simulatable even in the weak sense. We would like to note that a similar result is also obtained by Van den Nest in a rather different situation~\cite{WeakSim}. He showed that Toffoli-Diagonal circuits, which include quantum Fourier transformation for Shor's factorization algorithm, can be efficiently simulated if there is no basis change at the final round before the measurements in the computational basis. Thus also in Toffoli-Diagonal circuits single-qubit rotations play a very important role for them to be classically intractable. Another consequence of Theorem~\ref{freefermion} lies in the context of experimental verification of quantum benefits. When we utilize {\sf IQP} for the purpose of experimental verification of quantum benefits, we have to avoid planar-{\sf IQP}, since a malicious quantum device can cheat experimentalists by classically sampling the results instead of implementing the {\sf IQP} circuit. At the same time, the existence of efficient classical simulation for planar-{\sf IQP} implies that experimental verification of this class is much easier than any instance in ${\sf IQP}_{\sf UP}$. Thus when experimentalists realize {\sf IQP}, they should, at least, verify its validity in planar-{\sf IQP}, which can be done efficiently. It might be possible to efficiently ensure, under a plausible assumption, that two-qubit commuting gates are operated appropriately, since experimental devices are usually well known and not so malicious. Hopefully, classical intractability of quantum devices may be verified by an efficient experimental verification of planar-{\sf IQP} combined with other efficient witness or plausible assumptions. \section{Hardness of approximating Ising partition functions} \label{sec5} In this section, we utilize the established relationship between {\sf IQP} and Ising partition functions in an opposite direction; we show, by considering ${\sf IQP}_{\sf UP}$, that there is no FPRAS for a class of Ising model with imaginary coupling constants unless the {\sf PH} collapses at the third level. We further explore complexity of a multiplicative approximation of the Ising partition functions by considering strong simulation of {\sf IQP}. \begin{theo}[Hardness of approximating imaginary Ising partition functions] \label{FPRASHard} If there is a fully-polynomial randomized approximation scheme (FPRAS) to compute the partition functions of a class of Ising models on planar lattices that are associated with {\sf IQP} belonging to ${\sf IQP}_{\sf UP}$, the {\sf PH} collapses at the third level. That is, there is no FPRAS for such Ising models, unless the {\sf PH} collapses at the third level. \end{theo} Before proceeding to the proof, we define FPRAS. \begin{defi}[FPRAS] Given an input (instance) $\mathcal{I}$ and error parameter $\epsilon$, FPRAS returns an approximation $f_{ap}(\mathcal{I})$ of a function $f(\mathcal{I})$ within an approximation range \begin{eqnarray*} (1-\epsilon) f(\mathcal{I}) \leq f_{ap}(\mathcal{I}) \leq (1+\epsilon)f(\mathcal{I}) \end{eqnarray*} for an arbitrary $\mathcal{I}$ in polynomial time in both the size of the input and $1/\epsilon$. \end{defi} {\it Proof}: Due to Theorem~\ref{Main2}, the existence of an FPRAS for the Ising models on planar lattices implies an existence of an FPRAS for all marginal distributions of the associated {\sf IQP} with an appropriately chosen measurement order. Thus we can compute the conditional probability $p(s_{i_k} | \{s_i \}_{i \in M_k})$ within the following approximation range: \begin{eqnarray*} \frac{(1-\epsilon _k)}{(1+\epsilon_k)} p (s_{i_k} | \{s_i \}_{i \in M_k}) \leq p_{ap} (s_{i_k} | \{s_i \}_{i \in M_k}) \leq \frac{(1+\epsilon_k)}{(1-\epsilon _k)} p (s_{i_k} | \{s_i \}_{i \in M_k}). \end{eqnarray*} By recursively applying the FPRAS, we can obtain an approximation $P^{ap}_{IQP}$ of the probability distribution for {\sf IQP} within the following approximation range, \begin{eqnarray*} \prod _{k=1}^{|V_A|}\frac{1-\epsilon _{k}}{1+\epsilon _{k}}P_{IQP} \leq P^{ap}_{IQP} \leq \prod _{k=1}^{|V_A|}\frac{1+\epsilon _{k}}{1-\epsilon _{k}} P_{IQP}. \end{eqnarray*} Here we are considering the class of Ising models that are associated with ${\sf IQP}_{\sf UP}$. As in Remark~\ref{IQP}, we can simulate {\sf PP} by such an approximated {\sf IQP}, if the multiplicative error is smaller than $\sqrt{2}$. For simplicity, if we take $\epsilon_1=...=\epsilon_{n}\equiv\epsilon$, \begin{eqnarray*} \left(\frac{1+\epsilon }{1-\epsilon }\right)^{n} \leq \sqrt{2}. \end{eqnarray*} Accordingly, the error has to be \begin{eqnarray*} \epsilon < \frac{2^{1/(2n)}-1}{2^{1/(2n)}+1} \simeq \frac{\sqrt{2} -1}{n}. \end{eqnarray*} Since FPRAS takes only polynomial time in $1/\epsilon$, the above approximation can be done in polynomial time. This results in {\sf PP} $\subseteq$ post-{\sf BPP}, and hence the {\sf PH} collapses at the third level. \hfill $\square$ This result indicates that unless the {\sf PH} collapses, there is no FPRAS to compute a wide range of Ising partition functions with imaginary coupling constants and magnetic fields. Below we will show that this intractability holds for all most all imaginary coupling constants even on a 2D planar graph with a bounded degree. Let us consider the simplest case where the coupling constants including the magnetic fields are given homogeneously by the same imaginary constant $i \theta$. As shown in Ref.~\cite{IQP}, {\sf IQP} associated with planar graphs with $|S_j| \leq 2$ and a bounded degree belongs to ${\sf IQP}_{\sf UP}$ with the homogeneous coupling constant $i\theta = i\pi/8$. Thus the partition functions of two-body Ising models with the homogeneous coupling constant $i \theta = i \pi/8$ have no FPRAS even on the planar lattice unless the {\sf PH} collapses at the third level. The same result holds not only $i\theta =i \pi/8$ but also $i\theta = i(2l+1) \pi/(8m)$ for integers $l$ and $m$. Suppose the homogeneous coupling is given by an irrational angle i.e., $\theta = 2\nu \pi$ with $\nu \in [0,1) $ being an irrational number. Let $m$ be an integer. Since $2 m\nu \pi $ ($\textrm{mod } 2\pi$) is distributed in a uniform fashion, we can find an approximation with an additive error $\epsilon$ with some integer $m = \mathcal{O}(1/\epsilon)$~\cite{NielsenChuang}. An erroneous rotation $D(\theta _j + \epsilon, S_j)$ is sufficiently close to an ideal rotation $D(\theta _j +\epsilon,S_j)$ in the sense of an appropriately defined distance such as the diamond norm~\cite{Diamond}. In the present case, the erroneous rotation $D(\theta _j + \epsilon, S_j)$ is unitary, and hence the diamond norm is equivalent to the square of the operator norm, which is given by \begin{eqnarray*} || D (\theta _j, S_j) [ I - D (\epsilon ,S_j) ] ||^2 =|| I - D (\epsilon ,S_j) ||^2 = 2 (1-\cos \epsilon) = O(\epsilon). \end{eqnarray*} Post-{\sf IQP}${}_{\sf UP}$ can simulate universal MBQC and hence also simulate fault-tolerant quantum computation in MBQC. If the error $\epsilon$ is sufficiently smaller than the threshold value of fault-tolerant quantum computation~\cite{DoritBenor,RaussendorfPhD,Nielsen}, we can reliably simulate universal quantum computation (i.e., {\sf BQP}) and moreover {\sf PP} with the help of postselection. Thus there is no FPRAS for the partition functions of two-body Ising models on planar lattices with imaginary coupling constants of an arbitrary irrational angle $2 \nu \pi$, unless the PH collapses at the third level. This result contrasts with the existence of an FPRAS in ferromagnetic cases with magnetic fields shown by Jerrum and Sinclair~\cite{Jerrum93} and antiferromagnetic cases on a sort of lattices shown by Sinclair, Srivastava, and Thurley ~\cite{Sinclair}, and Sly and Sun~\cite{Sly}. It seems that an FPRAS is hard to obtain in the imaginary coupling regimes even on a planar lattice with a bounded degree except for a certain special rational angles such as $\pi, \pi/2$ and $\pi/4$ corresponding to the Clifford circuits. This result also contrasts with the recent studies on quantum computational complexity of Ising partition functions with imaginary coupling constants~\cite{Nest_circuit,Cuevas11,AharonovJones,AharonovJonesHard,AharonovTutte}. These quantum algorithms calculate the partition function of the Ising models or, more generally, Jones or Tutte polynomials with additive error $\epsilon$ in polynomial time of $1/\epsilon$: \begin{eqnarray*} || \mathcal{Z} - \mathcal{Z}_{ap} || \leq \epsilon \Delta, \end{eqnarray*} where $\mathcal{Z}$ and $\mathcal{Z}_{ap}$ are true and approximated values respectively, and $\Delta$ is a certain algorithmic scale. Furthermore, it has been shown that these additive approximations are enough to simulate {\sf BQP} (i.e., {\sf BQP}-hard). Hence these problems are {\sf BQP}-complete. Classical simulation of these problems is in general hard, unless {\sf BQP}={\sf BPP}, which also implies collapse of the {\sf PH} at the third level. In the present case, on the other hand, we obtain the hardness result with a multiplicative approximation. Since additive error makes sense when $||\mathcal{Z}||$ is known, a multiplicative approximation is much more meaningful than an additive approximation. However, as shown below, we should note that {\sf IQP} itself hardly approximate the values of the Ising partition functions within multiplicative error. This discrepancy lies in the difference between strong and weak simulations of quantum computation. An FPRAS of the Ising partition functions, meaning strong simulation of {\sf IQP}, is a sufficient condition for classical simulation of {\sf IQP} but much stronger than what is required to simulate {\sf IQP} (i.e., weak simulation). Strong simulation of {\sf IQP} with multiplicative error $c$ is equivalent to strong simulation of the output $P(\mathcal{U})$ of an arbitrary quantum circuit $\mathcal{U}$ with the same multiplicative error. This is because if we postselect an appropriate subset $\{s _i\}_{\rm post}$ of the measurement outcome, we can simulate an arbitrary quantum circuit in {\sf IQP}. Thus we obtain \begin{eqnarray*} P(\mathcal{U}) =\mathcal{N}P_{IQP}(\{s_i\}_{\notin \textrm{ post}} , \{s _i =0\}_{\rm post} ), \end{eqnarray*} where $\mathcal{N} = 1/{\rm prob}(\{ s_i =0 \}_{\rm post})$ is the normalization factor. We can always map an arbitrary quantum circuit $\mathcal{U}$ into an MBQC with which $\{ s_i \}_{\rm post}$ are distributed randomly with independent probability 1/2. The normalization factor can be calculated exactly as $\mathcal{N}=2^{|{\rm post}|}$, which leads to an equivalence between strong simulations of {\sf IQP} and universal quantum computation with the same multiplicative error. How about strong simulation of marginal distributions (partial measurements of a quantum circuit)? As shown in Theorem~\ref{Main2}, strong simulation of the output of the {\sf IQP} circuits with multiplicative error $c$ also provides that of its marginals with multiplicative error $\sqrt{c}$ as long as the associated graphs are closed under merging. On the other hand, an arbitrary marginal distribution of universal quantum computation can also viewed as a marginal distribution of {\sf IQP} due to the above equivalence. Thus not only the joint probability distribution of an arbitrary quantum circuit but also its marginal distributions (partial measurements) can be obtained from strong simulation of {\sf IQP} within the same or smaller multiplicative error. This result indirectly provides an equivalence between strong simulation of a universal class of quantum circuits and its marginal distributions (partial measurements). This fact indicates that, for almost any imaginary coupling constants, a multiplicative approximation of Ising partition functions on a planar graph of a bounded degree is as hard as a multiplicative approximation of the output distribution an arbitrary quantum circuit including its marginals. By considering the fact that strong simulation of a quantum circuit is \#{\sf P}-hard \cite{WeakSim}, this results indicates that a multiplicative approximation of Ising partition functions is highly intractable for almost all imaginary coupling constants even on a 2D planar graph with a bounded degree. \section{Conclusion and discussion \label{sec6}} We have investigated {\sf IQP} by relating it with computational complexity of Ising partition functions with imaginary coupling constants and magnetic fields. We found classes of {\sf IQP} that are classically simulatable in the strong sense. Specifically, the {\sf IQP} circuits consisting only of 2D nearest-neighbor two-qubit commuting gates, namely planar-{\sf IQP}, are classically simulatable. However, if single-qubit rotations are allowed, planar-{\sf IQP} becomes ${\sf IQP}_{\sf UP}$, which are as powerful as {\sf PP} under postselection. Thus single-qubit rotations make a drastic change of the {\sf IQP} circuits from strongly simulatable to intractable in the sense of weak simulation. The classical simulatability of planar-{\sf IQP} stems from the exact solvability of Ising models on planar lattices without magnetic fields. Both classical computational complexity of Ising models on nonplanar lattices~\cite{Barahona,Istrail} and quantum computation complexity of MBQC on nonplanar surface codes~\cite{RaussendorfNonplaner} have been studied already. While we did not addressed here, computational complexity of the {\sf IQP} circuits consisting of two-qubit commuting gates with a nonplanar geometry is an intriguing topic. By considering strong simulation of {\sf IQP}, we further explored hardness of a multiplicative approximation of the Ising partition functions with multiplicative error. We have shown that there exist no FPRAS of a class of Ising models for almost all imaginary coupling constants, unless the {\sf PH} collapses at the third level. This indicates that a multiplicative approximation of the class of Ising models is highly intractable. Actually, strong simulation of {\sf IQP} with multiplicative error can also provide strong simulation of an arbitrary quantum circuit. Furthermore, it also provides an arbitrary marginal distribution of the quantum circuit with the same or smaller multiplicative error. This means that a multiplicative approximation of the class of Ising models are at least as hard as strong simulation of quantum computation. By considering the fact that strong simulation of a quantum circuit is \#{\sf P}-hard \cite{WeakSim}, a multiplicative approximation of Ising partition functions is highly intractable for almost all imaginary coupling constants even on a 2D planar graph with a bounded degree. The results obtained in this work exhibit a rich structure of {\sf IQP}, ranging from classically simulatable to highly intractable problems in strong simulation. This suggests that {\sf IQP} is a good playground for studying quantum computational complexity and its intractability in classical computation. While we did not address, further investigations on computational complexity of weak simulation of {\sf IQP} should be made. Hopefully the relation between {\sf IQP} and Ising partition functions with imaginary coupling constants is also helpful in doing so. \section*{Acknowledgements} The authors thank S. Tamate for useful discussions. KF is supported by JSPS Grant-in-Aid for Research Activity Start-up 25887034. TM is supported by Tenure Track System by MEXT, Japan. \bibliographystyle{ieeetr}
\section*{} Cosmological scalar fields are among the simplest and most promising candidates for dark energy. Up to now, many different Lagrangians and potentials have been studied. Here we consider the classical scalar field with barotropic equation of state. Such class of models involves both quintessential and phantom subclasses. We include into analysis the subclasses of models without peculiarities in the past \cite{Novosyadlyj2013}: \begin{itemize} \item $w_0>-1$, $c_a^2>-1$; \item $w_0>-1$, $c_a^2<-1$; \item $w_0<-1$, $c_a^2<-1$, $c_a^2<w_0$. \end{itemize} We exclude the folowing subclasses of models, which can lead to $\rho_{tot}<0$ at some time in the past: \begin{itemize} \item $w_0<-1$, $c_a^2>-1$; \item $w_0<-1$, $c_a^2<-1$, $c_a^2>w_0$. \end{itemize} We have determined the best-fit values and confidence limits of the model parameters using the Markov chain Monte Carlo (MCMC) technique (implemented in the code CosmoMC)\cite{cosmomc} and the following data: \begin{itemize} \item \textit{CMB temperature fluctuations and polarization angular power spectra} from the 7-year WMAP observations (WMAP7) \cite{wmap7_1,wmap7_2}; \item \textit{Baryon acoustic oscillations} in the space distribution of galaxies from SDSS DR7 (BAO) \cite{bao_sdss}; \item \textit{Hubble constant measurements} from HST (HST) \cite{hst}; \item \textit{Big Bang Nucleosynthesis prior} on baryon abundance (BBN) \cite{bbn_1,bbn_2}; \item \textit{supernovae Ia luminosity distance moduli} from SDSS compilation \cite{sn_sdss} with MLCS2k2 \cite{Jha2007} (SN SDSS MLCS2k2) and the SALT2 \cite{Guy2007} (SN SDSS SALT2) methods of light curve fitting. \end{itemize} The results for the combined datasets WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS MLCS2k2 and WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS SALT2 are presented in Fig. \ref{fig} and in Table \ref{tab}. We see that the dataset WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS SALT2 prefers the phantom models of dark energy with barotropic equation of state, while the dataset WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS MLCS2k2 gives preference to the models with $w_0>-1$, $c_a^2<-1$ (in accordance with conclusions of [\refcite{Novosyadlyj2013}]). We have also used the newer data on SNe Ia distance moduli from \begin{itemize} \item SNLS3 compilation (SNLS3) \cite{sn_snls} and \item Union2.1 compilation (Union2.1) \cite{sn_union} \end{itemize} together with data on BAO from the WiggleZ Dark Energy Survey (BAO WiggleZ) \cite{bao_wigglez}. The analysis of combined datasets WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} BAO WiggleZ {+} SNLS3 and WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} BAO WiggleZ {+} Union2.1 has shown that these data prefer the phantom models. Allowing for non-zero masses of active neutrinos, non-zero curvature or non-zero contribution from the tensor mode of perturbations does not change this conclusion. The obtained constraints on the massive active neutrino fraction of dark matter $f_{\nu}$, the curvature of 3-space $\Omega_k$ and the contribution from the tensor mode of perturbations $r$ are consistent with zero values of these parameters. The forecast made for the Planck mock data (generated using the code FuturCMB \cite{bluebook}) suggests that the models with $c_a^2>-0.75$ may be ruled out at $2\sigma$ confidence level by the Planck data. \begin{figure}[htb] \begin{center} \psfig{file=postlike_combi_sdss.eps,width=0.75\textwidth} \end{center} \caption{One-dimensional marginalized posteriors (solid lines) and mean likelihoods (dotted lines) for $w_0$ (top panels) and $c_a^2$ (middle panels). Left: WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS MLCS2k2. Right: WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS SALT2. Bottom: the corresponding two-dimensional mean likelihood distributions in the plane $c_a^2-w_0$. Solid lines show the $1\sigma$ and $2\sigma$ confidence contours.} \label{fig} \end{figure} \begin{table}[htb] \tbl{The best-fit values, mean values and 2$\sigma$ marginalized confidence ranges for cosmological parameters determined by the MCMC technique using two observational datasets: WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS MLCS2k2 and WMAP7 {+} HST {+} BBN {+} BAO SDSS {+} SN SDSS SALT2. The rescaled energy density of the component $X$ is denoted by $\omega_X \equiv \Omega_Xh^2$.} { \begin{tabular}{ccccc} \toprule Parameter&\multicolumn{2}{c}{MLCS2k2}&\multicolumn{2}{c}{SALT2}\\ &best-fit&2$\sigma$ c.l.&best-fit&2$\sigma$ c.l.\\ \colrule $\Omega_{de}$& 0.702& 0.700$_{- 0.034}^{+ 0.031}$& 0.725& 0.725$_{- 0.030}^{+ 0.027}$\medskip\\ $w_0$&-0.758&-0.814$_{- 0.170}^{+ 0.228}$&-1.049&-1.010$_{- 0.171}^{+ 0.145}$\medskip\\ $c_a^2$&-1.295&-1.112$_{- 0.464}^{+ 0.672}$&-1.486&-1.139$_{- 0.423}^{+ 0.830}$\medskip\\ $10\omega_b$& 0.230& 0.227$_{- 0.011}^{+ 0.011}$& 0.224& 0.225$_{- 0.010}^{+ 0.010}$\medskip\\ $\omega_{cdm}$& 0.110& 0.110$_{- 0.009}^{+ 0.009}$& 0.114& 0.113$_{- 0.009}^{+ 0.009}$\medskip\\ $h$& 0.667& 0.665$_{- 0.028}^{+ 0.030}$& 0.704& 0.702$_{- 0.029}^{+ 0.029}$\medskip\\ $n_s$& 0.975& 0.974$_{- 0.026}^{+ 0.027}$& 0.966& 0.969$_{- 0.026}^{+ 0.026}$\medskip\\ $\log(10^{10}A_s)$& 3.075& 3.083$_{- 0.068}^{+ 0.071}$& 3.081& 3.086$_{- 0.066}^{+ 0.069}$\medskip\\ $\tau_{rei}$& 0.089& 0.090$_{- 0.024}^{+ 0.026}$& 0.080& 0.088$_{- 0.023}^{+ 0.025}$\medskip \colrule $-\log L$&\multicolumn{2}{c}{ 3857.113}&\multicolumn{2}{c}{ 3864.929}\\ \botrule \end{tabular} } \label{tab} \end{table} \bibliographystyle{ws-procs975x65}
\section{Introduction} The Wide-field Infrared Survey Explorer ({\it WISE}, Wright et al. 2010) has significantly advanced the study of brown dwarfs, stellar-like objects that have insufficient mass for stable hydrogen burning (Kumar 1963, Hayashi \& Nakano 1963). About the size of Jupiter, the cool brown dwarfs are intrinsically very faint. Prior to {\it WISE}, brown dwarfs as cool as $T_{\rm eff} = 500$~K had been found in near-infrared surveys undertaken by 4$\,$m-class ground-based telescopes (e.g. Lucas et al. 2010). Two brown dwarfs were also known with $T_{\rm eff} \approx 400$~K: CFBDSIR J145829+101343B, a companion to a warmer brown dwarf, discovered using laser guide star adaptive optics imaging (Liu et al. 2011), and GJ 3483B, a companion to a white dwarf, found using the Infrared Array Camera on the {\it Spitzer Telescope} in a proper motion search for faint companions (Luhman, Burgasser \& Bochanski 2011). These 400~K brown dwarfs were expected to have a later spectral type than all previously known T dwarfs. Observations beyond the near-infrared, such as those obtained with the {\it Spitzer Telescope}, are advantageous for studying cool brown dwarfs because objects with $T_{\rm eff} < 700$~K emit $> 50$\% of their energy at $\lambda > 3$ $\mu$m (Leggett et al. 2010a). In 2011 the mid-infrared 0.4$\,$m {\it WISE} telescope identified the first large sample of brown dwarfs with $T_{\rm eff} < 500$~K (Cushing et al. 2011, Kirkpatrick et al. 2011 and 2012). Kirkpatrick et al. (2011) presented the first hundred {\it WISE} brown dwarfs and this paper studies one of those objects, WISEPC J121756.91+162640.2 (hereafter WISE 1217$+$16), which was classified as a T9 spectral type by Kirkpatrick et al. WISE 1217$+$16 was found to be a binary system by Liu et al. (2012, hereafter Liu12) using Keck laser guide star adaptive optics. The pair is separated by $0\farcs 76$ at a position angle of 14.3$^\circ$ . It is an unusual binary, having a relatively wide separation and a large difference in near-infrared brightness between the two components (see Table 1). Liu12 obtained resolved $H$-band spectra of the two components and classified WISE 1217$+$16A as a T9 and WISE 1217$+$16B as a Y0, using the spectral classification scheme for the latest T-type and early Y-type brown dwarfs proposed by Cushing et al. (2011). Here we present resolved spectra for the system covering a wider wavelength range of 0.9 -- 2.5 $\mu$m and expand on the analysis presented by Liu12. \section{Observations} We used the Gemini near-infrared spectrograph (GNIRS, Elias et al. 2006) to obtain 0.9 -- 2.5 $\mu$m cross-dispersed spectra of WISE 1217$+$16AB via queue program GN-2012B-Q-28. In order to resolve the pair, observations were carried out only when the natural seeing full width half maximum (FWHM) was $0\farcs 45$ or better and only at airmasses less than 1.2. Both photometric and cloudy conditions were utilised. The $0\farcs 3$ slit was used with the $0\farcs 15$ pixel$^{-1}$ camera, resulting in a resolving power $\lambda/\delta\lambda \sim 1700$. The slit was placed at 14.3$^\circ$ so that both sources were in the slit. Individual exposure times of 300 seconds were used, with the target nodded along the slit. The A0 star HD 101060 and the F4 star HD 114072 were used as calibrators to remove telluric features. Observations were obtained on 2013 February 21, April 9, April 26 and May 8 UT. Time on source on these nights was 35, 40, 60 and 60 minutes, respectively. The data from 2013 February 21 were obtained in thicker cloud cover, and as the signal was around half that of the three other nights those data were omitted. Calibration lamps on the telescope provided data for wavelength calibration and flat fielding, as well as pinhole images for tracing the cross-dispersed data along the detector. Figure 1 shows the flat fielded, sky-subtracted and rectified $J$-band spectrum from 2013 May 8 as a two-dimensional image. The FWHM is around 2 pixels ($0\farcs 3$) and the separation between components is 5 pixels ($0\farcs 76$). The spectrum of both the A and B components were extracted using apertures centered at the respective peaks, with a lower limit of $-2.5$ pixels and an upper limit of 2.5 pixels. Where B is extremely faint, we used the known offset from A to place the aperture. The contribution of the brighter component to the fainter component's spectrum was determined by extracting the signal of the A component at the location of the B components aperture on the opposite side of the profile. Typically the contribution was 5\% of the signal of the A component, and 20\% of that of the B component. To avoid adding noise, we subtracted an appropriately scaled version of the A spectrum from the B spectrum, and not the spectrum extracted from the wing. The spectral orders for each component were combined, averaging the regions of overlap at 0.98 -- 0.95, 0.98 -- 1.06, 1.13 -- 1.25, 1.45 -- 1.52, 1.88 -- 1.90 $\mu$m. Each spectrum was flux-calibrated using the photometry presented in Liu12. The final spectrum for each component was constructed by averaging each night's data. Based on the scatter between the three measurements, the uncertainty in the flux for the A component is 7\% over the $Y$-band peak, 5\% over the $J$-band peak, 7\% over the $H$-band peak, and 15\% over the $K$-band peak. Similarly for the B component the uncertainty in the flux is 20\% over the $Y$-band peak, 15\% over the $J$-band peak, 10\% over the $H$-band peak, and 60\% over the $K$-band peak (where there is very little flux). Figure 2 shows the spectrum of each component derived here, together with a comparison of the summed spectrum to the spectrum of the unresolved pair presented by Kirkpatrick et al. (2011). The inset plot compares our $H$-band sections to the resolved spectra obtained previously at Keck by Liu12. The agreement with the Keck data is excellent, and that with the unresolved spectrum is reasonable, given the noise in each measurement. These comparisons validate our extraction of the one-dimensional spectra from the two-dimensional images. \section{Spectral Classification} Figure 3 compares our spectrum of WISE 1217$+$16A and WISE 1217$+$16B to T and Y dwarfs classified by Cushing et al. (2011) as T8.5, T9, Y0 and Y0.5. The comparison spectra have been scaled to the $J$-band flux peak of each component. Inset plots zoom in on the $J$-band flux peak, the width of which is a diagnostic of the spectral type (Warren et al. 2007, Burningham et al. 2008, Cushing et al. 2011, Kirkpatrick et al. 2012, Mace et al. 2013). It is clear visually that WISE 1217$+$16A is an excellent match to the T8.5 reference, and has a wider $J$-band flux peak than the T9. However WISE 1217$+$16A is much fainter at $K$ than either reference source, which we discuss in the following paragraph. We classify WISE 1217$+$16A as T8.5. The WISE 1217$+$16B comparison is less straightforward. The blue side of the $J$-band flux peak matches the Y0.5 template, however the red side is in better agreement with the Y0 template. Both the blue and red sides of the $J$-band peak are impacted by absorption by CH$_4$, H$_2$O and NH$_3$; the blue side is also affected by pressure-induced H$_2$ absorption (e.g. Figure 1 of Leggett et al. 2009). It is likely that metallicity and gravity, as well as temperature, impact the shape of the $J$-band peak and that we are seeing variations in these parameters for the three Y dwarfs. The fact that the $K$-band is suppressed for WISE 1217$+$16A implies that it either has a high gravity or low metallicity which enhances the H$_2$ absorption at 2 $\mu$m (e.g. Leggett et al. 2009); possibly we are also seeing enhanced H$_2$ absorption in the blue wing of the $J$-band peak. We define WISE 1217$+$16B as Y0 -- 0.5. Note that the near-infrared spectra of these components, which straddle the current T/Y classification boundary, are overall very similar. Increased molecular absorption narrows the flux peaks at lower $T_{\rm eff}$, but there is otherwise no strong spectral marker for the transition from T to Y in the near infrared. One difference that is apparent is the height and width of the $Y$-band peak around 1 $\mu$m --- relative to the $J$-band peak, the later spectral type and cooler object has a taller and broader $Y$-band peak. This is also seen in the $Y - J$ colors, which become bluer. Figure 4 is the {$Y - J$,$M_Y$} color-magnitude diagram, showing the trend to bluer $Y - J$. Also, this Figure shows about half the drop in absolute magnitude between T9 and Y0 than is seen for $M_J$ and $M_H$ (Dupuy \& Kraus 2013, Figure 5 and \S 4.3). For T9 types with $T_{\rm eff} \approx 500$~K, the alkali elements are condensing into chlorides and sulfides, weakening the strong $0.77 \mu$m K I resonance doublet, the red wing of which suppresses the $Y$-band flux. This likely explains the brightening at $Y$ (Liu12, Leggett et al. 2012). There is a lot of scatter in the $Y - J$ colors of early Y dwarfs (Figure 4), again suggesting that there are variations in metallicity and gravity, as well as temperature, within this small sample. This disparity can also be seen spectroscopically in Figure 3: WISE 1217$+$16B has a broader $Y$-band peak than the comparison Y dwarfs and is bluer in $Y - J$. The gap in $M_Y$ and $M_H$ at the T to Y transition is striking in Figures 4 and 5. This gap is not seen at mid-infrared wavelengths (Leggett et al 2013, hereafter Leg13; Dupuy \& Kraus 2013). The gap may disappear as more very late-type T and Y dwarfs are found and more distances are determined but it is tempting to associate the distinct drop in flux at 1 -- 2 $\mu$m at $T_{\rm eff} \approx 400$~K with a physical phenomenon. Possibilities include the appearance of water clouds or the impact of a newly significant opacity such as H$_2$. Improved models are required to explore this further. Finally we note that the object initially identified as the prototype Y dwarf, WISEP J182831.08+265037.8 (Cushing et al. 2011), appears to be different from the other Y dwarfs in terms of colors and luminosity (\S 4.3, Beichman et al. 2013, Dupuy \& Kraus 2013, Leg13, Kirkpatrick et al. 2013). Once a larger sample of Y dwarfs is known it is likely that the spectral classification scheme will have to be revisited. Also as the models are improved with more complete treatment of opacities, clouds and turbulence, we hope to disentangle the effects of temperature, metallicity and gravity on the near-infrared spectrum of these cold objects. \section{Comparison to the Models} \subsection{The Models} Leg13 compares near-infrared photometry to cloud-free models and to a new generation of models that include clouds consisting of sulfide and chloride condensates; we use the same models here. The cloud-free model atmospheres are as described in Saumon \& Marley (2008) and Marley et al. (2002), with updates to the line list of NH$_3$ and of the collision-induced absorption of H$_2$ as described in Saumon et al. (2012). The cloudy models are described in detail in Morley et al. (2012). Morley et al. have added absorption and scattering by condensates of Cr, MnS, Na$_2$S, ZnS and KCl to the cloud-free models. These condensates have been predicted to be present in low-$T_{\rm eff}$ atmospheres by Lodders (1999) and Visscher, Lodders \& Fegley (2006). Morley et al. use the Ackerman \& Marley (2001) cloud model to account for these previously neglected clouds. The vertical cloud extent is determined by balancing upward turbulent mixing and downward sedimentation. A parameter $f_{\rm sed}$ describes the efficiency of sedimentation, and is the ratio of the sedimentation velocity to the convective velocity; lower values of $f_{\rm sed}$ imply thicker (i.e. more vertically extended) clouds. We have found that models that include iron and silicate grains and which have $f_{\rm sed}$ of typically 2 -- 3 fit L dwarf spectra well, those with $f_{\rm sed}$ 2 -- 4 fit T0 to T3 spectral types well, and cloud-free models fit T4 -- T8 types well (e.g. Saumon \& Marley 2008, Stephens et al. 2009). However for the latest T-types significant discrepancies exist between the models and the observations, in the near-infrared (e.g. Leggett et al. 2009, 2012). The new models with the chloride and sulfide clouds help to resolve these discrepancies, because these clouds are significant for dwarfs with $T_{\rm eff} = 400$ -- 900~K (approximately T7 to Y1 spectral types), with a peak impact at around 600~K. Below $T_{\rm eff} \sim 400\,\rm K$ water clouds are expected to form (Burrows, Sudarsky \& Lunine 2003), which have not yet been incorporated into the models (although water condensation is accounted for in the gas opacity). The models used in the present analysis have solar metallicity and neglect departures from chemical equilibrium caused by vertical mixing. The mixing enhances the abundance of CO and CO$_2$ and reduces the 5 $\mu$m flux (Saumon et al. 2006). Vertical mixing also decreases the abundance of NH$_3$, which would otherwise produce stronger absorption features at 1.03 and 1.52 $\mu$m than are seen in the known Y dwarfs (Leg13). The mixing can be parameterized with the eddy diffusion coefficient $K_{zz}$ cm$^2$ s$^{-1}$, where values of log $K_{zz} = 2$ -- 6 corresponding to mixing timescales of $\sim 10$ yr to $\sim 1$ hr, respectively, reproduce the observations of T dwarfs (e.g. Saumon et al. 2007). Leggett et al. (2012) find that the $T_{\rm eff}= 500\,$~K dwarf UGPS J0722$-$0540 is undergoing vigorous mixing, with log $K_{zz} \approx 5.5$ -- 6.0, and this impacts the {\it WISE} 4.6 $\mu$m W2 band by $\gtrsim 0.3$ magnitude. Leg13 find that increasing the calculated W2 flux by 0.3 magnitude results in the model sequences reproducing the observed color trends in T and Y dwarfs quite well. \subsection{Previously Determined Properties of WISE 1217$+$16AB} Liu12 derived a photometric distance of 10.5 $\pm 1.7$ pc for WISE 1217$+$16A based on a spectral type assignment of T9 and a $J$-band bolometric correction for very late-type T dwarfs. The luminosity was combined with evolutionary models to derive physical properties for ages of 1 Gyr and 5 Gyr. For the younger age the inferred $T_{\rm eff}$ and mass are 550~K and 13 $M_{\rm Jup}$ for the primary, and 400~K and 7 $M_{\rm Jup}$ for the secondary. For the older age these values are 650~K and 33 $M_{\rm Jup}$ for the primary, and 400~K and 17 $M_{\rm Jup}$ for the secondary. Dupuy \& Kraus (2013) have recently published a trigonometric distance to the binary of $10.1_{-1.4}^{+1.9}\,$pc. Using this distance and model-based bolometric corrections to the summed observed flux given by the measured magnitudes, they determine, for an age of 5 Gyr, $T_{\rm eff}$ and mass of 600~K and 31 $M_{\rm Jup}$ for the primary, and 450~K and 19 $M_{\rm Jup}$ for the secondary. Leg13 compare the observed resolved near-infrared colors of the WISE 1217$+$16 components to the Morley et al. (2012) cloudy models. The W2 magnitudes for each component are estimated based on spectral type, constrained by the unresolved W2 value. Leg13 find that the model sequences are consistent with a single-age solution for the binary. Higher gravity solutions, corresponding to an age $\sim$5 Gyr, provided better fits than the lower gravity corresponding to 1 Gyr, because of the relatively blue $H - K$ color. The colors also suggested that the atmospheres of both components had thin cloud layers with $f_{\rm sed} \approx 5$. The proper motion for the binary is measured to be $1\farcs45 \pm 0\farcs04$ yr$^{-1}$ (Dupuy \& Kraus 2013), implying a tangential velocity of $70 \pm 10$ km s$^{-1}$. This velocity implies kinematics intermediate between the thin and thick disk populations (e.g. Brook et al. 2012; Dupuy \& Liu 2012) and therefore an age around 7 Gyr. This result is consistent with the findings of Leg13, that the higher gravity and therefore older age of 5 Gyr is favored over the younger 1 Gyr age. \subsection{Color-Magnitude and Spectral Energy Distribution} Figure 5 illustrates the location of WISE 1217$+$16A and WISE 1217$+$16B in a near-infrared color-magnitude diagram. Here $M_H$ is used as the luminosity indicator as $H$ is less sensitive to the clouds than $Y$ or $J$, and less sensitive to metallicity and gravity than $K$. The intrinsically faintest sources are identified. Photometry and parallaxes are taken from Leg13 and references therein, updated by measurements from Beichman et al. (2013), Dupuy \& Kraus (2013), Kirkpatrick et al. (2013), Mace et al. (2013), Marsh et al. (2013) and Wright et al. (2013). Note that the latest type Y dwarf currently identified, WISEP J182831.08+265037.8 (\#15), appears to be unusually bright in $H$ (Beichman et al. 2013, Dupuy \& Kraus 2013; see also \S 3). The typical spectral types and approximate $T_{\rm eff}$ at particular $M_H$ are shown along the right axis of Figure 5 (e.g. Burningham et al. 2010; Dupuy \& Kraus 2013; Leggett et al. 2009, 2010a,b, 2012, 2013; Liu12; Pinfield et al. 2012; Smart et al. 2010; Wright et al. 2013). The $T_{\rm eff}$ have been derived primarily from luminosity arguments and evolutionary models. There are three brown dwarfs with $T_{\rm eff} \approx 600$~K and $\log g \approx 5.0$ that are of particular interest: BD $+01^{\circ} 2920$B (2), SDSS J141624.08+134826.7B (3) and Wolf 940B (5). These are companions to a G, L and M dwarf, respectively. Although $T_{\rm eff}$ and $\log g$ are similar, Wolf 940B has a metallicity close to solar, while SDSS J141624.08+134826.7B and BD $+01^{\circ} 2920$B have [Fe/H] $\lesssim -0.3$ dex (Burgasser et al. 2010, Burningham et al. 2010, Leggett et al. 2010b, Pinfield et al. 2012; see also Burningham et al. 2013). The impact of the lower metallicity is clearly seen in the $H - K$ color in Figure 5, which also suggests that the WISE 1217$+$16 system may be slightly metal poor. If, instead, the blue $H - K$ is due to gravity only, the size of the shift ($\approx 0.3$ magnitudes) implies a gravity $\sim 1.0$ dex higher than typical for the type (e.g. Burningham et al. 2013, their Figure 10). A gravity this large is unlikely, given that at $T_{\rm eff} \approx 600$~K an increase in age from 1 to 10 Gyr corresponds to an increase in gravity of 0.6 dex (Saumon \& Marley 2008, their Figure 4). Hence we suggest that the system has a relatively high gravity combined with a metallicity of about $-0.1$ dex. It can be seen that the absolute $H$ magnitudes for the WISE 1217$+$16 components are consistent with the assigned spectral types of T8.5 and Y0--0.5. Using the previous studies of late T and early Y dwarfs, the figure suggests that the components have $T_{\rm eff}=550$ -- 600$\,$K and 400 -- 450$\,$K, respectively, which is also consistent with previous determinations (\S 4.2). Table 2 gives physical parameters for each component for these values of $T_{\rm eff}$, for a range in age of 4 Gyr to 10 Gyr, calculated using the evolutionary models of Saumon \& Marley (2008). Figure 6 shows the spectrum of each component and synthetic spectra generated by the Morley et al. (2012) models. The spectra have been scaled by the known distance to the binary (Table 1), and the radius of each component as given by the evolutionary models for an age of 6 Gyr (Table 2). If the system is younger the radius is larger and the synthetic spectrum would be brighter, and vice versa. For ages of 2 and 4 Gyr the scaling factor is 20\% and 8\% larger, while for ages of 8 and 10 Gyr the factor is 6\% and 8\% smaller, respectively. Strictly, if $T_{\rm eff}$ is kept constant and radius changed then gravity also changes, which would impact the spectral energy distribution. However for the purpose of constraining the model by the luminosity, the gravity impact is small. For example, if $T_{\rm eff} = 500$~K, and age is 6 Gyr, then radius and gravity are $R=0.0938\,R_\odot$ and $\log g {\rm (cgs)}=4.888$. If $T_{\rm eff} = 500\,$~K, and age is 10$\,$Gyr, then radius and gravity are 0.0896$\,R_\odot$ and $\log g=5.026$ (Saumon \& Marley 2008). The spectral change due to a change in gravity of 0.14 dex is small (e.g. Leggett et al. 2009), while the change in the flux scaling factor ($R^2$) of 10\% is significant. Another factor to bear in mind when examining Figure 6 is that the models do not include the vertical mixing that likely occurs in such atmospheres (e.g. Leg13, \S 4.1). The mixing is expected to increase the abundance of N$_2$ at the expense of that of NH$_3$. In the near-infrared this affects the $H$-band in particular; the NH$_3$ absorption is much reduced, making the blue wing and peak of the $H$-band brighter. The effect is $\sim$ 20\% at the peak of the $H$-band at these temperatures (based on preliminary Saumon \& Marley models). The model comparison in Figure 6 and the relative strengths of the $Y$, $J$ and $H$ peaks ($K$ is very faint) suggests that each component of the WISE 1217$+$16 binary has very thin to no sulfide/chloride clouds: $f_{\rm sed} \gtrsim 6$. The brightness of the flux peaks, especially considering that the $H$ peak is likely to be brighter when mixing is included, suggests that the age is unlikely to be less than 4 Gyr, as the model spectra will then be too bright. The system may be as old as 10 Gyr, as shown in the inset in Figure 6, although there is a discrepancy at $J$ for the B component in that case. Enhanced H$_2$ absorption, if the system is metal poor, could reduce the flux at $K$, and possibly at the blue wing of $J$ for the cooler dwarf, improving the fit. The relative height of the flux peaks are likely to also be sensitive to the detailed structure of any cloud decks, as is seen at the L/T dwarf transition (e.g. Marley et al. 2012, Apai et al. 2013). In summary, plausible fits are obtained for an age range of 4 to 8 Gyr. For the A component the best match to the models occurs if $550 \lesssim T_{\rm eff}$~(K) $\leq 600$ and there are no clouds. For the B component the best match to the models occurs if $T_{\rm eff} \approx 450$~K and there are no clouds or an extremely thin sulfide/chloride cloud layer with $f_{\rm sed} > 5$. Other spectral comparisons (not shown), where the synthetic spectra are scaled by the known distance and the evolutionary-determined radius, showed that we can exclude $T_{\rm eff}$ values of 500~K for either the A or B component, due to large discrepancies in brightness levels. Similarly temperatures as high as 650~K can be excluded for WISE 1217$+$16A. Although we do not have 350~K model spectra, luminosity arguments (see Figure 5 and Dupuy \& Kraus 2013, their Table S5) show that WISE 1217$+$16B cannot be as cool as 350~K. Table 3 summarises the likely values for the physical properties for the system. \section{Conclusion} We have used nights of excellent seeing on Mauna Kea to obtain resolved 0.9 -- 2.3 $\mu$m spectra for the T dwarf and Y dwarf binary WISEPC J121756.91$+$162640.2AB. The spectral extraction is confirmed to be accurate by comparison to the unresolved spectrum and resolved $H$-band spectra obtained previously. Comparison to the near-infrared spectra of (the small number of) very late-type T dwarfs and early-type Y dwarfs implies spectral types of T8.5 and Y0 -- Y0.5 for the primary and secondary, respectively. Thus the system straddles the currently defined T/Y spectral type boundary. Using synthetic spectra generated by model atmospheres that include chloride and sulfide clouds (Morley et al. 2012), and constrained by the distance to the system (Dupuy \& Kraus 2013) and the radius of each component based on evolutionary models (Saumon \& Marley 2008), we can determine a probable range of physical properties for the binary. The effective temperature of the primary is 550 -- 600~K, and that of the secondary is 450~K. Temperatures warmer or cooler by 50~K can be excluded as they result in significant discrepancies in brightness between the observed and calculated spectra. The shapes of the spectral distributions show that the atmospheres of both components have either very thin or no chloride/sulfide cloud layers, with a sedimentation parameter $f_{\rm sed} \gtrsim 6$. We find that the masses of the primary and secondary are around 30 and $22\,M_{\rm Jup}$, respectively, and that the age of the system is 4 -- 8 Gyr. This age is consistent with astrometric measurements by Dupuy \& Kraus (2013) which show that the system has kinematics intermediate between the thin and thick disk populations of the Galaxy. The system may be metal poor based on the $H - K$ colors of both components, which would also generally be consistent with an older age. Coeval binary systems such as WISEPC J121756.91$+$162640.2AB offer a powerful probe of the atmospheric changes that occur at very low temperatures. At $T_{\rm eff} \approx 500$~K, the alkali elements are condensing, and cloud decks of sulfides and chlorides form. This system offers an insight into the interplay between temperature, gravity, metallicity and cloud formation in cold atmospheres, and will provide a benchmark for the models as they are improved with more complete treatment of opacities, clouds and turbulence. In the near term, model atmospheres with a range of metallicity and mixing efficiency would enable a significant improvement in our understanding of the recently discovered Y dwarf population. \acknowledgments This research was supported by NSF grants AST09-09222 awarded to MCL. DS is supported by NASA Astrophysics Theory grant NNH11AQ54I. Based on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the Science and Technology Facilities Council (United Kingdom), the National Research Council (Canada), CONICYT (Chile), the Australian Research Council (Australia), Minist\'{e}rio da Ci\^{e}ncia, Tecnologia e Inova\c{c}\~{a}o (Brazil) and Ministerio de Ciencia, Tecnolog\'{i}a e Innovaci\'{o}n Productiva (Argentina). SKL's research is supported by Gemini Observatory. This publication makes use of data products from the Wide-field Infrared Survey Explorer, which is a joint project of the University of California, Los Angeles, and the Jet Propulsion Laboratory/California Institute of Technology, funded by the National Aeronautics and Space Administration. This research has made use of the NASA/ IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
\chapter{Introduction} \label{chap-intro} \def\Degsym{D}% In the Fall of 2012, my friend Kurt Reillag suggested that I should be ashamed about knowing so little about graph Laplacians and normalized graph cuts. These notes are the result of my efforts to rectify this situation. \medskip I begin with a review of basic notions of graph theory. Even though the graph Laplacian is fundamentally associated with an undirected graph, I review the definition of both directed and undirected graphs. For both directed and undirected graphs, I define the degree matrix $D$, the incidence matrix $\widetilde{D}$, and the adjacency matrix $A$. I also define weighted graphs (with nonnegative weights), and the notions of {\it volume\/}, $\mathrm{vol}(A)$ of a set of nodes $A$, of {\it links\/}, $\mathrm{links}(A, B)$ between two sets of nodes $A, B$, and of {\it cut\/}, $\mathrm{cut}(A) = \mathrm{links}(A, \overline{A})$ of a set of nodes $A$. These concepts play a crucial role in the theory of normalized cuts. Then, I introduce the (unnormalized) {\it graph Laplacian\/} $L$ of a directed graph $G$ in an ``old-fashion,'' by showing that for any orientation of a graph $G$, \[ \widetilde{D}\transpos{\widetilde{D}} = D - A = L \] is an invariant. I also define the (unnormalized) {\it graph Laplacian\/} $L$ of a weighted graph $(V, W)$ as $L = D - W$, and prove that \[ \transpos{x} L x = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} (x_i - x_j)^2 \quad\mathrm{for\ all}\> x\in \mathbb{R}^m. \] Consequently, $\transpos{x} L x$ does not depend on the diagonal entries in $W$, and if $w_{i\, j} \geq 0$ for all $i, j\in \{1, \ldots,m\}$, then $L$ is positive semidefinite. Then, if $W$ consists of nonnegative entries, the eigenvalues $0 = \lambda_1 \leq \lambda_2 \leq \ldots \leq \lambda_m$ of $L$ are real and nonnegative, and there is an orthonormal basis of eigenvectors of $L$. I show that the number of connected components of the graph $G = (V,W)$ is equal to the dimension of the kernel of $L$. \medskip I also define the normalized graph Laplacians $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$, given by \begin{align*} L_{\mathrm{sym}}& = \Degsym^{-1/2} L \Degsym^{-1/2} = I - \Degsym^{-1/2} W \Degsym^{-1/2} \\ L_{\mathrm{rw}}& = \Degsym^{-1} L = I - \Degsym^{-1} W, \end{align*} and prove some simple properties relating the eigenvalues and the eigenvectors of $L$, $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$. These normalized graph Laplacians show up when dealing with normalized cuts. \medskip Next, I turn to {\it graph drawings\/} (Chapter \ref{chap2}). Graph drawing is a very attractive application of so-called spectral techniques, which is a fancy way of saying that that eigenvalues and eigenvectors of the graph Laplacian are used. Furthermore, it turns out that graph clustering using normalized cuts can be cast as a certain type of graph drawing. \medskip Given an undirected graph $G = (V, E)$, with $|V| = m$, we would like to draw $G$ in $\mathbb{R}^n$ for $n$ (much) smaller than $m$. The idea is to assign a point $\rho(v_i)$ in $\mathbb{R}^n$ to the vertex $v_i\in V$, for every $v_i \in V$, and to draw a line segment between the points $\rho(v_i)$ and $\rho(v_j)$. Thus, a {\it graph drawing\/} is a function $\mapdef{\rho}{V}{\mathbb{R}^n}$. \medskip We define the {\it matrix of a graph drawing $\rho$ (in $\mathbb{R}^n$)\/} as a $m \times n$ matrix $R$ whose $i$th row consists of the row vector $\rho(v_i)$ corresponding to the point representing $v_i$ in $\mathbb{R}^n$. Typically, we want $n < m$; in fact $n$ should be much smaller than $m$. \medskip Since there are infinitely many graph drawings, it is desirable to have some criterion to decide which graph is better than another. Inspired by a physical model in which the edges are springs, it is natural to consider a representation to be better if it requires the springs to be less extended. We can formalize this by defining the {\it energy\/} of a drawing $R$ by \[ \s{E}(R) = \sum_{\{v_i, v_j\}\in E} \norme{\rho(v_i) - \rho(v_j)}^2, \] where $\rho(v_i)$ is the $i$th row of $R$ and $\norme{\rho(v_i) - \rho(v_j)}^2$ is the square of the Euclidean length of the line segment joining $\rho(v_i)$ and $\rho(v_j)$. \medskip Then, ``good drawings'' are drawings that minimize the energy function $\s{E}$. Of course, the trivial representation corresponding to the zero matrix is optimum, so we need to impose extra constraints to rule out the trivial solution. \medskip We can consider the more general situation where the springs are not necessarily identical. This can be modeled by a symmetric weight (or stiffness) matrix $W = (w_{i j})$, with $w_{i j} \geq 0$. In this case, our energy function becomes \[ \s{E}(R) = \sum_{\{v_i, v_j\}\in E} w_{i j} \norme{\rho(v_i) - \rho(v_j)}^2. \] Following Godsil and Royle \cite{Godsil}, we prove that \[ \s{E}(R) = \mathrm{tr}(\transpos{R} L R), \] where \[ L = \Degsym - W, \] is the familiar unnormalized Laplacian matrix associated with $W$, and where $\Degsym$ is the degree matrix associated with $W$. \medskip It can be shown that there is no loss in generality in assuming that the columns of $R$ are pairwise orthogonal and that they have unit length. Such a matrix satisfies the equation $\transpos{R} R = I$ and the corresponding drawing is called an {\it orthogonal drawing\/}. This condition also rules out trivial drawings. \medskip Then, I prove the main theorem about graph drawings (Theorem \ref{graphdraw}), which essentially says that the matrix $R$ of the desired graph drawing is constituted by the $n$ eigenvectors of $L$ associated with the smallest nonzero $n$ eigenvalues of $L$. We give a number examples of graph drawings, many of which are borrowed or adapted from Spielman \cite{Spielman}. \medskip The next chapter (Chapter \ref{chap3}) contains the ``meat'' of this document. This chapter is devoted to the method of normalized graph cuts for graph clustering. This beautiful and deeply original method first published in Shi and Malik \cite{ShiMalik}, has now come to be a ``textbook chapter'' of computer vision and machine learning. It was invented by Jianbo Shi and Jitendra Malik, and was the main topic of Shi's dissertation. This method was extended to $K \geq 3$ clusters by Stella Yu in her dissertation \cite{Yu}, and is also the subject of Yu and Shi \cite{YuShi2003}. \medskip Given a set of data, the goal of clustering is to partition the data into different groups according to their similarities. When the data is given in terms of a similarity graph $G$, where the weight $w_{i\, j}$ between two nodes $v_i$ and $v_j$ is a measure of similarity of $v_i$ and $v_j$, the problem can be stated as follows: Find a partition $(A_1, \ldots, A_K)$ of the set of nodes $V$ into different groups such that the edges between different groups have very low weight (which indicates that the points in different clusters are dissimilar), and the edges within a group have high weight (which indicates that points within the same cluster are similar). \medskip The above graph clustering problem can be formalized as an optimization problem, using the notion of cut mentioned earlier. If we want to partition $V$ into $K$ clusters, we can do so by finding a partition ($A_1, \ldots, A_K$) that minimizes the quantity \[ \mathrm{cut}(A_1, \ldots, A_K) = \frac{1}{2} \sum_{1 = 1}^K \mathrm{cut}(A_i). \] For $K = 2$, the mincut problem is a classical problem that can be solved efficiently, but in practice, it does not yield satisfactory partitions. Indeed, in many cases, the mincut solution separates one vertex from the rest of the graph. What we need is to design our cost function in such a way that it keeps the subsets $A_i$ ``reasonably large'' (reasonably balanced). \medskip A example of a weighted graph and a partition of its nodes into two clusters is shown in Figure \ref{ncg-fig4a}. \begin{figure}[http] \begin{center} \includegraphics[height=2.5truein,width=2.8truein]{ncuts-figs/ncuts-g-fig1.pdf} \hspace{0.5cm} \includegraphics[height=2.5truein,width=2.8truein]{ncuts-figs/ncuts-g-fig2.pdf} \end{center} \caption{A weighted graph and its partition into two clusters.} \label{ncg-fig4a} \end{figure} \medskip A way to get around this problem is to normalize the cuts by dividing by some measure of each subset $A_i$. A solution using the volume $\mathrm{vol}(A_i)$ of $A_i$ (for $K = 2$) was proposed and investigated in a seminal paper of Shi and Malik \cite{ShiMalik}. Subsequently, Yu (in her dissertation \cite{Yu}) and Yu and Shi \cite{YuShi2003} extended the method to $K > 2$ clusters. The idea is to minimize the cost function \[ \mathrm{Ncut}(A_1, \ldots, A_K) = \sum_{i = 1}^K \frac{\mathrm{links}(A_i, \overline{A_i})}{\mathrm{vol}(A_i)} = \sum_{i = 1}^K \frac{\mathrm{cut}(A_i, \overline{A_i})}{\mathrm{vol}(A_i)}. \] \medskip The first step is to express our optimization problem in matrix form. In the case of two clusters, a single vector $X$ can be used to describe the partition $(A_1, A_2) = (A, \overline{A})$. We need to choose the structure of this vector in such a way that \[ \mathrm{Ncut}(A, \overline{A}) = \frac{\transpos{X} L X}{\transpos{X} \Degsym X}, \] where the term on the right-hand side is a Rayleigh ratio. \medskip After careful study of the orginal papers, I discovered various facts that were implicit in these works, but I feel are important to be pointed out explicitly. \medskip First, I realized that it is important to pick a vector representation which is invariant under multiplication by a nonzero scalar, because the Rayleigh ratio is scale-invariant, and it is crucial to take advantage of this fact to make the denominator go away. This implies that {\it the solutions $X$ are points in the projective space $\mathbb{RP}^{N - 1}$\/}. This was my first revelation. \medskip Let $N = |V|$ be the number of nodes in the graph $G$. In view of the desire for a scale-invariant representation, it is natural to assume that the vector $X$ is of the form \[ X = (x_1, \ldots, x_N), \] where $x_i \in \{a, b\}$ for $i = 1, \ldots, N$, for any two distinct real numbers $a, b$. This is an indicator vector in the sense that, for $i = 1, \ldots, N$, \[ x_i = \begin{cases} a & \text{if $v_i \in A$} \\ b & \text{if $v_i \notin A$} . \end{cases} \] The choice $a = +1, b = -1$ is natural, but premature. The correct interpretation is really to view $X$ as a representative of a point in the real projective space $\mathbb{RP}^{N-1}$, namely the point $\mathbb{P}(X)$ of homogeneous coordinates $(x_1\colon \cdots \colon x_N)$. \medskip Let $d = \transpos{\mathbf{1}} \Degsym \mathbf{1}$ and $\alpha = \mathrm{vol}(A)$. I prove that \[ \mathrm{Ncut}(A, \overline{A}) = \frac{\transpos{X} L X}{\transpos{X} \Degsym X} \] holds iff the following condition holds: \begin{equation} a \alpha + b(d - \alpha) = 0. \tag{$\dagger$} \end{equation} Note that condition $(\dagger)$ applied to a vector $X$ whose components are $a$ or $b$ is equivalent to the fact that $X$ is orthogonal to $\Degsym \mathbf{1}$, since \[ \transpos{X} \Degsym \mathbf{1} = \alpha a+ (d - \alpha) b, \] where $\alpha = \mathrm{vol}(\{v_i\in V \mid x_i = a\})$. \medskip If we let \[ \s{X} = \big\{ (x_1, \ldots, x_N) \mid x_i \in \{a, b\}, \> a, b\in \mathbb{R},\> a, b\not = 0 \big\}, \] our solution set is \[ \s{K} = \big\{ X \in\s{X} \mid \transpos{X} \Degsym\mathbf{1} = 0 \big\}. \] Actually, to be perfectly rigorous, we are looking for solutions in $\mathbb{RP}^{N-1}$, so our solution set is really \[ \mathbb{P}(\s{K}) = \big\{ (x_1\colon \cdots\colon x_N) \in \mathbb{RP}^{N-1}\mid (x_1, \ldots, x_N) \in \s{K} \big\}. \] Consequently, our minimization problem can be stated as follows: \medskip\noindent {\bf Problem PNC1} \begin{align*} & \mathrm{minimize} & & \frac{\transpos{X} L X}{\transpos{X} \Degsym X} & & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym\mathbf{1} = 0, & & X\in \s{X}. \end{align*} It is understood that the solutions are points $\mathbb{P}(X)$ in $\mathbb{RP}^{N-1}$. \medskip Since the Rayleigh ratio and the constraints $\transpos{X}\Degsym\mathbf{1} = 0$ and $X\in \s{X}$ are scale-invariant, we are led to the following formulation of our problem: \medskip\noindent {\bf Problem PNC2} \begin{align*} & \mathrm{minimize} & & \transpos{X} L X & & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym X = 1, && \transpos{X} \Degsym\mathbf{1} = 0, && X\in \s{X}. \end{align*} \medskip Problem PNC2 is equivalent to problem PNC1 in the sense that they have the same set of minimal solutions as points $\mathbb{P}(X) \in\mathbb{RP}^{N-1}$ given by their homogenous coordinates $X$. More precisely, if $X$ is any minimal solution of PNC1, then $X/(\transpos{X} \Degsym X)^{1/2}$ is a minimal solution of PNC2 (with the same minimal value for the objective functions), and if $X$ is a minimal solution of PNC2, then $\lambda X$ is a minimal solution for PNC1 for all $\lambda\not= 0$ (with the same minimal value for the objective functions). \medskip Now, as in the classical papers, we consider the relaxation of the above problem obtained by dropping the condition that $X\in \s{X}$, and proceed as usual. However, having found a solution $Z$ to the relaxed problem, we need to find a discrete solution $X$ such that $d(X, Z)$ is minimum in $\mathbb{RP}^{N-1}$. All this presented in Section \ref{ch3-sec2}. \medskip If the number of clusters $K$ is at least $3$, then we need to choose a matrix representation for partitions on the set of vertices. It is important that such a representation be scale-invariant, and it is also necessary to state necessary and sufficient conditions for such matrices to represent a partition (to the best of our knowledge, these points are not clearly articulated in the literature). \medskip We describe a partition $(A_1, \ldots, A_K)$ of the set of nodes $V$ by an $N\times K$ matrix $X = [X^1 \cdots X^K]$ whose columns $X^1, \ldots, X^K$ are indicator vectors of the partition $(A_1, \ldots, A_K)$. Inspired by what we did when $K = 2$, we assume that the vector $X^j$ is of the form \[ X^j = (x_1^j, \ldots, x_N^j), \] where $x_i^j \in \{a_j, b_j\}$ for $j = 1, \ldots, K$ and $i = 1, \ldots, N$, and where $a_j, b_j$ are any two distinct real numbers. The vector $X^j$ is an indicator vector for $A_j$ in the sense that, for $i = 1, \ldots, N$, \[ x_i^j = \begin{cases} a_j & \text{if $v_i \in A_j$} \\ b_j & \text{if $v_i \notin A_j$} . \end{cases} \] The choice $\{a_j, b_j\} = \{0, 1\}$ for $j = 1, \ldots, K$ is natural, but premature. I show that if we pick $b_i = 0$, then we have \[ \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)} = \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} \quad j = 1, \ldots, K, \] which implies that \[ \mathrm{Ncut}(A_1, \ldots, A_K) = \sum_{j = 1}^K \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)} = \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j}. \] Then, I give necessary and sufficient conditions for a matrix $X$ to represent a partition. \medskip If we let \[ \s{X} = \Big\{[X^1\> \ldots \> X^K] \mid X^j = a_j(x_1^j, \ldots, x_N^j) , \> x_i^j \in \{1, 0\}, a_j\in \mathbb{R}, \> X^j \not= 0 \Big\} \] (note that the condition $X^j \not= 0$ implies that $a_j \not= 0$), then the set of matrices representing partitions of $V$ into $K$ blocks is \begin{align*} & & &\s{K} = \Big\{ X = [X^1 \> \cdots \> X^K] \quad \mid & & X\in\s{X}, &&\\ & & & & & \transpos{(X^i)} \Degsym X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, && \quad\quad\quad\quad\quad\\ & & & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}\Big\}. && \end{align*} As in the case $K = 2$, to be rigorous, the {\it solution are really $K$-tuples of points in $\mathbb{RP}^{N-1}$\/}, so our solution set is really \[ \mathbb{P}(\s{K}) = \Big\{(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K)) \mid [X^1 \> \cdots \> X^K] \in \s{K} \Big\}. \] \medskip In view of the above, we have our first formulation of $K$-way clustering of a graph using normalized cuts, called problem PNC1 (the notation PNCX is used in Yu \cite{Yu}, Section 2.1): \medskip\noindent {\bf $K$-way Clustering of a graph using Normalized Cut, Version 1: \\ Problem PNC1} \begin{align*} & \mathrm{minimize} & & \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j}& & & &\\ & \mathrm{subject\ to} & & \transpos{(X^i)} \Degsym X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, & & & & \\ & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}, & & X\in \s{X}. & & \end{align*} As in the case $K = 2$, the solutions that we are seeking are $K$-tuples $(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K))$ of points in $\mathbb{RP}^{N-1}$ determined by their homogeneous coordinates $X^1, \ldots, X^K$. \medskip Then, step by step, we transform problem PNC1 into an equivalent problem PNC2, which we eventually relax by dropping the condition that $X\in \s{X}$. \medskip Our second revelation is that the relaxation $(*_1)$ of version 2 of our minimization problem (PNC2), which is equivalent to version 1, reveals that that the solutions of the relaxed problem $(*_1)$ are members of the {\it Grassmannian\/} $G(K, N)$. \medskip This leads us to our third revelation: {\it we have two choices of metrics to compare solutions\/}: (1) a metric on $(\mathbb{RP}^{N - 1})^K$; (2) a metric on $G(K, N)$. We discuss the first choice, which is the choice implicitly adopted by Shi and Yu. \medskip Some of the most technical material on the Rayleigh ratio, which is needed for some proofs in Chapter \ref{chap2}, is the object of Appendix \ref{Rayleigh-Ritz}. Appendix \ref{ch3-sec6} may seem a bit out of place. Its purpose is to explain how to define a metric on the projective space $\mathbb{RP}^n$. For this, we need to review a few notions of differential geometry. \medskip I hope that these notes will be make it easier for people to become familiar with the wonderful theory of normalized graph cuts. As far as I know, except for a short section in one of Gilbert Strang's book, and von Luxburg \cite{Luxburg} excellent survey on spectral clustering, there is no comprehensive writing on the topic of normalized cuts. \chapter{Graphs and Graph Laplacians; Basic Facts} \label{chap1} \section[Directed Graphs, Undirected Graphs, Weighted Graphs] {Directed Graphs, Undirected Graphs, Incidence Matrices, Adjacency Matrices, Weighted Graphs} \label{ch1-sec1}% \begin{definition} \label{dirgraph} A {\it directed graph\/} is a pair $G = (V, E)$, where $V = \{v_1, \ldots, v_m\}$ is a set of {\it nodes\/} or {\it vertices\/}, and $E\subseteq V \times V$ is a set of ordered pairs of distinct nodes (that is, pairs $(u, v)\in V\times V$ with $u\not= v$), called {\it edges\/}. Given any edge $e = (u, v)$, we let $s(e) = u$ be the {\it source\/} of $e$ and $t(e) = v$ be the {\it target\/} of $e$. \end{definition} \medskip \bigskip\noindent{\bf Remark:}\enspace Since an edge is a pair $(u, v)$ with $u\not= v$, self-loops are not allowed. Also, there is at most one edge from a node $u$ to a node $v$. Such graphs are sometimes called {\it simple graphs\/}. \medskip An example of a directed graph is shown in Figure \ref{graphfig17}. \begin{figure}[http] \begin{center} \includegraphics[height=1.7truein,width=2.3truein]{ncuts-figs/cis160book-graphfig17.pdf} \end{center} \caption{Graph $G_1$.} \label{graphfig17} \end{figure} \medskip For every node $v\in V$, the {\it degree\/} $d(v)$ of $v$ is the number of edges leaving or entering $v$: \[ d(v) = |\{u\in V \mid (v, u)\in E\> \mathrm{or}\> (u, v)\in E\}|. \] The {\it degree matrix\/} $\Degsym(G)$, is the diagonal matrix \[ \Degsym(G) = \mathrm{diag}(d_1, \ldots, d_m). \] For example, for graph $G_1$, we have \[ \Degsym(G_1) = \begin{pmatrix} 2 & 0 & 0 & 0 & 0 \\ 0 & 4 & 0 & 0 & 0 \\ 0 & 0 & 3 & 0 & 0 \\ 0 & 0 & 0 & 3 & 0 \\ 0 & 0 & 0 & 0 & 2 \end{pmatrix}. \] Unless confusion arises, we write $\Degsym$ instead of $\Degsym(G)$. \begin{definition} \label{incidence-matrix1} Given a directed graph $G = (V, E)$, with $V = \{v_1, \ldots, v_m\}$, if $E = \{e_1, \ldots, e_n\}$, then the {\it incidence matrix\/} $\widetilde{D}(G)$ of $G$ is the $m\times n$ matrix whose entries $\widetilde{d}_{i\, j}$ are given by \[ \widetilde{d}_{i\, j} = \begin{cases} +1 & \text{if $e_j = (v_i, v_k)$ for some $k$} \\ -1 & \text{if $e_j = (v_k, v_i)$ for some $k$} \\ 0 & \text{otherwise}. \end{cases} \] \end{definition} Here is the incidence matrix of the graph $G_1$: \[ \widetilde{D} = \begin{pmatrix} 1 & 1 & 0 & 0 & 0 & 0 & 0 \\ -1 & 0 & -1 & -1 & 1 & 0 & 0 \\ 0 & -1 & 1 & 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 1 & 0 & -1 & -1 \\ 0 & 0 & 0 & 0 & -1 & 1 & 0 \end{pmatrix}. \] \medskip Again, unless confusion arises, we write $\widetilde{D}$ instead of $\widetilde{D}(G)$. \medskip Undirected graphs are obtained from directed graphs by forgetting the orientation of the edges. \begin{definition} \label{graph} A {\it graph\/} (or {\it undirected graph\/}) is a pair $G = (V, E)$, where $V = \{v_1, \ldots, v_m\}$ is a set of {\it nodes\/} or {\it vertices\/}, and $E$ is a set of two-element subsets of $V$ (that is, subsets $\{u, v\}$, with $u, v\in V$ and $u\not= v$), called {\it edges\/}. \end{definition} \medskip \bigskip\noindent{\bf Remark:}\enspace Since an edge is a set $\{u, v\}$, we have $u\not= v$, so self-loops are not allowed. Also, for every set of nodes $\{u, v\}$, there there is at most one edge between $u$ and $v$. As in the case of directed graphs, such graphs are sometimes called {\it simple graphs\/}. \medskip An example of a graph is shown in Figure \ref{graphfig5bis}. \begin{figure} \begin{center} \includegraphics[height=1.7truein,width=2.3truein]{ncuts-figs/cis160book-graphfig5bis.pdf} \end{center} \caption{The undirected graph $G_2$.} \label{graphfig5bis} \end{figure} \medskip For every node $v\in V$, the {\it degree\/} $d(v)$ of $v$ is the number of edges adjacent to $v$: \[ d(v) = |\{u\in V \mid \{u, v\}\in E\}|. \] The degree matrix $\Degsym$ is defined as before. The notion of incidence matrix for an undirected graph is not as useful as the in the case of directed graphs \begin{definition} \label{incidence-matrix2} Given a graph $G = (V, E)$, with $V = \{v_1, \ldots, v_m\}$, if $E = \{e_1, \ldots, e_n\}$, then the {\it incidence matrix\/} $\widetilde{D}(G)$ of $G$ is the $m\times n$ matrix whose entries $\widetilde{d}_{i\, j}$ are given by \[ \widetilde{d}_{i\, j} = \begin{cases} +1 & \text{if $e_j = \{v_i, v_k\}$ for some $k$} \\ 0 & \text{otherwise}. \end{cases} \] \end{definition} Unlike the case of directed graphs, the entries in the incidence matrix of a graph (undirected) are nonnegative. We usally write $\widetilde{D}$ instead of $\widetilde{D}(G)$. \medskip The notion of adjacency matrix is basically the same for directed or undirected graphs. \begin{definition} \label{adjacency} Given a directed or undirected graph $G = (V, E)$, with $V = \{v_1, \ldots, v_m\}$, the {\it adjacency matrix\/} $A(G)$ of $G$ is the symmetric $m\times m$ matrix $(a_{i\, j})$ such that \begin{enumerate} \item[(1)] If $G$ is directed, then \[ a_{i\, j} = \begin{cases} 1 & \text{if there is some edge $(v_i, v_j)\in E$ or some edge $(v_j, v_i)\in E$} \\ 0 & \text{otherwise}. \end{cases} \] \item[(2)] Else if $G$ is undirected, then \[ a_{i\, j} = \begin{cases} 1 & \text{if there is some edge $\{v_i, v_j\}\in E $} \\ 0 & \text{otherwise}. \end{cases} \] \end{enumerate} \end{definition} \medskip As usual, unless confusion arises, we write $A$ instead of $A(G)$. Here is the adjacency matrix of both graphs $G_1$ and $G_2$: \[ A = \begin{pmatrix} 0 & 1 & 1 & 0 & 0 \\ 1 & 0 & 1 & 1 & 1 \\ 1 & 1 & 0 & 1 & 0 \\ 0 & 1 & 1 & 0 & 1 \\ 0 & 1 & 0 & 1 & 0 \end{pmatrix}. \] \medskip In many applications, the notion of graph needs to be generalized to capture the intuitive idea that two nodes $u$ and $v$ are linked with a degree of certainty (or strength). Thus, we assign a nonnegative weights $w_{i\, j}$ to an edge $\{v_i, v_j\}$; the smaller $w_{i\, j}$ is, the weaker is the link (or similarity) between $v_i$ and $v_j$, and the greater $w_{i\, j}$ is, the stronger is the link (or similarity) between $v_i$ and $v_j$. \begin{definition} \label{graph-weighted} A {\it weighted graph\/} is a pair $G = (V, W)$, where $V = \{v_1, \ldots, v_m\}$ is a set of {\it nodes\/} or {\it vertices\/}, and $W$ is a symmetric matrix called the {\it weight matrix\/}, such that $w_{i\, j} \geq 0$ for all $i, j \in \{1, \ldots, m\}$, and $w_{i\, i} = 0$ for $i = 1, \ldots, m$. We say that a set $\{v_i, v_j\}$ is an edge iff $w_{i\, j} > 0$. The corresponding (undirected) graph $(V, E)$ with $E = \{\{e_i, e_j\} \mid w_{i\, j} > 0\}$, is called the {\it underlying graph\/} of $G$. \end{definition} \medskip \bigskip\noindent{\bf Remark:}\enspace Since $w_{i\, i} = 0$, these graphs have no self-loops. We can think of the matrix $W$ as a generalized adjacency matrix. The case where $w_{i\, j} \in \{0, 1\}$ is equivalent to the notion of a graph as in Definition \ref{graph}. \medskip We can think of the weight $w_{i\, j}$ of an edge $\{v_i, v_j\}$ as a degree of similarity (or affinity) in an image, or a cost in a network. An example of a weighted graph is shown in Figure \ref{ncg-fig1}. The thickness of the edges corresponds to the magnitude of its weight. \begin{figure}[http] \begin{center} \includegraphics[height=2.5truein,width=2.8truein]{ncuts-figs/ncuts-g-fig1.pdf} \end{center} \caption{A weighted graph.} \label{ncg-fig1} \end{figure} \medskip For every node $v_i\in V$, the {\it degree\/} $d(v_i)$ of $v_i$ is the sum of the weights of the edges adjacent to $v_i$: \[ d(v_i) = \sum_{j = 1}^m w_{i\, j}. \] Note that in the above sum, only nodes $v_j$ such that there is an edge $\{v_i, v_j\}$ have a nonzero contribution. Such nodes are said to be {\it adjacent\/} to $v_i$. The degree matrix $\Degsym$ is defined as before, namely by $\Degsym = \mathrm{diag}(d(v_1), \ldots, d(v_m))$. \medskip Following common practice, we denote by $\mathbf{1}$ the (column) vector whose components are all equal to $1$. Then, observe that $W \mathbf{1}$ is the (column) vector $(d(v_1), \ldots, d(v_m))$ consisting of the degrees of the nodes of the graph. \medskip Given any subset of nodes $A \subseteq V$, we define the {\it volume\/} $\mathrm{vol}(A)$ of $A$ as the sum of the weights of all edges adjacent to nodes in $A$: \[ \mathrm{vol}(A) = \sum_{v_i\in A} d(v_i) = \sum_{v_i \in A} \sum_{j = 1}^m w_{i\, j}. \] \bigskip\noindent{\bf Remark:}\enspace Yu and Shi \cite{YuShi2003} use the notation $\mathrm{degree}(A)$ instead of $\mathrm{vol}(A)$. The notions of degree and volume are illustrated in Figure \ref{ncg-fig2}. \begin{figure}[http] \begin{center} \includegraphics[height=2truein,width=2truein]{ncuts-figs/ncuts-g-fig3.pdf} \hspace{1cm} \includegraphics[height=2truein,width=2truein]{ncuts-figs/ncuts-g-fig4.pdf} \end{center} \caption{Degree and volume.} \label{ncg-fig2} \end{figure} \medskip Observe that $\mathrm{vol}(A) = 0$ if $A$ consists of isolated vertices, that is, if $w_{i\, j} = 0$ for all $v_i\in A$. Thus, it is best to assume that $G$ does not have isolated vertices. \medskip Given any two subset $A, B\subseteq V$ (not necessarily distinct), we define $\mathrm{links}(A, B)$ by \[ \mathrm{links}(A, B) = \sum_{v_i\in A, v_j\in B} w_{i\, j}. \] Since the matrix $W$ is symmetric, we have \[ \mathrm{links}(A, B) = \mathrm{links}(B, A), \] and observe that $\mathrm{vol}(A) = \mathrm{links}(A, V)$. \medskip The quantity $ \mathrm{links}(A, \overline{A}) = \mathrm{links}(\overline{A}, A)$, where $\overline{A} = V - A$ denotes the complement of $A$ in $V$, measures how many links escape from $A$ (and $\overline{A}$), and the quantity $\mathrm{links}(A,A)$ measures how many links stay within $A$ itself. The quantity \[ \mathrm{cut}(A) = \mathrm{links}(A, \overline{A}) \] is often called the {\it cut\/} of $A$, and the quantity \[ \mathrm{assoc}(A) = \mathrm{links}(A,A) \] is often called the {\it association\/} of $A$. Clearly, \[ \mathrm{cut}(A) + \mathrm{assoc}(A) = \mathrm{vol}(A). \] The notions of cut is illustrated in Figure \ref{ncg-fig3}. \begin{figure}[http] \begin{center} \includegraphics[height=2.2truein,width=2.2truein]{ncuts-figs/ncuts-g-fig5.pdf} \end{center} \caption{A Cut involving the set of nodes in the center and the nodes on the perimeter.} \label{ncg-fig3} \end{figure} \medskip We now define the most important concept of these notes: The Laplacian matrix of a graph. Actually, as we will see, it comes in several flavors. \section{Laplacian Matrices of Graphs} \label{ch1-sec2} Let us begin with directed graphs, although as we will see, graph Laplacians are fundamentally associated with undirected graph. The key proposition whose proof can be found in Gallier \cite{GallDiscmath} and Godsil and Royle \cite{Godsil} is this: \begin{proposition} \label{adjp2} Given any directed graph $G$ if $\widetilde{D}$ is the incidence matrix of $G$, $A$ is the adjacency matrix of $G$, and $\Degsym$ is the degree matrix such that $\Degsym_{i\, i} = d(v_i)$, then \[ \widetilde{D}\transpos{\widetilde{D}} = \Degsym - A. \] Consequently, $\widetilde{D}\transpos{\widetilde{D}}$ is independent of the orientation of $G$ and $\Degsym - A$ is symmetric, positive, semidefinite; that is, the eigenvalues of $\Degsym - A$ are real and nonnegative. \end{proposition} \medskip The matrix $L = \widetilde{D}\transpos{\widetilde{D}} = \Degsym - A$ is called the {\it (unnormalized) graph Laplacian\/} of the graph $G$. For example, the graph Laplacian of graph $G_1$ is \[ L = \begin{pmatrix} 2 & -1 & -1 & 0 & 0 \\ -1 & 4 & -1 & -1 & -1 \\ -1 & -1 & 3 & -1 & 0 \\ 0 & -1 & -1 & 3 & -1 \\ 0 & -1 & 0 & -1 & 2 \end{pmatrix}. \] \medskip The {\it (unnormalized) graph Laplacian\/} of an undirected graph $G = (V, E)$ is defined by \[ L = \Degsym - A. \] Since $L$ is equal to $\widetilde{D}\transpos{\widetilde{D}}$ for any orientation of $G$, it is also positive semidefinite. Observe that each row of $L$ sums to zero. Consequently, the vector $\mathbf{1}$ is in the nullspace of $L$. \bigskip\noindent{\bf Remark:}\enspace With the unoriented version of the incidence matrix (see Definition \ref{incidence-matrix2}), it can be shown that \[ \widetilde{D}\transpos{\widetilde{D}} = \Degsym + A. \] \medskip The natural generalization of the notion of graph Laplacian to weighted graphs is this: \begin{definition} \label{graphLaplacian} Given any weighted directed graph $G = (V, W)$ with $V = \{v_1, \ldots,v_m\}$, the {\it (unnormalized) graph Laplacian $L(G)$ of $G$\/} is defined by \[ L(G) = \Degsym(G) - W, \] where $\Degsym(G) = \mathrm{diag}(d_1, \ldots,d_m)$ is the degree matrix of $G$ (a diagonal matrix), with \[ d_i = \sum_{j = 1}^m w_{i \, j}. \] As usual, unless confusion arises, we write $L$ instead of $L(G)$. \end{definition} \medskip It is clear that each row of $L$ sums to $0$, so the vector $\mathbf{1}$ is the nullspace of $L$, but it is less obvious that $L$ is positive semidefinite. An easy way to prove this is to evaluate the quadratic form $\transpos{x} L x$. \begin{proposition} \label{Laplace1} For any $m\times m$ symmetric matrix $W$, if we let $L = \Degsym - W$ where $D$ is the degree matrix of $W = (w_{i j})$, then we have \[ \transpos{x} L x = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} (x_i - x_j)^2 \quad\mathrm{for\ all}\> x\in \mathbb{R}^m. \] Consequently, $\transpos{x} L x$ does not depend on the diagonal entries in $W$, and if $w_{i\, j} \geq 0$ for all $i, j\in \{1, \ldots,m\}$, then $L$ is positive semidefinite. \end{proposition} \begin{proof} We have \begin{align*} \transpos{x} L x & = \transpos{x} \Degsym x - \transpos{x} W x \\ & = \sum_{i = 1}^m d_i x_i^2 - \sum_{i, j = 1}^m w_{i\, j} x_i x_j \\ & = \frac{1}{2}\left( \sum_{i = 1}^m d_i x_i^2 - 2 \sum_{i, j = 1}^m w_{i\, j} x_i x_j + \sum_{i = 1}^m d_i x_i^2 \right) \\ & = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} (x_i - x_j)^2. \end{align*} Obviously, the quantity on the right-hand side does not depend on the diagonal entries in $W$, and if if $w_{i\, j} \geq 0$ for all $i, j$, then this quantity is nonnegative. \end{proof} \medskip Proposition \ref{Laplace1} immediately implies the following facts: For any weighted graph $G = (V, W)$, \begin{enumerate} \item The eigenvalues $0 = \lambda_1 \leq \lambda_2 \leq \ldots \leq \lambda_m$ of $L$ are real and nonnegative, and there is an orthonormal basis of eigenvectors of $L$. \item The smallest eigenvalue $\lambda_1$ of $L$ is equal to $0$, and $\mathbf{1}$ is a corresponding eigenvector. \end{enumerate} \medskip It turns out that the dimension of the nullspace of $L$ (the eigenspace of $0$) is equal to the number of connected components of the underlying graph of $G$. \begin{proposition} \label{Laplace2} Let $G = (V, W)$ be a ,weighted graph. The number $c$ of connected components $K_1, \ldots, K_c$ of the underlying graph of $G$ is equal to the dimension of the nullspace of $L$, which is equal to the multiplicity of the eigenvalue $0$. Furthermore, the nullspace of $L$ has a basis consisting of indicator vectors of the connected components of $G$, that is, vectors $(f_1, \ldots, f_m)$ such that $f_j = 1$ iff $v_j\in K_i$ and $f_j = 0$ otherwise. \end{proposition} \begin{proof} A complete proof can be found in von Luxburg \cite{Luxburg}, and we only give a sketch of the proof. \medskip First, assume that $G$ is connected, so $c = 1$. A nonzero vector $x$ is in the kernel of $L$ iff $L x = 0$, which implies that \[ \transpos{x} L x = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} (x_i - x_j)^2 = 0. \] This implies that $x_i = x_j$ whenever $w_{i\, j} > 0$, and thus, $x_i = x_j$ whenever nodes $v_i$ and $v_j$ are linked by an edge. By induction, $x_i = x_j$ whenever there is a path from $v_i$ to $v_j$. Since $G$ is assumed to be connected, any two nodes are linked by a path, which implies that $x_i = x_j$ for all $i\not= j$. Therefore, the nullspace of $L$ is spanned by $\mathbf{1}$, which is indeed the indicator vector of $K_1 = V$, and this nullspace has dimension $1$. \medskip Let us now assume that $G$ has $c \geq 2$ connected components. If so, by renumbering the rows and columns of $W$, we may assume that $W$ is a block matrix consisting of $c$ blocks, and similarly $L$ is a block matrix of the form \[ L = \begin{pmatrix} L_1 & & & \\ & L_2 & & \\ & & \ddots & \\ & & & L_c \end{pmatrix}, \] where $L_i$ is the graph Laplacian associated with the connected component $K_i$. By the induction hypothesis, $0$ is an eigenvalue of multiplicity $1$ for each $L_i$, and so the nullspace of $L$ has dimension $c$. The rest is left as an exercise (or see von Luxburg \cite{Luxburg}). \end{proof} \medskip Proposition \ref{Laplace2} implies that if the underlying graph of $G$ is connected, then the second eigenvalue, $\lambda_2$, of $L$ is strictly positive. \medskip Remarkably, the eigenvalue $\lambda_2$ contains a lot of information about the graph $G$ (assuming that $G = (V, E)$ is an undirected graph). This was first discovered by Fiedler in 1973, and for this reason, $\lambda_2$ is often referred to as the {\it Fiedler number\/}. For more on the properties of the Fiedler number, see Godsil and Royle \cite{Godsil} (Chapter 13) and Chung \cite{Chung}. More generally, the spectrum $(0, \lambda_2, \ldots, \lambda_m)$ of $L$ contains a lot of information about the combinatorial structure of the graph $G$. Leverage of this information is the object of {\it spectral graph theory\/}. \medskip It turns out that normalized variants of the graph Laplacian are needed, especially in applications to graph clustering. These variants make sense only if $G$ has no isolated vertices, which means that every row of $W$ contains some strictly positive entry. In this case, the degree matrix $\Degsym$ contains positive entries, so it is invertible and $\Degsym^{-1/2}$ makes sense; namely \[ \Degsym^{-1/2} = \mathrm{diag}(d_1^{-1/2}, \ldots, d_m^{-1/2}), \] and similarly for any real exponent $\alpha$. \begin{definition} \label{graphLaplacian2} Given any weighted directed graph $G = (V, W)$ with no isolated vertex and with $V = \{v_1, \ldots,v_m\}$, the {\it (normalized) graph Laplacians $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$ of $G$\/} are defined by \begin{align*} L_{\mathrm{sym}} & = \Degsym^{-1/2} L \Degsym^{-1/2} = I - \Degsym^{-1/2} W \Degsym^{-1/2} \\ L_{\mathrm{rw}} & = \Degsym^{-1} L = I - \Degsym^{-1} W. \end{align*} \end{definition} \medskip Observe that the Laplacian $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$ is a symmetric matrix (because $L$ and $\Degsym^{-1/2}$ are symmetric) and that \[ L_{\mathrm{rw}} = \Degsym^{-1/2} L_{\mathrm{sym}} \Degsym^{1/2}. \] The reason for the notation $L_{\mathrm{rw}}$ is that this matrix is closely related to a random walk on the graph $G$. There are simple relationships between the eigenvalues and the eigenvectors of $L_{\mathrm{sym}}$, and $L_{\mathrm{rw}}$. There is also a simple relationship with the generalized eigenvalue problem $Lx = \lambda \Degsym x$. \begin{proposition} \label{Laplace3} Let $G = (V, W)$ be a weighted graph without isolated vertices. The graph Laplacians, $L, L_{\mathrm{sym}}$, and $L_{\mathrm{rw}}$ satisfy the following properties: \begin{enumerate} \item[(1)] The matrix $ L_{\mathrm{sym}}$ is symmetric, positive, semidefinite. In fact, \[ \transpos{x} L_{\mathrm{sym}} x = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} \left(\frac{x_i}{\sqrt{d_i}} - \frac{x_j}{\sqrt{d_j}}\right)^2 \quad\mathrm{for\ all}\> x\in \mathbb{R}^m. \] \item[(2)] The normalized graph Laplacians $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$ have the same spectrum \\ $(0 = \nu_1 \leq \nu_2 \leq \ldots\leq \nu_m)$, and a vector $u\not= 0$ is an eigenvector of $L_{\mathrm{rw}}$ for $\lambda$ iff $\Degsym^{1/2} u$ is an eigenvector of $L_{\mathrm{sym}}$ for $\lambda$. \item[(3)] The graph Laplacians, $L, L_{\mathrm{sym}}$, and $L_{\mathrm{rw}}$ are symmetric, positive, semidefinite. \item[(4)] A vector $u\not = 0$ is a solution of the generalized eigenvalue problem $L u = \lambda \Degsym u$ iff $\Degsym^{1/2} u$ is an eigenvector of $L_{\mathrm{sym}}$ for the eigenvalue $\lambda$ iff $u$ is an eigenvector of $L_{\mathrm{rw}}$ for the eigenvalue $\lambda$. \item[(5)] The graph Laplacians, $L$ and $L_{\mathrm{rw}}$ have the same nullspace. \item[(6)] The vector $\mathbf{1}$ is in the nullspace of $L_{\mathrm{rw}}$, and $\Degsym^{1/2} \mathbf{1}$ is in the nullspace of $L_{\mathrm{sym}}$. \end{enumerate} \end{proposition} \begin{proof} (1) We have $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$, and $\Degsym^{-1/2}$ is a symmetric invertible matrix (since it is an invertible diagonal matrix). It is a well-known fact of linear algebra that if $B$ is an invertible matrix, then a matrix $S$ is symmetric, positive semidefinite iff $B S\transpos{B}$ is symmetric, positive semidefinite. Since $L$ is symmetric, positive semidefinite, so is $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$. The formula \[ \transpos{x} L_{\mathrm{sym}} x = \frac{1}{2}\sum_{i, j = 1}^m w_{i\, j} \left(\frac{x_i}{\sqrt{d_i}} - \frac{x_j}{\sqrt{d_j}}\right)^2 \quad\mathrm{for\ all}\> x\in \mathbb{R}^m \] follows immediately from Proposition \ref{Laplace1} by replacing $x$ by $\Degsym^{-1/2} x$, and also shows that $L_{\mathrm{sym}}$ is positive semidefinite. \medskip (2) Since \[ L_{\mathrm{rw}}= \Degsym^{-1/2} L_{\mathrm{sym}} \Degsym^{1/2}, \] the matrices $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$ are similar, which implies that they have the same spectrum. In fact, since $\Degsym^{1/2}$ is invertible, \[ L_{\mathrm{rw}} u = \Degsym^{-1} L u = \lambda u \] iff \[ \Degsym^{-1/2} L u = \lambda \Degsym^{1/2} u \] iff \[ \Degsym^{-1/2} L \Degsym^{-1/2} \Degsym^{1/2} u = L_{\mathrm{sym}} \Degsym^{1/2} u = \lambda \Degsym^{1/2} u, \] which shows that a vector $u\not= 0$ is an eigenvector of $L_{\mathrm{rw}}$ for $\lambda$ iff $\Degsym^{1/2} u$ is an eigenvector of $L_{\mathrm{sym}}$ for $\lambda$. \medskip (3) We already know that $L$ and $L_{\mathrm{sym}}$ are positive semidefinite, and (2) shows that $L_{\mathrm{rw}}$ is also positive semidefinite. \medskip (4) Since $\Degsym^{-1/2}$ is invertible, we have \[ Lu = \lambda \Degsym u \] iff \[ \Degsym^{-1/2} Lu = \lambda \Degsym^{1/2} u \] iff \[ \Degsym^{-1/2} L\Degsym^{-1/2} \Degsym^{1/2} u = L_{\mathrm{sym}} \Degsym^{1/2} u = \lambda \Degsym^{1/2} u, \] which shows that a vector $u\not = 0$ is a solution of the generalized eigenvalue problem $L u = \lambda \Degsym u$ iff $\Degsym^{1/2} u$ is an eigenvector of $L_{\mathrm{sym}}$ for the eigenvalue $\lambda$. The second part of the statement follows from (2). \medskip (5) Since $\Degsym^{-1}$ is invertible, we have $L u = 0$ iff $\Degsym^{-1}u = L_{\mathrm{rw}} u = 0$. \medskip (6) Since $L\mathbf{1} = 0$, we get $L_{\mathrm{rw}} u = \Degsym^{-1} L \mathbf{1} = 0$. That $\Degsym^{1/2} \mathbf{1}$ is in the nullspace of $L_{\mathrm{sym}}$ follows from (2). \end{proof} \medskip A version of Proposition \ref{Laplace4} also holds for the graph Laplacians $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$. The proof is left as an exercise. \begin{proposition} \label{Laplace4} Let $G = (V, W)$ be a weighted graph. The number $c$ of connected components $K_1, \ldots, K_c$ of the underlying graph of $G$ is equal to the dimension of the nullspace of both $L_{\mathrm{sym}}$ and $L_{\mathrm{rw}}$, which is equal to the multiplicity of the eigenvalue $0$. Furthermore, the nullspace of $L_{\mathrm{rw}}$ has a basis consisting of indicator vectors of the connected components of $G$, that is, vectors $(f_1, \ldots, f_m)$ such that $f_j = 1$ iff $v_j\in K_i$ and $f_j = 0$ otherwise. For $L_{\mathrm{sym}}$, a basis of the nullpace is obtained by multipying the above basis of the nullspace of $L_{\mathrm{rw}}$ by $\Degsym^{1/2}$. \end{proposition} \chapter{Spectral Graph Drawing} \label{chap2} \section{Graph Drawing and Energy Minimization} \label{ch2-sec1} Let $G = (V, E)$ be some undirected graph. It is often desirable to draw a graph, usually in the plane but possibly in 3D, and it turns out that the graph Laplacian can be used to design surprisingly good methods. Say $|V| = m$. The idea is to assign a point $\rho(v_i)$ in $\mathbb{R}^n$ to the vertex $v_i\in V$, for every $v_i \in V$, and to draw a line segment between the points $\rho(v_i)$ and $\rho(v_j)$. Thus, a {\it graph drawing\/} is a function $\mapdef{\rho}{V}{\mathbb{R}^n}$. \medskip We define the {\it matrix of a graph drawing $\rho$ (in $\mathbb{R}^n$)\/} as a $m \times n$ matrix $R$ whose $i$th row consists of the row vector $\rho(v_i)$ corresponding to the point representing $v_i$ in $\mathbb{R}^n$. Typically, we want $n < m$; in fact $n$ should be much smaller than $m$. A representation is {\it balanced\/} iff the sum of the entries of every column is zero, that is, \[ \transpos{\mathbf{1}} R = 0. \] If a representation is not balanced, it can be made balanced by a suitable translation. We may also assume that the columns of $R$ are linearly independent, since any basis of the column space also determines the drawing. Thus, from now on, we may assume that $n \leq m$. \medskip \bigskip\noindent{\bf Remark:}\enspace A graph drawing $\mapdef{\rho}{V}{\mathbb{R}^n}$ is not required to be injective, which may result in degenerate drawings where distinct vertices are drawn as the same point. For this reason, we prefer not to use the terminology {\it graph embedding\/}, which is often used in the literature. This is because in differential geometry, an embedding always refers to an injective map. The term {\it graph immersion\/} would be more appropriate. \medskip As explained in Godsil and Royle \cite{Godsil}, we can imagine building a physical model of $G$ by connecting adjacent vertices (in $\mathbb{R}^n$) by identical springs. Then, it is natural to consider a representation to be better if it requires the springs to be less extended. We can formalize this by defining the {\it energy\/} of a drawing $R$ by \[ \s{E}(R) = \sum_{\{v_i, v_j\}\in E} \norme{\rho(v_i) - \rho(v_j)}^2, \] where $\rho(v_i)$ is the $i$th row of $R$ and $\norme{\rho(v_i) - \rho(v_j)}^2$ is the square of the Euclidean length of the line segment joining $\rho(v_i)$ and $\rho(v_j)$. \medskip Then, ``good drawings'' are drawings that minimize the energy function $\s{E}$. Of course, the trivial representation corresponding to the zero matrix is optimum, so we need to impose extra constraints to rule out the trivial solution. \medskip We can consider the more general situation where the springs are not necessarily identical. This can be modeled by a symmetric weight (or stiffness) matrix $W = (w_{i j})$, with $w_{i j} \geq 0$. Then our energy function becomes \[ \s{E}(R) = \sum_{\{v_i, v_j\}\in E} w_{i j} \norme{\rho(v_i) - \rho(v_j)}^2. \] It turns out that this function can be expressed in terms of the matrix $R$ and a diagonal matrix $\widehat{W}$ obtained from $W$. Let $p = |E|$ be the number of edges in $E$ and pick any enumeration of these edges, so that every edge $\{v_i, v_j\}$ is uniquely represented by some index $e$. Then, let $\widehat{W}$ be the $p\times p$ diagonal matrix such that \[ \widehat{w}_{e e} = w_{i j},\quad \text{where $e$ correspond to the edge $\{v_i, v_j\}$}. \] We have the following proposition from Godsil and Royle \cite{Godsil}. \begin{proposition} \label{energyprop1} Let $G = (V, E)$ be an undirected graph, with $|V| = m$, $|E| = p$, let $W$ be a $m\times m$ weight matrix, and let $R$ be the matrix of a graph drawing $\rho$ of $G$ in $\mathbb{R}^n$ (a $m\times n$ matrix). If $\widetilde{D}$ is the incidence matrix associated with any orientation of the graph $G$, and $\widehat{W}$ is the $p\times p$ diagonal matrix associated with $W$, then \[ \s{E}(R) = \mathrm{tr}(\transpos{R}\widetilde{D} \widehat{W} \transpos{\widetilde{D}} R). \] \end{proposition} \begin{proof} Observe that the rows of $\transpos{\widetilde{D}} R$ are indexed by the edges of $G$, and if $\{v_i, v_j\}\in E$, then the $e$th row of $\transpos{\widetilde{D}} R$ is \[ \pm(\rho(v_i) - \rho(v_j)), \] where $e$ is the index corresponding to the edge $\{v_i, v_j\}$. As a consequence, the diagonal entries of $\transpos{\widetilde{D}} R \transpos{R} \widetilde{D}$ have the form $\norme{\rho(v_i) - \rho(v_j)}^2$, where $\{v_i, v_j\}$ ranges over the edges in $E$. Hence, \[ \s{E}(R) = \mathrm{tr}(\widehat{W} \transpos{\widetilde{D}} R \transpos{R} \widetilde{D}) = \mathrm{tr}(\transpos{R}\widetilde{D} \widehat{W} \transpos{\widetilde{D}} R), \] since $\mathrm{tr}(AB) = \mathrm{tr}(BA)$ for any two matrices $A$ and $B$. \end{proof} \medskip The matrix \[ L = \widetilde{D} \widehat{W} \transpos{\widetilde{D}} \] may be viewed as a weighted Laplacian of $G$. Observe that $L$ is a $m\times m$ matrix, and that \[ L_{i j} = \begin{cases} - w_{i j} & \text{if $i \not= j$} \\ \sum_{\{v_i, v_k\}\in E} w_{i k} & \text{if $i = j$}. \end{cases} \] Therefore, \[ L = \Degsym - W, \] the familiar unnormalized Laplacian matrix associated with $W$, where $\Degsym$ is the degree matrix associated with $W$, and so \[ \s{E}(R) = \mathrm{tr}(\transpos{R} L R). \] Note that \[ L \mathbf{1} = 0, \] as we already observed. \medskip Since the matrix $\transpos{R}\widetilde{D} \widehat{W} \transpos{\widetilde{D}} R = \transpos{R} L R $ is symmetric, it has real eigenvalues. Actually, since $L = \widetilde{D} \widehat{W} \transpos{\widetilde{D}}$ is positive semidefinite, so is $\transpos{R} L R$. Then, the trace of $\transpos{R} L R$ is equal to the sum of its positive eigenvalues, and this is the energy $\s{E}(R)$ of the graph drawing. \medskip If $R$ is the matrix of a graph drawing in $\mathbb{R}^n$, then for any invertible matrix $M$, the map that assigns $v_i$ to $\rho(v_i)M$ is another graph drawing of $G$, and these two drawings convey the same amount of information. From this point of view, a graph drawing is determined by the column space of $R$. Therefore, it is reasonable to assume that the columns of $R$ are pairwise orthogonal and that they have unit length. Such a matrix satisfies the equation $\transpos{R} R = I$, and the corresponding drawing is called an {\it orthogonal drawing\/}. This condition also rules out trivial drawings. The following result tells us how to find minimum energy graph drawings, provided the graph is connected. \begin{theorem} \label{graphdraw} Let $G = (V, W)$ be a weigted graph with $|V| = m$. If $L = \Degsym - W$ is the (unnormalized) Laplacian of $G$, and if the eigenvalues of $L$ are $0 = \lambda_1 < \lambda_2 \leq \lambda_3 \leq \ldots \leq \lambda_m$, then the minimal energy of any balanced orthogonal graph drawing of $G$ in $\mathbb{R}^n$ is equal to $\lambda_2 + \cdots + \lambda_{n + 1}$ (in particular, this implies that $n < m$). The $m \times n$ matrix $R$ consisting of any unit eigenvectors $u_2, \ldots, u_{n+1}$ associated with $\lambda_2 \leq \ldots \leq \lambda_{n + 1}$ yields an orthogonal graph drawing of minimal energy; it satisfies the condition $\transpos{R} R = I$. \end{theorem} \begin{proof} We present the proof given in Godsil and Royle \cite{Godsil} (Section 13.4, Theorem 13.4.1). The key point is that the sum of the $n$ smallest eigenvalues of $L$ is a lower bound for $\mathrm{tr}(\transpos{R} L R)$. This can be shown using an argument using the Rayleigh ratio; see Proposition \ref{interlace}. Then, any $n$ eigenvectors $(u_1, \ldots, u_n)$ associated with $\lambda_1, \ldots, \lambda_n$ achieve this bound. Because the first eigenvalue of $L$ is $\lambda_1 = 0$ and because we are assuming that $\lambda_2 > 0$, we have $u_1 = \mathbf{1}/\sqrt{m}$, and by deleting $u_1$ we obtain a balanced orthogonal graph drawing in $\mathbb{R}^{n - 1}$ with the same energy. The converse is true, so the minimum energy of an orthogonal graph drawing in $\mathbb{R}^n$ is equal to the minimum energy of an orthogonal graph drawing in $\mathbb{R}^{n+1}$, and this minimum is $\lambda_2 + \cdots + \lambda_{n + 1}$. The rest is clear. \end{proof} \medskip Observe that for any orthogonal $n\times n$ matrix $Q$, since \[ \mathrm{tr}(\transpos{R} L R) = \mathrm{tr}(\transpos{Q}\transpos{R} L R Q), \] the matrix $RQ$ also yields a minimum orthogonal graph drawing. \medskip In summary, if $\lambda_2 > 0$, an automatic method for drawing a graph in $\mathbb{R}^2$ is this: \begin{enumerate} \item Compute the two smallest nonzero eigenvalues $\lambda_2 \leq \lambda_3$ of the graph Laplacian $L$ (it is possible that $\lambda_3 = \lambda_2$ if $\lambda_2$ is a multiple eigenvalue); \item Compute two unit eigenvectors $u_2, u_3$ associated with $\lambda_2$ and $\lambda_3$, and let $R = [u_2\> u_3]$ be the $m\times 2$ matrix having $u_2$ and $u_3$ as columns. \item Place vertex $v_i$ at the point whose coordinates is the $i$th row of $R$, that is, $(R_{i 1}, R_{i 2})$. \end{enumerate} \medskip This method generally gives pleasing results, but beware that there is no guarantee that distinct nodes are assigned distinct images, because $R$ can have identical rows. This does not seem to happen often in practice. \section{Examples of Graph Drawings} \label{ch2-sec2} We now give a number of examples using {\tt Matlab}. Some of these are borrowed or adapted from Spielman \cite{Spielman}. \medskip\noindent {\it Example\/} 1. Consider the graph with four nodes whose adjacency matrix is \[ A = \begin{pmatrix} 0 & 1 & 1 & 0 \\ 1 & 0 & 0 & 1 \\ 1 & 0 & 0 & 1 \\ 0 & 1 & 1 & 0 \end{pmatrix}. \] We use the following program to compute $u_2$ and $u_3$: \begin{verbatim} A = [0 1 1 0; 1 0 0 1; 1 0 0 1; 0 1 1 0]; D = diag(sum(A)); L = D - A; [v, e] = eigs(L); gplot(A, v(:,[3 2])) hold on; gplot(A, v(:,[3 2]),'o') \end{verbatim} The graph of Example 1 is shown in Figure \ref{graph1}. The function {\tt eigs(L)} computes the six largest eigenvalues of $L$ in decreasing order, and corresponding eigenvectors. It turns out that $\lambda_2 = \lambda_3 = 2$ is a double eigenvalue. \begin{figure}[http] \begin{center} \includegraphics[height=2.2truein,width=2.4truein]{ncuts-figs/graph-fig1.pdf} \end{center} \caption{Drawing of the graph from Example 1.} \label{graph1} \end{figure} \medskip\noindent {\it Example\/} 2. Consider the graph $G_2$ shown in Figure \ref{graphfig5bis} given by the adjacency matrix \[ A = \begin{pmatrix} 0 & 1 & 1 & 0 & 0 \\ 1 & 0 & 1 & 1 & 1 \\ 1 & 1 & 0 & 1 & 0 \\ 0 & 1 & 1 & 0 & 1 \\ 0 & 1 & 0 & 1 & 0 \end{pmatrix}. \] We use the following program to compute $u_2$ and $u_3$: \begin{verbatim} A = [0 1 1 0 0; 1 0 1 1 1; 1 1 0 1 0; 0 1 1 0 1; 0 1 0 1 0]; D = diag(sum(A)); L = D - A; [v, e] = eig(L); gplot(A, v(:, [2 3])) hold on gplot(A, v(:, [2 3]),'o') \end{verbatim} The function {\tt eig(L)} (with no {\tt s} at the end) computes the eigenvalues of $L$ in increasing order. The result of drawing the graph is shown in Figure \ref{graph1b}. Note that node $v_2$ is assigned to the point $(0, 0)$, so the difference between this drawing and the drawing in Figure \ref{graphfig5bis} is that the drawing of Figure \ref{graph1b} is not convex. \begin{figure}[http] \begin{center} \includegraphics[height=2.2truein,width=2.4truein]{ncuts-figs/graph-fig2b.pdf} \end{center} \caption{Drawing of the graph from Example 2.} \label{graph1b} \end{figure} \medskip\noindent {\it Example\/} 3. Consider the ring graph defined by the adjacency matrix $A$ given in the {\tt Matlab} program shown below: \begin{verbatim} A = diag(ones(1, 11),1); A = A + A'; A(1, 12) = 1; A(12, 1) = 1; D = diag(sum(A)); L = D - A; [v, e] = eig(L); gplot(A, v(:, [2 3])) hold on gplot(A, v(:, [2 3]),'o') \end{verbatim} \begin{figure}[http] \begin{center} \includegraphics[height=2.2truein,width=2.4truein]{ncuts-figs/graph-fig2.pdf} \end{center} \caption{Drawing of the graph from Example 3.} \label{graph2} \end{figure} Observe that we get a very nice ring; see Figure \ref{graph2}. Again $\lambda_2 = 0.2679$ is a double eigenvalue (and so are the next pairs of eigenvalues, except the last, $\lambda_{12} = 4$). \medskip\noindent {\it Example\/} 4. In this example adpated from Spielman, we generate $20$ randomly chosen points in the unit square, compute their Delaunay triangulation, then the adjacency matrix of the corresponding graph, and finally draw the graph using the second and third eigenvalues of the Laplacian. \begin{verbatim} A = zeros(20,20); xy = rand(20, 2); trigs = delaunay(xy(:,1), xy(:,2)); elemtrig = ones(3) - eye(3); for i = 1:length(trigs), A(trigs(i,:),trigs(i,:)) = elemtrig; end A = double(A >0); gplot(A,xy) D = diag(sum(A)); L = D - A; [v, e] = eigs(L, 3, 'sm'); figure(2) gplot(A, v(:, [2 1])) hold on gplot(A, v(:, [2 1]),'o') \end{verbatim} The Delaunay triangulation of the set of $20$ points and the drawing of the corresponding graph are shown in Figure \ref{graph3}. The graph drawing on the right looks nicer than the graph on the left but is is no longer planar. \begin{figure}[http] \begin{center} \includegraphics[height=2.2truein,width=2.4truein]{ncuts-figs/delau-fig1.pdf} \hspace{1cm} \includegraphics[height=2.2truein,width=2.4truein]{ncuts-figs/graph-fig3.pdf} \end{center} \caption{Delaunay triangulation (left) and drawing of the graph from Example 4 (right).} \label{graph3} \end{figure} \medskip\noindent {\it Example\/} 5. Our last example, also borrowed from Spielman \cite{Spielman}, corresponds to the skeleton of the ``Buckyball,'', a geodesic dome invented by the architect Richard Buckminster Fuller (1895--1983). The Montr\'eal Biosph\`ere is an example of a geodesic dome designed by Buckminster Fuller. \begin{verbatim} A = full(bucky); D = diag(sum(A)); L = D - A; [v, e] = eig(L); gplot(A, v(:, [2 3])) hold on; gplot(A,v(:, [2 3]), 'o') \end{verbatim} Figure \ref{graph4} shows a graph drawing of the Buckyball. This picture seems a bit squashed for two reasons. First, it is really a $3$-dimensional graph; second, $\lambda_2 = 0.2434$ is a triple eigenvalue. (Actually, the Laplacian of $L$ has many multiple eigenvalues.) What we should really do is to plot this graph in $\mathbb{R}^3$ using three orthonormal eigenvectors associated with $\lambda_2$. \begin{figure}[http] \begin{center} \hspace{1cm} \includegraphics[height=2truein,width=2.2truein]{ncuts-figs/bucky-fig1.pdf} \end{center} \caption{Drawing of the graph of the Buckyball.} \label{graph4} \end{figure} \medskip A $3$D picture of the graph of the Buckyball is produced by the following {\tt Matlab} program, and its image is shown in Figure \ref{graph5}. It looks better! \begin{verbatim} [x, y] = gplot(A, v(:, [2 3])); [x, z] = gplot(A, v(:, [2 4])); plot3(x,y,z) \end{verbatim} \begin{figure}[http] \begin{center} \hspace{1cm} \includegraphics[height=2.7truein,width=4truein]{ncuts-figs/bucky3D-fig.pdf} \end{center} \caption{Drawing of the graph of the Buckyball in $\mathbb{R}^3$.} \label{graph5} \end{figure} \chapter{Graph Clustering} \label{chap3} \section{Graph Clustering Using Normalized Cuts} \label{ch3-sec1} Given a set of data, the goal of clustering is to partition the data into different groups according to their similarities. When the data is given in terms of a similarity graph $G$, where the weight $w_{i\, j}$ between two nodes $v_i$ and $v_j$ is a measure of similarity of $v_i$ and $v_j$, the problem can be stated as follows: Find a partition $(A_1, \ldots, A_K)$ of the set of nodes $V$ into different groups such that the edges between different groups have very low weight (which indicates that the points in different clusters are dissimilar), and the edges within a group have high weight (which indicates that points within the same cluster are similar). \medskip The above graph clustering problem can be formalized as an optimization problem, using the notion of cut mentioned at the end of Section \ref{ch1-sec1}. \medskip Given a subset $A$ of the set of vertices $V$, recall that we define $\mathrm{cut}(A)$ by \[ \mathrm{cut}(A) = \mathrm{links}(A, \overline{A})= \sum_{v_i\in A, v_j\in \overline{A}} w_{i\, j}, \] and that \[ \mathrm{cut}(A)= \mathrm{links}(A, \overline{A}) = \mathrm{links}(\overline{A}, A) = \mathrm{cut}(\overline{A}). \] If we want to partition $V$ into $K$ clusters, we can do so by finding a partition ($A_1, \ldots, A_K$) that minimizes the quantity \[ \mathrm{cut}(A_1, \ldots, A_K) = \frac{1}{2} \sum_{1 = 1}^K \mathrm{cut}(A_i). \] The reason for introducing the factor $1/2$ is to avoiding counting each edge twice. In particular, \[ \mathrm{cut}(A,\overline{A}) = \mathrm{links} (A, \overline{A}). \] For $K = 2$, the mincut problem is a classical problem that can be solved efficiently, but in practice, it does not yield satisfactory partitions. Indeed, in many cases, the mincut solution separates one vertex from the rest of the graph. What we need is to design our cost function in such a way that it keeps the subsets $A_i$ ``reasonably large'' (reasonably balanced). \medskip A example of a weighted graph and a partition of its nodes into two clusters is shown in Figure \ref{ncg-fig4}. \begin{figure}[http] \begin{center} \includegraphics[height=2.5truein,width=2.8truein]{ncuts-figs/ncuts-g-fig1.pdf} \hspace{0.5cm} \includegraphics[height=2.5truein,width=2.8truein]{ncuts-figs/ncuts-g-fig2.pdf} \end{center} \caption{A weighted graph and its partition into two clusters.} \label{ncg-fig4} \end{figure} \medskip A way to get around this problem is to normalize the cuts by dividing by some measure of each subset $A_i$. One possibility if to use the size (the number of elements) of $A_i$. Another is to use the volume $\mathrm{vol}(A_i)$ of $A_i$. A solution using the second measure (the volume) (for $K = 2$) was proposed and investigated in a seminal paper of Shi and Malik \cite{ShiMalik}. Subsequently, Yu (in her dissertation \cite{Yu}) and Yu and Shi \cite{YuShi2003} extended the method to $K > 2$ clusters. We will describe this method later. The idea is to minimize the cost function \[ \mathrm{Ncut}(A_1, \ldots, A_K) = \sum_{i = 1}^K \frac{\mathrm{links}(A_i, \overline{A_i})}{\mathrm{vol}(A_i)} = \sum_{i = 1}^K \frac{\mathrm{cut}(A_i, \overline{A_i})}{\mathrm{vol}(A_i)}. \] \medskip We begin with the case $K = 2$, which is easier to handle. \section{Special Case: $2$-Way Clustering Using Normalized Cuts} \label{ch3-sec2} Our goal is to express our optimization problem in matrix form. In the case of two clusters, a single vector $X$ can be used to describe the partition $(A_1, A_2) = (A, \overline{A})$. We need to choose the structure of this vector in such a way that $\mathrm{Ncut}(A, \overline{A})$ is equal to the Rayleigh ratio \[ \frac{\transpos{X} L X}{\transpos{X} \Degsym X}. \] It is also important to pick a vector representation which is invariant under multiplication by a nonzero scalar, because the Rayleigh ratio is scale-invariant, and it is crucial to take advantage of this fact to make the denominator go away. \medskip Let $N = |V|$ be the number of nodes in the graph $G$. In view of the desire for a scale-invariant representation, it is natural to assume that the vector $X$ is of the form \[ X = (x_1, \ldots, x_N), \] where $x_i \in \{a, b\}$ for $i = 1, \ldots, N$, for any two distinct real numbers $a, b$. This is an indicator vector in the sense that, for $i = 1, \ldots, N$, \[ x_i = \begin{cases} a & \text{if $v_i \in A$} \\ b & \text{if $v_i \notin A$} . \end{cases} \] The correct interpretation is really to view $X$ as a representative of a point in the real projective space $\mathbb{RP}^{N-1}$, namely the point $\mathbb{P}(X)$ of homogeneous coordinates $(x_1\colon \cdots \colon x_N)$. Therefore, from now on, we view $X$ as a vector of homogeneous coordinates representing the point $\mathbb{P}(X)\in \mathbb{RP}^{N-1}$. \medskip Let $d = \transpos{\mathbf{1}} \Degsym \mathbf{1}$ and $\alpha = \mathrm{vol}(A)$. Then, $\mathrm{vol}(\overline{A}) = d - \alpha$. By Proposition \ref{Laplace1}, we have \[ \transpos{X} L X = (a - b)^2\, \mathrm{cut}(A, \overline{A}), \] and we easily check that \[ \transpos{X} \Degsym X = \alpha a^2 + (d - \alpha)b^2. \] Since $\mathrm{cut}(A, \overline{A}) = \mathrm{cut}(\overline{A}, A)$, we have \[ \mathrm{Ncut}(A, \overline{A}) = \frac{\mathrm{cut}(A, \overline{A})}{\mathrm{vol}(A)} + \frac{\mathrm{cut}(\overline{A}, A)}{\mathrm{vol}(\overline{A})} = \left(\frac{1}{\mathrm{vol}(A)} + \frac{1}{\mathrm{vol}(\overline{A})} \right) \mathrm{cut}(A, \overline{A}), \] so we obtain \[ \mathrm{Ncut}(A, \overline{A}) = \left(\frac{1}{\alpha} + \frac{1}{d - \alpha} \right) \mathrm{cut}(A, \overline{A}) = \frac{d}{\alpha(d - \alpha)}\,\mathrm{cut}(A, \overline{A}). \] Since \[ \frac{\transpos{X} L X}{\transpos{X} \Degsym X} = \frac{(a - b)^2}{\alpha a^2 + (d - \alpha)b^2}\, \mathrm{cut}(A, \overline{A}), \] in order to have \[ \mathrm{Ncut}(A, \overline{A}) = \frac{\transpos{X} L X}{\transpos{X} \Degsym X}, \] we need to find $a$ and $b$ so that \[ \frac{(a - b)^2}{\alpha a^2 + (d - \alpha)b^2} = \frac{d}{\alpha(d - \alpha)}. \] The above is equivalent to \[ (a - b)^2 \alpha(d - \alpha) = \alpha d a^2 + (d - \alpha)d b^2, \] which can be rewritten as \[ a^2(\alpha d - \alpha(d - \alpha)) + b^2(d^2 - \alpha d - \alpha(d - \alpha)) + 2 \alpha(d - \alpha) a b = 0. \] The above yields \[ a^2 \alpha^2 + b^2(d^2 - 2\alpha d + \alpha^2) + 2\alpha(d - \alpha) a b = 0, \] that is, \[ a^2\alpha^2 + b^2(d - \alpha)^2 + 2\alpha(d - \alpha) a b = 0, \] which reduces to \[ (a \alpha + b(d - \alpha))^2 = 0. \] Therefore, we get the condition \begin{equation} a \alpha + b(d - \alpha) = 0. \tag{$\dagger$} \end{equation} Note that condition $(\dagger)$ applied to a vector $X$ whose components are $a$ or $b$ is equivalent to the fact that $X$ is orthogonal to $\Degsym \mathbf{1}$, since \[ \transpos{X} \Degsym \mathbf{1} = \alpha a+ (d - \alpha) b, \] where $\alpha = \mathrm{vol}(\{v_i\in V \mid x_i = a\})$. \medskip We claim the following two facts. For any nonzero vector $X$ whose components are $a$ or $b$, if $\transpos{X} \Degsym \mathbf{1} = \alpha a+ (d - \alpha) b = 0$, then \begin{enumerate} \item[(1)] $\alpha\not= 0$ and $\alpha \not= d$ iff $a\not= 0$ and $b\not= 0$. \item[(2)] if $a, b\not= 0$, then $a\not = b$. \end{enumerate} \medskip (1) First assume that $a \not= 0$ and $b \not= 0$. If $\alpha = 0$, then $\alpha a + (d - \alpha) b = 0$ yields $d b = 0$ with $d\not= 0$, which implies $b = 0$, a contradiction. If $d - \alpha = 0$, then we get $d a = 0$ with $d\not= 0$, which implies $a = 0$, a contradiction. \medskip Conversely, assume that $\alpha\not= 0$ and $\alpha \not= d$. If $a = 0$, then from $\alpha a+ (d - \alpha) b = 0$ we get $(d - \alpha) b = 0$, which implies $b = 0$, contradicting the fact that $X\not = 0$. Similarly, if $b = 0$, then we get $\alpha a = 0$, which implies $a = 0$, contradicting the fact that $X\not = 0$. \medskip (2) If $a, b\not= 0$, $a = b$ and $\alpha a+ (d - \alpha) b = 0$, then $\alpha a+ (d - \alpha) a = 0$, and since $a\not= 0$, we deduce that $d = 0$, a contradiction. \medskip If $\transpos{X} \Degsym \mathbf{1} = \alpha a+ (d - \alpha) b = 0$ and $a, b\not= 0$, then \[ b = -\frac{\alpha}{(d - \alpha)}\, a, \] so we get \begin{align*} \alpha a^2 + (d - \alpha)b^2& = \alpha \frac{(d - \alpha)^2}{\alpha^2} b^2+ (d - \alpha)b^2 \\ & = (d - \alpha)\left( \frac{d - \alpha}{\alpha} + 1 \right)b^2 = \frac{(d - \alpha) d b^2}{\alpha}, \end{align*} and \begin{align*} (a - b)^2 & = \left( -\frac{(d - \alpha)}{\alpha}\, b - b \right)^2 \\ & = \left( \frac{d - \alpha}{\alpha} + 1 \right)^2b^2 = \frac{d^2 b^2}{\alpha^2}. \end{align*} Since \begin{align*} \transpos{X} \Degsym X & = \alpha a^2 + (d - \alpha)b^2 \\ \transpos{X} L X & = (a - b)^2\, \mathrm{cut}(A, \overline{A}), \end{align*} we obtain \begin{align*} \transpos{X} \Degsym X & = \frac{(d - \alpha)d b^2}{\alpha} = \frac {\alpha d a^2}{(d - \alpha)} \\ \transpos{X} L X & = \frac{d^2 b^2}{\alpha^2} \, \mathrm{cut}(A, \overline{A}) = \frac{d^2 a^2}{(d - \alpha)^2} \, \mathrm{cut}(A, \overline{A}) . \end{align*} If we wish to make $\alpha$ disappear, we pick \[ a = \sqrt{\frac{d - \alpha}{\alpha}}, \quad b = -\sqrt{\frac{\alpha}{d - \alpha}}, \] and then \begin{align*} \transpos{X} \Degsym X & = d \\ \transpos{X} L X & = \frac{d^2}{\alpha(d - \alpha)} \, \mathrm{cut}(A, \overline{A}) = d\, \mathrm{Ncut}(A, \overline{A}). \end{align*} In this case, we are considering indicator vectors of the form \[ \left\{ (x_1, \ldots, x_N) \mid x_i \in \left\{ \sqrt{\frac{d - \alpha}{\alpha}}, -\sqrt{\frac{\alpha}{d - \alpha}} \right\}, \alpha = \mathrm{vol}(A) \right\}, \] for any nonempty proper subset $A$ of $V$. This is the choice adopted in von Luxburg \cite{Luxburg}. Shi and Malik \cite{ShiMalik} use \[ a = 1, \quad b = -\frac{\alpha}{d - \alpha} = - \frac{k}{1 - k}, \] with \[ k = \frac{\alpha}{d}. \] Another choice found in the literature (for example, in Belkin and Niyogi \cite{Belkin-Niyogi}) is \[ a = \frac{1}{\alpha}, \quad b = -\frac{1}{d - \alpha} . \] However, there is no need to restrict solutions to be of either of these forms. So, let \[ \s{X} = \big\{ (x_1, \ldots, x_N) \mid x_i \in \{a, b\}, \> a, b\in \mathbb{R},\> a, b\not = 0 \big\}, \] so that our solution set is \[ \s{K} = \big\{ X \in\s{X} \mid \transpos{X} \Degsym\mathbf{1} = 0 \big\}, \] because by previous observations, since vectors $X\in \s{X}$ have nonzero components, $\transpos{X}\Degsym\mathbf{1} = 0$ implies that $\alpha \not= 0$, $\alpha \not= d$, and $a\not= b$, where $\alpha = \mathrm{vol}(\{v_i\in V \mid x_i = a\})$. Actually, to be perfectly rigorous, we are looking for solutions in $\mathbb{RP}^{N-1}$, so our solution set is really \[ \mathbb{P}(\s{K}) = \big\{ (x_1\colon \cdots\colon x_N) \in \mathbb{RP}^{N-1}\mid (x_1, \ldots, x_N) \in \s{K} \big\}. \] Consequently, our minimization problem can be stated as follows: \medskip\noindent {\bf Problem PNC1} \begin{align*} & \mathrm{minimize} & & \frac{\transpos{X} L X}{\transpos{X} \Degsym X} & & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym\mathbf{1} = 0, & & X\in \s{X}. \end{align*} It is understood that the solutions are points $\mathbb{P}(X)$ in $\mathbb{RP}^{N-1}$. \medskip Since the Rayleigh ratio and the constraints $\transpos{X}\Degsym\mathbf{1} = 0$ and $X\in \s{X}$ are scale-invariant (for any $\lambda \not= 0$, the Rayleigh ratio does not change if $X$ is replaced by $\lambda X$, $X\in \s{X}$ iff $\lambda X \in \s{X}$, and $\transpos{ (\lambda X)}\Degsym\mathbf{1} = \lambda \transpos{ X}\Degsym\mathbf{1} = 0$), we are led to the following formulation of our problem: \medskip\noindent {\bf Problem PNC2} \begin{align*} & \mathrm{minimize} & & \transpos{X} L X & & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym X = 1, && \transpos{X} \Degsym\mathbf{1} = 0, && X\in \s{X}. \end{align*} \medskip Problem PNC2 is equivalent to problem PNC1 in the sense that if $X$ is any minimal solution of PNC1, then $X/(\transpos{X} \Degsym X)^{1/2}$ is a minimal solution of PNC2 (with the same minimal value for the objective functions), and if $X$ is a minimal solution of PNC2, then $\lambda X$ is a minimal solution for PNC1 for all $\lambda\not= 0$ (with the same minimal value for the objective functions). Equivalently, problems PNC1 and PNC2 have the same set of minimal solutions as points $\mathbb{P}(X) \in\mathbb{RP}^{N-1}$ given by their homogenous coordinates $X$. \medskip Unfortunately, this is an NP-complete problem, as shown by Shi and Malik \cite{ShiMalik}. As often with hard combinatorial problems, we can look for a {\it relaxation\/} of our problem, which means looking for an optimum in a larger continuous domain. After doing this, the problem is to find a discrete solution which is close to a continuous optimum of the relaxed problem. \medskip The natural relaxation of this problem is to allow $X$ to be any nonzero vector in $\mathbb{R}^N$, and we get the problem: \[ \mathrm{minimize} \quad \transpos{X} L X \quad\mathrm{subject\ to} \quad \transpos{X} \Degsym X = 1, \quad \transpos{X} \Degsym\mathbf{1} = 0. \] \medskip As usual, let $Y = \Degsym^{1/2} X$, so that $X = \Degsym^{-1/2} Y$. Then, the condition $\transpos{X}\Degsym X = 1$ becomes \[ \transpos{Y} Y = 1, \] the condition \[ \transpos{X} \Degsym \mathbf{1} = 0 \] becomes \[ \transpos{Y} \Degsym^{1/2} \mathbf{1} = 0, \] and \[ \transpos{X} L X = \transpos{Y} \Degsym^{-1/2} L \Degsym^{-1/2} Y. \] We obtain the problem: \[ \mathrm{minimize} \quad \transpos{Y} \Degsym^{-1/2} L \Degsym^{-1/2} Y \quad\mathrm{subject\ to} \quad \transpos{Y} Y = 1, \quad \transpos{Y} \Degsym^{1/2}\mathbf{1} = 0. \] \medskip Because $L\mathbf{1} = 0$, the vector $ \Degsym^{1/2} \mathbf{1}$ belongs to the nullspace of the symmetric Laplacian $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$. By Proposition \ref{PCAlem1}, minima are achieved by any unit eigenvector $Y$ of the second eigenvalue $\nu_2$ of $L_{\mathrm{sym}}$. Then, $Z = \Degsym^{-1/2} Y$ is a solution of our original relaxed problem. Note that because $Z$ is nonzero and orthogonal to $\Degsym \mathbf{1}$, a vector with positive entries, it must have negative and positive entries. \medskip The next question is to figure how close is $Z$ to an exact solution in $\s{X}$. Actually, because solutions are points in $\mathbb{RP}^{N-1}$, the correct statement of the question is: Find an exact solution $\mathbb{P}(X) \in\mathbb{P}(\s{X})$ which is the closest (in a suitable sense) to the approximate solution $\mathbb{P}(Z)\in \mathbb{RP}^{N-1}$. However, because $\s{X}$ is closed under the antipodal map, as explained in Appendix \ref{ch3-sec6}, minimizing the distance $d(\mathbb{P}(X), \mathbb{P}(Z))$ on $\mathbb{RP}^{N-1}$ is equivalent to minimizing the Euclidean distance $\norme{X - Z}_2$ (if we use the Riemannian metric on $\mathbb{RP}^{N-1}$ induced by the Euclidean metric on $\mathbb{R}^N$). \medskip We may assume $b < 0$, in which case $a > 0$. If all entries in $Z$ are nonzero, due to the projective nature of the solution set, it seems reasonable to say that the partition of $V$ is defined by the signs of the entries in $Z$. Thus, $A$ will consist of nodes those $v_i$ for which $x_i > 0$. Elements corresponding to zero entries can be assigned to either $A$ or $\overline{A}$, unless additional information is available. \medskip Now, using the fact that \[ b = - \frac{\alpha a}{d - \alpha}, \] a better solution is to look for a vector $X\in \mathbb{R}^N$ with $X_i \in \{a, b\}$ which is closest to a minimum $Z$ of the relaxed problem. Here is a proposal for an algorithm. \medskip For any solution $Z$ of the relaxed problem, let $I_{Z}^{+} = \{i \mid Z_i > 0\}$ be the set of indices of positive entries in $Z$, $I_{Z}^{-} = \{i \mid Z_i < 0\}$ the set of indices of negative entries in $Z$, $I_{Z}^{0} = \{i \mid Z_i = 0\}$ the set of indices of zero entries in $Z$, and let $Z^+$ and $Z^-$ be the vectors given by \begin{align*} Z^+_i & = \begin{cases} Z_i & \text{if $i \in I_Z^+$} \\ 0 & \text{if $i \notin I_Z^+$} \end{cases} & Z^-_i & = \begin{cases} Z_i & \text{if $i \in I_Z^-$} \\ 0 & \text{if $i \notin I_Z^-$} \end{cases}. \end{align*} Also let $n_a = |I_Z^{+} |$, $n_b = |I_Z^{-} |$, let $\overline{a}$ and $\overline{b}$ be the average of the positive and negative entries in $Z$ respectively, that is, \begin{align*} \overline{a} & = \frac{\sum_{i\in I_Z^+} Z_i}{n_a} & \overline{b} & = \frac{\sum_{i\in I_Z^-} Z_i}{n_b}, \end{align*} and let $\overline{Z^+}$ and $\overline{Z^-}$ be the vectors given by \begin{align*} (\overline{Z^+})_i & = \begin{cases} \overline{a} & \text{if $i \in I_Z^+$} \\ 0 & \text{if $i \notin I_Z^+$} \end{cases} & (\overline{Z^-})_i & = \begin{cases} \overline{b} & \text{if $i \in I_Z^-$} \\ 0 & \text{if $i \notin I_Z^-$} \end{cases}. \end{align*} If $\norme{\overline{Z^+} - Z^+} > \norme{\overline{Z^-} - Z^-}$, then replace $Z$ by $-Z$. Then, perform the following steps: \begin{enumerate} \item[(1)] Let \begin{align*} n_a & = |I_Z^{+} |, & \alpha & = \mathrm{vol}(\{v_i \mid i \in I_Z^{+}\}), & \beta & = \frac{\alpha}{d - \alpha} , \end{align*} and form the vector $\overline{X}$ with \[ \overline{X}_i = \begin{cases} a & \text{if $i \in I_Z^{+}$} \\ -\beta a & \text{otherwise}, \end{cases} \] such that $\norme{\overline{X} - Z}$ is minimal; the scalar $a$ is determined by finding the solution of the equation \[ \s{Z} a =Z, \] in the least squares sense, where \[ \s{Z} = \begin{cases} 1 & \text{if $i \in I_Z^{+}$} \\ -\beta & \text{otherwise}, \end{cases} \] and is given by \[ a = (\transpos{\s{Z}} \s{Z})^{-1}\transpos{\s{Z}}Z; \] that is, \[ a = \frac{\sum_{i \in I_Z^{+}} Z_i}{n_a + \beta^2(N - n_a)} - \frac{\sum_{i \in I_{Z}^{-}} \beta Z_i}{n_a + \beta^2(N - n_a)} . \] \item[(2)] While $I_Z^0 \not= \emptyset$, pick the smallest index $i\in I_Z^0$, compute \begin{align*} \widetilde{I}_{Z}^{+} & = I_{Z}^{+} \cup \{i\} \\ \widetilde{n}_a & = n_a + 1 \\ \widetilde{\alpha} & = \alpha + d(v_i) \\ \widetilde{\beta} & = \frac{\widetilde{\alpha}}{d - \widetilde{\alpha}}, \end{align*} and then $\widetilde{X}$ with \[ \widetilde{X}_j = \begin{cases} \widetilde{a} & \text{if $j \in \widetilde{I}_Z^{+}$} \\ -\widetilde{\beta} \widetilde{a} & \text{otherwise} , \end{cases} \] and \[ \widetilde{a} = \frac{\sum_{j \in \widetilde{I}_Z^{+}} Z_j}{\widetilde{n}_a + \widetilde{\beta}^2(N - \widetilde{n}_a)} - \frac{\sum_{j \in I_{Z}^{-}} \widetilde{\beta} Z_j}{\widetilde{n}_a + \widetilde{\beta}^2(N - \widetilde{n}_a)} . \] If $\smnorme{\widetilde{X} - Z} < \smnorme{\overline{X} - Z}$, then let $\overline{X} = \widetilde{X}$, $I_Z^+ = \widetilde{I}_Z^+$, $n_a = \widetilde{n}_a$, $\alpha = \widetilde{\alpha}$, and $I_Z^0 = I_Z^0 - \{i\}$; go back to (2). \item[(3)] The final answer if $\overline{X}$. \end{enumerate} \section{$K$-Way Clustering Using Normalized Cuts} \label{ch3-sec3} We now consider the general case in which $K \geq 3$. Two crucial issues need to be addressed (to the best of our knowledge, these points are not clearly articulated in the literature). \begin{enumerate} \item The choice of a matrix representation for partitions on the set of vertices. It is important that such a representation be scale-invariant. It is also necessary to state necessary and sufficient conditions for such matrices to represent a partition. \item The choice of a metric to compare solutions. It turns out that the space of discrete solutions can be viewed as a subset of the $K$-fold product $(\mathbb{RP}^{N - 1})^K$ of the projective space $\mathbb{RP}^{N - 1}$. Version 1 of the formulation of our minimization problem (PNC1) makes this point clear. However, the relaxation $(*_1)$ of version 2 of our minimization problem (PNC2), which is equivalent to version 1, reveals that that the solutions of the relaxed problem $(*_1)$ are members of the {\it Grassmannian\/} $G(K, N)$. Thus, we have two choices of metrics: (1) a metric on $(\mathbb{RP}^{N - 1})^K$; (2) a metric on $G(K, N)$. We discuss the first choice, which is the choice implicitly adopted by Shi and Yu. \end{enumerate} \medskip We describe a partition $(A_1, \ldots, A_K)$ of the set of nodes $V$ by an $N\times K$ matrix $X = [X^1 \cdots X^K]$ whose columns $X^1, \ldots, X^K$ are indicator vectors of the partition $(A_1, \ldots, A_K)$. Inspired by what we did in Section \ref{ch3-sec2}, we assume that the vector $X^j$ is of the form \[ X^j = (x_1^j, \ldots, x_N^j), \] where $x_i^j \in \{a_j, b_j\}$ for $j = 1, \ldots, K$ and $i = 1, \ldots, N$, and where $a_j, b_j$ are any two distinct real numbers. The vector $X^j$ is an indicator vector for $A_j$ in the sense that, for $i = 1, \ldots, N$, \[ x_i^j = \begin{cases} a_j & \text{if $v_i \in A_j$} \\ b_j & \text{if $v_i \notin A_j$} . \end{cases} \] When $\{a_j, b_j\} = \{0, 1\}$ for $j = 1, \ldots, K$, such a matrix is called a {\it partition matrix\/} by Yu and Shi. However, such a choice is premature, since it is better to have a scale-invariant representation to make the denominators of the Rayleigh ratios go away. \medskip Since the partition $(A_1, \ldots, A_K)$ consists of nonempty pairwise disjoint blocks whose union is $V$, some conditions on $X$ are required to reflect these properties, but we will worry about this later. \medskip Let $d = \transpos{\mathbf{1}} \Degsym \mathbf{1}$ and $\alpha_j = \mathrm{vol}(A_j)$, so that $\alpha_1 + \cdots + \alpha_K = d$. Then, $\mathrm{vol}(\overline{A_j}) = d - \alpha_j$, and as in Section \ref{ch3-sec2}, we have \begin{align*} \transpos{(X^j)} L X^j & = (a_j - b_j)^2\, \mathrm{cut}(A_j, \overline{A_j}), \\ \transpos{(X^j)} \Degsym X^j & = \alpha_j a_j^2 + (d - \alpha_j)b_j^2. \end{align*} When $K \geq 3$, unlike the case $K = 2$, in general we have $\mathrm{cut}(A_j, \overline{A_j}) \not= \mathrm{cut}(A_k, \overline{A_k}) $ if $j \not= k$, and since \[ \mathrm{Ncut}(A_1, \ldots, A_K) = \sum_{j = 1}^K \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)}, \] we would like to choose $a_j , b_j$ so that \[ \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)} = \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} \quad j = 1, \ldots, K, \] because this implies that \[ \mu(X) = \mathrm{Ncut}(A_1, \ldots, A_K) = \sum_{j = 1}^K \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)} = \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j}. \] Since \[ \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} = \frac{(a_j - b_j)^2\, \mathrm{cut}(A_j, \overline{A_j})} {\alpha_j a_j^2 + (d - \alpha_j)b_j^2} \] and $\mathrm{vol}(A_j) = \alpha_j$, in order to have \[ \frac{\mathrm{cut}(A_j, \overline{A_j})}{\mathrm{vol}(A_j)} = \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} \quad j = 1, \ldots, K, \] we need to have \[ \frac{(a_j - b_j)^2} {\alpha_j a_j^2 + (d - \alpha_j)b_j^2} = \frac{1}{\alpha_j} \quad j = 1, \ldots, K. \] Thus, we must have \[ (a_j^2 - 2a_jb_j + b_j^2)\alpha_j = \alpha_j a_j^2 + (d - \alpha_j)b_j^2, \] which yields \[ 2 \alpha_jb_j(b_j - a_j) = db_j^2. \] The above equation is trivially satisfied if $b_j = 0$. If $b_j\not= 0$, then \[ 2 \alpha_j(b_j - a_j) = db_j, \] which yields \[ a_j = \frac{2\alpha_j - d}{2\alpha_j} b_j. \] This choice seems more complicated that the choice $b_j = 0$, so we will opt for the choice $b_j = 0$, $j = 1, \ldots, K$. With this choice, we get \[ \transpos{(X^j)} \Degsym X^j = \alpha_j a_j^2. \] Thus, it makes sense to pick \[ a_j = \frac{1}{\sqrt{\alpha_j}} = \frac{1}{\sqrt{\mathrm{vol}(A_j)}}, \quad j = 1, \ldots, K, \] which is the solution presented in von Luxburg \cite{Luxburg}. This choice also corresponds to the scaled partition matrix used in Yu \cite{Yu} and Yu and Shi \cite{YuShi2003}. \medskip When $N = 10$ and $K = 4$, an example of a matrix $X$ representing the partition of $V = \{v_1, v_2, \ldots, v_{10}\}$ into the four blocks \[ \{A_1, A_2, A_3, A_4\} = \{\{v_2, v_4, v_6\}, \{v_1, v_5\}, \{v_3, v_8, v_{10}\}, \{v_7, v_9\}\} , \] is shown below: \[ X = \begin{pmatrix} 0 & a_2 & 0 & 0 \\ a_1 & 0 & 0 & 0 \\ 0 & 0 & a_3 & 0 \\ a_1 & 0 & 0 & 0 \\ 0 & a_2 & 0 & 0 \\ a_1 & 0 & 0 & 0 \\ 0 & 0 & 0 & a_4 \\ 0 & 0 & a_3 & 0 \\ 0 & 0 & 0 & a_4 \\ 0 & 0 & a_3 & 0 \end{pmatrix}. \] \medskip Let us now consider the problem of finding necessary and sufficient conditions for a matrix $X$ to represent a partition of $V$. \medskip When $b_j = 0$, the pairwise disjointness of the $A_i$ is captured by the orthogonality of the $X^i$: \begin{equation} \transpos{(X^i)}X^j = 0, \quad 1 \leq i, j \leq K, \> i \not= j. \tag{$*$} \end{equation} This is because, for any matrix $X$ where the nonzero entries in each column have the same sign, for any $i\not= j$, the condition \[ \transpos{(X^i)} X^j = 0 \] says that for every $k = 1,\ldots, N$, if $x^i_k \not= 0$ then $x^j_k = 0$. \medskip When we formulate our minimization problem in terms of Rayleigh ratios, conditions on the quantities $\transpos{(X^i)} \Degsym X^i$ show up, and it is more convenient to express the orthogonality conditions using the quantities $\transpos{(X^i)} \Degsym X^j$ instead of the $\transpos{(X^i)} X^j$, because these various conditions can be combined into a single condition involving the matrix $\transpos{X} \Degsym X$. Now, because $\Degsym$ is a diagonal matrix with positive entries and because the nonzero entries in each column of $X$ have the same sign, for any $i\not= j$, the condition \[ \transpos{(X^i)} X^j = 0 \] is equivalent to \begin{equation} \transpos{(X^i)} \Degsym X^j = 0, \tag{$**$} \end{equation} since, as above, it means that for $k = 1, \ldots, N$, if $x^i_k \not= 0$ then $x^j_k = 0$. Observe that the orthogonality conditions $(*)$ (and $(**)$) are equivalent to the fact that every row of $X$ has at most one nonzero entry. \medskip \bigskip\noindent{\bf Remark:}\enspace The disjointness condition \[ X \mathbf{1}_K= \mathbf{1}_N \] is used in Yu \cite{Yu}. However, this condition does guarantee the disjointness of the blocks. For example, it is satisfied by the matrix $X$ whose first column is $\mathbf{1}_N$, with $0$ everywhere else. \medskip Each $A_j$ is nonempty iff $X^j \not= 0$, and the fact that the union of the $A_j$ is $V$ is captured by the fact that each row of $X$ must have some nonzero entry (every vertex appears in some block). It is not obvious how to state conveniently this condition in matrix form. \medskip Observe that the diagonal entries of the matrix $X\transpos{X}$ are the square Euclidean norms of the rows of $X$. Therefore, we can assert that these entries are all nonzero. Let $\mathrm{DIAG}$ be the function which returns the diagonal matrix (containing the diagonal of $A$), \[ \mathrm{DIAG}(A) = \mathrm{diag}(a_{1\, 1}, \ldots, a_{n\, n}), \] for any square matrix $A = (a_{i \, j})$. Then, the condition for the rows of $X$ to be nonzero can be stated as \begin{equation*} \det(\mathrm{DIAG}(X\transpos{X})) \not= 0. \end{equation*} Observe that the matrix \[ \mathrm{DIAG}(X\transpos{X})^{-1/2} X \] is the result of normalizing the rows of $X$ so that they have Euclidean norm $1$. This normalization step is used by Yu \cite{Yu} in the search for a discrete solution closest to a solution of a relaxation of our original problem. For our special matrices representing partitions, normalizing the rows will have the effect of rescaling the columns (if row $i$ has $a_j$ in column $j$, then all nonzero entries in column $j$ are equal to $a_j$), but for a more general matrix, this is false. Since our solution matrices are invariant under rescaling the columns, but not the rows, rescaling the rows does not appear to be a good idea. \medskip A better idea which leads to a scale-invariant condition stems from the observation that since every row of any matrix $X$ representing a partition has a single nonzero entry $a_j$, we have \[ \transpos{X}\mathbf{1} = \begin{pmatrix} n_1a_1 \\ \vdots \\ n_K a_K \end{pmatrix}, \quad \transpos{X}X = \mathrm{diag}\left(n_1a_1^2, \ldots, n_K a_K^2\right), \] where $n_j$ is the number of elements in $A_j$, the $j$th block of the partition, which implies that \[ (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \begin{pmatrix} \frac{1}{a_1} \\ \vdots \\ \frac{1}{a_K} \end{pmatrix}, \] and thus, \begin{equation} X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}. \tag{$\dagger$} \end{equation} When $a_j = 1$ for $j = 1, \ldots, K$, we have $(\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}$, and condition $(\dagger)$ reduces to \[ X\mathbf{1}_K = \mathbf{1}_N. \] Note that because the columns of $X$ are linearly independent, $(\transpos{X} X)^{-1} \transpos{X}$ is the pseudo-inverse of $X$. Consequently, condition $(\dagger)$, can also be written as \[ XX^+ \mathbf{1} = \mathbf{1}, \] where $X^+ = (\transpos{X} X)^{-1} \transpos{X}$ is the pseudo-inverse of $X$. However, it is well known that $XX^+$ is the orthogonal projection of $\mathbb{R}^K$ onto the range of $X$ (see Gallier \cite{Gallbook2}, Section 14.1), so the condition $XX^+ \mathbf{1} = \mathbf{1}$ is equivalent to the fact that $\mathbf{1}$ belongs to the range of $X$. In retrospect, this should have been obvious since the columns of a solution $X$ satisfy the equation \[ a_1^{-1} X^1 + \cdots + a_K^{-1} X^K = \mathbf{1}. \] \medskip We emphasize that it is important to use conditions that are invariant under multiplication by a nonzero scalar, because the Rayleigh ratio is scale-invariant, and it is crucial to take advantage of this fact to make the denominators go away. \medskip If we let \[ \s{X} = \Big\{[X^1\> \ldots \> X^K] \mid X^j = a_j(x_1^j, \ldots, x_N^j) , \> x_i^j \in \{1, 0\}, a_j\in \mathbb{R}, \> X^j \not= 0 \Big\} \] (note that the condition $X^j \not= 0$ implies that $a_j \not= 0$), then the set of matrices representing partitions of $V$ into $K$ blocks is \begin{align*} & & &\s{K} = \Big\{ X = [X^1 \> \cdots \> X^K] \quad \mid & & X\in\s{X}, &&\\ & & & & & \transpos{(X^i)} \Degsym X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, && \quad\quad\quad\quad\quad\\ & & & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}\Big\}. && \end{align*} As in the case $K = 2$, to be rigorous, the solution are really $K$-tuples of points in $\mathbb{RP}^{N-1}$, so our solution set is really \[ \mathbb{P}(\s{K}) = \Big\{(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K)) \mid [X^1 \> \cdots \> X^K] \in \s{K} \Big\}. \] \medskip In view of the above, we have our first formulation of $K$-way clustering of a graph using normalized cuts, called problem PNC1 (the notation PNCX is used in Yu \cite{Yu}, Section 2.1): \medskip\noindent {\bf $K$-way Clustering of a graph using Normalized Cut, Version 1: \\ Problem PNC1} \begin{align*} & \mathrm{minimize} & & \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j}& & & &\\ & \mathrm{subject\ to} & & \transpos{(X^i)} \Degsym X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, & & & & \\ & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}, & & X\in \s{X}. & & \end{align*} As in the case $K = 2$, the solutions that we are seeking are $K$-tuples $(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K))$ of points in $\mathbb{RP}^{N-1}$ determined by their homogeneous coordinates $X^1, \ldots, X^K$. \medskip \bigskip\noindent{\bf Remark:}\enspace Because \[ \transpos{(X^j)} L X^j = \transpos{(X^j)} \Degsym X^j - \transpos{(X^j)} W X^j = \mathrm{vol}(A_j) - \transpos{(X^j)} W X^j, \] Instead of minimizing \[ \mu(X^1, \ldots, X^K) = \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j}, \] we can maximize \[ \epsilon(X^1, \ldots, X^K) = \sum_{j = 1}^K \frac{\transpos{(X^j)} W X^j}{\transpos{(X^j)}\Degsym X^j}, \] since \[ \epsilon(X^1, \ldots, X^K) = K - \mu(X^1, \ldots, X^K). \] This second option is the one chosen by Yu \cite{Yu} and Yu and Shi \cite{YuShi2003} (actually, they work with $\frac{1}{K}(K - \mu(X^1, \ldots, X^K))$, but this doesn't make any difference). \medskip Let us now show how our original formulation (PNC1) can be converted to a more convenient form, by chasing the denominators in the Rayleigh ratios, and by expressing the objective function in terms of the {\it trace\/} of a certain matrix. \medskip For any $N\times N$ matrix $A$, because \begin{align*} \transpos{X} A X & = \begin{bmatrix} \transpos{(X^1)} \\ \vdots \\ \transpos{(X^K)} \end{bmatrix} A [X^1 \cdots X^K] \\ & = \begin{pmatrix} \transpos{(X^1)} A X^1 & \transpos{(X^1)} A X^2 & \cdots & \transpos{(X^1)} A X^K \\ \transpos{(X^2)} A X^1 & \transpos{(X^2)} A X^2 & \cdots & \transpos{(X^2)} A X^K \\ \vdots & \vdots & \ddots & \vdots \\ \transpos{(X^K)} A X^1 & \transpos{(X^K)} A X^2 & \cdots & \transpos{(X^K)} A X^K \end{pmatrix}, \end{align*} we have \[ \mathrm{tr}(\transpos{X} A X) = \sum_{j = 1}^K \transpos{(X^j)} A X^j , \] and the conditions \[ \transpos{(X^i)} A X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, \] are equivalent to \[ \transpos{X} A X= \mathrm{diag}(\transpos{(X^1)} A X^1, \ldots, \transpos{(X^K)} A X^K). \] As a consequence, if we assume that \[ \transpos{(X^1)} A X^1 = \cdots = \transpos{(X^K)} A X^K = \alpha^2, \] then we have \[ \transpos{X} A X = \alpha^2 I, \] and if $R$ is any orthogonal $K\times K$ matrix, then by multiplying on the left by $\transpos{R}$ and on the right by $R$, we get \[ \transpos{R}\transpos{X} A XR = \transpos{R} \alpha^2 I R = \alpha^2 \transpos{R} R = \alpha^2 I. \] Therefore, if \[ \transpos{X} A X = \alpha^2 I, \] then \[ \transpos{(XR)} A (XR) = \alpha^2 I, \] for any orthogonal $K\times K$ matrix $R$. Furthermore, because $\mathrm{tr}(AB) = \mathrm{tr}(BA)$ for all matrices $A, B$, we have \[ \mathrm{tr}(\transpos{R}\transpos{X} A X R) = \mathrm{tr}(\transpos{X} A X). \] Since the Rayleigh ratios \[ \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} \] are invariant under rescaling by a nonzero number, we have \begin{align*} \mu(X) & = \mu(X^1, \ldots, X^K) = \sum_{j = 1}^K \frac{\transpos{(X^j)} L X^j}{\transpos{(X^j)}\Degsym X^j} \\ & = \mu((\transpos{(X^1)} \Degsym X^1)^{-1/2} X^1, \ldots, (\transpos{(X^K)} \Degsym X^K)^{-1/2} X^K) \\ & = \sum_{j = 1}^K (\transpos{(X^j)} \Degsym X^j)^{-1/2} \transpos{(X^j)} L \, (\transpos{(X^j)} \Degsym X^j)^{-1/2} X^j \\ & = \mathrm{tr}(\Lambda^{-1/2}\transpos{X} L X \Lambda^{-1/2}), \end{align*} where \[ \Lambda = \mathrm{diag}( \transpos{(X^1)} \Degsym X^1, \ldots, \transpos{(X^K)} \Degsym X^K). \] If $\transpos{(X^1)} \Degsym X^1 = \cdots = \transpos{(X^K)} \Degsym X^K = \alpha^2$, then $\Lambda = \alpha^2 I_K$, so $\Lambda$ commutes with any $K\times K$ matrix which implies that \[ \mathrm{tr}(\Lambda^{-1/2}\transpos{R}\transpos{X} L X R \Lambda^{-1/2}) = \mathrm{tr}(\transpos{R}\Lambda^{-1/2}\transpos{R}\transpos{X} L X \Lambda^{-1/2} R) = \mathrm{tr}(\Lambda^{-1/2}\transpos{X} L X \Lambda^{-1/2}), \] and thus, \[ \mu(X) = \mu(XR), \] for any orthogonal $K\times K$ matrix $R$. \medskip The condition \[ X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1} \] is also invariant if we replace $X$ by $XR$, where $R$ is any invertible matrix, because \begin{align*} XR (\transpos{(XR)} (XR))^{-1} \transpos{(XR)} \mathbf{1} & = XR (\transpos{R}\transpos{X} X R)^{-1} \transpos{R}\transpos{X} \mathbf{1}\\ & = XR R^{-1}(\transpos{X} X)^{-1}(\transpos{R})^{-1} \transpos{R}\transpos{X} \mathbf{1} \\ & = X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}. \end{align*} In summary we proved the following proposition: \begin{proposition} \label{Kway1} For any orthogonal $K\times K$ matrix $R$, any symmetric $N\times N$ matrix $A$, and any $N\times K$ matrix $X = [X^1 \> \cdots \> X^K]$, the following properties hold: \begin{enumerate} \item[(1)] $\mu(X) = \mathrm{tr}(\Lambda^{-1/2}\transpos{X} L X \Lambda^{-1/2})$, where \[ \Lambda = \mathrm{diag}( \transpos{(X^1)} \Degsym X^1, \ldots, \transpos{(X^K)} \Degsym X^K). \] \item[(2)] If $\transpos{(X^1)} \Degsym X^1 = \cdots = \transpos{(X^K)} \Degsym X^K = \alpha^2$, then $\mu(X) = \mu(XR)$. \item[(3)] The condition $\transpos{X} A X = \alpha^2 I$ is preserved if $X$ is replaced by $XR$. \item[(4)] The condition $X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}$ is preserved if $X$ is replaced by $XR$. \end{enumerate} \end{proposition} \medskip Now, by Proposition \ref{Kway1}(1) and the fact that the conditions in PNC1 are scale-invariant, we are led to the following formulation of our problem: \begin{align*} & \mathrm{minimize} & & \mathrm{tr}(\transpos{X} L X) & & & &\\ & \mathrm{subject\ to} & & \transpos{(X^i)} \Degsym X^j = 0, \quad 1\leq i, j \leq K,\> i\not= j, & & & & \\ & & & \transpos{(X^j)}\Degsym X^j = 1, \quad 1\leq j \leq K, & & & &\\ & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}, & & X\in \s{X}. & & \end{align*} Conditions on lines 2 and 3 can be combined in the equation \[ \transpos{X} \Degsym X = I, \] and, we obtain the following formulation of our minimization problem: \medskip\noindent {\bf $K$-way Clustering of a graph using Normalized Cut, Version 2: \\ Problem PNC2} \begin{align*} & \mathrm{minimize} & & \mathrm{tr}(\transpos{X} L X)& & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym X = I, & & & & \\ & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}, & & X\in \s{X}. & & \end{align*} \medskip Problem PNC2 is equivalent to problem PNC1 is the sense that for every minimal solution $(X^1, \ldots, X^K)$ of PNC1, $((\transpos{(X^1)} D X^1)^{-1/2} X^1, \ldots, (\transpos{(X^K)} D X^K)^{-1/2} X^K)$ is a minimal solution of PNC2 (with the same minimum for the objective functions), and that for every minimal solution $(Z^1, \ldots, Z^k)$ of PNC2, $(\lambda_1 Z^1, \ldots, \lambda_K Z^K)$ is a minimal solution of PNC1, for all $\lambda_i \not= 0$, $i = 1, \ldots, K$ (with the same minimum for the objective functions). In other words, problems PNC1 and PNC2 have the same set of minimal solutions as $K$-tuples of points $(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K))$ in $\mathbb{RP}^{N-1}$ determined by their homogeneous coordinates $X^1, \ldots, X^K$. \medskip Formulation PNC2 reveals that finding a minimum normalized cut has a geometric interpretation in terms of the graph drawings discussed in Section \ref{ch2-sec1}. Indeed, PNC2 has the following equivalent formulation: Find a minimal energy graph drawing $X$ in $\mathbb{R}^K$ of the weighted graph $G = (V, W)$ such that: \begin{enumerate} \item The matrix $X$ is orthogonal with respect to the inner product $\left\langle -, - \right\rangle_{\Degsym}$ in $\mathbb{R}^N$ induced by $\Degsym$, with \[ \left\langle x, y \right\rangle_{\Degsym} = \transpos{x} \Degsym y, \quad x, y \in \mathbb{R}^N. \] \item The rows of $X$ are nonzero; this means that no vertex $v_i\in V$ is assigned to the origin of $\mathbb{R}^K$ (the zero vector $0_K$). \item Every vertex $v_i$ is assigned a point of the form $(0, \ldots, 0,a_j,0,\ldots,0)$ on some axis (in $\mathbb{R}^K$). \item Every axis in $\mathbb{R}^K$ is assigned at least some vertex. \end{enumerate} \medskip Condition 1 can be reduced to the standard condition for graph drawings ($\transpos{R}R = I$) by making the change of variable $Y = \Degsym^{1/2} X$ or equivalently $X = \Degsym^{-1/2} Y$. Indeed, \[ \mathrm{tr}(\transpos{X} LX) = \mathrm{tr} (\transpos{Y}\Degsym^{-1/2} L \Degsym^{-1/2} Y), \] so we use the normalized Laplacian $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$ instead of $L$, \[ \transpos{X} \Degsym X = \transpos{Y} Y = I, \] and conditions (2), (3), (4) are preserved under the change of variable $Y = \Degsym^{1/2} X$, since $\Degsym^{1/2} $ is invertible. However, conditions (2), (3), (4) are ``hard'' constraints, especially condition (3). In fact, condition (3) implies that the columns of $X$ are orthogonal with respect to both the Euclidean inner product and the inner product $\left\langle -, - \right\rangle_{\Degsym}$, so condition (1) is redundant, except for the fact that it prescribes the norm of the columns, but this is not essential due to the projective nature of the solutions. \medskip The main problem in finding a good relaxation of problem PNC2 is that it is very difficult to enforce the condition $X\in \s{X}$. Also, the solutions $X$ are not preserved under arbitrary rotations, but only by very special rotations which leave $\s{X}$ invariant (they exchange the axes). The first natural relaxation of problem PNC2 is to drop the condition that $X\in \s{X}$, and we obtain the \medskip\noindent {\bf Problem $(*_1)$} \begin{align*} & \mathrm{minimize} & & \mathrm{tr}(\transpos{X} L X)& & & &\\ & \mathrm{subject\ to} & & \transpos{X} \Degsym X = I, & & & & \\ & & & X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}. & & & & \end{align*} \medskip By Proposition \ref{Kway1}, for every orthogonal matrix $R\in \mathbf{O}(K)$ and for every $X$ minimizing $(*_1)$, the matrix $XR$ also minimizes $(*_1)$. As a consequence, as explained below, we can view the solutions of problem $(*_1)$ as elements of the {\it Grassmannian\/} $G(N, K)$. \medskip Recall that the {\it Stiefel manifold\/} $St(k, n)$ consists of the set of orthogonal $k$-frames in $\mathbb{R}^n$, that is, the $k$-tuples of orthonormal vectors $(u_1, \ldots, u_k)$ with $u_i\in \mathbb{R}^n$. For $k = n$, the manifold $St(n, n)$ is identical to the orthogonal group $\mathbf{O}(n)$. For $1 \leq n \leq n - 1$, the group $\mathbf{SO}(n)$ acts transitively on $St(k, n)$, and $St(k, n)$ is isomorphic to the coset manifold $\mathbf{SO}(n)/\mathbf{SO}(n - k)$. The {\it Grassmann manifold\/} $G(k, n)$ consists of all (linear) $k$-dimensional subspaces of $\mathbb{R}^n$. Again, the group $\mathbf{SO}(n)$ acts transitively on $G(k, n)$, and $G(k, n)$ is isomorphic to the coset manifold $\mathbf{SO}(n)/S(\mathbf{SO}(k)\times\mathbf{SO}(n - k))$. The group $\mathbf{O}(k)$ acts on the right on the Stiefel manifold $St(k, n)$ (by multiplication), and the orbit manifold $St(k,n)/\mathbf{O}(k)$ is isomorphic to the Grassmann manifold $G(k, n)$. Furthermore, both $St(k, n)$ and $G(k, n)$ are {\it naturally reductive homogeneous manifolds\/} (for the Stiefel manifold, when $n \geq 3$), and $G(k, n)$ is even a {\it symmetric space\/} (see O'Neill \cite{Oneill}). The upshot of all this is that to a large extent, the differential geometry of these manifolds is completely determined by some subspace $\mfrac{m}$ of the Lie algebra $\mfrac{so}(n)$, such that we have a direct sum \[ \mfrac{so}(n) = \mfrac{m} \oplus \mfrac{h}, \] where $\mfrac{h} = \mfrac{so}(n - k)$ in the case of the Stiefel manifold, and $\mfrac{h} = \mfrac{so}(k) \times \mfrac{so}(n - k)$ in the case of the Grassmannian manifold (some additional condition on $\mfrac{m}$ is required). In particular, the geodesics in both manifolds can be determined quite explicitly, and thus we obtain closed form formulae for distances, {\it etc}. \medskip The Stiefel manifold $St(k, n)$ can be viewed as the set of all $n \times k$ matrices $X$ such that \[ \transpos{X} X = I_k. \] In our situation, we are considering $N\times K$ matrices $X$ such that \[ \transpos{X} D X = I. \] This is not quite the Stiefel manifold, but if we write $Y = D^{1/2} X$, then we have \[ \transpos{Y} Y = I, \] so the space of matrices $X$ satisfying the condition $\transpos{X} D X = I$ is the image $\s{D}(St(K, N))$ of the Stiefel manifold $St(K, N)$ under the linear map $\s{D}$ given by \[ \s{D}(X) = D^{1/2} X. \] Now, the right action of $\mathbf{O}(K)$ on $\s{D}(St(K, N))$ yields a coset manifold $\s{D}(St(K, N))/\mathbf{O}(K)$ which is obviously isomorphic to the Grassmann manidold $G(K, N)$. \medskip Therefore, {\it the solutions of problem $(*_1)$ can be viewed as elements of the {\it Grassmannian\/} $G(N, K)$\/}. We can take advantage of this fact to find a discrete solution of our original optimization problem PNC2 approximated by a continuous solution of $(*_1)$. \medskip Recall that condition $X (\transpos{X} X)^{-1} \transpos{X} \mathbf{1} = \mathbf{1}$ is equivalent to $XX^+ \mathbf{1} = \mathbf{1}$, which is also equivalent to the fact that $\mathbf{1}$ is in the range of $X$. If we make the change of variable $Y = \Degsym^{1/2} X$ or equivalently $X = \Degsym^{-1/2} Y$, the condition that $\mathbf{1}$ is in the range of $X$ becomes the condition that $\Degsym^{1/2} \mathbf{1}$ is in the range of $Y$, which is equivalent to \[ Y Y^+ \Degsym^{1/2} \mathbf{1} = \Degsym^{1/2} \mathbf{1}. \] However, since $\transpos{Y} Y = I$, we have \[ Y^+ = \transpos{Y}, \] so we get the equivalent problem \medskip\noindent {\bf Problem $(**_1)$} \begin{align*} & \mathrm{minimize} & & \mathrm{tr}(\transpos{Y}\Degsym^{-1/2} L \Degsym^{-1/2} Y)& & & &\\ & \mathrm{subject\ to} & & \transpos{Y} Y = I, & & & & \\ & & & Y \transpos{Y} \Degsym^{1/2} \mathbf{1} = \Degsym^{1/2} \mathbf{1}. & & & & \end{align*} \medskip This time, the matrices $Y$ satisfying condition $\transpos{Y} Y = I$ do belong to the Stiefel manifold $St(K, N)$, and again, {\it we view the solutions of problem $(**_1)$ as elements of the Grassmannian $G(K, N)$\/}. We pass from a solution $Y$ of problem $(**_1)$ in $G(K, N)$ to a solution $Z$ of of problem $(*_1)$ in $G(K, N)$ by the linear map $\s{D}^{-1}$; namely, $Z = \s{D}(Y) = D^{-1/2} Y$. \medskip The Rayleigh--Ritz Theorem (see Proposition \ref{PCAlem1}) tells us that if we temporarily ignore the second constraint, minima of problem $(**_1)$ are obtained by picking any $K$ unit eigenvectors $(u_1, \ldots, u_k)$ associated with the smallest eigenvalues \[ 0 = \nu_1\leq \nu_2 \leq \ldots \leq \nu_K \] of $L_{\mathrm{sym}} = \Degsym^{-1/2} L \Degsym^{-1/2}$. We may assume that $\nu_2 > 0$, namely that the underlying graph is connected (otherwise, we work with each connected component), in which case $Y^1 = \Degsym^{1/2}\mathbf{1}/\norme{\Degsym^{1/2}\mathbf{1}}_2$, because $\mathbf{1}$ is in the nullspace of $L$. Since $Y^1 = \Degsym^{1/2}\mathbf{1}/\norme{\Degsym^{1/2}\mathbf{1}}_2$, the vector $\Degsym^{1/2}\mathbf{1}$ is in the range of $Y$, so the condition \[ Y \transpos{Y} \Degsym^{1/2} \mathbf{1} = \Degsym^{1/2} \mathbf{1} \] is also satisfied. Then, $Z = \Degsym^{-1/2}Y$ with $Y = [u_1\> \ldots\> u_K]$ yields a minimum of our relaxed problem $(*_1)$ (the second constraint is satisfied because $\mathbf{1}$ is in the range of $Z$). \medskip By Proposition \ref{Laplace3}, the vectors $Z^j$ are eigenvectors of $L_{\mathrm{rw}}$ associated with the eigenvalues $0 = \nu_1 \leq \nu_2 \leq \ldots \leq \nu_K$. Recall that $\mathbf{1}$ is an eigenvector for the eigenvalue $\nu_1 = 0$, and $Z^1 = \mathbf{1}/\norme{\Degsym^{1/2}\mathbf{1}}_2$. Because, $\transpos{(Y^i)}Y^j = 0$ whenever $i \not= j$, we have \[ \transpos{(Z^i)}\Degsym Z^j = 0, \quad\text{whenever $i \not= j$}. \] This implies that $Z^2, \ldots, Z^K$ are all orthogonal to $\Degsym\mathbf{1}$, and thus, that each $Z^j$ has both some positive and some negative coordinate, for $j = 2, \ldots, K$. \medskip The conditions $\transpos{(Z^i)}\Degsym Z^j = 0$ do not necessarily imply that $Z^i$ and $Z^j$ are orthogonal (w.r.t. the Euclidean inner product), but we can obtain a solution of Problem $(*_1)$ achieving the same minimum for which distinct columns $Z^i$ an $Z^j$ are simultaneoulsy orthogonal and $\Degsym$-orthogonal, by multiplying $Z$ by some $K\times K$ orthogonal matrix $R$ on the right. Indeed, the $K\times K$ symmetric matrix $\transpos{Z}Z$ can be diagonalized by some orthogonal $K\times K$ matrix $R$ as \[ \transpos{Z}Z = R \Sigma \transpos{R}, \] where $\Sigma$ is a diagonal matrix, and thus, \[ \transpos{R}\transpos{Z} Z R = \transpos{(ZR)} ZR = \Sigma, \] which shows that the columns of $ZR$ are orthogonal. By Proposition \ref{Kway1}, $ZR$ also satisfies the constraints of $(*_1)$, and $\mathrm{tr}(\transpos{(ZR)} L (ZR)) = \mathrm{tr}(\transpos{Z} L Z)$. \medskip \bigskip\noindent{\bf Remark:}\enspace Since $Y$ has linearly independent columns (in fact, orthogonal) and since $Z = D^{-1/2} Y$, the matrix $Z$ also has linearly independent columns, so $\transpos{Z} Z$ is positive definite and the entries in $\Sigma$ are all positive. \medskip In summary, we should look for a solution $X$ that corresponds to an element of the Grassmannian $G(K, N)$, and hope that for some suitable orthogonal matrix $R$, the vectors in $XR$ are close to a true solution of the original problem. \section[$K$-Way Clustering; Using The Dependencies Among $X^1, \ldots, X^K$] {$K$-Way Clustering; Using The Dependencies \\ Among $X^1, \ldots, X^K$} \label{ch3-sec4} At this stage, it is interesting to reconsider the case $K = 2$ in the light of what we just did when $K \geq 3$. When $K = 2$, $X^1$ and $X^2$ are not independent, and it is convenient to assume that the nonzero entries in $X^1$ and $X^2$ are both equal to some positive real $c\in \mathbb{R}$, so that \[ X^1 + X^2 = c \mathbf{1}. \] To avoid subscripts, write $(A, \overline{A})$ for the partition of $V$ that we are seeking, and as before let $d = \transpos{\mathbf{1}} \Degsym \mathbf{1}$ and $\alpha = \mathrm{vol}(A)$. We know from Section \ref{ch3-sec2} that \begin{align*} \transpos{(X^1)} \Degsym X^1& = \alpha c^2 \\ \transpos{(X^2)} \Degsym X^2 &= (d - \alpha) c^2, \end{align*} so we normalize $X^1$ and $X^2$ so that $\transpos{(X^1)} \Degsym X^1 = \transpos{(X^2)} \Degsym X^2 = c^2$, and we consider \[ X = \left[\frac{X^1}{\sqrt{\alpha}} \> \frac{X^2}{\sqrt{d - \alpha}}\right]. \] Now, we claim that there is an orthogonal matrix $R$ so that if $X$ as above is a solution to our discrete problem, then $X R$ contains a multiple of $\mathbf{1}$ as a first column. A similar observation is made in Yu \cite{Yu} and Yu and Shi \cite{YuShi2003} (but beware that in these works $\alpha = \mathrm{vol}(A)/\sqrt{d}$). In fact, \[ R = \frac{1}{\sqrt{d}} \begin{pmatrix} \sqrt{\alpha} & \sqrt{d - \alpha} \\[6pt] \sqrt{d - \alpha} & - \sqrt{\alpha} \end{pmatrix}. \] Indeed, we have \begin{align*} X R & = \left[\frac{X^1}{\sqrt{\alpha}} \> \frac{c \mathbf{1} - X^1}{\sqrt{d - \alpha}}\right] R \\ & = \left[\frac{X^1}{\sqrt{\alpha}} \> \frac{c \mathbf{1} - X^1}{\sqrt{d - \alpha}}\right] \frac{1}{\sqrt{d}} \begin{pmatrix} \sqrt{\alpha} & \sqrt{d - \alpha} \\[6pt] \sqrt{d - \alpha} & - \sqrt{\alpha} \end{pmatrix} \\ & = \frac{1}{\sqrt{d}} \left[ c \mathbf{1} \>\> \sqrt{\frac{d - \alpha}{\alpha}}\, X^1 - \sqrt{\frac {\alpha}{d - \alpha}}\,(c \mathbf{1} - X^1) \right]. \end{align*} If we let \[ a = c\sqrt{\frac{d - \alpha}{\alpha}}, \quad b = - c\sqrt{\frac {\alpha}{d - \alpha}}, \] then we check that \[ \alpha a + b (d - \alpha) = 0, \] which shows that the vector \[ Z = \sqrt{\frac{d - \alpha}{\alpha}}\, X^1 - \sqrt{\frac {\alpha}{d - \alpha}}\,(c \mathbf{1} - X^1) \] is a potential solution of our discrete problem in the sense of Section \ref{ch3-sec2}. Furthermore, because $L \mathbf{1} = 0$, \[ \mathrm{tr}(\transpos{X} L X) = \mathrm{tr}(\transpos{(XR)} L (XR)) = \transpos{Z} L Z, \] the vector $Z$ is indeed a solution of our discrete problem. Thus, we reconfirm the fact that the second eigenvector of $L _{\mathrm{rw}} = \Degsym^{-1} L$ is indeed a continuous approximation to the clustering problem when $K = 2$. This can be generalized for any $K \geq 2$. \medskip Again, we may assume that the nonzero entries in $X^1, \ldots, X^K$ are some positive real $c\in \mathbb{R}$, so that \[ X^1 + \cdots + X^K = c \mathbf{1}, \] and if $(A_1, \ldots, A_K)$ is the partition of $V$ that we are seeking, write $\alpha_j = \mathrm{vol}(A_j)$. We have $\alpha_1 + \cdots + \alpha_K = d = \transpos{\mathbf{1}} \Degsym \mathbf{1}$. Since \[ \transpos{(X^j)} \Degsym X^j = \alpha_j c^2, \] we normalize the $X^j$ so that $\transpos{(X^j)} \Degsym X^j = \cdots = \transpos{(X^K)} \Degsym X^K = c^2$, and we consider \[ X = \left[\frac{X^1}{\sqrt{\alpha_1}} \> \frac{X^2}{\sqrt{\alpha_2}} \>\cdots \> \> \frac{X^K}{\sqrt{\alpha_K}}\right]. \] Then, we have the following result. \begin{proposition} \label{propdep1} If $X = \left[\frac{X^1}{\sqrt{\alpha_1}} \> \frac{X^2}{\sqrt{\alpha_2}} \>\cdots \> \> \frac{X^K}{\sqrt{\alpha_K}}\right]$ is a solution of our discrete problem, then there is an orthogonal matrix $R$ such that its first column $R^1$ is \[ R^1 = \frac{1}{\sqrt{d}} \begin{pmatrix} \sqrt{\alpha_1} \\ \sqrt{\alpha_2} \\ \vdots \\ \sqrt{\alpha_K} \end{pmatrix} \] and \[ XR = \left[\frac{c}{\sqrt{d}} \mathbf{1} \> Z^2 \>\cdots \> Z^{K}\right]. \] Furthermore, \[ \transpos{(XR)}\Degsym (XR) = c^2 I \] and \[ \mathrm{tr} (\transpos{(XR)} L (XR)) = \mathrm{tr}(\transpos{Z} L Z), \] with $Z = [Z^2 \> \cdots \> Z^{K}]$. \end{proposition} \begin{proof} Apply Gram--Schmidt to $(R^1, e_2, \ldots, e_K)$ (where $(e_1, \ldots, e_K)$ is the canonical basis of $\mathbb{R}^K$) to form an orthonormal basis. The rest follows from Proposition \ref{Kway1}. \end{proof} Proposition \ref{propdep1} suggests that if $Z = [\mathbf{1}\> Z^2 \>\cdots\> Z^K]$ is a solution of the relaxed problem $(*_1)$, then there should be an orthogonal matrix $R$ such that $Z\transpos{R}$ is an approximation of a solution of the discrete problem PNC1. \medskip The next step is to find an exact solution $(\mathbb{P}(X^1), \ldots, \mathbb{P}(X^K)) \in \mathbb{P}(\s{K})$ which is the closest (in a suitable sense) to our approximate solution $(Z^1, \ldots, Z^K)\in G(K, N)$. The set $\s{K}$ is not necessarily closed under all orthogonal transformations in $\mathbf{O}(K)$, so we can't view $\s{K}$ as a subset of the Grassmannian $G(K, N)$. However, we can think of $\s{K}$ as a subset of $G(K, N)$ by considering the subspace spanned by $(X^1, \ldots, X^K)$ for every $[X^1\> \cdots X^K\> ] \in \s{K}$. Then, we have two choices of distances. \begin{enumerate} \item We view $\s{K}$ as a subset of $(\mathbb{RP}^{N-1})^K$. Because $\s{K}$ is closed under the antipodal map, as explained in Appendix \ref{ch3-sec6}, minimizing the distance $d(\mathbb{P}(X^j), \mathbb{P}(Z^j))$ on $\mathbb{RP}^{N-1}$ is equivalent to minimizing the Euclidean distance $\norme{X^j - Z^j}_2$, for $j = 1, \ldots, K$ (if we use the Riemannian metric on $\mathbb{RP}^{N-1}$ induced by the Euclidean metric on $\mathbb{R}^N$). Then, minimizing the distance $d(X, Z)$ in $(\mathbb{RP}^{N-1})^K$ is equivalent to minimizing $\norme{X - Z}_F$, where \[ \norme{X - Z}_F^2 = \sum_{j = 1}^K \norme{X^j - Z^j}_2^2 \] is the Frobenius norm. This is implicitly the choice made by Yu. \item We view $\s{K}$ as a subset of the Grassmannian $G(K, N)$. In this case, we need to pick a metric on the Grassmannian, and we minimize the corresponding Riemannian distance $d(X, Z)$. A natural choice is the metric on $\mfrac{se}(n)$ given by \[ \left\langle X, Y\right\rangle = \mathrm{tr}(\transpos{X} Y). \] This choice remains to be explored, and will be the subject of a forthcoming report. \end{enumerate} \section[Discrete Solution Close to a Continuous Approximation] {Finding a Discrete Solution Close to a Continuous Approximation} \label{ch3-sec5} Inspired by Yu \cite{Yu} and the previous section, given a solution $Z_0$ of problem $(*_1)$, we look for pairs $(X, R) \in \s{K}\times \mathbf{O}(K)$ (where $R$ is a $K\times K$ orthogonal matrix), with $\norme{X^j} = \norme{Z^j_0}$ for $j = 1, \ldots, K$, that minimize \[ \varphi(X, R) = \norme{X - Z_0R}_F. \] Here, $\norme{A}_F$ is the Frobenius norm of $A$, with $\norme{A}_F^2 = \mathrm{tr}(\transpos{A} A)$. \medskip It may seem desirable to look for discrete solutions $X\in \s{K}$ whose entries are $0$ or $1$, in which case \[ X \mathbf{1}_K = \mathbf{1}_N. \] Therefore, we begin by finding a diagonal matrix $\Lambda = \mathrm{diag}(\lambda_1, \ldots, \lambda_K)$ such that \[ \norme{Z_0\Lambda\mathbf{1}_K - \mathbf{1}_N}_2 \] is minimal in the least-square sense. As we remarked earlier, since the columns of $Z_0$ are orthogonal with respect to the inner product $\left\langle u, v\right\rangle_{\Degsym} = \transpos{x}\Degsym y$, they are linearly independent, thus the pseudo-inverse of $Z_0$ is $(\transpos{Z_0} Z_0)^{-1} \transpos{Z_0}$, and the best solution $(\lambda_1, \ldots, \lambda_K)$ of least Euclidean norm is given by \[ (\transpos{Z_0} Z_0)^{-1} \transpos{Z_0}\mathbf{1}_N. \] Therefore, we form the (column-rescaled) matrix \[ Z = Z_0 \,\mathrm{diag}( (\transpos{Z_0} Z_0)^{-1} \transpos{Z_0}\mathbf{1}_N). \] \bigskip\noindent{\bf Remark:}\enspace In Yu \cite{Yu} and Yu and Shi \cite{YuShi2003}, the rows of $Z_0$ are normalized by forming the matrix \[ \mathrm{DIAG}(Z_0\transpos{Z_0})^{-1/2} Z_0. \] However, this does not yield a matrix whose columns are obtained from those of $Z_0$ by rescaling, so the resulting matrix is no longer a rescale of a correct solution of problem $(*_1)$, which seems undesirable. \medskip Actually, even though the columns of $Z$ are $\Degsym$-orthogonal, the matrix $Z$ generally does not satisfy the condition $\transpos{Z}\Degsym Z = I$, so $ZR$ may not have $\Degsym$-orthogonal columns (with $R \in \mathbf{O}(K)$), yet $\mathrm{tr}(\transpos{Z} L Z) = \mathrm{tr}(\transpos{(ZR)} L (ZR))$ holds! The problem is that the conditions $\transpos{Z}\Degsym Z = I$ and $Z \mathbf{1} = \mathbf{1}$ are antagonistic. If we try to force condition $Z \mathbf{1} = \mathbf{1}$, we modify the $\Degsym$-norm of the columns of $Z$, and then $ZR$ may no longer have $\Degsym$-orthogonal columns. Unless these methods are implemented and tested, it seems almost impossible to tell which option yields the best result. We will proceed under the assumption that $Z_0$ has been rescaled as explained above, but the method described next also applies if we pick $Z = Z_0$ . In this latter case, by Proposition \ref{Kway1}, $Z$ satisfies the condition $\transpos{Z}\Degsym Z = I$, and so does $ZR$. \medskip The key to minimizing $\norme{X - ZR}_F$ rests on the following computation: \begin{align*} \norme{X - ZR}_F^2 & = \mathrm{tr}(\transpos{(X - ZR)}(X - ZR)) \\ & = \mathrm{tr}((\transpos{X} -\transpos{R}\transpos{Z})(X - ZR)) \\ & = \mathrm{tr}(\transpos{X}X- \transpos{X}ZR - \transpos{R}\transpos{Z}X + \transpos{R}\transpos{Z}ZR)\\ & = \mathrm{tr}(\transpos{X}X) - \mathrm{tr}(\transpos{X}ZR) - \mathrm{tr}(\transpos{R}\transpos{Z}X) + \mathrm{tr}(\transpos{R}\transpos{Z}ZR)\\ & = \mathrm{tr}(\transpos{X}X) - \mathrm{tr}(\transpos{(\transpos{R}\transpos{Z}X)}) - \mathrm{tr}(\transpos{R}\transpos{Z}X) + \mathrm{tr}(\transpos{Z}ZR \transpos{R})\\ & = \mathrm{tr}(\transpos{X}X) - 2\mathrm{tr}(\transpos{R}\transpos{Z}X) + \mathrm{tr}(\transpos{Z}Z). \end{align*} Therefore, minimizing $\norme{X - ZR}_F^2$ is equivalent to maximizing $\mathrm{tr}(\transpos{R}\transpos{Z}X)$. This will be done by alternating steps during which we minimize $\varphi(X, R) = \norme{X - ZR}_F$ with respect to $X$ holding $R$ fixed, and steps during which we minimize $\varphi(X, R) = \norme{X - ZR}_F$ with respect to $R$ holding $X$ fixed. For this second step, we need the following proposition. \begin{proposition} \label{Kway2} For any $n \times n$ matrix $A$ and any orthogonal matrix $Q$, we have \[ \max\{\mathrm{tr}(QA) \mid Q\in \mathbf{O}(n)\} = \sigma_1 + \cdots + \sigma_n, \] where $\sigma_1 \geq \cdots \geq \sigma_n$ are the singular values of $A$. Furthermore, this maximum is achieved by $Q = V\transpos{U}$, where $A = U \Sigma \transpos{V}$ is any SVD for $A$. \end{proposition} \begin{proof} Let $A = U \Sigma \transpos{V}$ be any SVD for $A$. Then we have \begin{align*} \mathrm{tr}(QA) & = \mathrm{tr}(Q U \Sigma \transpos{V})\\ & = \mathrm{tr}(\transpos{V} Q U \Sigma). \end{align*} The matrix $Z = \transpos{V} Q U$ is an orthogonal matrix so $|z_{i j}| \leq 1$ for $1\leq i, j \leq n$, and $\Sigma$ is a diagonal matrix, so we have \[ \mathrm{tr}(Z\Sigma) = z_{1 1} \sigma_1 + \cdots + z_{n n } \sigma_n \leq \sigma_1 + \cdots + \sigma_n, \] which proves the first statement of the proposition. For $Q = V\transpos{U}$, we get \begin{align*} \mathrm{tr}(QA) & = \mathrm{tr}(Q U \Sigma \transpos{V}) \\ & = \mathrm{tr}(V\transpos{U} U \Sigma \transpos{V}) \\ & = \mathrm{tr}(V\Sigma \transpos{V}) = \sigma_1 + \cdots + \sigma_n, \end{align*} which proves the second part of the proposition. \end{proof} As a corollary of Proposition \ref{Kway2} (with $A = \transpos{Z} X$ and $Q = \transpos{R}$), we get the following result (see Golub and Van Loan \cite{Golub}, Section 12.4.1): \begin{proposition} \label{Kway3} For any two fixed $N\times K$ matrices $X$ and $Z$, the minimum of the set \[ \{\norme{X - ZR}_F \mid R\in \mathbf{O}(K)\} \] is achieved by $R = U \transpos{V}$, for any SVD decomposition $U \Sigma \transpos{V} = \transpos{Z} X$ of $\transpos{Z} X$. \end{proposition} \medskip We now deal with step 1. The solutions $Z$ of the relaxed problem $(*_1)$ have columns $Z^j$ of norm $\rho_j$. Then, for fixed $Z$ and $R$, we would like to find some $X\in \s{K}$ with $\norme{X^j} = \norme{Z^j} = \rho_j$ for $j = 1, \ldots, K$, so that $\norme{X - ZR}_F$ is minimal. Without loss of generality, we may assume that the entries $a_1, \ldots, a_K$ occurring in the matrix $X$ are positive. To find $X\in \s{K}$, first we find the shape $\widehat{X}$ of $X$, which is the matrix obtained from $X$ by rescaling the columns of $X$ so that $\widehat{X}$ has entries $+1, 0$. Then, we rescale the columns of $\widehat{X}$ so that $\norme{X^j} = \rho_j$ for $j = 1, \ldots, K$. \medskip Since \[ \norme{X - ZR}_F^2 = \norme{X}_F^2 + \norme{Z}_F^2 - \mathrm{tr}(\transpos{R}\transpos{Z}X) = 2\sum_{j = 1}^K\rho_j^2 - \mathrm{tr}(\transpos{R}\transpos{Z}X), \] minimizing $\norme{X - ZR}_F$ is equivalent to maximizing \[ \mathrm{tr}(\transpos{R}\transpos{Z}X) = \mathrm{tr}(\transpos{(ZR)}X) =\mathrm{tr}(X\transpos{(ZR)}), \] and since the $i$th row of $X$ contains a single nonzero entry, say $a_{j_i}$ (in column $j_i$, $1\leq j_i \leq K$), if we write $Y = ZR$, then \begin{equation} \mathrm{tr}(X \transpos{Y}) = \sum_{i = 1}^N a_{j_i} y_{i\, j_i }. \tag{$*$} \end{equation} By $(*)$, $\mathrm{tr}(X \transpos{Y})$ is maximized iff $a_{j_i}y_{i\, j_i }$ is maximized for $i = 1, \ldots, N$. Since the $a_k$ are positive, this is achieved if, for the $i$th row of $X$, we pick a column index $\ell$ such that $y_{i\, \ell}$ is maximum. \medskip Observe that if we change the $\rho_j$s, minimal solutions for these new values of the $\rho_j$ are obtained by rescaling the $a_\ell$'s. Thus, to find the shape $\widehat{X}$ of $X$, we may assume that $a_{\ell} = 1$. \medskip Actually, to find the shape $\widehat{X}$ of $X$, we first find a matrix $\overline{X}$ according to the following method. If we let \begin{align*} \mu_i & = \max_{1 \leq j \leq K} y_{i j} \\ J_i & = \{j \in \{1, \ldots, K\} \mid y_{i j} = \mu_i\}, \end{align*} for $i = 1, \ldots, N$, then \[ \overline{x}_{i j} = \begin{cases} +1 & \text{for some chosen $j \in J_i$,} \\ 0 & \text{otherwise}. \end{cases} \] Of course, a single column index is chosen for each row. Unfortunately, the matrix $\overline{X}$ may not be a correct solution, because the above prescription does not guarantee that every column of $\overline{X}$ is nonzero. Therefore, we may have to reassign certain nonzero entries in columns having ``many'' nonzero entries to zero columns, so that we get a matrix in $\s{K}$. When we do so, we set the nonzero entry in the column from which it is moved to zero. This new matrix is $\widehat{X}$, and finally we normalize each column of $\widehat{X}$ to obtain $X$, so that $\norme{X^j} = \rho_j$, for $j = 1, \ldots, K$. This last step may not be necessary since $Z$ was chosen so that $\norme{Z\mathbf{1}_K - \mathbf{1}_N}_2$ is miminal. A practical way to deal with zero columns in $\overline{X}$ is to simply decrease $K$. Clearly, further work is needed to justify the soundness of such a method. \medskip The above method is essentially the method described in Yu \cite{Yu} and Yu and Shi \cite{YuShi2003}, except that in these works (in which $X, Z$ and $Y$ are denoted by $X^*, \widetilde{X}^*$, and $\widetilde{X}$, respectively) the entries in $X$ belong to $\{0, 1\}$; as described above, for row $i$, the index $\ell$ corresponding to the entry $+1$ is given by \[ \arg \max_{1\leq j \leq K} \widetilde{X}(i, j). \] The fact that $\overline{X}$ may have zero columns is not addressed by Yu. Furthermore, it is important to make sure that each column of $X$ has the same norm as the corresponding column of $ZR$, but this normalization step is not performed in the above works. On the hand, the rows of $Z$ are normalized, but the resulting matrix may no longer be a correct solution of the relaxed problem. Only a comparison of tests obtained by implementating both methods will reveal which method works best in practice. \medskip The method due to Yu and Shi (see Yu \cite{Yu} and Yu and Shi \cite{YuShi2003}) to find $X\in \s{K}$ and $R\in \mathbf{O}(K)$ that minimize $\varphi(X, R) = \norme{X - ZR}_F$ is to alternate steps during which either $R$ is held fixed (step PODX) or $X$ is held fixed (step PODR). \begin{enumerate} \item[(1)] In step PODX, the next discrete solution $X^*$ is obtained fom the previous pair $(R^*, Z)$ by computing $\overline{X}$ and then $X^* = \widehat{X}$ from $Y = ZR^*$, as just explained above. \item[(2)] In step PODR, the next matrix $R^*$ is obtained from the previous pair $(X^*, Z)$ by \[ R^* = U \transpos{V}, \] for any SVD decomposition $U \Sigma \transpos{V}$ of $\transpos{Z} X^*$. \end{enumerate} \medskip It remains to initialize $R^*$ to start the process, and then steps (1) and (2) are iterated, starting with step (1). The method advocated by Yu \cite{Yu} is to pick $K$ rows of $Z$ that are as orthogonal to each other as possible. This corresponds to a $K$-means clustering strategy with $K$ nearly orthogonal data points as centers. Here is the algorithm given in Yu \cite{Yu}. \medskip Given the $N\times K$ matrix $Z$ (whose columns all have the same norm), we compute a matrix $R$ whose columns are certain rows of $Z$. We use a vector $c\in \mathbb{R}^N$ to keep track of the inner products of all rows of $Z$ with the columns $R^1, \ldots, R^{k-1}$ that have been constructed so far, and initially when $k = 1$, we set $c = 0$. \medskip The first column $R^1$ of $R$ is any chosen row of $Z$. \medskip Next, for $k = 2, \ldots, K$, we compute all the inner products of $R^{k-1}$ with all rows in $Z$, which are recorded in the vector $ZR^{k-1}$, and we update $c$ as follows: \[ c = c + \mathtt{abs}(ZR^{k - 1}). \] We take the absolute values of the entries in $ZR^{k - 1}$ so that the $i$th entry in $c$ is a score of how orthogonal is the $i$th row of $Z$ to $R^1, \ldots, R^{k-1}$. Then, we choose $R^k$ as any row $Z_i$ of $Z$ for which $c_i$ is minimal (the customary (and ambiguous) $i = \arg \min c$).
\section{Introduction} \label{sec:introduction} {\let\thefootnote\relax\footnote{R code for cMDS is available on github \href{https://github.com/ginagruenhage/cmdsr}{(here)}.}} \IEEEPARstart{T}{he} notion of distance is at the core of data analysis, pattern recognition and machine learning: most methods need to know how similar two datapoints are. The choice of distance metric is often a hidden assumption in algorithms. For complex data, distance or similarity are not uniquely defined. On the contrary, they can be arbitrary to some extent \cite{Carlsson2009}. It is, for example, often possible to describe signals on different temporal or spatial scales, and distance functions will give a certain scale more weight than another. Each datapoint might describe several features, and there is often no unique, optimal way to weigh the features when computing a distance measure: are two individuals more alike if they have similar eye colour or hair colour, or do we think the shape of the nose matters most? There are ways around that problem. One is to select the distance function that is best adapted to the task at hand, for example the one that gives the best performance in classification (this is effectively what is done in kernel hyperparameter selection \cite{Scholkopf2002}). Another is to give up on distance and rely instead on the weaker notion of neighbourhood \cite{Lum2013}. We argue here that a third option is available. One may study how the shape of the data evolves under a change in the distance metric by representing the data in lower dimension. We suppose that a family of distance functions $d_\alpha(x,y)$ is defined by varying a hyperparameter $\alpha \in [0,1]$. Suppose that for a given level of $\alpha$ the relative distances between datapoints are well described by representing the datapoints as points on the line. As we vary $\alpha$ the points will move, so that each point now describes a curve. Many scenarios are possible, and we sketch them in Fig.~\ref{fig:sketch}. We may have full or partial \emph{invariance}: patterns in the data that hold regardless of the value of the hyperparameter (Fig.~\ref{fig:sketch}A). On the other hand, the structure in the data may appear only for certain values of $\alpha$ (Fig.~\ref{fig:sketch}B and C), indicating that these values are more useful than others for characterizing the data. Analyzing the evolution of structures in the data might reveal interesting dependencies, for example, declustering (Fig.~\ref{fig:sketch}C) or loss of information (Fig.~\ref{fig:sketch}D). \begin{figure}[h] \centering \includegraphics[width=3.25in]{sketch_structures.pdf} \caption{Sketches of different effects on the data structure that emerge when varying a hyperparameter in a distance function. A) Invariance: patterns hold independent of the hyperparameter. B/C) Structure emerges only for certain values of the hyperparameter. C) Declustering: clusters are lost with increasing hyperparameter. D) Information loss: Structure collapses with increasing hyperparameter.} \label{fig:sketch} \end{figure} To visualize the effects of varying the distance function we suggest to embed data into a space of smooth curves, forming what we call continuous embeddings: in continuous embeddings each datapoint is embedded as a smooth curve in $\ensuremath{\mathbb{R}}^d$. We will show that this approach is quite general. Our implementation of continuous embeddings is based on mulit-dimensional scaling (MDS), one of the most widely-used tools for dimensionality reduction \cite{Buja2002,Buja2008}. MDS builds on the pairwise relation between single data points and has an intuitive way of characterizing the structure in high-dimensional data. MDS supposes that one has distance information available, that is, we can characterize the data by a distance matrix. MDS seeks to find a set of points in a low dimensional Euclidean space, such that the Euclidean distances between points approximate the original distances. MDS goes back to the 1950s, when it was first introduced as classical scaling \cite{Torgerson1952}. In classical scaling, the distance matrix is transformed to a matrix of inner products from which an embedding can be computed using eigendecompositions \cite{Torgerson1952,Torgerson1958,Gower1966}. Classical scaling finds a perfect embedding when the data can indeed be embedded exactly, but in all realistic cases distance matrices are not exactly Euclidean and \emph{distance} scaling is more appropriate. \cite{Kruskal1964} introduced distance scaling by defining a cost function, \emph{Stress}, that directly measures the error between original and embedding distances. This cost function is then optimized over the space of embedding matrices which can be done using gradient descent. Since the early work on MDS many other variants and optimization solutions have been discussed. So called non-metric variants of MDS seek to only recover the ranks of distances \cite{Shepard1962}. \cite{Ramsay1977,Ramsay1978} introduce a statistical model for MDS, allowing for a maximum likelihood estimate. This approach is implemented in {\em Multiscale} \cite{Ramsay1978b}. Other MDS variants based on Stress include Sammon's mapping \cite{Sammon1969}, elastic stress \cite{McGee1966}, multidimensional unfolding \cite{Borg2005} and local MDS \cite{Chen2009}. Isomap \cite{Tenenbaum2000} is also related to MDS. Here, distances are computed as geodesic distances on a manifold, which are then embedded with classical scaling. In terms of optimization one of the most popular approaches is SMACOF, a majorization method for MDS \cite{Guttman1968,DeLeeuw1977,DeLeeuw1977b,DeLeeuw1988}. Here, we introduce a continuous version of MDS (cMDS) by adding a smoothing penalty to the MDS cost function. Similar ideas have been used in the visualization of dynamic networks. A network is commonly represented as a graph. A 2D embedding of a static graph is often constructed using MDS or similar methods \cite{Kamada1989,Gansner2005}. In the dynamical context, where a graph is measured over time, it is important to preserve the so-called ``mental map'' when jumping from one timepoint to the next \cite{Misue1995}. Early work on such {\em controlled stability} was done by \cite{Boehringer1990,North1996}. \cite{Brandes1997} developed a more rigorous formulation of controlled stability based on regularization in a bayesian framework. There have been three different approaches to the problem of preservation of the mental map: aggregation, anchoring and linking. In aggregation methods, the graph is aggregated into an average graph which is then visualized with a static layout algorithm \cite{Brandes2003,Moody2005}. Anchoring methods use auxiliary edges which connect nodes to stationary reference positions \cite{Brandes1997,Diehl2002,Frishman2008}. In linking, edges are created that connect instances of a single vertex over time. The resulting graph is then visualized using standard methods \cite{Erten2004,Erten2004b,Dwyer2006}. Linking has been formulated in more rigorous ways in terms of regularized cost functions \cite{Baur2008,Brandes2012a}. \cite{Xu2013} introduce an additional grouping penalty. \cite{Brandes2012a} provide a good overview on dynamic graph layout. Another approach to dynamic embeddings is an extension of the Hoff latent space model \cite{Sarkar2005}. All these approaches and contributions in the field visualize temporal developments. We show here that continuous embeddings can be applied to a great variety of data, going far beyond the visualization of temporal dynamics in graphs. In particular, the continuous variable can be used to visualize artificial dynamics. This makes the method very general and is especially useful in the analysis of families of distance functions and their effects on data structure. Continuous embeddings are thus a tool for making an informed choice of the distance metric for use in further analyses. We show how continuous embeddings can be efficiently computed using the Concave-Convex Procedure (CCCP) for optimization \cite{Yuille2003}. The resulting algorithm is a simple iterative procedure, in which the inner loops are nothing more than least squares regression with smoothing splines. We prove that the algorithm always leads to a stationary point. This goes further than other proofs \cite{Sriperumbudur2009,Yen2012} because the cost function is nondifferentiable at certain points and doesn't share the directional derivative with the upper bound at those points. We illustrate the results of cMDS with several examples. We compare cDMS to a method based on $k$-means to exemplify that cMDS provides a more informative way of understanding the data structure. We show that cMDS leads to novel forms of data visualization and enhances the analysis of various meta-effects in data, such as hierarchy levels in hierarchical clustering, weighting of different distance measures and consensus requirements across subjects. Furthermore, we point out that quantitative analyses on cMDS results are possible and useful. We show that cMDS is especially well-suited to dynamic and interactive contexts \cite{Cook2007}. We provide several examples of interactive, web-based visualizations based on cMDS. \section{Methods} In order to present the cMDS algorithm, we introduce some notations and definitions. We describe the objective of the algorithm and present the cost function that we need to optimize. Optimization can be done in a coordinate-wise manner and we present the optimization of single coordinates via the Concave-Convex Procedure and pseudocode of the full algorithm in Section \ref{sec:findConfig}. In the subsequent section, we proof that the optimization of single coordinates is first-order optimal, i.e. it leads to stationary points of the original cost. We also show that this directly entails first order optimality of the full cost. We start by setting the notations. The original data or objects are denoted by $\mathbf{s}_1(\alpha) \,\ldots\, \mathbf{s}_N(\alpha)$ and are defined in an arbitrary metric space, e.g. $\ensuremath{\mathbb{R}}^D$. The parameter $\alpha$ measures a continuous dimension. We will see that objects, such as images or networks, can be endowed with a continuous dimension, for example by examining them at different scales, so that scale plays the role of the continuous parameter. At this point we would like to note that we discretize all equations from the beginning, because ultimately, in the implementation, the hyperparameter has to be discretized. It would certainly be possible to develop the mathematics in a continuous matter. We prefer to present the mathematics in such a way that the equations in the paper can be used as a direct reference for implementation. Thus, $\alpha$ is represented on a grid $\alpha_1 \,\ldots\, \alpha_T$. With $f(\cdot)$ we refer to function values at all grid points while $f(k)$ refers to a function value at a specific time. The objects are endowed with a distance function $d\left(x,y\right)$. The appropriate distance function depends on the data-space and on the nature of the problem. Given a distance measure, we can define a distance array, $\bDt{(\cdot)} \in \ensuremath{\mathbb{R}}^{T\times N\times N}$, where the entry $d_{ij}^k$ holds the distance between objects $\mathbf{s}_i$ and $\mathbf{s}_j$ at $\alpha_k$. We assume that the distances between datapoints give a good summary of the patterns in the data. The goal of cMDS will be to extract these patterns. \subsection{Objective} The objective of cMDS is to retrieve curves or manifolds in $\ensuremath{\mathbb{R}}^d$, $d<<D$ which we denote as $\bxt{(\cdot)}_1 \,\ldots\, \bxt{(\cdot)}_N$, such that the evolution of distances between the curves represents the evolution of distances between the datapoints. The curves are represented as a configuration array $\bXt{(\cdot)} \in \ensuremath{\mathbb{R}}^{T\times d\times N}$ and for a given time-point $\alpha_k$ $\bXt{k}$ is a $d \times N$ matrix in which each column represents the coordinates of a curve at time $\alpha_k$. In $\ensuremath{\mathbb{R}}^d$ we measure the distance between two curves at time $\alpha_k$ with the Euclidean distance. Thus, for each configuration, we have a distance array $\bDXt{(\cdot)} \in \ensuremath{\mathbb{R}}^{T\times N\times N}$ such that $\tilde{d}_{ij}^k = \norm{ \bxt{k}_i - \bxt{k}_j}_2$. These are the approximate distances given by our embedding, and the objective is to make the approximate distances as close as possible to the real distances. A natural expression of that objective is the following cost function, which quantifies the distortion of the embedding: \begin{equation}\label{eq:distortion} \mathcal{L}\left(\bDXt{(\cdot)},\bDt{(\cdot)} \right) = \sum_{k=1}^T \left(\norm{\bDXt{k}-\bDt{k}}_F\right)^2 \end{equation} where $\norm{\cdot}_F$ denotes the Frobenius norm. This is the MDS cost function. In practice and for real datasets, MDS is a highly non-convex optimization problem with multiple local minima. Additionally, since distances are invariant to rotations, translations and symmetries, so are the MDS embeddings. Minimizing (\ref{eq:distortion}) is equivalent to solving MDS problems for different values of the hyperparameter individually. The individual problems are known as \emph{Kruskal-Shephard} scaling \cite{Kruskal1964}. This would result in solutions that are independent for different $\alpha_k$ and thus might lie in quite different local minima. However, one would expect that slight changes in the hyperparameter lead to only slight changes in the embedding. To solve this problem, we will require a priori that each curve is continuous and smooth, which can be achieved by adding a suitable penalty function $\Omega(\bXt{(\cdot)})$ to the cost. The effect of the smoothing penalty can be interpreted as the goal of tracing one particular local minimum across different values of the hyperparameter. This results in the cost function \begin{equation}\label{eq:cMDS-objective} \mathcal{C}\left(\bXt{(\cdot)},\bDt{(\cdot)}\right) = {\mathcal{L}\left(\bXt{(\cdot)},\bDt{(\cdot)}\right) +\lambda \, \Omega\left(\bXt{(\cdot)}\right) } \end{equation} Classical spline penalties \cite{Ramsay2002} are particularly convenient. In practice we work with discrete one-dimensional curves for which the penalty reads \begin{IEEEeqnarray}{rCl} \Omega\left(\bXt{(\cdot)}\right) & = & \sum_i (\bxt{(\cdot)}_i)^T\, (\mathcal{D}^{(2)})^T \mathcal{D}^{(2)}\, \bxt{(\cdot)}_i \nonumber\\ & = & \sum_i (\bxt{(\cdot)}_i)^T\, \mathbf{M} \, \bxt{(\cdot)}_i \end{IEEEeqnarray} where $\mathcal{D}^{(2)}$ denotes the discrete second order differential operator. The parameter $\lambda$ controls how strongly the roughness of the curves is penalized. For $\lambda = 0$, we recover the classical cost function of MDS, with the extension that we have $T$ seperate MDS problems, one for each time step. For $\lambda \rightarrow \infty $ the resulting curves are straight lines, independent of the original data. The value of $\lambda$ is easy to set by visual inspection. It should simply be large enough to result in fairly smooth curves, while keeping the distortion reasonably small. A reasonable strategy for setting $\lambda$ automatically is then to maximize $\lambda$, under a constraint on the quality of the embedding. This quality can be measured in various ways \cite{Kaski2011,Mokbel2013}. \subsection{Optimization via the Concave-Convex Procedure (CCCP)} \label{sec:findConfig} The cMDS algorithm optimizes the cost in a curve-by-curve manner. This is possible since the cost is a sum over costs per curve. Thus, we have an outer loop over curves and an inner loop that performs conditional optimization on $\bxt{(\cdot)}_i$. That is, we assume that all curves $\bxt{(\cdot)}_j, j\ne i$ are fixed. The inner loop is a Maximization Minimization procedure. Specifically, we use the Concave-Convex Procedure (CCCP) \cite{Yuille2003}. This way, each minimization step is a simple spline regression. In effect, the algorithm only needs to compute spline regressions with surrogate data. We now outline the usage of CCCP for the optimization of a single curve. We expand the first term of the cost function to identify the convex and concave parts. We denote the cost of a single curve as $f(\cdot)$. \begin{IEEEeqnarray}{rCl} f\bigl(\bxt{(\cdot)}_i\bigr) & = & \sum_{kj} \left( \left\| \bxt{k}_i - \bxt{k}_j\right\|_2 - d_{ij}^k \right)^2+ \lambda \sum_i (\bxt{(\cdot)}_i)^T\, \mathbf{M}\, \bxt{(\cdot)}_i \nonumber\\ & = & \sum_{kj}(\bxt{k}_i - \bxt{k}_j)^T(\bxt{k}_i-\bxt{k}_j) \nonumber \\ &&- \sum_{kj} 2\,d_{ij}^k \left\|\bxt{k}_i - \bxt{k}_j\right\| + \sum_{kj} (d_{ij}^k)^2 \nonumber\\ && +\> \lambda\, \bigl(\bxt{(\cdot)}_i\bigr)^T\, \mathbf{M}\, \bxt{(\cdot)}_i \nonumber\\ & = & f_{vex}\bigl(\bxt{(\cdot)}_i\bigr) + f_{cave}\bigl(\bxt{(\cdot)}_i\bigr) + \text{const} \end{IEEEeqnarray} where \begin{align} &f_{vex}\bigl(\bxt{(\cdot)}_i\bigr) = \sum_{kj} (\bxt{k}_i- \bxt{k}_j)^T(\bxt{k}_i-\bxt{k}_j) + \lambda\, (\bxt{(\cdot)}_i)^T\, \mathbf{M}\, \bxt{(\cdot)}_i\\ &f_{cave}\bigl(\bxt{(\cdot)}_i\bigr) = - \sum_{kj} 2\,d_{ij}^k \left\|\bxt{k}_i - \bxt{k}_j\right\| \end{align} In the optimization, we can omit the constant term that doesn't depend on $\bxt{(\cdot)}_i$. The iterative CCCP algorithm $(\bxt{(\cdot)}_i)^{t-1} \mapsto (\bxt{(\cdot)}_i)^{t}$ is given by \begin{equation} (\bxt{(\cdot)}_i)^{t} = \argmin_{\bxt{(\cdot)}_i}\;\; u \bigl(\bxt{(\cdot)}_i, (\bxt{(\cdot)}_i)^{t-1}\bigr) \label{eq:cccp-def} \end{equation} where the convex upper bound is computed by taking the first-order Taylor expansion of the concave part of $f$. The concave part is, however, non-differentiable at $\bxt{k}_i = \bxt{k}_j$. We thus work with a modified subdifferential. \begin{IEEEeqnarray}{rCl} \IEEEeqnarraymulticol{3}{l}{\vec\nabla^{SD} f_{cave}\left((\bxt{k}_i)^{t-1}\right)}\nonumber \\ \quad & = & 2 \sum_k\sum_i d_{ij}^k \begin{cases} \frac{(\bxt{k}_i)^{t-1} - \bxt{k}_j}{\left\|(\bxt{k}_i)^{t-1} - \bxt{k}_j\right\|} & \text{if $ (\bxt{k}_i)^{t-1} \ne \bxt{k}_j$} \\ \left\{\mathbf{t} : \left\| \mathbf{t} \right\| = 1 \right\} & \text{if $(\bxt{k}_i)^{t-1} = \bxt{k}_j$} \end{cases} \label{eq:Subdiv} \end{IEEEeqnarray} In the latter case, $\mathbf{t}$ is chosen randomly. The usual definition of the subdifferential \cite{Rockafellar1970} would yield a less strong condition on $\mathbf{t}$, namely $\left\| \mathbf{t} \right\| \le 1$. We choose the stronger variant to achieve maximum descent in each optimization step. We introduce surrogate points $\bxht{(\cdot)}_j$. \begin{equation} \bxht{k}_j = \bxt{k}_j + d_{ij}^k \begin{cases} \frac{(\bxt{k}_i)^{t-1} - \bxt{k}_j}{\left\|(\bxt{k}_i)^{t-1} - \bxt{k}_j\right\|} & \text{if $ (\bxt{k}_i)^{t-1} \ne \bxt{k}_j$} \\ \left\{\mathbf{t} : \norm{\mathbf{t}} = 1 \right\} & \text{if $(\bxt{k}_i)^{t-1} = \bxt{k}_j$} \end{cases} \end{equation} The resulting upper bound for the original cost is \begin{IEEEeqnarray}{rCl} u\bigl( (\bxt{(\cdot)}_i)^{t}, (\bxt{(\cdot)}_i)^{t-1}\bigr) & =& \sum_{kj} \bigl((\bxt{k}_i)^t - \bxt{k}_j\bigr)^T\bigl((\bxt{k}_i)^t-\bxt{k}_j\bigr) \nonumber\\ && -\> 2 \sum_{kj} d_{ij} \norm{(\bxt{k}_i)^{t-1} -\bxt{k}_j} \nonumber\\ && -\> 2 \sum_{kj} \bigl((\bxt{k}_i)^t - (\bxt{k}_i)^{t-1}\bigr)\bigl(\bxht{k}_j - \bxt{k}_j\bigr) \nonumber\\ &&+\> \lambda\, \bigl((\bxt{(\cdot)}_i)^t\bigr)^T\, \mathbf{M}\, (\bxt{(\cdot)}_i)^t \label{eq:upperbound} \end{IEEEeqnarray} The updating step is thus a simple spline regression and can be performed analytically \footnote{Spline regression involves a matrix inversion with cost $\mathcal{O}(T^3)$, but the inverse needs to be computed only once. Once the inverse has been obtained the cost of the spline regression becomes $\mathcal{O}(T^2)$, with further savings possible using sparse matrix techniques. }. For simplicity of notation, we rewrite the embedding points $\bxt{(\cdot)}$ as column vectors in $\ensuremath{\mathbb{R}}^{T\cdot D}$, $\vec\mathbf{x}_i = \mathbf{vec} \, \bxt{(\cdot)}_i$. \begin{equation} \label{eq:update} \vec\mathbf{x}_i^{t} = \left(N\cdot \mathbf{E}_{T \cdot D} + \lambda \mathbf{E}_D \otimes \mathbf{M}\right)^{-1}\left(\sum_j \vec\hat{\mathbf{x}}_j\right) \end{equation} \cite{Agarwal2010} introduce the same surrogate points based on geometrical considerations for metric (non-continuous) MDS. There, they are defined as projecting $\mathbf{x}_i$ on the Sphere $\mathcal{S}_j$ with center $\mathbf{x}_j$ and radius $d_{ij}$. In our formulation, they arise naturally from a natural splitting of the cost function into convex and concave parts. Additionally, using the subgradient circumvents the problem of the surrogate points being undefined for $\mathbf{x}_i = \mathbf{x}_j$, which \cite{Agarwal2010} do not address. Together with the iteration over the curves $\bxt{(\cdot)}_i$ we have everything we need for the cMDS algorithm (see Fig.~\ref{alg:cmds}). The complexity of the algorithm is $\mathcal{O}(dT^2 N^2)$ per iteration of the outer loop, since its function \textsc{MM} has complexity $\mathcal{O}(dT^2N)$ due to the matrix product in the spline regressions. \begin{figure} \begin{algorithmic}[0] \Function{cMDS}{$\bDt{(\cdot)},\mathbf{X}^{(\cdot)}_0,\lambda,\mathbf{M},\delta$} \While{$k < \text{maxIter}$} \State $\epsilon \gets \bXt{(\cdot)}$ \For{($i = 1:N$)} \State $\bxt{(\cdot)}_i \gets$ \Call{MM}{$\bDt{(\cdot)},\bXt{(\cdot)},i,\text{params}$} \EndFor \State $k \gets k+1$ \EndWhile \State \textbf{return} $\bXt{(\cdot)}$ \EndFunction \vspace{0.5cm} \Function{MM}{$\bDt{(\cdot)},\bXt{(\cdot)},i$, params} \Repeat \State $\epsilon \gets \bxt{(\cdot)}_i$ \State $\bxht{(\cdot)} \gets $ \Call{Majorize}{$\bXt{(\cdot)},\bDt{(\cdot)},i,\text{params}$} \State $\bXt{(\cdot)} \gets$ \Call{Minimize}{$\bXt{(\cdot)},\bxht{(\cdot)},i,\text{params}$} \Until{($\left(\overline{\epsilon - \bxt{(\cdot)}_i}\right) / \left(\overline{\bxt{(\cdot)}_i}\right) > \delta$)} \State \textbf{return} $\bxt{(\cdot)}_i$ \EndFunction \end{algorithmic} \caption{The cMDS algorithm.} \label{alg:cmds} \end{figure} The resulting algorithm can also be interpreted as a majorization-minimization algorithm \cite{DeLeeuw1977,DeLeeuw1994,Hunter2004} which is true for any CCCP algorithm \cite{Sriperumbudur2009}. The majorization entails the computation of the surrogate points, while minimization is the computation of the updating step. \subsection{Proof of first-order optimality of the cMDS algorithm} \label{sec:proof} Since CCCP has become an important tool in machine learning, there is a vast literature on convergence proofs for CCCP in general. \cite{Sriperumbudur2009} analyze the convergence of CCCP, for smooth and differentiable cost functions. In this case, the authors prove global convergence, in the sense that the algorithm arrives at a stationary point, i.e. a local minimum or saddle point, from any initialization point. The cMDS cost function, however, is nonsmooth and nondifferentiable at $\bxt{k}_i = \bxt{k}_j$. \cite{Yen2012} focus on convergence rate of CCCP and work with nonsmooth objective functions. They connect CCCP to more general block coordinate descent methods. They use results on convergence of coordinate descent on nonconvex and nonsmooth problems \cite{Tseng2009} by showing that CCCP is an instance of block coordinate descent. However, the authors constrain the class of objective functions to which their proof applies. The nonsmooth part should be convex piecewise linear, which is not the case for cMDS. \cite{Razaviyayn2013} extend block coordinate descent methods to inexact block coordinate descend. This allows the authors to treat nondifferentiable and nonconvex objective functions. The authors unify convergence results for various methods such as CCCP and Expectation Maximization. We follow ideas of this paper to show that a limit point of the optimization chain of a single curve in cMDS via CCCP is a stationary, first-order optimal point of the cost. For readability we denote $x = \bxt{(\cdot)}_i$ in the following. It is easy to see that, by optimizing single curves via minimization of an upper bound, we monotonically decrease the original cost. \begin{equation} \label{eq:decr.seq} f\left(x^{1}\right) \geq f\left(x^{2}\right) \geq f\left(x^{3}\right) \geq \dots \end{equation} However, from this it is not yet clear (assuming convergence of the algorithm) that the resulting configuration is a stationary point of the original cost, since the algorithm could produce a non-increasing sequence that does not tend to an optimal point. Therefore, we need to show that the limit point $z$ of the inner loop is first-order optimal, i.e. $\vec{\nabla} f(x) \bigl|_{x=z} = 0$. Then, we prove that this entails first-order optimality of the entire algorithm. Strictly speaking, this only holds with probability one, with respect to the randomisation step in (\ref{eq:Subdiv}), which chooses the direction of the next step when an iterate ends on one of the other data points. We will first show that with probability one, such a point will never be a fixed point of the algorithm. For simplicity, we consider this for the standard MDS cost function without the penalty term. The updating step is then \begin{equation} \mathbf{x}_i^t = \frac{1}{N} \sum_j \hat{\mathbf{x}}_j \end{equation} Now we assume, without loss of generality, that the current iterate $\mathbf{x}_i^{t-1} = \mathbf{x}_N$. Using the definition of the auxiliary points, we have: \begin{equation} \mathbf{x}_i^t = \frac{1}{N}\left( \sum_j \mathbf{x}_j + \sum_{j=1}^{N-1} d_{ij}\frac{\mathbf{x}_N-\mathbf{x}_j}{\left\| \mathbf{x}_N-\mathbf{x}_j \right\|} + d_{iN} \,\mathbf{t}\right) \label{eq:lin_t} \end{equation} If this point was a fixed point of the algorithm, i.e. $\mathbf{x}_i^{t} = \mathbf{x}_i^{t-1}$, there would be only one single direction $\mathbf{t}$ which fulfills the resulting linear equation (\ref{eq:lin_t}). But this event has probability zero when $\mathbf{t}$ is chosen randomly. We thus know that for any fixed point, with probability one, $u(\cdot,z)$ and $f(\cdot)$ are differentiable at $x=z$. Hence, with probability one, for any fixed point $z$ of the algorithm the following condition on the derivative holds. \begin{equation} \vec{\nabla} u(x, z)\biggl|_{x = z} = \vec{\nabla} f\left(z\right) \label{eq:con_deriv} \end{equation} \begin{lem} \label{lem:innerloop} Every limit point of the iterates generated by (\ref{eq:update}) is first-order optimal with respect to the problem \begin{equation} \text{min}\, f\left(x\right) \end{equation}. \end{lem} \begin{proof} Let us assume that a subsequence ${x^{t_j}}$ exists which converges to a limit point $z$. \begin{equation*} u(x^{t_{j+1}}, x^{t_{j+1}}) = f\left(x^{t_{j+1}}\right) \leq u(x^{t_{j+1}}, x^{t_{j}}) \leq u(x, x^{t_{j}}) \; \forall x \end{equation*} The equality follows from the fact that $u(y, y) = f\left(y\right)$. The first bound follows from $u(x,y) \geq f(x)$ and the last inequality is due to the optimality of $x^{t_{j+1}}$. We now perform the limit $j \rightarrow \infty$ and arrive at \begin{equation*} u(z,z) \leq u(x,z) \; \forall x \end{equation*} Thus $u(\cdot,z)$ has a local minimum at $x=z$. This implies that \begin{equation*} \vec{\nabla} u(x,z)\biggl|_{x = z} = 0 \end{equation*} since $u(\cdot,z)$ is a convex differentiable function. Combining this with (\ref{eq:con_deriv}) we obtain \begin{equation*} \vec{\nabla}f(z) = 0 \end{equation*} \end{proof} This proof does not exclude the possibility of having several limit points with the same cost between which the algorithm jumps. However, in the algorithm, we determine convergence based on the stationarity of the configuration and not of the cost. Thus, if the inner loop converges, it converges to a unique configuration which is thus a stationary point. If the algorithm did indeed jump between several limit points, it would not converge and stop after a certain number of iterations. However, this has never happened in our experience. \begin{thm} The cMDS algorithm has first-order optimal limit points. \end{thm} \begin{proof} It is easy to see that, since the cost function is symmetric, coordinate-wise first-order optimality yields overall first-order optimality. With Lemma \ref{lem:innerloop} we know that the inner loop of the algorithm yields such coordinate-wise optimality. Thus, we get first-order optimality of the limit points generated by the cMDS algorithm. \end{proof} Beyond the fact that the algorithm does converge to an appropriate point, it would be interesting to prove some results on the speed at which it converges. Unfortunately, local convergence is very difficult to stydy in our case since most techniques rely on the assumption that the cost function becomes quadratic in a neighbourhood of an optimum. Due to the invariances inherent in the MDS cost function, it is not clear at all that a local quadratic model is in any way appropriate. \subsection{cMDS with more than one hyperparameter} In certain cases it might be interesting to look at the effects of varying more than one hyperparameter. An extension of cMDS is thus the use of two or more hyperparameters, $\a$ and $\b$. This allows to consider, for example, time plus an additional hyperparameter such as scale or weighting. Having multiple hyperparameters effectively only changes the penalty matrix used in the spline regressions: instead of a penalty matrix appropriate for parametric curves in $\mathbb{R}\rightarrow\mathbb{R}^d$, we need a penalty appropriate for curves in $\mathbb{R}^m\rightarrow\mathbb{R}^d$, where $m$ is the number of parameters. As an example, take the case where we have two hyperparameters, $m = 2$. Suppose we have a grid $\a_1 \,\ldots\, \a_{T_\a} \times \b_1 \,\ldots\, \b_{T_\b}$. Then, our configuration matrix is $\bXt{(\cdot)\rep} \in \ensuremath{\mathbb{R}}^{T_\a \cdot T_\b \times d \times N}$. If $T_\a \neq T_\b$, we have two matrices for the discrete second order differential operator, $\mathbf{M}_\a$ and $\mathbf{M}_\b$, of appropriate dimensionality. Thus, we get the following penalty matrix for two hyperparameters: \begin{equation} \label{eq:two_hyperparameters} \mathbf{M} = \mathbf{M}_\b \otimes \mathbf{E}_{T_\a} + \mathbf{E}_{T_\b}\otimes \mathbf{M}_\a \end{equation} defining a separable penalty across the two hyperparameters. For more on penalty matrices, see for example \cite{Ramsay2005}. The results can be visualised by selecting certain slices of the grid. In the case of 1-dimensional embeddings, selecting a certain value for one hyperparameter leads to the visualization of curves that depend on the other hyperparameter. In case of 2-dimensional embeddings, one can select a certain value of one hyperparameter and then analyze the corresponding, possibly animated, scatterplot that varies with the second one. \subsection{Initialization} \label{sec:init} The cMDS algorithm needs to be seeded with a starting configuration. The optimization works with a random initialization, but performance can be improved with a more structured approach. One possibility is to perform classical scaling \cite{Kruskal1964,Torgerson1952} or other standard MDS methods such as SMACOF \cite{DeLeeuw1977,DeLeeuw2009} for each timestep. However, the separate solutions might be difficult for cMDS to "glue" together and thus form a poor initialisation. A more robust variant is initialization with an aggregated solution. To obtain this solution we average each curve $\bxt{(\cdot)}_i$ over time and then perform classical scaling on $\bar{\mathbf{X}}$. If some patterns are very strong for only a certain range of $\a$, they might influence the entire embedding via the aggregrated initialization. Thus, one should consider which kind of initialization is suitable for the data at hand. \subsection{Variants of cMDS} There exists multiple variants of the standard MDS problem. All of them can be implemented in cMDS. Most variants are defined over weights $\mathbf{W}^{(\cdot)}$ in the cost function: \begin{equation*} \label{eq:cost} \begin{split} \mathcal{C}\left(\bXt{(\cdot)},\bDt{(\cdot)}\right) =& \sum_{k=1}^T \norm{\mathbf{W}^k \circ \left( \bDXt{k} - \bDt{k} \right)^2}_F \\ &+ \lambda \sum_i (\bxt{(\cdot)}_i)^T\, (\mathcal{D}^{(2)})^T \mathcal{D}^{(2)}\, \bxt{(\cdot)}_i \end{split} \end{equation*} {\em Sammon's mapping} \cite{Sammon1969} can be implemented by setting $w_{ij}^k = (d_{ij}^k)^{-1}$. In {\em elastic stress}, on the other hand, we have weights $w_{ij}^k = (d_{ij}^k)^{-2}$ \cite{McGee1966}. This is equivalent to a Kamada-Kawai layout in graph visualization \cite{Kamada1989}. {\em Multidimensional unfolding} is useful when the data separates into groups \cite{Borg2005}. Then, the weights corresponding to between-group distances are set to zero. Local MDS (LMDS) \cite{Chen2009} is a slightly more complex variant. Here, the weights are set depending on local neighborhoods: If object $\mathbf{s}_j^k$ is a k-nearest neighbor of $\mathbf{s}_i^k$ (which we denote as $j \in \mathcal{N}_i$), then the corresponding term in the cost gets weight $w_{ij}=1$ and the original distance $d_{ij}$ is used. In all other cases the weight $w$ is set to a small value (not dependent on $i,j$) and $d_{ij}=D_{\infty}$, where $D_{\infty}$ is a large constant. This leads to a focus on local neighborhood structure and adds a repulsive force to avoid a 'crumbling together' of the embedding. The corresponding cMDS cost function is \begin{equation*} \label{eq:cost} \begin{split} \mathcal{C}\left(\bXt{(\cdot)},\bDt{(\cdot)}\right) =& \sum_{k=1}^T \sum_{i=1}^N\sum_{j \in \mathcal{N}_i} \left( \norm{\bxt{(\cdot)}_i-\bxt{(\cdot)}_j}_2 - d_{ij}^k \right)^2 \\ &+ \sum_{k=1}^T \sum_{i=1}^N\sum_{j \not\in \mathcal{N}_i} w^k \left( \norm{\bxt{(\cdot)}_i - \bxt{(\cdot)}_j}_2 - D_{\infty}^k \right)^2 \\ &+ \lambda \sum_i (\bxt{(\cdot)}_i)^T\, (\mathcal{D}^{(2)})^T \mathcal{D}^{(2)}\, \bxt{(\cdot)}_i \end{split} \end{equation*} LMDS introduces an additional hyperparameter via $w$, that needs to be optimized for the data at hand. It is also possible to use ISOMAP \cite{Tenenbaum2000} to visualize changes in manifold structure that are induced by a change in metric. To this end, one constructs geodesic distances based on different distances $d_{\alpha}\left(x,y\right)$ and subsequently applies cMDS. \subsection{Distance families} We would like to discuss some general properties of distance families. \paragraph{Weighted distances} A very common case is that data can be described using different features or sets of features. In that case, we can build two distance matrices $\mathbf{D}_1$ and $\mathbf{D}_2$ on one feature respectively. Then, we can construct a weighted metric using a convex combination of the two. Using a convex combination ensures that each resulting matrix is again a distance matrix. \begin{equation} \bDt{(\cdot)} = \sqrt{ \alpha \mathbf{D}_1^2 + (1-\alpha) \mathbf{D}_2^2} \end{equation} We present examples of weighted distance in Section \ref{sec:examples}. \paragraph{Changing inherent dimensionality} An interesting issue is a change in intrinsic dimensionality in the distance function. In classical scaling, low dimensional embeddings are projections of higher dimensional embeddings. For example, a 2d embedding is the projection of the 3d embedding. In reverse, this means that embedding high dimensional data in lower dimensions leads to larger distortion. In continuous embeddings, the dimensions are not stacked, as in the classical case. However, it is still true that a larger difference in dimensionality between original and embedded data leads to larger distortion. We illustrate this by mixing a distance matrix based on low dimensional data ($d=2$) with one based on high dimensional data ($d=12$). We embed this data in $d=2$. Plotting the distortion per timestep , as defined in (\ref{eq:distortion}), shows that, as expected, embedding the 2d data in $d=2$ yields zero distortion. The distortion increases as the high dimensional data gets more weight in the mixture. \begin{figure}[h] \centering \includegraphics{change_inherent_dimension.pdf} \caption{Change of inherent dimensionality: $\alpha=0$ represents a distance matrix based on two dimensional data, $\alpha=1$ is based on 12-dimensional data. The mixture was embedded in $d=2$. It is clearly visible that as the influence of the high-dimensional data increases, the distortion also increases. } \label{fig:inherent_dimension} \end{figure} \section{Examples} \label{sec:examples} In the following examples we show how cMDS can be used to analyze the effect of hyperparameters of distance functions on the data. The first example is a toy example, where we illustrate how cMDS can visualize cluster structure in data and how it compares to a more basic method based on $k$-means. In the second example, we use a weighted metric: distances evolve according to the relative weight given to some of the dimensions. In the third example, we work with brain connectivity data. We vary the distance function between networks according to different thresholding rules. In the fourth example, we derive a multiscale representation of data using hierarchical clustering and visualize changes in distance over scale. \subsection{Comparison of cMDS to clustering} We compare cMDS to a rudimentary method of tracking the influence of changing distances on the data structure. An approach based on traditional methods is to perform clustering on the original data for multiple instances of the distance family $d_{\alpha}(x,y)$. To track whether clusters in the data emerge independently of the distance measure, we use the set of cluster indices $c_i$ of a specific level of $\alpha$ for all levels of $\alpha$ and compute the corresponding quality of the clustering result, for example, by comparing within and between cluster variance. If the quality breaks down, this should be visible in the cMDS result on first glance. We demonstrate this with an artificial example. Suppose we draw five random cluster centers in $\ensuremath{\mathbb{R}}^5$ and then let those cluster centers linearly collapse to zero, such that for $\alpha=0$ we have well separated cluster centers and at $\alpha=0$ all cluster centers are at the origin. We then simulate Gaussian data at all time points, using the respective cluster centers and with 10 points per cluster. We then compute the corresponding distance matrices. Now, let us assume we don't know anything about the data. We compute a $k$-means clustering for $\alpha=0$. We use the resulting cluster indices for all levels of $\alpha$ and compute cluster quality. We do this ten times to get an average measure for cluster quality. The results are shown in Fig.~\ref{fig:clustering_kmeans}. We see that the quality of the clustering decreases with $\alpha$. From this we can get only that the clustering in the beginning is not present at the end. What we do not know is what the data actually looks like. Are there no clusters in the end or maybe different clusters than in the beginning? Let us now look at the cMDS embedding of the distance matrices (Fig.~\ref{fig:clustering_cmds}). Here we see that the clustering structure slowly degenerates with increasing $\alpha$ and that no other clusters arise for large $\alpha$. Thus, we gain a lot more insight about the data with a simpler method. \begin{figure*} \centering \subfloat[]{\includegraphics{clustering_kmeans.pdf} \label{fig:clustering_kmeans}} \hfil \subfloat[]{\includegraphics{clustering_cmds.pdf} \label{fig:clustering_cmds}} \caption{(a) Clustering quality based on $k$-means clustering. We performed clustering at $\alpha=0$ and used the resulting indeces at all levels of $\alpha$. One can see that the quality decreases, thus indicating that the cluster structure that is present in the beginning is not invariant. However, one cannot know what happens with the data exactly. (b) cMDS on the toy example. $\alpha=0$ represents the distance matrix based on data in five clusters in $\ensuremath{\mathbb{R}}^5$. $\alpha=1$ represents a random distance matrix. The results provide at a glance that clusters dissolvev at larger values of $\alpha$. This information is difficult to obtain based on $k$-means clustering alone.} \end{figure*} \subsection{Economic and demographic descriptors of EU countries} \label{sec:EU} \begin{figure*}[tbph] \centering \includegraphics[width=7.16in]{Embedding_all_8.pdf} \caption{Effects of changes in a weighted metric. Economic and demographic distance measures are weighted according to $\mathbf{D}(\alpha) = \sqrt{\alpha \cdot \mathbf{D}_E^2 + (1-\alpha) \cdot \mathbf{D}_D^2}$. Thus, the first panel represents an embedding that is solely based on the demographic metric, while the last one is based on the economic one. In between are samples of different weights, denoted by $\alpha$. The red labels depict the neighborhood relation between the Netherlands (NLD) and Finland (FIN) which is invariant with respect to $\alpha$, while the green ones, Bulgaria (BGR) and Estland (EST), are an example of a strong dependency on $\alpha$.} \label{fig:EUcountries} \end{figure*} For this example we use economic and demographic descriptors of the 27 EU countries. The data are publicly available from the {\bf gapminder} website. As economic variables we use income per capita (2008), CO$\mbox{}_2$ emissions per capita (2008) and number of granted patents per capita (2002). Demographic variables are total fertility rate, life expectancy at birth and the fraction of urban population. We scale all variables logarithmically. We then build two distance matrices, one solely based on economic variables, $\mathbf{D}_E$, the other on demographic variables, $\mathbf{D}_D$. Now, we put different weights on both variable groups and have a continuous range with one extreme only considering demographic variables and the other extreme only considering economic ones: \[\mathbf{D}(\alpha) = \sqrt{\alpha \cdot \mathbf{D}_E^2 + (1-\alpha) \cdot \mathbf{D}_D^2}\] where $\alpha$ is between 0 and 1. We sample this continuum at $N$ different values of $\alpha$ and thus get a distance array in $R^{N \times 27 \times 27}$. For this example, a 1D embedding is not sufficient to capture relevant trends in the data. Thus, we give snapshots of the 2D results in Fig.~\ref{fig:EUcountries}. In this case, an interactive presentation of the results is much easier to read and thus an advantageous choice. An interactive visualization is viewable \href{http://tinyurl.com/cMDS-demo}{here}\footnote{\url{http://tinyurl.com/cMDS-demo}}. We also implemented an interactive web application with the R package {\bf shiny}, which is online \href{http://ginagruenhage.shinyapps.io/EU-App}{here}\footnote{\url{http://ginagruenhage.shinyapps.io/EU-App}}. The 2D embedding shows that the changes in weighting have significantly different effects on the individual neighborhood relations. For example, some demographically very similar countries start diverging when economic variables are taken into account and end up far apart under the economic distance metric (e.g. Bulgaria and Estonia). Other countries stay similar, independent of the weighting of different distances (e.g. Finland and the Netherlands). These patterns are lost when deciding a distance measure a priori. Using cMDS to visualize the effects of the hyperparameter makes them easier to discover and understand. In the web application, one can also toggle quantitative analyses of the cMDS output, such as a vector graphic showing local stability of various countries. It is also possible to color countries according to their respective penalty, to judge overall stability. \subsection{Diffusion Tensor Imaging} \label{sec:DTI} \begin{figure*}[tbph] \centering \includegraphics[width=7.16in]{network_thresholding_LH.pdf} \caption{Embedding of thresholded networks aquired by DTI and tractography \cite{Hagmann2008}. The hyperparameter $\alpha$ corresponds to the quantile of weakest connections that is ignored for inter-subject comparisons. Measurements are compared based on the resulting binary connectivity matrix using the Hamming distance between graphs. The analysis shows that the two measurements for subject A, A1 and A2, differ about as much as other pairwise comparisons. This is interesting because it highlights the relatively low reliability of DTI tractography.} \label{fig:dti_network} \end{figure*} \begin{figure*} \centering \includegraphics[width=7.16in]{agreement_embedding_LH.pdf} \caption{Embedding of regional networks aquired by DTI and tractography \cite{Hagmann2008}. The different panels correspond to different threshold levels $\alpha$ when building consensus networks. The superior frontal cortex (SF) and superior parietal cortex (SP) are stable regions in the core of the network, while the parahippocampal cortex (PARH), the caudal anterior cingulate (CAC), and the transverse temporal cortex (TT) are stable peripheral regions. Some regions are less stable, e.g. the pars orbitalis (PORB), the temporal pole (TP), the medial orbitofrontal cortex (MOF) and the postcentral gyrus (PSTS).} \label{fig:dti_concensus} \end{figure*} Brain regions are linked by white matter tracts, forming a network called the Connectome \cite{Koetter2005}. Diffusion Tensor Imaging (DTI) is a form of magnetic resonance imaging that can be used to find connections between brain regions (using tractography, \cite{Hagmann2003}). Here we use data obtained by \cite{Hagmann2008}, available \href{http://www.cmtk.org/datasets/homo_sapiens_01.cff}{here} \footnote{\url{http://www.cmtk.org/datasets/homo_sapiens_01.cff}}. DTI produces noisy results and it is difficult to compare individual subjects. Consequently, connectivity must often be averaged over individual subjects. There is by necessity some arbitrariness in the averaging and we show here how cMDS can be used to visually compare different subjects and to visualize changes in network structure as the averaging criterion is varied. In the original data, the brain is segmented into 998 regions of interests that cover about $1.5~\textrm{cm}^2$ each and belong to one of 66 anatomical regions (33 per hemisphere). Here, we do not work with the full network, but rather with regionally aggregated data for the left hemisphere. That is, the data are adjacency matrices in $\ensuremath{\mathbb{R}}^{33\times 33}$, where $A^{(s)}_{ij}$ is the connection strength between regions $i$ and $j$ for subject $s$. Connections are measured for five different subjects, with two separate measurements for subject A. To compare different subjects and to evaluate whether strong connections are more stable across subjects than weak ones, we perform a thresholding analysis. Our hyperparameter is the quantile of weak connections that we ignore in the comparisons. Thus, $\alpha = 0.1$ corresponds to setting the weakest 10\% of connections to zero. We then binarize the adjacency matrix, $B^{(s)} > A^{(s)}$, such that we have $B^{(s)}_{ij} \in {0,1}$ and compute the Hamming distance between the graphs corresponding to different measurements. We use embeddings in $\ensuremath{\mathbb{R}}^2$ to represent the data (Fig.~\ref{fig:dti_network}). Surprisingly, the approximate distances between the two measurements of subject A, A1 and A2, are of the same order of magnitude compared to the distances between distinct subjects. This is true across thresholding levels. This is a curious finding, since these two measurements are averaged in \cite{Hagmann2008} before inter-subject averaging is performed. Our findings suggest that these two measurements are not easily distinguishable from other pair comparisons. According to \cite{Hagmann2008}, the correlation between the regional connectivity of A1 and A2 is $r^2 = 0.78$, compared to $r^2 = 0.65$ between distinct subjects. The cMDS visualization suggests that A1 and A2 are not notably more similar than other pairwise comparisons. Because of the relatively high noise in the data some form of averaging (over subjects) may be useful to perform structural analyses on the regional network. Due to our findings in the network thresholding analysis we treat measurements A1 and A2 separately. One approach to perform averaging is to produce consensus networks out of the individual networks, with the rule that a link is introduced in the 50\% consensus network if and only if it is present in at least 50\% of the individual networks. Thus, we look at the average adjacency matrix $\bar{B}$, where $\bar{B}_{ij}= \left( 1/N \right) \sum_s B^{(s)}_{ij}$ and $B^{(s)} = A^{(s)}>0$. We then threshold it at different levels to produce different consensus networks. We use the threshold level $\alpha$ as the continuous parameter in cMDS, such that $\bar{B}_{ij}(\alpha) = \bar{B}_{ij}>\alpha$. If we take the shortest-path distance on the graph that is defined by $\bar{A}(\alpha)$ as the distance measure between two regions, then changing the threshold level is exactly the same as changing the distance measure, as some links will start disappearing with higher threshold values. We can therefore apply cMDS to visualize the changes in network structure as the averaging rule is changed. For this example, we implement a weighted version of the algorithm, to mirror the standard Kamada-Kawai layout methods \cite{Kamada1989}. The results are shown in Fig.~\ref{fig:dti_concensus} for regions in the left hemisphere. We also present these results interactively with a web application using the R package {\bf shiny} which is viewable \href{http://ginagruenhage.shinyapps.io/DTI-App}{here}\footnote{\url{http://ginagruenhage.shinyapps.io/DTI-App}}. A first (and unsurprising) result is that the network density decreases significantly with the threshold level. That is, as we start requiring higher levels of consistency among the subjects, a lot of connections are rejected. We would like to note, that in this case, integer distances are embedded in a continuous space. However, aspects such centrality are recovered in the visualization: regions with dense connections and high values of centrality and betweenness are placed at the center of the configuration. Results show that some regions are very stable: for example, the superior frontal cortex (SF) and the superior parietal cortex remain at the core of the network, while the parahippocampal cortex (PARH), the caudal anterior cingulate (CAC) and the temporal cortex (TT) are examples of stable regions at the periphery of the network. What is even more interesting is that some regions are rather unstable and change their role in the network. The postcentral gyrus (PSTS) starts out in the periphery and then moves to the center. Other regions move from the core to the periphery, e.g. the pars orbitalis (PORB), the temporal pole (TP) and the medial orbitofrontal cortex (MOF). Since core-periphery relationships are central to the interpretation of connectome data, it is crucial to know which regions can be reliably called peripheral and others central \cite{Hagmann2008}. cMDS provides this information at a glance. \subsection{Hierarchical clustering} \label{sec:hclust} We mentioned in the introduction that distance is often computed relative to a certain scale. For spatial or temporal data, scale corresponds to a concrete spatial or temporal window, but there are other ways to obtain a multiscale representation. Hierarchical clustering \cite{Kaufman1990,Hastie2009} is such a technique. We focus here on agglomerative clustering, where the algorithm starts with each observation in one cluster. At each level, the algorithm merges the two closest clusters until only one cluster remains. Thus, there are $N-1$ levels in the hierarchy, which gives a view of the data going from the roughest to the most detailed level. cMDS provides an interesting visualization of the results. We first define a distance function at each level of the hierarchy. For each level, we build the distance matrix for the N datapoints as follows: if two datapoints are in the same cluster we assign a very small positive distance, otherwise we assign the distance between the respective cluster centers. Here, we use the publicly available {\bf USArrests} dataset from the R {\bf datasets} package. It contains data on murder arrests, assault arrests, rape arrests and urban population for the different US states in 1973. We picked this dataset because it is used as an example for the R {\bf hclust} function. We use cMDS to embed these data (Fig.~\ref{fig:hclust}). The result is a tree structure with the property that, at each level of the tree, distances between branches are representative of distances between clusters. This enables an immediate understanding of the hierarchical clustering results. We also developed an interactive visualization based on a 2D embedding, which better captures the potential of cMDS for such applications. We invite readers to have a look at these results (online \href{http://tinyurl.com/cMDS-demo}{here}\footnote{http://tinyurl.com/cMDS-demo}). \begin{figure}[h] \centering \includegraphics[width=3.5in]{embedding_centered_trial1.pdf} \caption{Embedding of an hierarchical clustering results for the {\bf USArrests} data set which is publicly available as part of the R \textbf{datasets} package. We selected a sample of 17 states. We ran hierarchical clustering using the centroid of clusters for agglomeration. Then, we built the distance matrix for each level: two datapoints are assigned a small positive distance if they are in the same cluster and the distance between cluster centers otherwise. The cMDS result is a visualization of the tree structure. At each level, the distances between branches represents the distances between cluster centers.} \label{fig:hclust} \end{figure} \section{Discussion} We introduced an easy to implement and flexible version of continuous (or dynamic) multi-dimensional scaling, namely cMDS. We showed that cMDS provides a fast and informative way of understanding the data structure by presenting four examples. In a toy example, we compared the results to a method based on $k$-means which gave only an approximate idea of what was going on in the data, while cMDS yielded a very concise representation of the data. With the second example on EU data we showed the effects of changing a weighted metric, putting different weights on two feature classes. Visualization revealed countries whose neighborhood relations are invariant to the change in weights as well as countries whose relative position strongly depends on the weights. In a brain connectivity example, where averaging over subjects is not straightforward, we found a way to compare different subjects giving more and more weight to strong connections. We found that cMDS suggests two measurements of the same subjects to be as different as measurements of two different subjects. In a second analysis we found that the shape of the network changes according to the averaging rule. With cMDS, identifying stable and unstable regions turned out to be straightforward. Finally, we showed how hierarchical clustering can be thought of as a multiscale representation of data, and we used cMDS to visualize how the structure of the data changes across scale. We only considered small datasets here, for which cMDS is very fast (with a runtime of a few seconds at most). Like other MDS methods, cMDS has $\mathcal{O}(n^3)$ scaling in its naive version. Runtime remains reasonable with a few hundred datapoints, and straightforward extensions for larger datasets are possible. One idea is to embed a subset of landmark points, as in landmark MDS Another is to use sparse weighting matrices, tying each datapoint to a random subset of neighbours. By visual inspection, cMDS immediately reveals qualitative structures in the neighborhood dynamics for various datasets. Furthermore, quantitative analyses are possible. For example, performing clustering on cMDS output can yield results that are robust to changes in the distance measure. We leave these extensions to future work. \section{Conclusion} With cMDS, we address a fundamental problem in pattern recognition and machine learning: the initial choice of a distance metric. This is a hidden assumption in various methods. We argue that this choice should be addressed explicitly. Our suggestion is to use continuous MDS techniques, which visualize the dynamics that (so far) arbitrary choices in distance functions introduce in data. cMDS can deal with numerous sources of arbitrariness in the distance metric, examples of which are varying scale or weighting. We show that interesting and important dynamics, such as invariance and declustering, are readily revealed by cMDS. Finally, we provide a cMDS algorithm that is straightforward to implement, use and extend. \section*{Acknowledgments} This work was partially supported by the Deutsche Forschungsgemeinschaft (GRK1589/1). \newpage \bibliographystyle{IEEEtran}
\section{Introduction} Let $\mathcal{D}\subset \mathbb{R}^d$, $d\le 3$, be a spatial domain with smooth boundary $\partial\cD$ and consider the stochastic partial differential equation written in the abstract It\^o form \begin{equation}\label{eq:sac} \dd u+Au\,\dd t+f(u)\,\dd t=\dd W,~t\in(0,T];\quad u(0)=u_0, \end{equation} where $\{W(t)\}_{t\ge 0}$ is an $L^2(\cD)$-valued $Q$-Wiener process on a filtered probability space $(\Omega, \cF, \mathbb{P}, \{\cF_t\}_{t\ge 0})$ with respect to the normal filtration $ \{\cF_t\}_{t\ge 0}$. We use the notation $H=L^2(\cD)$ with inner product $\langle\cdot\,,\cdot\rangle$ and induced norm $\|\cdot\|$ and $V=H^1_0(\cD)$. Moreover, $A\colon V\to V'$ denotes the linear elliptic operator $Au=-\nabla\cdot(\kappa\nabla u)$ for $u\in V$, where $\kappa(x)>\kappa_0>0$ is smooth. As usual we consider the bilinear form $a\colon V\times V\to\mathbb{R}$ defined by $a(u,v) = ( Au,v)$ for $u,v\in V$, and $(\cdot\, ,\cdot)$ denotes the duality pairing of $V'$ and $V$. We denote by $\{E(t)\}_{t\ge 0}$ the analytic semigroup in $H$ generated by the realization of $-A$ in $H$ with $D(A)=H^2(\mathcal D)\cap H^1_0(\mathcal{D})$. Finally, $f\colon \cD_f\subset H\to H$ is given by $(f(u))(x)=F'(u(x))$, where $F(s)=c(s^2-\beta^2)^2$ ($c>0$) is a double well potential. Note that $f$ is only locally Lipschitz and does not satisfy a linear growth condition. It does, however, satisfy a global one-sided Lipschitz condition, which is a key property for proving uniform moment bounds. We consider a fully implicit Backward Euler discretization of \eqref{eq:sac} via the iteration \begin{equation}\label{eq:be} u^j-u^{j-1}+\Delta t\, A u^j+\Delta t\, f(u^j)=\Delta W^j,~j=1,2,\dots, N; \quad u^0=u_0, \end{equation} where $\Delta t>0$. Note that this scheme is implicit also in the drift term $f$. In return, the scheme preserves key qualitative aspects of the solution of \eqref{eq:sac} such as moment bounds. The following two results constitute the main results of the paper. For notation we refer to Section \ref{sec:prelim}. Let $N\in \mathbb{N}$, $T=N\Delta t$ and $t_n=n\Delta t$, $n=1,2,\dots, N$. In Theorem~\ref{thm:pwc} (pathwise convergence) we show that if $\|A^{\frac12+\varepsilon} Q^{\frac12}\|_{\HS}<\infty$ for some small $\varepsilon>0$, $\mathbb{E}\|u_0\|_1^2<\infty$, and $0\le \gamma <\frac12$, then there are finite random variables $K\ge 0$ and $\Delta t_0>0$ such that, almost surely, $$ \sup_{t_n\in [0,T]}\|u(t_n)-u^n\|\le K \Delta t^\gamma,\quad \Delta t\le \Delta t_0. $$ In Theorem~\ref{thm:strc} (strong convergence) we prove that if $p\ge 1$ and $\mathbb{E}\|u_0\|_1^{2p}<\infty$, then $$ \lim_{\Delta t\to 0}\mathbb{E}\sup_{t_{n}\in [0,T]}\|u(t_n)-u^n\|^p= 0. $$ Since the method of proof uses a priori bounds obtained via energy arguments together with a pathwise error analysis based on the mild formulation of the equation, a strong rate cannot be obtained via this line of argument. We would like to point out that the strong convergence of the Backward Euler scheme is somewhat surprising given the superlinearly growing character of $f$, see also the discussion in \cite{Jentzen1}. We do not know of any results where strong convergence results, with or without rates, are obtained for a time-discretization scheme for an SPDE with non-global Lipschitz nonlinearity without linear growth (for SODEs, we refer to \cite{HMS}). There are many results on pathwise and strong convergence of the Backward Euler scheme (usually explicit in the drift term $f$) under global Lipschitz conditions (or local Lipschitz with linear growth conditions), see, for example, \cite{CoxVN,GyM,Haus1,Haus2} and the references therein. For non-global Lipschitz nonlinearities the relatively recent method developed in \cite{Jentzen1} uses a scheme which is based on the mild formulation of the SPDE. This is also employed, for example, in \cite{Blomker_et_al}. In that setting pathwise error estimates are derived but strong convergence results would be rather difficult to obtain as the method loses the information about the one-sided Lipschitz condition on $f$, which can only be exploited in a variational or weak solution approach. We also mention \cite{P2001} where convergence in probability is obtained without global Lipschitz conditions for the Backward Euler scheme. Spatial pathwise convergence results for certain semilinear SPDEs with non-global Lipschitz $f$ without linear growth are obtained in \cite{Blomker_Jentzen,Blomker_et_al}, both using spectral Galerkin approximation. Concerning spatial strong convergence we only know of \cite{Liu-1} and \cite{SS}, both with rates, based on a spectral Galerkin method and a finite difference method, respectively. In the latter two papers the authors use energy type arguments, and hence they can fully exploit the one-sided Lipschitz character of $f$. Finally, we would like to note that \eqref{eq:be} is also referred to as Rothe's method. Since we can prove both pathwise and strong convergence, one can set up a nonlinear wavelet-based adaptive algorithm to solve the elliptic equation in each time-step and obtain a implementable scheme, which converges both path-wise and strongly in a similar way as in \cite{Dahlkeetal2} and \cite{KLU}. The paper is organized as follows. In Section \ref{sec:prelim} we collect frequently used results from infinite dimensional analysis and introduce some notation. In Section \ref{sec:reg} we discuss the spatial Sobolev regularity of the solution and the H\"older regularity in time. In Section \ref{sec:apriori} we prove maximal type $p$-th moment bounds on $u^n$ (Propositions \ref{prop:ul} and \ref{prop:ulp}), which are in fact the exact analogues of the ones on $u(t)$ (Proposition \ref{prop:eb}). Here we highlight that for $p=2$ the bounds only grow linearly in $T$, while for $p>2$ exponentially because of a Gronwall argument. In Section \ref{sec:mr} we state and prove the main results of the paper on the convergence of \eqref{eq:be}. An important part of the proof is a maximal type error estimate for the linear part (Proposition \ref{prop:wa}), where we employ a discrete version of the celebrated factorization method. \section{Preliminaries}\label{sec:prelim} Throughout the paper we will use various norms for linear operators on a Hilbert space. We denote by $\mathcal{L}(H)$, the space of bounded linear operators on $H$ with the usual operator norm denoted by $\|\cdot\|$. If for a positive semidefinite operator $T\colon H\to H$, the sum $$ \Tr T:=\sum_{k=1}^\infty\langle Te_k,e_k\rangle<\infty $$ for an orthonormal basis (ONB) $\{e_k\}_{k\in \mathbb{N}}$ of $H$, then we say that $T$ is trace-class. In this case $\Tr T$, the trace of $T$, is independent of the choice of the ONB. If for an operator $T\colon H\to H$, the sum $$ \|T\|_{\HS}^2:=\sum_{k=1}^\infty\|Te_k\|^2<\infty $$ for an ONB $\{e_k\}_{k\in \mathbb{N}}$ of $H$, then we say that $T$ is Hilbert-Schmidt and call $\|T\|_{\HS}$ the Hilbert-Schmidt norm of $T$. The Hilbert-Schmidt norm of $T$ is independent of the choice of the ONB. We have the following well-known properties of the trace and Hilbert-Schmidt norms, see, for example, \cite[Appendix C]{DPZ}, \begin{align} \label{eq:hs} \|T\|&\le \|T\|_{\HS},\quad \|TS\|_{\HS}\le \|T\|_{\HS}\|S\|, \quad\|ST\|_{\HS}\le \|S\|\,\|T\|_{\HS}, \\ \label{eq:tr} \Tr Q&=\|Q^{\frac12}\|_{\HS}^2=\|T\|^2_{\HS}=\|T^*\|_{\HS}^2, \quad\text{ if $Q=TT^*$.} \end{align} Next, we introduce fractional order spaces and norms. It is well known that our assumptions on $A$ and on the spatial domain $\mathcal{D}$ imply the existence of a sequence of nondecreasing positive real numbers $\{\lambda_k\}_{k\geq 1}$ and an orthonormal basis $\{e_k\}_{k\geq 1}$ of $H$ such that \begin{equation*}\label{eq:spectral} Ae_k = \lambda_k e_k, \quad \lim_{k\rightarrow +\infty} \lambda_k = +\infty. \end{equation*} Using the spectral functional calculus for $A$ we introduce the fractional powers $A^s$, $s \in \mathbb{R}$, of $A$ as \begin{equation*}\label{eq:fp} A^s v=\sum_{k=1}^{\infty}\lambda_k^s(v,e_k)e_k,\quad D(A^s)=\Big\{v\in H:\|A^sv\|^2=\sum_{k=1}^{\infty}\lambda_k^{2s}(v,e_k)^2<\infty\Big\} \end{equation*} and spaces $\dot{H}^\beta=D(A^{\beta/2})$ with inner product $\langle u , v\rangle_{\beta}=\langle A^{\frac{\beta}{2}}u , A^{\frac{\beta}{2}}v\rangle$ and induced norms $\norm[\beta]{v}=\norm{A^{\beta/2} v}$. It is well-known that if $0\le \beta < 1/2$, then $\dot{H}^\beta=H^\beta$ and if $1/2<\beta\le 2$, then $\dot{H}^\beta=\{u\in H^\beta:u|_{\partial \mathcal{D}}=0\}$, where $H^\beta$ denotes the standard Sobolev space of order $\beta$. We recall the fact that the semigroup $\{E(t)\}_{t\ge 0}$ generated by $-A$ is analytic and therefore it follows from \cite[Theorem 6.13]{Pazy} that for $t> s > 0$, \begin{align} &\|A^{\beta}E(t)v\|\le Ct^{-\beta}\|v\|,\quad\beta\ge 0,\label{eq:anal0}\\ \label{eq:anal} &\|A^{\beta}(E(t)-E(s))v\| \le Cs^{-(\beta+\gamma)}|t-s|^{\gamma+\rho}\|A^{\rho}v\|, ~\beta \ge 0,~0\le \gamma+\rho \le 1. \end{align} We will also use It\^o's Isometry and the Burkholder-Davies-Gundy inequality for It\^o-integrals of the form $\int_0^t\langle \eta(s),\dd \tilde{W}(s)\rangle$, where $\tilde{W}$ is a $\tilde{Q}$-Wiener process. For this kind of integral, It\^o's Isometry, \cite[Proposition 4.5]{DPZ} reads as \begin{equation}\label{eq:rito} \mathbb{E}\left|\int_0^t\langle \eta(s),\dd \tilde{W}(s)\rangle\right|^2 =\mathbb{E}\int_0^t\|\tilde{Q}^{\frac12}\eta(s)\|^2\,\dd s, \end{equation} and the Burkholder-Davies-Gundy inequality, \cite[Lemma 7.2]{DPZ}, takes the form \begin{equation}\label{eq:rbdg} \mathbb{E}\sup_{t\in [0,t_0]} \left|\int_0^t\langle \eta(s),\dd \tilde{W}(s)\rangle\right|^p \le C_p \mathbb{E}\left(\int_0^{t_0}\|\tilde{Q}^{\frac12}\eta(s)\|^2\,\dd s\right)^{\frac{p}{2}}, ~p\ge 2. \end{equation} Finally, if $Y$ is an $H$-valued Gaussian random variable with covariance operator $\tilde{Q}$, then, by \cite[Corollary 2.17]{DPZ}, we can bound its $p$-th moments via its covariance operator as \begin{equation}\label{eq:es1a} \mathbb{E}\|Y\|^{2p}\le C_p (\mathbb{E}\|Y\|^2)^{p} =C_p (\Tr \tilde{Q})^{p}=\|\tilde{Q}^{\frac12}\|_{\HS}^{2p}. \end{equation} \section{Regularity of the solution}\label{sec:reg} The following existence, uniqueness, and regularity result can essentially be found in \cite[Example 3.5]{Liu} for $\mathcal{D}=[0,1]$, where it is stated with $\essup$ instead of $\sup$ for the second term in \eqref{eq:esup}. It is remarked there, \cite[Remark 3.4]{Liu}, that the result can be proved in higher dimensions by using \cite[Example 3.2]{LR}, where domains with smooth boundary are considered. Finally, by \cite[Theorem 1.1]{Liu0}, the $\essup$ can be replaced by $\sup$ in the second term as stated below in \eqref{eq:esup}. We also note that for the equation considered in this paper, this result can be obtained by using the deterministic Ljapunov functional $J(u)=\|\nabla u\|_1^2+\int_{{\mathcal D}}F(u)\,\dd x$ and It\^o's formula in a way analogous to \cite[Theorem 3.1 and Corollary 3.2]{KML}, see also \cite{Prato}. For the definition of \emph{variational solution} we refer to \cite[Definition 4.2.1]{PR}. \begin{prop}\label{prop:eb} If $\|A^{\frac12}Q^{\frac12}\|_{\HS}<\infty$ and $\mathbb{E}\|u_0\|_1^p<\infty$ for some $p\ge 2$, then there is a unique variational solution $u$ of \eqref{eq:sac}. Furthermore, there is $C_T>0$ such that \begin{equation}\label{eq:esup} \mathbb{E}\sup_{t\in [0,T]}\|u(t)\|^p+\mathbb{E}\sup_{t\in [0,T]}\|u(t)\|_{1}^p\le C_T. \end{equation} \end{prop} In this case, $u$ is also a mild solution, see \cite[Proposition F.0.5 and Remark F.0.6]{PR}; that is, $u$ satisfies the integral equation \begin{equation}\label{eq:mild} u(t)=E(t)u_0+\int_0^tE(t-s)f(u(s))\,\dd s+W_A(t),\quad t\in[0,T], \end{equation} almost surely, where the stochastic convolution $W_A$ is defined by $$ W_A(t)=\int_0^tE(t-s)\,\dd W(s). $$ This ultimately follows from the fact that the noise is additive trace class and that, by Sobolev's inequality, \begin{equation}\label{eq:lg} \|f(u(t))\|\le C (\|u(t)\|+\|u(t)\|_{L^6}^3) \le C (\|u(t)\|+\|u(t)\|_{1}^3), \end{equation} which is bounded almost surely for $t\in [0,T]$ by Proposition \ref{prop:eb}. Note that here, in order to be able to use Sobolev's inequality, it is crucial that $d\le 3$ and that the nonlinearity $f$ is at most cubic. Next we look at the pathwise H\"older regularity of $u$. First we consider the stochastic convolution $W_A$. \begin{lem}\label{lem:wlip} Let $0<\beta\le 1$, $\|A^{\frac{\beta-1}{2}}Q^{\frac12}\|_{\HS}<\infty$ and $p> \frac{2}{\beta}$. Then, there is a nonnegative real random variable $K$ with $\mathbb{E}K^p<\infty$ such that, almost surely, $$ \sup_{t\neq s\in [0,T]}\frac{\|W_A(t)-W_A(s)\|}{|t-s|^{\gamma}}\le K \quad\text{for $0\le\gamma<\frac{\frac{\beta p}{2}-1}{p}$.} $$ \end{lem} \begin{proof} Let $t> s\ge 0$. Note that the stochastic integrals below are Gaussian random variables and hence we can use \eqref{eq:es1a} to bound their $p$-th moments. Therefore, \begin{equation*} \begin{split} \mathbb{E}\|W_A(t)-W_A(s)\|^p & \le C_p\mathbb{E}\left\|\int_s^t E(t-\sigma)\, \dd W(\sigma)\right\|^p \\ &\quad +C_p\mathbb{E}\left\|\int_0^s E(t-\sigma)-E(s-\sigma)\, \dd W(\sigma)\right\|^p\\ & \le C_p\left(\int_s^t \|E(t-\sigma)Q^{\frac12}\|^2_{\HS}\, \dd \sigma\right)^{\frac{p}{2}} \\ &\quad +C_p\left( \int_0^s \|E(t-\sigma)-E(s-\sigma)Q^{\frac12}\|_{\HS}^2\, \dd\sigma\right)^{\frac{p}{2}}\le C|t-s|^{\frac{\beta p}{2}}, \end{split} \end{equation*} where the last inequality is shown in the proof of \cite[Theorem 4.2]{KLU}. Then the statement follows from Kolmogorov's criterion, see, for example, \cite[Theorem 1.4.1]{Kunita}. \end{proof} With the above preparations, we now prove the H\"older continuity of $u$. Note that the result is suboptimal compared to the corresponding result for $W_A$ in Lemma~\ref{lem:wlip}, which requires only $\beta=1$ to get the same H\"older exponent, while here we assume $\beta=2$. This is a consequence of the fact that we use the mild formulation here and hence cannot exploit the one-sided Lipschitz condition on $f$ but only its cubic growth. \begin{prop}\label{propo:ulip} Let $\|A^{\frac{1}{2}}Q^{\frac12}\|_{\HS}<\infty$, $\mathbb{E}\|u_0\|_1^2<\infty$ and $T>0$. Then, for all $\gamma\in[0,\frac12)$, there is a finite nonnegative random variable $K$ such that, almost surely, $$ \sup_{t\neq s\in [0,T]}\frac{\|u(t)-u(s)\|}{|t-s|^{\gamma}}\le K. $$ \end{prop} \begin{proof} Let $T>0$, $0\le s< t\le T$, and $0\le \gamma<\frac{1}{2}$. We use the mild formulation \eqref{eq:mild} to represent $u(t)-u(s)$ as follows: \begin{align*} u(t)-u(s)&=(E(t)-E(s))u_0 +\int_s^tE(t-r)f(u(r))\,\dd r\\ & \quad +\int_0^s(E(t-r)-E(s-r)f(u(r))\,\dd r +W_A(t)-W_A(s). \end{align*} The estimate in \eqref{eq:anal}, with $\beta=\gamma=0$ and $\rho=\frac12$, implies that $\|(E(t)-E(s))u_0\|\le C|t-s|^{\frac12}\|u_0\|_1$. The second term can be bounded, using Proposition \ref{prop:eb} and \eqref{eq:lg}, \begin{align*} \left\|\int_s^tE(t-r)f(u(r))\right\|&\le\int_s^t\|E(t-r)\|\,\|f(u(r))\|\,\dd r\\ &\le C |t-s| \sup_{r\in [0,T]}(\|u(r)\|+\|u(r)\|_1^3)\le K |t-s|. \end{align*} In a similar fashion, using this time \eqref{eq:anal} with $\beta=\rho=0$ and $\frac12\le\gamma<1$, \begin{align*} &\left\|\int_0^s(E(t-r)-E(s-r))f(u(r))\,\dd r\right\|\\ &\qquad\le \int_0^s\|(E(t-r)-E(s-r))\|\,\dd r\sup_{r\in [0,T]}(\|u(r)\|+\|u(r)\|_1^3)\\ &\qquad\le K |t-s|^\gamma \int_0^sr^{-\gamma}\dd r\le KT^{1-\gamma}|t-s|^{\gamma}. \end{align*} Finally, we note that $\|Q^{\frac12}\|_{\HS}\le C\|A^{\frac{1}{2}}Q^{\frac12}\|_{\HS}<\infty$ by \eqref{eq:hs} as $A^{-\frac12}\in \mathcal{L}(H)$, so that we can use Lemma \ref{lem:wlip} with $\beta=1$ to conclude the proof. \end{proof} \section{A priori moment bounds}\label{sec:apriori} Our first result bounds the second moment of the Euler iterates in \eqref{eq:be}. The proof uses a kind of bootstrapping argument and as a result we avoid Gronwall's lemma. Therefore, we are able to obtain bounds that only grow linearly with $T$ instead of exponentially. Since these bounds will be used in the Gronwall step in the pathwise convergence analysis, the constants appearing there will grow exponentially with time instead of double-exponentially. We have to use test functions in the energy arguments below that are different from the ones used in the deterministic setting, for example in \cite{Stig}, because of the presence of a non-differentiable right hand side. This ultimately forces the choice of a scheme implicit also in the drift in order to be able to use the one-sided Lipschitz property of $f$. \begin{prop}\label{prop:ul} Let $I_N=\{1,2,\dots, N\}$ and $T=N\Delta t$. If $\|A^{\frac12}Q^{\frac12}\|_{\HS}<\infty$ and $\mathbb{E}\|u_0\|_1^2<\infty$, then there is $C>0$ independent of $T$ such that $$ \mathbb{E}\sup_{l\in I_N}\|u^l\|^2+\mathbb{E}\sup_{l\in I_N}\|u^l\|^2_1\le C(1+T). $$ \end{prop} \begin{proof} First note that it is enough to bound the second term on the left hand side since $\|\cdot\|\le C\|\cdot\|_1$. Taking the inner product of \eqref{eq:be} with $u^j$, we get $$ \langle u^j-u^{j-1}, u^j\rangle +\Delta t\, \|u^j\|_{1}^2 +\Delta t\, \langle f(u^j),u^j\rangle =\langle \Delta W^j, u^j\rangle. $$ Using the identity $\langle x-y,x\rangle=\frac{1}{2}(\|x\|^2-\|y\|^2)+\frac12\|x -y\|^2$ and the fact that for some $C>0$ we have $sf(s)\ge -C$ for all $s\in \mathbb{R}$ we get \begin{align*} &\frac{1}{2}(\|u^j\|^2-\|u^{j-1}\|^2)+\frac12\|u^j -u^{j-1}\|^2+\Delta t\, \|u^j\|_{1}^2 \\ &\qquad \le C\Delta t +\langle \Delta W^j, u^j-u^{j-1}\rangle+\langle \Delta W^j, u^{j-1}\rangle. \end{align*} Using a kick back with the second term on the right and summing from $1$ to $n$ ($1\le n\le N$) gives \begin{align*} &\|u^n\|^2+\sum_{j=1}^n\|u^j -u^{j-1}\|^2+\Delta t\, \sum_{j=1}^n\|u^j\|_{1}^2 \\ &\qquad \le C\left(T+\|u_0\|^2+\sum_{j=1}^n\left(\|\Delta W^j\|^2+\langle \Delta W^j, u^{j-1}\rangle\right)\right). \end{align*} Taking expectation, using that $\Delta W^j$ is Gaussian with covariance operator $\Delta t\, Q$ and hence $\mathbb{E}\|\Delta W^j\|^2=\Delta t\, \Tr Q=\Delta t\, \|Q^{\frac12}\|^2_{\HS}$, and that $\mathbb{E}\sum_{j=1}^n\langle \Delta W^j, u^{j-1}\rangle=0$, we conclude \begin{equation}\label{eq:iter1} \begin{aligned} &\mathbb{E}\left(\|u^n\|^2+ \sum_{j=1}^n\|u^j -u^{j-1}\|^2+\Delta t\, \sum_{j=1}^n\|u^j\|_{1}^2 \right)\\ &\qquad \le C\left(T+\mathbb{E}\|u_0\|^2+T\|Q^{\frac12}\|^2_{\HS}\right). \end{aligned} \end{equation} Next, we take the inner product of \eqref{eq:be} with $Au^j$ and obtain similarly as above \begin{align*} &\frac{1}{2}(\|u^j\|_{1}^2-\|u^{j-1}\|_{1}^2)+\frac{1}{2}\|u^j-u^{j-1}\|_{1}^2+\Delta t\, \|u^j\|_{2}^2+\Delta t\,\langle f(u^j),u^j\rangle_{1} \\ &\qquad = \langle \Delta W^j, u^j\rangle_{1}. \end{align*} Since $f'(s)\ge -C$, we have \begin{equation*} \langle f(u^j),u^j\rangle_{1}=\langle \nabla f(u^j),\nabla u^j\rangle =\langle f'(u^j)\nabla u^j,\nabla u^j\rangle \ge -C\|u^j\|_1^2. \end{equation*} Hence, \begin{align*} &\frac{1}{2}(\|u^j\|_{1}^2-\|u^{j-1}\|_{1}^2)+\frac{1}{2}\|u^j-u^{j-1}\|_{1}^2+\Delta t\,\|u^j\|_{2}^2\\ &\qquad \le \Delta t\, C\|u^j\|_{1}^2+\langle\Delta W^j, u^j-u^{j-1}\rangle_{1}+\langle \Delta W^j, u^{j-1}\rangle_{1}. \end{align*} Thus, using a kick back with the second term, we obtain \begin{align}\label{eq:1norm} \begin{split} &\|u^l\|_{1}^2+\sum_{j=1}^l\|u^j-u^{j-1}\|_1^2+\Delta t\,\sum_{j=1}^l\|u^j\|^2_2\\ &\qquad\le C\left(\|u_0\|^2_1+\sum_{j=1}^l\left(\Delta t\,\|u^j\|^2_1 +\| \Delta W^j\|_1^2 +\langle \Delta W^j, u^{j-1}\rangle_{1}\right)\right). \end{split} \end{align} Therefore, \begin{equation}\label{eq:iter2} \begin{aligned} &\mathbb{E}\sup_{l\in I_N}\left(\|u^l\|_{1}^2+\sum_{j=1}^l\|u^j-u^{j-1}\|_1^2+\Delta t\, \sum_{j=1}^l\|u^j\|^2_2\right)\\ &\quad\le C\mathbb{E}\|u_0\|^2_1+C\mathbb{E}\sup_{l\in I_N}\left(\sum_{j=1}^l\left(\Delta t\,\|u^j\|^2_1+\|\Delta W^j\|_1^2 +\langle\Delta W^j, u^{j-1}\rangle_{1}\right)\right)\\ &\quad\le C\mathbb{E}\|u_0\|^2_1+C\mathbb{E}\left(\sum_{j=1}^N\big(\Delta t\,\|u^j\|^2_1+\|\Delta W^j\|_1^2\big)\right)+C\mathbb{E}\sup_{l\in I_N}\sum_{j=1}^l\langle\Delta W^j, u^{j-1}\rangle_{1}. \end{aligned} \end{equation} Since $A^{\frac12}\Delta W^j$ is a Gaussian random variable with covariance operator $$\tilde{Q}:=\Delta t\, A^{\frac12}Q^{\frac12}(A^{\frac12}Q^{\frac12})^*,$$ it follows, by \eqref{eq:tr}, that \begin{align*} \mathbb{E}\|\Delta W^j\|_1^2=\Delta t\,\Tr\tilde Q=\Delta t\, \|A^{\frac12}Q^{\frac12}\|^2_{\HS}. \end{align*} Next note that $\sum_{j=1}^l\langle\Delta W^j, u^{j-1}\rangle_{1}$ is an It\^o integral the form $\int_0^{t_l}\langle \eta(t),\dd A^{\frac12}W(t)\rangle$, where $\eta$ is a piecewise continuous process, and hence also a martingale. Then, using H\"older's inequality, the martingale inequality \cite[Theorem 3.8]{DPZ}, It\^o's Isometry \eqref{eq:rito}, \eqref{eq:hs}, \eqref{eq:tr}, and \eqref{eq:iter1}, \begin{align*} &\left(\mathbb{E}\sup_{l\in I_N}\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\right)^2\le \mathbb{E}\sup_{l\in I_N}\left(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\right)^2\\ &\qquad\le 4 \sup_{l\in I_N} \mathbb{E}\left(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\right)^2=4\mathbb{E}\Delta t\,\sum_{j=1}^N\|\tilde{Q}^{\frac12}A^{\frac12}u^{j-1}\|^2\\ &\qquad\le 4\|\tilde{Q}^{\frac12}\|^2\Delta t\, \sum_{j=1}^N\mathbb{E}\|u^{j-1}\|_1^2\le 4 \|\tilde{Q}^{\frac12}\|_{\HS}^2 \Delta t\,\sum_{j=1}^N\mathbb{E}\|u^{j-1}\|_1^2\\ &\qquad\le C\|A^{\frac12}Q^{\frac12}\|_{\HS}^2 (T+\mathbb{E}\|u_0\|^2+T\|Q^{\frac12}\|^2_{\HS}). \end{align*} Therefore, by \eqref{eq:iter2}, using also \eqref{eq:iter1}, we conclude that \begin{align*} \mathbb{E}\sup_{l\in I_N}\left(\|u^l\|_{1}^2+\sum_{j=1}^l\|u^j-u^{j-1}\|_1^2+\Delta t\,\sum_{j=1}^l\|u^j\|^2_2\right) \le C(1+T) \end{align*} and the proof is complete. \end{proof} When proving strong convergence, even without rate, one needs bounds on higher moments of the time discretization. This will be achieved via a discrete Gronwall inequality, resulting in a bound that grows exponentially with time. However, since our approach does not provide rates for the strong error, this is not a major drawback. Note also, that this result is the exact time-discrete analogue of the bounds on the solution from Proposition \ref{prop:eb}. \begin{prop}\label{prop:ulp} Let $p\ge 2$, $I_n=\{1,2,\dots, n\}$, $1\le n\le N$, and $ T=N\Delta t$. If $\|A^{\frac12}Q^{\frac12}\|_{\HS}<\infty$, $\mathbb{E}\|u_0\|_1^p<\infty$, and $T^{p-1}\Delta t\le \frac12$, then $$ \mathbb{E}\sup_{l\in I_n}\|u^l\|^p+\mathbb{E}\sup_{l\in I_n}\|u^l\|^p_1\le C(T,p,u_0). $$ \end{prop} \begin{proof} As noted in the proof of the previous proposition it is enough to bound the second term on the left hand side. We start from \eqref{eq:1norm} and take the $p$th power of both sides for $p\ge 1$ to get \begin{equation*} \begin{aligned} \|u^l\|_{1}^{2p}&\le C\left(\|u_0\|^{2p}_1+\Big(\sum_{j=1}^l\Delta t\, \|u^j\|^2_1\Big)^p\right.\\ &\quad+\left.\Big(\sum_{j=1}^l\| \Delta W^j\|_1^2\Big)^{p} +\Big(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\Big)^p\right)\\ & \le C\left(\|u_0\|^{2p}_1+\Delta t^{p-1}l^{p-1}\Delta t\, \sum_{j=1}^l\|u^j\|^{2p}_1\right.\\ &\quad \left. +l^{p-1}\sum_{j=1}^l\| \Delta W^j\|_1^{2p} +\Big(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\Big)^p\right). \end{aligned} \end{equation*} Therefore, \begin{equation}\label{eq:es1} \begin{aligned} \mathbb{E}\sup_{l\in I_n}\|u^l\|^{2p}_1&\le C \left(\mathbb{E}\|u_0\|^{2p}_1+T^{p-1}\Delta t\,\sum_{j=1}^n\mathbb{E}\sup_{l\in I_j}\|u^l\|^{2p}_1\right.\\ &\quad+ \left. n^{p-1}\sum_{j=1}^n\mathbb{E}\| \Delta W^j\|_1^{2p}+\mathbb{E}\sup_{l\in I_n}\Big(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\Big)^p\right). \end{aligned} \end{equation} Next, we bound the last two terms in \eqref{eq:es1}. We already noted that $A^{\frac12}\Delta W^j$ is a Gaussian random variable with covariance operator $\tilde{Q}=\Delta t\, A^{\frac12}Q^{\frac12}(A^{\frac12}Q^{\frac12})^*$. Hence we use \eqref{eq:tr} and \eqref{eq:es1a} to bound its $2p$-th moment as \begin{equation*} \mathbb{E}\| \Delta W^j\|_1^{2p}\le C_p (\Tr\tilde{Q})^p=C_p\Delta t^p\|A^{\frac12}Q^{\frac12}\|^{2p}_{\HS}. \end{equation*} Therefore, it follows that \begin{equation}\label{eq:es2} n^{p-1}\sum_{j=1}^n\mathbb{E}\| \Delta W^j\|_1^{2p}\le C_pn^{p-1}\Delta t^p\sum_{j=1}^n\|A^{\frac12}Q^{\frac12}\|^{2p}_{\HS}\le C_pT^p\|A^{\frac12}Q^{\frac12}\|^{2p}_{\HS}. \end{equation} For the last term in \eqref{eq:es1} we use the Burkholder-Davies-Gundy inequality \eqref{eq:rbdg}, \eqref{eq:hs}, and \eqref{eq:tr} to conclude that \begin{equation}\label{eq:es3} \begin{aligned} &\mathbb{E}\sup_{l\in I_n}\left(\sum_{j=1}^l\langle \Delta W^j, u^{j-1}\rangle_{1}\right)^p\le C_p\mathbb{E}\left( \Delta t\,\sum_{j=1}^n \|\tilde{Q}^{\frac12}A^{\frac12}u^{j-1}\|^2\right)^{p/2}\\ &\qquad\le C \|\tilde{Q}^{\frac12}\|^p \Delta t^{p/2} n^{p/2-1} \sum_{j=1}^n \mathbb{E}\|u^{j-1}\|_1^{p}\\ &\qquad\le C \|\tilde{Q}^{\frac12}\|^p_{\HS}T^{p/2-1} \Delta t\, \sum_{j=0}^{n-1}\left(\frac12+\frac{1}{2}\mathbb{E}\sup_{l\in I_j}\|u^l\|^{2p}_1\right)\\ &\qquad=C\|A^{\frac12}Q^{\frac12}\|^{p}_{\HS}T^{p/2}+ C\|A^{\frac12}Q^{\frac12}\|^{p}_{\HS}T^{p/2-1}\Delta t\, \sum_{j=0}^{n-1}\left(\mathbb{E}\sup_{l\in I_j}\|u^l\|^{2p}_1\right). \end{aligned} \end{equation} Finally, substituting \eqref{eq:es2} and \eqref{eq:es3} into \eqref{eq:es1} yields the desired bound by using the discrete Gronwall inequality. Before applying the discrete Gronwall inequality we kick back the last term from the sum $T^{p-1}\Delta t\sum_{j=1}^n\mathbb{E}\sup_{l\in I_j}\|u^l\|^{2p}_1$ in \eqref{eq:es1} using the condition $T^{p-1}\Delta t\le \frac12$. \end{proof} \section{The convergence results}\label{sec:mr} We begin by showing a maximal type error estimate for the linear problem. Define the Backward Euler approximation of the stochastic convolution $W_A(t_n)$ by $$ W_A^n:=\sum_{k=1}^nE^{n-k+1}\Delta W^k=\sum_{k=1}^n\int_{t_{k-1}}^{t_k}E^{n-k+1}\,\dd W(s), \ \text{where $E^n=(I+\Delta t\, A)^{-n}$.} $$ The following result has been proved in a larger generality for multiplicative noise in Banach spaces using heavy machinery in the range $0\le \beta<1$. This would be enough for the purposes of the semilinear problem with additive noise. However, it is possible to obtain the range $0\le \beta \le 2$ because the noise is additive and the approximation of the noise is exact at the mesh points. Since this result is interesting on its own, and the proof presented here is rather elementary based on a discrete version of the factorization method, we present the result and the proof for the full range $0\le \beta \le 2$. \begin{prop}\label{prop:wa} Let $\varepsilon\in (0,\frac12)$, $p>\frac{1}{\varepsilon}$, $0\le \beta \le 2$, and $T=N\Delta t$. Then there is $C=C(p,\varepsilon,T)$ such that $$ \left(\mathbb{E}\sup_{t_n\in [0,T]}\|W_A(t_n)-W_A^n\|^p\right)^\frac{1}{p}\le C\Delta t^{\frac{\beta }{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS},\quad t_n=n\Delta t. $$ \end{prop} \begin{proof} Define the deterministic error operator $F_n$ by $F_n=E(t_n)-E^n$. It is well known that the following error estimate holds \begin{equation}\label{eq:dete} \|A^{\frac{\rho}{2}}F_nv\|\le C\Delta t^\frac{\beta}{2} t_n^{-\frac{\beta-\gamma+\rho}{2}}\|A^{\frac{\gamma}{2}}v\|, \quad 0\le \gamma\le \beta+\rho,~\rho,\gamma\ge 0,~\beta\in [0,2]. \end{equation} Next, we consider the decomposition \begin{align*} W_A(t_n)-W_A^n&=\sum_{k=1}^n\int_{t_{k-1}}^{t_k}(E(t_n-\sigma)-E^{n-k+1})\,\dd W(\sigma)\\ &=\sum_{k=1}^n\int_{t_{k-1}}^{t_k}(E(t_n-\sigma)-E(t_{n-k+1}))\,\dd W(\sigma)\\ &\quad+\sum_{k=1}^n\int_{t_{k-1}}^{t_k}(E(t_{n-k+1})-E^{n-k+1})\,\dd W(\sigma)=:e^n_1+e^n_2. \end{align*} To estimate $e_1$ we first write \begin{align*} e^n_1=\sum_{k=1}^n\int_{t_{k-1}}^{t_k}E(t_n-\sigma)(I-E(\sigma-t_{k-1}))\,\dd W(\sigma) =\int_0^{t_n}E(t_n-\sigma)\Psi(\sigma)\,dW(\sigma) \end{align*} with $\Psi(\sigma)=(I-E(\sigma-t_{k-1}))$ for $\sigma\in (t_{k-1},t_k]$. Next we use the factorization method from \cite[Chapter 5]{DPZ} to write \begin{align*} e^n_1&=c_{\alpha}\int_0^{t_n}E(t_n-\sigma)\int_{\sigma}^{t_n}(t_n-s)^{-1+\alpha}(s-\sigma)^{-\alpha}\,\dd s\,\dd W(\sigma)\\ &=c_{\alpha}\int_0^{t_n}(t_n-s)^{-1+\alpha}E(t_n-s)\int_0^s(s-\sigma)\Psi(\sigma)E(s-\sigma)\,\dd W(\sigma)\,\dd s\\ &=c_{\alpha}\int_0^{t_n}(t_n-s)^{-1+\alpha}E(t_n-s)Y(s)\,\dd s, \end{align*} where $\alpha\in (0,\frac12)$, $c_\alpha^{-1}=\int_{\sigma}^t(t-s)^{-1+\alpha}(s-\sigma)^{-\alpha}\,\dd s$ and $$ Y(s)=\int_0^s(s-\sigma)\Psi(\sigma)E(s-\sigma)\,\dd W(\sigma). $$ Therefore, by H\"older's inequality and that $\|E(t)\|\le 1$ for all $t\geq 0$, \begin{align*} \mathbb{E}\sup_{t_n\in [0,T]}\|e_2^n\|^p\le c_\alpha \left(\int_0^T s^{(-1+\alpha)\frac{p}{p-1}}\,\dd s\right)^{p-1}\int_0^T\mathbb{E}\|Y(s)\|^p\,\dd s. \end{align*} The first integral is finite for $p>\frac{1}{\alpha}$. To bound the second integral, notice that $Y(s)$ is a Gaussian random variable for all $s\in [0,T]$ and therefore, we use \eqref{eq:es1a} to bound its $p$-th moment, \eqref{eq:hs}, \eqref{eq:anal0} with $\beta=\frac12 -\varepsilon$, and \eqref{eq:anal} with $\beta=\gamma=0$ and $\rho=\frac{\beta}{2}$, to obtain \begin{align*} \mathbb{E}\|Y(s)\|^p&\le C_p\left(\int_0^s(s-\sigma)^{-2\alpha}\|\Psi(\sigma)E(s-\sigma)Q^{\frac12}\|^2_{\HS}\,\dd \sigma\right)^{\frac{p}{2}}\\ &=C_p\left(\int_0^s(s-\sigma)^{-2\alpha}\|\Psi(\sigma)A^{-\frac{\beta}{2}}A^{\frac12-\varepsilon}E(s-\sigma)A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|^2_{\HS}\,\dd \sigma\right)^{\frac{p}{2}}\\ &\le C_p\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^p\left(\int_0^s(s-\sigma)^{-2\alpha -1+2\varepsilon}\|\Psi(\sigma)A^{-\frac{\beta}{2}}\|^2\,\dd \sigma\right)^{\frac{p}{2}} \\ &\le C\Delta t^{\frac{\beta p}{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^p\left(\int_0^s(s-\sigma)^{-2\alpha -1+2\varepsilon}\,\dd s \right)^{\frac{p}{2}}\\ &\le C_{T,p,\alpha,\varepsilon}\Delta t^{\frac{\beta p}{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^p, \end{align*} provided that $\alpha<\varepsilon$. Given $p>1/\varepsilon$, we thus need to choose $\alpha\in(\frac1p,\varepsilon)$. We conclude $$ \int_0^T\mathbb{E}\|Y(s)\|^p\,\dd s\le TC_{T,p,\alpha,\varepsilon}\Delta t^{\frac{\beta p}{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^p, $$ which proves the bound on $e_1^n$. To bound $e_2^n$ we use a discrete version of the factorization method. First introduce the constants $$ c_{n,k}:=\Big(\Delta t\,\sum_{l=k}^nt_{n-l+1}^{-1+\alpha}t_{l-k+1}^{-\alpha}\Big)^{-1}. $$ It is not difficult to see that $c_{n,k}\le C$ for all $1\le k\le n$. Then we have \begin{align*} e_2^n&=\sum_{k=1}^nE(t_{n-k+1})c_{n,k}\left(\Delta t\,\sum_{l=k}^nt_{n-l+1}^{-1+\alpha}t_{l-k+1}^{-\alpha}\right)\Delta W^k\\ &\quad-\sum_{k=1}^nE^{n-k+1}c_{n,k}\left(\Delta t\,\sum_{l=k}^nt_{n-l+1}^{-1+\alpha}t_{l-k+1}^{-\alpha}\right)\Delta W^k\\ &=\Delta t\, \sum_{l=1}^nt_{n-l+1}^{-1+\alpha}E(t_{n-l})\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha} E(t_{l-k+1})\Delta W^k\\ &\quad - \Delta t\, \sum_{l=1}^nt_{n-l+1}^{-1+\alpha}E^{n-l}\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha} E^{l-k+1}\Delta W^k\\ &=\Delta t\, \sum_{l=1}^nt_{n-l+1}^{-1+\alpha}E(t_{n-l})Y^l-\Delta t\, \sum_{l=1}^nt_{n-l+1}^{-1+\alpha}E^{n-l}\tilde{Y}^l\\ &=\Delta t\, \sum_{l=1}^nt_{n-l+1}^{-1+\alpha}F_{n-l}Y^l+\Delta t\,\sum_{l=1}^n t_{n-l+1}^{-1+\alpha} E^{n-l}(Y^l-\tilde{Y}^l)=:e^n_{21}+e^n_{22}, \end{align*} where $$ Y_l=\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha} E(t_{l-k+1})\Delta W^k, \quad \tilde{Y}_l=\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha} E^{l-k+1}\Delta W^k. $$ Next, we bound $e^n_{21}$, by H\"older's inequality and \eqref{eq:dete} with $\rho=0$ and $\gamma=\beta$, as follows \begin{align*} \mathbb{E}\sup_{t_n\in [0,T]}\|e^n_{21}\|^p &\le \left(\Delta t\, \sum_{l=1}^N\big(t_{l}^{-1+\alpha}\|F_{l}A^{-\frac{\beta}{2}}\|\big)^{\frac{p}{p-1}}\right)^{p-1}\mathbb{E}\Delta t\,\sum_{l=1}^N\|A^{\frac{\beta}{2}}Y^l\|^p\\ &\le C \Delta t^{\frac{\beta p}{2}} \left(\Delta t\, \sum_{l=1}^nt_l^{(-1+\alpha)\frac{p}{p-1}}\right)^{p-1} \mathbb{E}\Delta t\,\sum_{l=1}^N\|A^{\frac{\beta}{2}}Y^l\|^p, \end{align*} where the first sum is finite if $p>\frac{1}{\alpha}$. To estimate the last sum, note that $A^{\beta/2}Y^l$ is a Gaussian random variable and hence, as before, we use \eqref{eq:es1a} to bound its $p$-th moment. Therefore, using also \eqref{eq:hs} and \eqref{eq:anal0}, \begin{align*} &\mathbb{E}\Delta t\,\sum_{l=1}^N\|A^{\frac{\beta}{2}}Y^l\|^p=\Delta t\,\sum_{l=1}^N\mathbb{E}\left\|\left(\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha} A^{\frac{\beta}{2}}E(t_{l-k+1})\Delta W^k\right)\right\|^p\\ &\qquad =\Delta t\,\sum_{l=1}^N \left(\Delta t\,\sum_{k=1}^l c^2_{n,k} t_{l-k+1}^{-2\alpha} \|A^{\frac{\beta}{2}}E(t_{l-k+1})Q^{\frac12}\|_{\HS}^2\right)^{\frac{p}{2}} \\ &\qquad \le C \Delta t\,\sum_{l=1}^N \left(\Delta t\,\sum_{k=1}^N t_{k}^{-2\alpha} \|E(t_{k})A^{\frac12-\varepsilon}A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^2\right)^{\frac{p}{2}}\\ &\qquad \le CT \|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|^p_{\HS}\left(\Delta t\, \sum_{k=1}^N t_{k}^{-1-2\alpha+2\varepsilon}\right)^{\frac{p}{2}}\le C_{T,p,\alpha,\varepsilon} \|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|^p_{\HS}, \end{align*} provided that $\alpha<\varepsilon$. Finally, we estimate $e_{22}^n$. By H\"older's inequality we first get \begin{align*} &\mathbb{E}\sup_{t_n\in [0,T]}\|e_{22}^n\|^p\le \left(\Delta t\,\sum_{l=1}^N t_{l}^{-1+\alpha} \|E^{l}\|^{\frac{p}{p-1}}\right)^{p-1}\mathbb{E} \sum_{l=1}^N\|Y(l)-\tilde{Y}(l)\|^p\\ &\qquad \le \left(\Delta t\,\sum_{l=1}^N t_{l}^{(-1+\alpha)\frac{p}{p-1}}\right)^{p-1}\mathbb{E} \sum_{l=1}^N\|Y(l)-\tilde{Y}(l)\|^p\le C_{\alpha,p}\sum_{l=1}^N\|Y(l)-\tilde{Y}(l)\|^p, \end{align*} if $p>\frac{1}{\alpha}$. To estimate the last term, we use \eqref{eq:es1a} to bound the $p$-th moment of a Gaussian random variable and also \eqref{eq:hs} and \eqref{eq:dete} with $\rho=1-2\varepsilon$ and $\gamma=\beta$ to get \begin{align*} &\sum_{l=1}^N\|Y(l)-\tilde{Y}(l)\|^p=\sum_{l=1}^N\left\|\sum_{k=1}^l c_{n,k} t_{l-k+1}^{-\alpha}F_{l-k+1}\Delta W^k\right\|^{p}\\ &\qquad\le C_p \sum_{l=1}^N \left(\Delta t\, \sum_{k=1}^N t_k^{-2\alpha} \|F_{k}Q^{\frac12}\|_{\HS}^2\right)^{\frac{p}{2}}\\ &\qquad=C_p \sum_{l=1}^N \left(\Delta t\, \sum_{k=1}^N t_k^{-2\alpha} \|A^{\frac{1}{2}-\varepsilon}F_{k}A^{-\frac{\beta}{2}}A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|_{\HS}^2\right)^{\frac{p}{2}}\\ &\qquad \le C_p \Delta t^{\frac{\beta p}{2}} \left(\Delta t\, \sum_{k=1}^N t_k^{-1-2\alpha+2\varepsilon}\right)^{\frac{p}{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|^p_{\HS}\\ &\qquad \le C_{T,p,\alpha,\varepsilon}\Delta t^{\frac{\beta p}{2}}\|A^{\frac{\beta-1}{2}+\varepsilon}Q^{\frac12}\|^p_{\HS}, \end{align*} whenever $\alpha<\varepsilon$, which finishes the proof. \end{proof} Next we state a Lipschitz estimate for $f(u)$. Here we use Sobolev's inequality and, similarly to \eqref{eq:lg}, it is crucial that $d\le 3$ and that the nonlinearity $f$ is at most cubic. For a proof we refer to \cite[Lemma 2.5]{KML}. \begin{lem}\label{lem:lip} For all $u,v\in \dot{H}^1$ we have $$ \|A^{-\frac{1}{2}}(f(u)-f(v))\|\le C(\|u\|^2_1+\|v\|^2_1)\|u-v\|. $$ \end{lem} We are now ready to state and prove the pathwise convergence of the Backward Euler scheme defined in \eqref{eq:be}. \begin{thm}\label{thm:pwc} Let $\varepsilon>0$, $\|A^{\frac12+\varepsilon} Q^{\frac12}\|_{\HS}<\infty$, $\mathbb{E}\|u_0\|_1^2<\infty$, $0\le \gamma <\frac12$, and $T=N\Delta t$. Then, there are finite random variables $K\ge 0$ and $\Delta t_0>0$ such that, almost surely, $$ \sup_{t_n\in [0,T]}\|u(t_n)-u^n\|\le K \Delta t^\gamma,\quad t_n=n\Delta t,\quad \Delta t\le \Delta t_0. $$ \end{thm} \begin{proof} Since the arguments are pathwise and hence basically deterministic, we omit standard details. Let $e^n=u(t_n)-u^n$ and $0\le \gamma <\frac12$. We decompose the error, using the mild formulation of \eqref{eq:be} and \eqref{eq:mild}, as follows \begin{align*} e^n&=(E(t_n)u_0-E^n u_0) +(W_A(t_n)-W_A^n)\\ &\quad +\sum_{k=1}^n \int_{t_{k-1}}^{t_k} E(t_n-s)f(u(s))-E^{n-k+1}f(u^{k})\,\dd s =:e^n_1+e^n_2+e^n_3. \end{align*} By \eqref{eq:dete} we may estimate $e_1$ as $$ \|e^n_1\|\le C \Delta t^{\frac12}\|u_0\|_1. $$ For $e_2^n$, by Proposition \ref{prop:wa} with $\beta=2$, we have that $$ \|e^n_2\|\le L \Delta t\, \|A^{\frac12+\varepsilon} Q^{\frac12}\|_{\HS} $$ almost surely for some finite nonnegative random variable $L$. Next, we can further decompose $e_3$ as \begin{align*} e^n_3&=\sum_{k=1}^n\int_{t_{k-1}}^{t_k}E^{n-k+1}(f(u(t_{k}))-f(u^{k}))\,\dd s\\ &\quad +\sum_{k=1}^n\int_{t_{k-1}}^{t_k}(E(t_{n-k+1})-E^{n-k+1})f(u(t_{k}))\,\dd s\\ &\quad +\sum_{k=1}^n\int_{t_{k-1}}^{t_k}E(t_{n-k+1})(f(u(s))-f(u(t_{k})))\,\dd s\\ &\quad +\sum_{k=1}^n\int_{t_{k-1}}^{t_k}(E(t_n-s)-E(t_{n-k+1}))f(u(s))\,\dd s =:e_{31}^n+e_{32}^n+e_{33}^n+e_{34}^n. \end{align*} To bound $e_{31}^n$ we use Propositions \ref{prop:eb} and \ref{prop:ul} together with Lemma \ref{lem:lip} to conclude that for some finite nonnegative random variable $L_1$ we have, almost surely, \begin{align*} \|e_{31}^n\| &=\left\|\sum_{k=1}^n\int_{t_{k-1}}^{t_k}A^{\frac{1}{2}}E^{n-k+1}A^{-\frac12}(f(u(t_{k}))-f(u^{k}))\,\dd s\right\|\\ &\le L_1\sum _{k=1}^n\int_{t_{k-1}}^{t_k} t_{k}^{-\frac 12} \|e^k\|\,\dd s =L_1 \Delta t\, \sum _{k=1}^n t_{k}^{-\frac 12} \|e^k\|, \end{align*} where we used the well known fact that $\|A^{1/2}E^k\|\le Ct_k^{-\frac12}$ (see, for example, \cite[Lemma 7.3]{Thomeebook}). Next we use Proposition \ref{prop:eb}, Lemma \ref{lem:lip}, and \eqref{eq:dete} with $\gamma=0$, $\rho=1$ and $\beta=2\gamma$ to estimate $e_{32}^n$ as \begin{align*} \|e_{32}^n\| &=\left\|\sum_{k=1}^n\int_{t_{k-1}}^{t_k}A^{\frac12}(E(t_{n-k+1})-E^{n-k+1})A^{-\frac12}f(u(t_{k}))\,\dd s\right\|\\ &\le\Delta t^{\gamma} L_2 \sum_{k=1}^n\int_{t_{k-1}}^{t_k}t_k^{-\frac12-\gamma}\,\dd s=\Delta t^{\gamma} L_2 \Delta t\,\sum_{k=1}^n t_k^{-\frac12-\gamma}, \end{align*} almost surely for some finite nonnegative random variable $L_2$. For $e_{33}^n$ we use the H\"older continuity of $u$ from Proposition \ref{propo:ulip} together with Proposition \ref{prop:eb}, Lemma \ref{lem:lip}, and \eqref{eq:anal0} with $\beta=\frac12$, and obtain \begin{align*} \|e_{33}^n\| &=\left\|\sum_{k=1}^n\int_{t_{k-1}}^{t_k}A^{\frac12}E(t_{n-k+1})A^{-\frac12}(f(u(s))-f(u(t_{k})))\,\dd s\right\|\\ &\le \Delta t^{\gamma} L_3\sum_{k=1}^n\int_{t_{k-1}}^{t_k}t_k^{-\frac12}\,\dd s=\Delta t^{\gamma} L_3 \Delta t\, \sum_{k=1}^nt_k^{-\frac12}, \end{align*} almost surely for some finite nonnegative random variable $L_3$. Finally, by Proposition \ref{prop:eb}, Lemma \ref{lem:lip} and \eqref{eq:anal} with $\beta=\frac12$ and $\rho=0$, we have \begin{align*} \|e_{34}\| &=\left\|\sum_{k=1}^n\int_{t_{k-1}}^{t_k}A^{\frac12}(E(t_n-s)-E(t_{n-k+1}))A^{-\frac12}f(u(s))\,\dd s\right\|\\ &\le \Delta t^{\gamma} L_4 \sum_{k=1}^n\int_{t_{k-1}}^{t_k}t_k^{-\frac12-\gamma}\,\dd s=\Delta t^{\gamma} L_4 \Delta t\,\sum_{k=1}^n t_k^{-\frac12-\gamma}, \end{align*} almost surely for some finite nonnegative random variable $L_2$. Putting together the estimates and using a generalized discrete Gronwall lemma \cite[Lemma 7.1]{Stig} finishes the proof. \end{proof} Finally, we show strong convergence in $L^p$, albeit without rate. \begin{thm}\label{thm:strc} Let $\varepsilon>0$, $p\ge 1$, and $N\Delta t=T$. If $\|A^{\frac12+\varepsilon} Q^{\frac12}\|_{\HS}<\infty$ and $\mathbb{E}\|u_0\|_1^{2p}<\infty$, then \begin{align*} \lim_{\Delta t\to 0}\mathbb{E}\sup_{t_{n}\in [0,T]}\|u(t_n)-u^n\|^p= 0,\quad t_n=n\Delta t. \end{align*} \end{thm} \begin{proof} Let \begin{align*} Y_N:=\sup_{t_{n}\in [0,T]}\|u(t_n)-u^n\|^p. \end{align*} By Theorem \ref{thm:pwc} it follows that $Y_N\to 0$ almost surely, and hence in probability, as $N\to \infty$. By Propositions \ref{prop:eb} and \ref{prop:ulp}, there is $M>0$ such that, for $T^{2p-1}\Delta t\le \frac12$, \begin{align*} \mathbb{E}\, Y_N^{2}\le C\mathbb{E}\sup_{t_{n}\in [0,T]}\left(\|u(t_n)\|^{2p}+\|u^n\|^{2p}\right)\le M. \end{align*} Therefore, it follows that $\{Y_N\}_{N\in \mathbb{N}}$ is uniformly integrable. Being convergent in probability and uniformly integrable, it converges in $L^1$; that is, \begin{align*} \lim_{\Delta t\to 0}\mathbb{E}\sup_{t_{n}\in [0,T]}\|u(t_n)-u^n\|^p=\lim_{N\to \infty}\mathbb{E}\,Y_N= 0, \end{align*} see \cite[Proposition 3.12]{Kal}. \end{proof} \def\cprime{$'$} \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace} \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR } \providecommand{\MRhref}[2] \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2} } \providecommand{\href}[2]{#2}
\section{Introduction} Let $G=(V,E)$ be a simple graph with $n=|V|$ vertices and $n$ edges and $L(G)=D(G)-A(G)$ be its {\it Laplacian matrix,} where $A(G)$ and $D(G)$ are its adjacency and degree diagonal matrices, respectively. The {\it Laplacian polynomial} $\mathcal{L}(G,\lambda)$ of $G$ is the characteristic polynomial of its Laplacian matrix $L(G)$, i.e., $$\mathcal{L}(G,\lambda)=det(\lambda I_{n}-L(G))=\sum_{k=0}^n (-1)^{k}c_{k}(G)\lambda^{n-k}.$$ Then $L(G)$ has nonnegative eigenvalues $\mu_{1}\geq\mu_{2}\geq\cdots\geq\mu_{n-1}\geq\mu_{n}=0$. From Viette's formula, $c_{k}=\sigma_{k}(\mu_{1},\mu_{2}\cdots\mu_{n-1})$ is a symmetric polynomial of order $n-1$. In particular, we have $c_{0}=1,c_{n}=0,c_{1}=2|E(G)|,c_{n-1}=n\tau(G)$, where $\tau(G)$ is the number of spanning trees of $G$ (see \cite{Merris1995}). If $G$ is a tree, the Laplacian coefficient $c_{n-2}$ is equal to its Wiener index, which is the sum of all distances between unordered pairs of vertices of $G$ (see \cite{Dobrynin2001}, \cite{Yan2006}). $$c_{n-2}(T)=W(T)=\sum_{u,v\in V} d(u,v).$$ In general, Laplacian coefficients $c_{k}$ can be expressed in terms of subtree structures of $G$. \begin{theorem}\cite{Kelmans1974}\label{theorem1.1} Let $\mathcal {F}_{k}$ be the set of all spanning forests of $G$ with exactly $k$ components. The Laplacian coefficient $c_{n-k}$ of a graph $G$ is expressed by $c_{n-k}(G)=\sum_{F\in \mathcal {F} _{k}} \gamma(F)$, where $F$ has $k$ components $T_{i}$ with $n_{i}$ vertices, $i=1,2,\cdots,k$ and $\gamma(F)=\prod_{i=1}^k n_{i}$. \end{theorem} Recently, the study on the Laplacian coefficients has attracted much attention. Let $G$ and $H$ be two graphs of order $n$. We write $G\preceq H$ if $c_{k}(G)\leq c_{k}(H)$ for all $0\leq k\leq n$, and write $G\prec H$ if $G\preceq H$ and $c_{k}(G)<c_{k}(H)$ for some $k\in \{0,1,\cdots,n\}$. Mohar \cite{Mohar2007} first investigated properties of the poset (partially ordered set) of acyclic graphs with the partial order $\preceq$ and proposed some problems. Later, Ili\'{c}\cite{Ilic2009} and Zhang et al. \cite{Zhang2009} investigated ordering trees by the Laplacian coefficients. Stevanovi\'{c} and Ili\'{c} \cite{Stevanovic2009b} investigated and characterized the minimum and maximum elements in the poset of unicyclic graphs of order $n$ with $\preceq$. Tan \cite{Tan2011} proved that the poset of unicyclic graphs of order $n$ and fixed matching number with $\preceq$ has only one minimal element. He and Shan \cite{He2010} studied the properties of the poset of bicyclic graphs of order $n$. {\it The Laplacian-like energy} \cite{Liu2008} of $G$ , $LEL$ for short, is defined as follows: $$LEL(G)=\sum_{k=1}^{n-1} \sqrt{\mu_{k}},$$ since it has similar features as molecular graph energy defined by Gutman \cite{Gutman1978}. $LEL$ describes well the properties which have a close relation with the molecular structures and was proved to be as good as the Randi\'{c} index, better than Wiener index in some areas (see \cite{Stevanovic2009c}). Further, Stevanovi\'{c} in \cite{Stevanovic2009} established a connection between $LEL$ and Laplacian coefficients. \begin{theorem}\cite{Stevanovic2009c} \label{theorem1.2} Let $G$ and $H$ be two graphs with $n$ vertices. If $G\preceq H$, then $LEL(G)\leq LEL(H)$. Furthermore, if $G\prec H$, then $LEL(G)<LEL(H)$. \end{theorem} Let \begin{eqnarray*} \mathcal{U}_{n,l}&=& \{G\ |\ G \mbox{ is a $n$-vertex unicyclic graph with fixed $l$ leaves }\},\\ \mathcal{U}_{n,l}^g&=& \{G\ |\ G\in \mathcal{U}_{n,l} \mbox{ with fixed girth $g$ }\}. \end{eqnarray*} Let $BST_{n,l}$ be a balanced starlike tree of order $n$ with $l$ leaves which is obtained by identifying one end of each of the $l$ paths of orders $\lfloor\tfrac{n-1}{l}\rfloor+1$ or $\lceil\tfrac{n-1}{l}\rceil+1$. Moreover, let $U_{n,l}^{g,p}$ (see Fig.1) be a balanced starlike unicyclic graph of order $n$ with $l$ leaves and girth $g$, which is obtained by identifying one end of a path $P_{p+1}$ of order $p+1$ and one vertex of a cycle of order $g$, the other end of $P_{p+1}$ and the center vertex of a balanced starlike tree $BST_{n-p-g+1, l}$, respectively. \begin{figure} \centering \begin{tikzpicture} \node(v900)[label=0:$G$]at (-1.3,0){}; \node (v0)[draw,shape=circle,inner sep=1.5pt,label=87:$u_1$] at (0:0){}; \node (v1)[draw,shape=circle,inner sep=1.5pt,label=90:$u_{2}$] at(-60:-1){}; \node (v2)[draw,shape=circle,inner sep=1.5pt,label=below:$u_{g}$] at(60:-1){}; \node(v3) [draw,shape=circle,inner sep=1.5pt,label=90:$w_{1}$] at (1,0){}; \node(v4) [draw,shape=circle,inner sep=1.5pt,label=90:$w_{p-1}$] at (2,0){}; \node(v5) [draw,shape=circle,inner sep=1.5pt,label=90:$w_{p}$] at (3,0){}; \node(v6) [draw,shape=circle,inner sep=1.5pt] at (4,1){}; \node(v7) [draw,shape=circle,inner sep=1.5pt,label=-90:$v_1$]at (4,-1){}; \node(v8) [draw,shape=circle,inner sep=1.5pt] at (5,1){}; \node(v9)[draw,shape=circle,inner sep=1.5pt,label=-90:$v_t$] at (5,-1){}; \draw (-1,0) circle(1); \draw [loosely dotted,thick] (v8) --(v6) (4,0)--(5,0) (v7)--(v9) (v3)--(v4); \draw (v0)--(v3) (v4)--(v5) (v5)--(v6) (v5)--(v7); \end{tikzpicture} \caption{Graph $U_{n,l}^{g,p}$} \label{fig:pepper} \end{figure} Ili\'{c} and Ili\'{c} in \cite{Ilic2009b} proposed the following conjecture: \begin{conjecture}\cite{Ilic2009b}\label{conjecture} Among all $n$-vertex unicyclic graphs, the graph $U_{n,l}^{3,0}$ has the minimum Laplacian coefficients $c_k,$ $k=0, \dots,n,$ i.e., $U_{n,l}^{3,0}$ is the only one minimal element in the poset $(\mathcal{U}_{n,l}, \preceq)$. \end{conjecture} However, this conjecture is, in general, not true. Let $G_1$ and $G_2 $ be the following two unicyclic graphs of orders 10 (see Fig. 2). \begin{figure} \centering \begin{tikzpicture}[scale=1] \tikzstyle{every node}=[draw,shape=circle,inner sep=1.5pt]; \node (v0) at (0:0){}; \node(v1) at (0:1){}; \node(v2) at(60:1){}; \node(v3) at (2,1){}; \node(v4) at (2,-1){}; \node(v5) at (3,1){}; \node(v6) [draw,shape=circle,inner sep=1.5pt,label=below:$G_1$]at (3,-1){}; \node(v7) at (4,1){}; \node(v8) at (4,-1){}; \node(v10) at (5,-1){}; \draw (v0) --(v1) (v1) --(v2) (v2)--(v0) (v1)--(v4) (v1)--(v3) (v8)--(v10) (v5)--(v7) (v4)--(v6) (v6)--(v8) (v5)--(v3); \end{tikzpicture} \begin{tikzpicture}[scale=1] \tikzstyle{every node}=[draw,shape=circle,inner sep=1.5pt]; \node (v0) at (0:0){}; \node(v1) at (0:1){}; \node(v2) at(60:1){}; \node(v3) at (2,0){}; \node(v4) at (3,1){}; \node(v5) at (3,-1){}; \node(v6) at (4,1){}; \node(v7) [draw,shape=circle,inner sep=1.5pt,label=below:$G_2$] at(4,-1){}; \node(v8) at (5,1){}; \node(v9) at (5,-1){}; \draw (v0) --(v1) (v1) --(v2) (v2)--(v0) (v1)--(v3) (v3)--(v5) (v3)--(v4) (v7)--(v9) (v5)--(v7) (v4)--(v6) (v6)--(v8); \end{tikzpicture} \caption{Counter-example} \label{fig:pepper} \end{figure} Then their Laplacian characteristic polynomials are $$\mathcal{L}(G_1,x)=x^{10}-20x^9+167x^8-758x^7+2036x^6-3296x^5+3130x^4-1612x^3+382x^2-30x,$$ $$\mathcal{L}(G_2,x)=x^{10}-20x^9+168x^8-770x^7+2091x^6-3414x^5+3243x^4-1642x^3+373x^2-30x.$$ Hence $G_1$ does not have minimal Laplacian coefficients $c_k,$ $k=0, \dots,10.$ So this conjecture is, in general, not true. But it is proven that $G_1$ is still a minimal element in the poset $\mathcal{U}_{10,2}$. In fact, there are many minimal elements in the poset $\mathcal{U}_{n,l}$. This paper is organized as follows: in section 2, we investigate some properties of minimal elements in the poset $\mathcal{U}_{n,l}^g$. In sections 3 and 4, all minimal elements in four special posets of $\mathcal{U}_{n,l}^g$ are characterized, respectively. Finally, in section 5, we conclude this paper with some conjectures. \section{The minimal elements in $(\mathcal{U}_{n,l}^g, \preceq)$} Let $v$ be a vertex of a connected graph $G$ and let $N_{G}(v)$ denote the set of the neighbors of $v$ in $G$. Let $d_{G}(v)$ denote the degree of $v$ in $G$, if $d_{G}(v)=1$, $v$ is called a leaf or a pendent vertex. Say that $P=vv_{1}\cdots v_{k}$ is a pendant path of length $k$ attached at vertex $v$ if its interval vertices $v_1\cdots v_{k-1}$ have degree two and $v_k$ is a leaf. If $k=1$, then $v_1$ is a {\it leaf} and $vv_1$ is called a {\it pendent edge}. A {\it branch vertex} is a vertex having degree more than two. Moreover, let $d(u,v)$ denote the distance between vertices $u$ and $v$. For a $n$-vertex unicyclic graph $G\in \mathcal{U}_{n,l}^g$, $G$ can be obtained from a cycle $C_g=u_{1}\cdots u_{g}$ of order $g$ by attaching trees $T_1\cdots T_g$ rooted at $u_{1}, \cdots, u_{g}$, respectively. So $G$ may be written to be $C_{T_1, \cdots, T_g}$. \begin{lemma}\label{theorem2.1} (\cite{Ilic2010}). Let $v$ be a vertex of a nontrivial connected graph $G$ and for nonnegative integers $p$ and $q$, let $G(p,q)$ denote the graph obtained from $G$ by adding two pendent paths of lengths $p$ and $q$ at $v$, respectively, $p \geq q \geq 1$. Then $c_k(G(p,q))\leq c_k(G(p+1,q-1)), k=0,1,\cdots,n$. \end{lemma} \begin{definition}\label{definition 2.2} Let $G$ be an arbitrary connected graph with $n$ vertices. If $uv$ is a non-pendent cut edge of $G$ with $d_{G}(v)\geq 3$ and $d_{G}(u)\geq 2$ such that there's at least one pendent path $P_{t+1}=vv_{1}\cdots v_{t}$ attached at $v$, then the graph $G^\prime=\xi(G,uv)$ obtained from $G$ by changing all edges (except $uv, vv_1$) incident with $v$ into new edges between $u$ and $N_G(v)\setminus \{u,v_1\}$. In other words, $$G^\prime=G-\{vx| x\in N_G(v)\setminus \{u,v_1\}\}+\{ux| x\in N_G(v)\setminus \{u,v_1\}\}.$$ We say that $G^\prime$ is a $\xi$-transformation of $G$. (See Fig. 3). \end{definition} \begin{figure} \centering \begin{tikzpicture} \node(v900)[label=0:$H$]at (-1.3,0){}; \node (v0)[draw,shape=circle,inner sep=1.5pt,label=45:$u$] at (0:0){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v$](v1) at (1,0){}; \node[draw,shape=circle,inner sep=1.5pt](v2) at (2,1){}; \node[draw,shape=circle,inner sep=1.5pt](v3) at (2,0){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_1$](v4) at (2,-1){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_2$](v5) at (3,-1){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_t$](v6) at (4,-1){}; \draw [loosely dotted,thick] (v2) -- (v3) (v5) --(v6); \draw (v0)--(v1)(v1)--(v2) (v1)--(v3)(v1)--(v4)(v4)--(v5); \draw (2,0.5) ellipse (0.6 and 1); \draw (-1,0) circle(1); \node(v900)[label=0:$H$]at (6,0){}; \node (v0)[draw,shape=circle,inner sep=1.5pt,label=180:$u$] at (7.5,0){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v$](v1) at (8.5,-1){}; \node[draw,shape=circle,inner sep=1.5pt](v2) at (8.5,1){}; \node[draw,shape=circle,inner sep=1.5pt](v3) at (8.5,0){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_1$](v4) at (9.5,-1){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_2$](v5) at (10.5,-1){}; \node[draw,shape=circle,inner sep=1.5pt,label=below:$v_t$](v6) at (11.5,-1){}; \draw [loosely dotted,thick] (v2) -- (v3) (v5) --(v6); \draw (v0)--(v2)(v0)--(v3) (v0)--(v1)(v1)--(v4)(v4)--(v5); \draw (8.5,0.5) ellipse (0.6 and 1); \draw (6.5,0) circle(1); \draw[->] (4,0)--(5,0); \node(v900)[label=0:$G$]at (1,-2){}; \node(v900)[label=0:$G'$]at (8,-2){}; \end{tikzpicture} \caption{$\xi$-transformation } \label{fig:pepper} \end{figure} Clearly, $\xi$-transformation preserves the number of leaves in $G$. \begin{lemma}\label{lemma2.3} Let $G$ be a connected graph of order $n$ and $G^\prime$ be obtained from $G$ by $\xi$-transformation. If there exists a path $P_{s+1}=uu_1^p\cdots u_{s-1}^pu_s^p$ of order $s+1$ in the component of $G-uv$ and a pendent path $P_{t+1}=vv_{1}\cdots v_{t}$ attached at $v$ with $s\ge t$, then $G^\prime\preceq G$, i.e, $$c_k(G)\geq c_k(G^{\prime}), k=0, 1,\dots, n,$$ with equality if and only if $k \in \{0,1,n-1,n\}$. \end{lemma} \begin{proof} Clearly, if $k\in \{0,1,n\}$, $c_k(G)=c_k(G^\prime)$. Since $uv$ is a cut edge, then every spanning tree of $G$ and $G^\prime$ includes edge $uv$, which implies $\tau(G)=\tau(G^\prime)$. Hence $c_{n-1}(G)=c_{n-1}(G^\prime)$. Now assume that $2 \leq k \leq n-2$ and consider the coefficients $c_{n-k}(G)$. Let $\mathcal {F}^\prime$ (resp. $\mathcal {F}$) be the set of all spanning forests of $G'$ (resp. $G$) with exactly $k$ components. For an arbitrary spanning forest $F' \in \mathcal {F}'$, denote by $T'$ the component of $F'$ containing $u$. Let $f: \mathcal{F}^\prime\rightarrow \mathcal{F}, F=f(F^\prime)$, where $V(F)=V(F^\prime)$ and $$E(F)=E(F')-\{ux| x\in N_{T^\prime}(u)\cap N_{G}(v)\}+\{vx| x\in N_{T^\prime}(u)\cap N_{G}(v)\}.$$ Then $f$ is injective. Let $\mathcal{F}^\prime =\mathcal{F}_{(1)}^\prime \bigcup \mathcal{F}_{(2)}^\prime $, where $\mathcal{F}_{(1)}^\prime=\{F'\in \mathcal{F}^\prime \ |\ uv\in E(F')\} $ and $\mathcal{F}_{(2)}^\prime=\{F'\in \mathcal{F}^\prime \ |\ uv\notin E(F')\} $. If $F^\prime\in \mathcal{F}_{(1)}^\prime,$ then $F'$ and $F=f(F')$ have the same components except $T^\prime$. Moreover, $T'$ and $f(T^{\prime})$ have the same vertices. Hence $\gamma(F)=\gamma(F')$. If $F^\prime\in \mathcal{F}_{(2)}^\prime,$ let $S^\prime$ be the component of $F'$ containing $v$. Then $F^\prime$ and $F=f(F^\prime)$ have the same components except $T^\prime$ and $S^\prime$ in $F^\prime$. Assume $T^\prime$ contains $a$ vertices in the component of $G-uv$ containing $v$, $|V(T^\prime)|-a$ vertices in the component of $G-uv$ containing $u$. Then $F$ has two components $f(T^{\prime})=T$ with $|V(T^\prime)|-a$ vertices and $f(S^{\prime})=S$ with $a+|V(S^\prime)|$ vertices corresponding to $T^\prime$ and $S'$, respectively. Denote by $N$ the product of the orders of all components of $F^\prime$ except $T^\prime$ and $S^\prime$. Then \begin{align*} \gamma(f(F^{'}))-\gamma(F^{'}) =&[(|V(T^\prime)|-a)(a+|V(S^\prime)|)-|V(T^\prime)|\cdot|V(S^\prime)|]N\\ =&(|V(T^\prime)|-a-|V(S^\prime)|)\cdot a\cdot N. \end{align*} Further let $\mathcal{F}_{(2)}^\prime= \mathcal{F}_{20}^\prime\cup\mathcal{F}_{21}^\prime\cup\mathcal{F}_{22}^\prime$, where \begin{eqnarray*} \mathcal{F}_{20}^\prime&=&\{F^\prime\in \mathcal{F}_{(2)}^\prime\ | \ |V(T^\prime)|-a=|V(S^\prime)| \mbox{ or } a=0\},\\ \mathcal{F}_{21}^\prime&=& \{F^\prime\in \mathcal{F}_{(2)}^\prime\ | \ |V(T^\prime)|-a<|V(S^\prime)|, a>0\},\\ \mathcal{F}_{22}^\prime&=&\{F^\prime\in \mathcal{F}_{(2)}^\prime\ |\ |V(T^\prime)|-a>|V(S^\prime)|, a>0\}. \end{eqnarray*} Hence it follows that $$\forall F^\prime\in \mathcal{F}_{20}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})=0,$$ $$\forall F^\prime\in \mathcal{F}_{21}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})<0,$$ $$\forall F^\prime\in \mathcal{F}_{22}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})>0.$$ For every spanning forest $F_1^\prime\in \mathcal{F}_{21}^\prime$, let $T^\prime$ and $S^\prime$ be two components of $F_1^\prime$ containing $u$ and $ v$, respectively. Assume that $ u, u_1^p, \cdots, u_{r-1}^p\in V(T')$ and $u_r^p\notin V(T')$. Moreover, let $R^\prime$ be a component of $F_1^\prime$ containing $u_r^p$ with $b$ vertices. Thus let $T^{\prime\prime}$ be a tree obtained from $T^\prime$ and $R^\prime$ by joining $u_{r-1}^p$ and $u_r^p$ with the edge $u_{r-1}^pu_r^p$, $S^{\prime\prime}$ be the path $vv_1\cdots v_{|V(T^\prime)|-a-1}$ and $R^{\prime\prime}$ be the path $v_{|V(T^\prime)|-a}\cdots v_{|V(S^\prime)|-1}$. Then $F_2^\prime=(F_1^\prime-\{T^\prime, S^\prime, R^\prime\})\cup \{T^{\prime\prime}, S^{\prime\prime}, R^{\prime\prime}\}$ is a spanning forest of $G^\prime$ with exactly $k$ components and $F_2^\prime\in \mathcal{F}_{22}^\prime$. Hence there exists an injective (not bijective) map from $\mathcal{F}_{21}^\prime$ to $ \mathcal{F}_{22}^\prime$, i.e., $$\varphi: \mathcal{F}_{21}^\prime\rightarrow \mathcal{F}_{22}^\prime: F_1^{\prime}\rightarrow F_2^{\prime}=\varphi(F_1^\prime),$$ where $F_2^{\prime}=\varphi(F_1^\prime)=(F_1^\prime-\{T^\prime, S^\prime, R^\prime\})\cup \{T^{\prime\prime}, S^{\prime\prime}, R^{\prime\prime}\}$. Note that $|V(T^{\prime\prime})|=|V(T^\prime)|+b,$ $|V(S^{\prime\prime})|=|V(T^\prime)|-a,$ and $ |V(R^{\prime\prime})|= |V(S^\prime)|-|V(T^\prime)|+a$. It is easy to see that for $F^\prime\in\mathcal{F}_{21}^\prime$, \begin{eqnarray*} &&\gamma(f(\varphi(F^\prime)))-\gamma(\varphi(F^\prime))\\ &=& [(|V(T^\prime)|+b-a)\cdot (|V(T^\prime)|-a+a)-(|V(T^\prime)|+b)\cdot (|V(T^\prime)|-a)]\\ && \cdot (|V(S^\prime)|-|V(T^\prime)|+a)\cdot \frac{N}{b} \\ &=& -(|V(T^\prime)|-a-|V(S^\prime)|)\cdot a N \\ &=&-(\gamma(f(F^\prime))-\gamma(F^\prime)) \end{eqnarray*} Therefore, \begin{align*} &\sum_{F^{\prime}\in \mathcal {F}_{21}^{\prime} \cup \mathcal {F}_{22}^\prime} [ \gamma(f(F^{\prime}))-\gamma(F^{\prime})]\\ =&\sum_{F^{\prime}\in \mathcal {F}_{21}^{\prime}} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})+\gamma(f(\varphi(F^\prime)))-\gamma(\varphi(F^\prime))]+\sum_{F^\prime\in \mathcal {F}_{22}^{\prime}\setminus \varphi(\mathcal {F}_{21}^{\prime})} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})]\\ =&\sum_{F^\prime\in \mathcal {F}_{22}^{\prime}\setminus \varphi(\mathcal {F}_{21}^{\prime})} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})]>0. \end{align*} It follows from Theorem~\ref{theorem1.1} that $$c_{n-k}(G^{\prime})=\sum_{F^{\prime}\in \mathcal {F}^{\prime}}\gamma(F^{\prime}) < \sum_{F^{\prime}\in \mathcal {F}^{\prime}}\gamma(f(F^{\prime}))\leq \sum_{F\in \mathcal {F}}\gamma(F)=c_{n-k}(G).$$ So the assertion holds.\end{proof} Now we are ready to present the main result in this section. \begin{theorem}\label{theorem2.5} If $C_{T_1,\cdots,T_g}\in \mathcal{U}_{n,l}^g$ is obtained from the cycle $C_g=u_1\cdots u_g$ by attaching $g$ trees $T_1, \cdots, T_g$ at the roots $u_1, \cdots, u_g$, respectively, where $|V(T_i)|=n_i$ and the number of leaves in $T_i$ is $l_i$, then $$C_{B_1,\cdots,B_g}\preceq C_{T_1,\cdots,T_g},$$ where $B_i$ is a tree with root $u_i$ obtained by identifying one end $u_{i,p_i}$ of the path $P_{p_i}: u_iu_{i,2}\cdots u_{i,p_i}$ and the center of $BST_{n_i-p_i+1, l_i}$ for $i=1, \cdots, g$ and $p_1, \cdots, p_g$ are equal to 1 except at most one $p_j$. Moreover, the equality holds if and only if $C_{B_1,\cdots,B_g}\cong C_{T_1,\cdots,T_g}$. \end{theorem} \begin{proof} Denote $G= C_{T_1,\cdots,T_g}$. We prove this result by two steps. {\bf Step 1:} Let $P$ be the longest path among all paths which start at $u_i$ in $T_i$ for $i=1, \cdots, g$. Without loss of generality, assume $P$ belongs to $T_1$. Let $v$ be a farthest branch vertex from vertex $u_i$ in $T_i$ for $i=2, \cdots, g$. Thus there exists a cut edge $uv$ in $T_i$ such that $d(u_i,u)+1=d(u_i,v)$. By $\xi$-transformation on $uv$ and Lemma~\ref{lemma2.3}, we obtain $G_1=\xi(G,uv)=C_{T_1,T_2, \cdots, T_i', \cdots, T_g}\preceq G$ such that the number of branch vertices in $T_i$ is non-increasing. Hence after a series of $\xi$-transformations on the cut edge $xy$ with $y$ being a branch vertex and $d(u_i, x)+1=d(u_i, y)$ in $T_i'$, $G_2=C_{T_1, T_2, \cdots, T_i^{\prime,\prime}, \cdots, T_g}$ is obtained, where $T_i^{\prime,\prime}$ is a starlike tree of order $n_i$ and leaves $l_i$ with root $u_i$. Hence for $i=2, \cdots, g$, after performing a series of $\xi$-transformations and repeatedly using Lemma~\ref{theorem2.1}, there exists an unicyclic graph $G_3=C_{T_1, B_2, \cdots, B_g} $ such that $G_3\preceq G_1$ where $B_i$ is $ BST_{n_i, l_i}$ with a center vertex $u_i$ for $i=2, \cdots, g$. {\bf Step 2:} If $T_1$ is a path or a starlike tree of order $n_1$, then the assertion holds by using Lemma~\ref{theorem2.1}. Assume that $T_1$ has at least one branch vertex except the root $u_1$. Let $v$ be the branch vertex which is nearest to some pendent vertices in $T_1$ (maybe $v$ is not unique). Then there exists a cut edge $uv$, $d(u_1,u)+1=d(u_1,v)$. If there exists $uv$ as defined above which satisfies the conditions of Lemma~\ref{lemma2.3}, then $G_4=\xi(G_3,uv)\preceq G_3$. Moreover, the number of branch vertices in $G_4$ is no more than that in $G_3$. After performing a series of this type of $\xi$-transformations and repeatedly using Lemma~\ref{theorem2.1}, there exists an unicyclic graph $G_5=C_{T_1^\prime, B_2, \cdots, B_g} $ such that $G_5\preceq G_4$, where $T_1^\prime$ is a tree rooted at $u_1$ obtained by attaching $l_1-1$ pendent paths at some vertices of the longest path $P^\prime$ of $T_1^\prime$. If all pendent paths are attached at the only one vertex $u_1$ or $u_{1,x}$ of $P^\prime$, then the result holds. Otherwise, let $v^\prime$ be the branch vertex which is nearest to $u_1$ in $T_1^\prime$ (when $d_{G_5}(u_1)> 3$, $v^\prime=u_1$). Then there exists a cut edge $u^\prime v^\prime$ which satisfies $d(u_1,v^\prime)+1=d(u_1,u^\prime)$. By $\xi$-transformation on $u^\prime v^\prime$ and Lemma~\ref{lemma2.3}, we obtain $G_6=\xi(G_5,u^\prime v^\prime)\preceq G_5$. Further, the number of branch vertices in $G_6$ is no more than that in $G_5$. Hence by performing a series of this type of $\xi$-transformations, $G_7=C_{B_1, \cdots, B_g}$ is obtained, where $B_1$ has exactly one branch vertex $u_1$ or $B_1$ has exactly one branch vertex $u\neq u_1$ with $d_{G_7}(u_1)=3$. If $u=u_1$, then by Lemma~\ref{theorem2.1}, the assertion holds. If $u\neq u_1$ and $d_{G_7}(u_1)=3$, then applying Lemma~\ref{theorem2.1} to all pendent paths in $G_7$ yields the desired result. \end{proof} \begin{corollary}\label{cor2.6} Let $C_{T_1,\cdots,T_g}\in \mathcal{U}_{n,l}^g$ be obtained from the cycle $C_g=u_1\cdots u_g$ by attaching $g$ trees $T_1, \cdots, T_g$ at the roots $u_1, \cdots, u_g$, respectively. If $d(u_1)> 3$ and $|V(T_i)|=1$ for $i=2, \cdots, g$, then $$C_{B_1,\cdots,B_g}\preceq C_{T_1,\cdots,T_g},$$ where $B_1$ is $BST_{n-g+1, l}$ with a center vertex $u_1$ and $|V(B_i)|=1$ for $i=2, \cdots, g.$ Moreover, the equality holds if and only if $C_{B_1,\cdots,B_g}\cong C_{T_1,\cdots,T_g}$. \end{corollary} \begin{proof} It is obvious that the assertion follows from the proof of Theorem~\ref{theorem2.5}.\end{proof} \section{The minimal elements in two subsets of $\mathcal{U}_{n,l}^g$} In this section, we characterize all extremal graphs which have minimal Laplacian coefficients in the following two special subsets of $\mathcal{U}_{n,l}^g$. Denote \begin{eqnarray*} \mathcal{U}_{n,l}^{g,1}&=&\{C_{T_1,\cdots,T_g}| \ |V(T_i)|=1\ \mbox{ for}\ i=2, \cdots, g\},\\ \mathcal{U}_{n,l}^{g,2}&=&\{C_{T_1,\cdots,T_g}| \ |V(T_1)|>1, |V(T_i)|>1, |V(T_j)|=1 \ \mbox{for } j\neq 1, i \}. \end{eqnarray*} Clearly, $U_{n, l}^{g, p}$ is in $\mathcal{U}_{n,l}^{g,1}$. For convenience, denote $U_{n,l}^{g,0}=U^0, U_{n,l}^{g,1}=U^1, \cdots, U_{n,l}^{g,p}=U^p$. \begin{lemma}\label{lemma3.1} For $1\le p\le \lfloor\frac{n-g-gl+l}{l+1}\rfloor$, $U^p$ and $U^{p-1}$ are incomparable in the poset $(\mathcal{U}_{n,l}^{g}, \preceq)$. (See Fig.1). \end{lemma} \begin{proof} We first show that $c_{n-2}(U^p)<c_{n-2}(U^{p-1})$. It is obvious that $U^{p-1}$ can be regarded as $U^{p-1}=\xi(U^p, w_{p-1}w_p)$. For convenience, we denote $U^p$ and $U^{p-1}$ by $G$ and $G^\prime$, respectively. Let $\mathcal{F}_2$ (resp. $\mathcal{F}_2^{\prime}$) be the set of all spanning forests of $G$ (resp. $G^\prime$) with exactly $2$ components. For an arbitrary spanning forest $F\in \mathcal{F}_{2}$ (resp. $F^\prime\in \mathcal{F}_{2}^\prime$), $F$ (resp. $F^\prime$) can be obtained by deleting two edges $\{e_1, e_2\}$ in $E(G)$ (resp. $E(G^\prime)$) with $e_1$ belonging to the cycle in $G$ (resp. $G^\prime$). If $e_2\neq w_{p-1}w_p$, then $F$ and $F^\prime$ have the same components, which implies $\gamma(F)=\gamma(F^\prime)$. If $e_2= w_{p-1}w_p$, then $$\gamma(F)-\gamma(F^\prime)=(g+p-1)(n-g-p+1)-(\lfloor\frac{n-g-p}{l}\rfloor+1)(n-\lfloor\frac{n-g-p}{l}\rfloor-1)<0,$$ Therefore, $$c_{n-2}(G)-c_{n-2}(G^\prime)=\sum _{F\in \mathcal{F}_2}\gamma (F) -\sum _{F^\prime\in \mathcal{F}_2^{\prime}}\gamma (F^\prime)<0.$$ We next show $c_{m}(G)> c_{m}(G^\prime)$ for $2p\leq m\leq 2(p+g)-3$. Clearly $P_{p+g-1}=w_{p-1}w_{p-2}\cdots w_1u_1u_2\cdots u_g$ (denote $w_i=u_{-i+1}, 1\leq i\leq p-1$ for convenience) is the longest path of order $p+g-1$ in the component of $G-w_{p-1}w_p$. Let $\mathcal{F}^{\prime}$ (resp. $\mathcal{F}$) be the set of all spanning forests of $G^\prime$ (resp. $G$) with exactly $n-m$ components, in other words, $\mathcal{F}^{\prime}$ (resp. $\mathcal{F}$) is the set of all spanning forests of $G^\prime$ (resp. $G$) with exactly $m$ edges. For an arbitrary spanning forest $F^\prime\in \mathcal{F}^\prime$, denote by $T'$ the component of $F'$ containing $w_{p-1}$. Let $f: \mathcal{F}^\prime\rightarrow \mathcal{F}, F=f(F^\prime)$, where $V(F)=V(F^\prime)$ and $$E(F)=E(F')-\{w_{p-1}x| x\in N_{T^\prime}(w_{p-1})\cap N_{G}(w_p)\}+\{w_px| x\in N_{T^\prime}(w_{p-1})\cap N_{G}(w_p)\}.$$ Then $f$ is injective. Let $\mathcal{F}^\prime =\mathcal{F}_{(1)}^\prime \bigcup \mathcal{F}_{(2)}^\prime$, where $\mathcal{F}_{(1)}^\prime=\{F'\in \mathcal{F}^\prime \ |\ w_{p-1}w_p\in E(F')\} $ and $\mathcal{F}_{(2)}^\prime=\{F'\in \mathcal{F}^\prime \ |\ w_{p-1}w_p\not\in E(F')\} $. If $F^\prime\in \mathcal{F}_{(1)}^\prime,$ then $F'$ and $F=f(F')$ have the same components except $T^\prime$. Moreover, $T'$ and $f(T^{\prime})$ have the same vertices. Hence $\gamma(F)=\gamma(F')$. If $F^\prime\in \mathcal{F}_{(2)}^\prime,$ let $S^\prime$ be the component of $F'$ containing $w_p$. Then $F^\prime$ and $F=f(F^\prime)$ have the same components except $T^\prime$ and $S^\prime$ in $F^\prime$. Assume $T^\prime$ contains $a$ vertices in the component of $G-w_{p-1}w_p$ containing $w_p$, $|V(T^\prime)|-a$ vertices in the component of $G-w_{p-1}w_p$ containing $w_{p-1}$. Then $F$ has two components $f(T^{\prime})=T$ with $|V(T^\prime)|-a$ vertices and $f(S^{\prime})=S$ with $a+|V(S^\prime)|$ vertices corresponding to $T^\prime$ and $S'$, respectively. Denote by $N$ the product of the orders of all components of $F^\prime$ except $T^\prime$ and $S^\prime$. Then \begin{align*} \gamma(f(F^{'}))-\gamma(F^{'}) =&[(|V(T^\prime)|-a)(a+|V(S^\prime)|)-|V(T^\prime)|\cdot|V(S^\prime)|]N\\ =&(|V(T^\prime)|-a-|V(S^\prime)|)\cdot a\cdot N. \end{align*} Further let $\mathcal{F}_{(2)}^\prime= \mathcal{F}_{20}^\prime\cup\mathcal{F}_{21}^\prime\cup\mathcal{F}_{22}^\prime$, where \begin{eqnarray*} \mathcal{F}_{20}^\prime&=&\{F^\prime\in \mathcal{F}_{(2)}^\prime\ | \ |V(T^\prime)|-a=|V(S^\prime)| \mbox{ or } a=0\},\\ \mathcal{F}_{21}^\prime&=& \{F^\prime\in \mathcal{F}_{(2)}^\prime\ | \ |V(T^\prime)|-a<|V(S^\prime)|, a>0\},\\ \mathcal{F}_{22}^\prime&=&\{F^\prime\in \mathcal{F}_{(2)}^\prime\ |\ |V(T^\prime)|-a>|V(S^\prime)|, a>0\}. \end{eqnarray*} Hence it follows that $$\forall F^\prime\in \mathcal{F}_{20}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})=0,$$ $$\forall F^\prime\in \mathcal{F}_{21}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})<0,$$ $$\forall F^\prime\in \mathcal{F}_{22}^\prime, \gamma(f(F^{'}))-\gamma(F^{'})>0.$$ For every spanning forest $F_1^\prime\in \mathcal{F}_{21}^\prime$, let $T^\prime$ and $S^\prime$ be two components of $F_1^\prime$ containing $w_{p-1}$ and $w_p$, respectively. Then $T^\prime$ does not contain all the vertices of the path $P_{p+g-1}$. In fact, if all the vertices of $P_{p+g-1}$ belong to $T^\prime$, then by the definition of $\mathcal{F}_{21}^\prime$, we have $|V(T^\prime)|=p+g-1+a$, $|E(T^\prime)|\geq p+g-1$, and $|V(S^\prime)|>|V(T^\prime)|-a=p+g-1$, $|E(S^\prime)|>p+g-2$, which implies $m\geq |E(T^\prime)|+|E(S^\prime)|> p+g-1+p+g-2=2(p+g)-3$. It is a contradiction to $m\leq 2(p+g)-3$. Therefore, assume that $u_{-(p-1)+1}, u_{-(p-2)+1}, \cdots, u_{r-1}\in V(T')$ and $u_r\notin V(T')$. Moreover, let $R^\prime$ be a component of $F_1^\prime$ containing $u_r$ with $b$ vertices. Thus let $T^{\prime\prime}$ be a tree obtained from $T^\prime$ and $R^\prime$ by joining $u_{r-1}$ and $u_r$ with edge $u_{r-1}u_r$, $S^{\prime\prime}$ be the path $w_pv_1\cdots v_{|V(T^\prime)|-a-1}$ and $R^{\prime\prime}$ be the path $v_{|V(T^\prime)|-a}\cdots v_{|V(S^\prime)|-1}$. Then $F_2^\prime=(F_1^\prime-\{T^\prime, S^\prime, R^\prime\})\cup \{T^{\prime\prime}, S^{\prime\prime}, R^{\prime\prime}\}$ is a spanning forest of $G^\prime$ with exactly $m$ edges and $F_2^\prime\in \mathcal{F}_{22}^\prime$. Hence there exists an injective map from $\mathcal{F}_{21}^\prime$ to $ \mathcal{F}_{22}^\prime$, i.e., $$\varphi: \mathcal{F}_{21}^\prime\rightarrow \mathcal{F}_{22}^\prime: F_1^{\prime}\rightarrow F_2^{\prime}=\varphi(F_1^\prime),$$ where $F_2^{\prime}=\varphi(F_1^\prime)=(F_1^\prime-\{T^\prime, S^\prime, R^\prime\})\cup \{T^{\prime\prime}, S^{\prime\prime}, R^{\prime\prime}\}$. Note that $|V(T^{\prime\prime})|=|V(T^\prime)|+b,$ $|V(S^{\prime\prime})|=|V(T^\prime)|-a,$ and $ |V(R^{\prime\prime})|= |V(S^\prime)|-|V(T^\prime)|+a$. It is easy to see that for $F^\prime\in\mathcal{F}_{21}^\prime$, \begin{eqnarray*} &&\gamma(f(\varphi(F^\prime)))-\gamma(\varphi(F^\prime))\\ &=& [(|V(T^\prime)|+b-a)\cdot (|V(T^\prime)|-a+a)-(|V(T^\prime)|+b)\cdot (|V(T^\prime)|-a)]\\ && \cdot (|V(S^\prime)|-|V(T^\prime)|+a)\cdot \frac{N}{b} \\ &=& -(|V(T^\prime)|-a-|V(S^\prime)|)\cdot a N \\ &=&-(\gamma(f(F^\prime))-\gamma(F^\prime)) \end{eqnarray*} Therefore, \begin{align*} &\sum_{F^{\prime}\in \mathcal {F}_{21}^{\prime} \cup \mathcal {F}_{22}^\prime} [ \gamma(f(F^{\prime}))-\gamma(F^{\prime})]\\ =&\sum_{F^{\prime}\in \mathcal {F}_{21}^{\prime}} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})+\gamma(f(\varphi(F^\prime)))-\gamma(\varphi(F^\prime))]+\sum_{F^\prime\in \mathcal {F}_{22}^{\prime}\setminus \varphi(\mathcal {F}_{21}^{\prime})} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})]\\ =&\sum_{F^\prime\in \mathcal {F}_{22}^{\prime}\setminus \varphi(\mathcal {F}_{21}^{\prime})} [\gamma(f(F^{\prime}))-\gamma(F^{\prime})]\geq 0. \end{align*} It follows from Theorem~\ref{theorem1.1} that for $m\leq 2(p+g)-3$, $$c_{m}(G^{\prime})=\sum_{F^{\prime}\in \mathcal {F}^{\prime}}\gamma(F^{\prime}) \leq \sum_{F^{\prime}\in \mathcal {F}^{\prime}}\gamma(f(F^{\prime}))\leq \sum_{F\in \mathcal {F}}\gamma(F)=c_{m}(G).$$ Further we show that the above inequality is strict for $m\geq 2p$. In other words, we show that $\varphi$ is not a bijective map for $2p\leq m\leq 2(p+g)-3$. Hence it is sufficient to find a spanning forest $\bar{F}_2^{\prime}\in \mathcal {F}_{22}^{\prime}$, such that $\bar{F}_2^{\prime}\not\in \varphi(\mathcal {F}_{21}^{\prime})$. Let $\bar{T}^{\prime\prime}, \bar{S}^{\prime\prime}\in \bar{F}_2^{\prime}$, where $\bar{T}^{\prime\prime}$ is the path: $x_1w_{p-1}w_{p-2}\cdots w_1u_1u_g (x_1\in N_G(w_p)\setminus \{w_{p-1}, v_1)\})$, $\bar{S}^{\prime\prime}$ is the path: $w_pv_1\cdots v_{p-1}$. The rest $m-2p$ edges are chosen from the edge set $E(G)\setminus (\{w_{p-1}w_p,u_1u_2, u_gu_{g-1}, v_{p-1}v_{p}\}\cup E(\bar{T}^{\prime\prime})\cup E(\bar{S}^{\prime\prime}))$ of order $n-2p-4$, it is obvious that $n-2p-4\geq m-2p$. Suppose that there exists $\bar{F}_1^{\prime}\in \mathcal {F}_{21}^{\prime}$, and $\varphi(\bar{F}_1^{\prime})=\bar{F}_2^{\prime}$. Let $\bar{T}^{\prime}$ be the component of $\bar{F}_1^{\prime}$ which contains $w_{p-1}$. By the definition of $\varphi$, $|V(\bar{S}^{\prime\prime})|=|V(\bar{T}^{\prime})|-a,$ so $u_2$ is the first vertex on $P_{p+g-1}$ which does not belong to $\bar{T}^{\prime}$, then $u_1u_2\in \varphi(\bar{F}_1^{\prime})$. Since $u_1u_2\not\in \bar{F}_2^{\prime}$, then $\varphi(\bar{F}_1^{\prime})\neq \bar{F}_2^{\prime}$, a contradiction. This completes the proof of Lemma~\ref{lemma3.1}. \end{proof} \begin{lemma}\label{lemma3.2} For $1\le p\neq q\le \lfloor\frac{n-g-gl+l}{l+1}\rfloor$, $U^p$ and $U^{q}$ are incomparable in the poset $(\mathcal{U}_{n,l}^{g}, \preceq)$. \end{lemma} \begin{proof} Without loss of generality, assume $p-q= h \geq1$. By using Lemma~\ref{lemma3.1} repeatedly, we have $$c_{n-2}(U^{p})< c_{n-2}(U^{p-1})<\cdots<c_{n-2}(U^{q}),$$ $$ c_{m}(U^{p})\geq c_{m}(U^{p-1})\geq\cdots c_{m}(U^{q+1})>c_{m}(U^{q}), \mbox{ for } 2(q+1)\leq m\leq 2(q+g+1)-3.$$ Thus the assertion holds. \end{proof} Now we are ready to characterize all minimal elements in $\mathcal{U}_{n,l}^{g,1}$. \begin{theorem}\label{theorem3.3} There are exactly $p+1$ minimal elements $U^0, \cdots, U^p$ in $\mathcal{U}_{n,l}^{g,1}$, where $p=\lfloor\frac{n-g-gl+l}{l+1}\rfloor$. \end{theorem} \begin{proof} For any $G=C_{T_1, T_2, \cdots,T_g}\in \mathcal{U}_{n,l}^{g,1}$ with $|V(T_i)|=1$ for $i=2, \cdots, g$, by Theorem~\ref{theorem2.5}, there exists a graph $G\succeq G_1= C_{B_1, B_2, \cdots, B_g}\in \mathcal{U}_{n,l}^{g,1}$, where $B_1$ is a tree with root $u_1$ obtained by identifying one end $u_{1,p_1}$ of the path $P_{p_1}: u_1u_{1,2}\cdots u_{1,p_1}$ and the center of $BST_{n_1-p_1+1, l_1}$ and $|V(B_i)|=1$ for $i=2, \cdots, g$. If $p_1> \lfloor\frac{n-g-gl+l}{l+1}\rfloor$, then by Lemma~\ref{lemma2.3}, there exists an cut edge $uv$ such that $G_1\succeq G_2=\xi(G_1, uv)=C_{ B_1', B_2, \cdots, B_g}\in \mathcal{U}_{n,l}^{g,1}$, where $B_1'$ is a tree with root $u_1$ obtained by identifying one end $u_{1,p_1-1}$ of the path $P_{p_1-1}: u_1u_{1,2}\cdots u_{1,p_1-1}$ and the center of $BST_{n_1-p_1+2, l_1}$. After a series of $\xi$-transformations, there exists a $U^p$ such that $G\succeq U^p$. If $p_1\leq\lfloor\frac{n-g-gl+l}{l+1}\rfloor$, then $G\succeq G_1=U^{p_1}.$ On the other hand, by Lemma~\ref{lemma3.2}, $U^0, \cdots, U^p$ are incomparable in the poset $(\mathcal{U}_{n,l}^{g,1}, \preceq)$. Hence $U^0, \cdots, U^p$ are exactly all minimal elements in the poset $(\mathcal{U}_{n,l}^{g,1}, \preceq)$. \end{proof} \begin{theorem}\label{theorem3.4} For any $G=C_{T_1,\cdots,T_g}\in \mathcal{U}_{n,l}^{g,2}$, $U^{0}\prec G$ \end{theorem} \begin{proof} By Theorem~\ref{theorem2.5}, we may assume that $G=C_{B_1, \cdots, B_i, \cdots, B_g}\in \mathcal{U}_{n,l}^{g,2}$, where $B_1$ is a tree with root $u_1$ obtained by identifying one end $u_{1,p_1}$ of the path $P_{p_1}: u_1u_{1,2}\cdots u_{1,p_1}$ and the center vertex of $BST_{n_1-p_1+1, l_1}$, $B_i$ is $BST_{n_i, l_i}$ with a center vertex $u_i$, and $|V(B_j)|=1 $ for $j\neq 1, i.$ Let $$G^{\prime}=G-\{u_ix| x\in N_G(u_i)\setminus \{u_{i+1}, u_{i-1}\}\}+ \{u_1x| x\in N_G(u_i)\setminus \{u_{i+1}, u_{i-1}\}\}.$$ We will prove $c_k(G)\geq c_k(G^\prime)$ with at least one strict inequality. Clearly, when $k\in \{0,1,n-1,n\}$, $c_k(G)=c_k(G^{\prime})$. For $2\leq k\leq n-2$, let $\mathcal {F}$ and $\mathcal{F}^{\prime}$ be the spanning forests of $G$ and $G^{\prime}$ with exactly $k$ components, respectively. For an arbitrary spanning forest $F^{\prime}\in \mathcal{F}^{\prime}$ with $T^{\prime}$ being the component of $F^{\prime}$ containing $u_1$, let $f: \mathcal{F}^\prime\rightarrow \mathcal{F}, F^\prime\rightarrow F=f(F^\prime)$, where $V(F)=V(F^\prime)$, and $$E(F)=E(F^{\prime})-\{u_1x|x\in N_{G}(u_i)\cap N_{T^{\prime}}(u_1)\setminus V(C_g)\}+\{u_ix|x\in N_{G}(u_i)\cap N_{T^{\prime}}(u_1)\setminus V(C_g)\}.$$ Then $f$ is injective from $\mathcal{F}^\prime$ to $\mathcal{F}$. Denote by $N$ the product of the orders of all components containing no $u_1, u_i$. If $u_i \in T^{\prime}$, then $F^\prime$ and $F$ have the same components except for $T^\prime$. Moreover, $F$ has a component containing $u_1$ which corresponds to $T^\prime$ in $F^\prime$. Clearly, the two components have the same orders. Hence $\gamma(F)=\gamma(F^{\prime})$. If $u_i\not\in T^{\prime}$, assume $u_i$ is in a component $S^\prime$ of $F^\prime$. Moreover, there are $b\geq 0$ vertices in the connected component containing $u_2$ in $T^\prime-u_1u_2$, and $d\geq 0$ vertices in the connected component containing $u_g$ in $T^\prime-u_1u_g$, $e_1\geq 1$ vertices (including $u_1$) in the vertex set $V(T_1)$ and $e_2\geq 0$ vertices in the vertex set $V(T_i)\setminus \{u_i\}$. Furthermore, the tree $S^{\prime}$ contains $a\geq 1$ vertices in the connected component containing $u_{i}$ in $S^\prime-u_iu_{i+1}$ and $c\geq 0$ vertices in the connected component containing $u_{i+1}$ in $S^\prime-u_iu_{i+1}$. Then $F^\prime$ and $F$ have the same components except for $T^\prime$ and $S^\prime$. Moreover, $F$ have two trees $T$ containing $u_1$ and $S$ containing $u_i$ which correspond to $T^\prime$ and $S^\prime$ in $F^\prime$, respectively. Hence $$\gamma(f(F^\prime))-\gamma(F^{\prime})=[(a+c+e_2)(b+d+e_1)-(a+c)(b+d+e_1+e_2)]N= e_2(b+d+e_1-a-c)N.$$ Consider the subset $\bar{\mathcal{F}}^{\prime}$ of those spanning forests $F^{\prime}$ with $k$ components which coincide on $G^{\prime}\backslash (T^{\prime}\cup S^{\prime})$ with fixed values $e_1, e_2$. Since the two parts of $F^{\prime}$ on the cycle between $T^{\prime}$ and $S^{\prime}$ may be translated, let $a+b=M_1, c+d=M_2$ be fixed. Then \begin{eqnarray*} &&\sum_{F^{\prime}\in \bar{\mathcal{F}}^{\prime}, a+b=M_1, c+d=M_2} (\gamma(f(F^\prime))-\gamma(F^{\prime}))\\ &=&\sum_{ a+b=M_1, c+d=M_2} e_2(b+d+e_1-a-c)N\\ &=&e_2N\sum_{c=0}^{M_2} \sum_{b=0}^{M_1-1} (2b+M_2+e_1-2c-M_1)\\ &=&e_2 N M_1 \sum_{c=0}^{M_2} (e_1+M_2-2c-1)\\ &=&e_2 N M_1(e_1-1)(M_2+1)\\ &\geq& 0. \end{eqnarray*} Hence $$\sum_{F^{\prime}\in \mathcal{F}^{\prime}}( \gamma(f(F^\prime))-\gamma(F^{\prime})) =\sum_{e_1}\sum_{e_2}\sum_{M_1}\sum_{M_2}\sum_{F^{\prime}\in \bar{\mathcal{F}}^{\prime},a+b=M_1, c+d=M_2} (\gamma(f(F^\prime))-\gamma(F^{\prime}))\ge 0.$$ Since $|V(T_1)|>1, |V(T_i)|>1$, there exists one forest $F^{\prime}$ such that $e_1 >1$ and $e_2>0$. Therefore $$c_{n-k}(G^{\prime})=\sum_{F^{\prime}\in\mathcal {F}^{\prime}}\gamma(F^\prime)< \sum_{F^\prime\in \mathcal {F}^\prime}\gamma(f(F^\prime))\leq \sum_{F\in \mathcal {F}}\gamma(F)=c_{n-k}(G), 2\leq k\leq n-2.$$ Hence by Corollary~\ref{cor2.6}, we have $U^0\preceq G^\prime$ with equality if and only if $G^\prime\cong U^0$. Therefore, $U^0\preceq G^\prime\preceq G$, the assertion holds. \end{proof} It follows from Theorems~\ref{theorem1.2}, \ref{theorem3.3} and \ref{theorem3.4} that the following results hold. \begin{corollary}\label{corollary3.5} Let $G=C_{T_1,\cdots,T_g}$ be an arbitrary unicyclic graph in $\mathcal{U}_{n,l}^{g,1}$. Then for $p= \lfloor\frac{n-g-gl+l}{l+1}\rfloor$, $$LEL(G)\geq \mbox{min}\{LEL(U^0), LEL(U^1), \cdots, LEL(U^p)\}.$$ \end{corollary} \begin{corollary}\label{corollary3.6} Let $G=C_{T_1,\cdots,T_g}$ be an arbitrary unicyclic graph in $\mathcal{U}_{n,l}^{g,2}$. Then $LEL(G)> LEL(U^0)$. \end{corollary} \section{The minimal elements in $\mathcal{U}_{n,l}^3$ and $\mathcal{U}_{n,l}^4$} In this section, we determine all the minimal elements in the posets $(\mathcal{U}_{n,l}^3, \preceq)$ and $(\mathcal{U}_{n,l}^4, \preceq)$. Before stating our results, we need the following definitions. \begin{definition}\label{definition4.1} For any $ G\in\mathcal{U}_{n,l}^3$, let $G^\ast$ be the graph obtained from $G$ by changing all the edges (except $E(C_3)$) incident with $u_2, u_3$ into new edges between $u_1$ and $N_G(u_2)\cup N_G(u_3)\setminus V(C_3)$. In other words, \begin{eqnarray*} G^\ast&=&G-\{u_2x| x\in N_G(u_2)\setminus V(C_3)\}-\{u_3x| x\in N_G(u_3)\setminus V(C_3)\}\\ &&+\{u_1x| x\in N_G(u_2)\cup N_G(u_3)\setminus V(C_3)\}. \end{eqnarray*} We say $G^\ast$ is a $\eta$-transformation of $G$. (See Fig. 4). \end{definition} \begin{figure}\label{3} \centering \begin{tikzpicture}[scale=1] \tikzstyle{every node}=[draw,shape=circle,inner sep=1.5pt]; \node (v0)[label=-180:$u_2$] at (0:0){}; \node(v1) [label=0:$u_1$]at (0:1){}; \node(v2) [label=0:$u_3$] at(60:1){}; \node(v3) at (1,2){}; \node(v4) at (0,2){} \node(v5) at (-1,-1){} \node(v6) at (0,-1){}; \node(v7) at (1,-1){} \node(v8) at (2,-1){}; \draw (v0) --(v1) (v1) --(v2) (v2)--(v0) (v2)--(v4) (v2)--(v3) (v0)--(v5) (v0)--(v6) (v1)--(v7) (v1)--(v8); \draw [loosely dotted,thick] (v3) --(v4) (v5)--(v6) (v7)--(v8); \draw (0.5,2) ellipse (1.5 and 0.5); \draw (-0.6,-1) ellipse (1 and 0.6); \draw (1.6,-1) ellipse( 1 and 0.6); \tikzstyle{every node}=[draw,shape=circle,inner sep=1.5pt]; \node (v0)[label=-180:$u_2$] at (6,0){}; \node(v1) [label=88:$u_1$]at (7,0){}; \node(v2) [label=0:$u_3$] at(6.5,0.9){}; \node(v3) at (8,1){}; \node(v4) at (8,0){} \node(v5) at (5,-1){} \node(v6) at (6,-1){}; \node(v7) at (7,-1){} \node(v8) at (8,-1){}; \draw (v0) --(v1) (v1) --(v2) (v2)--(v0) (v1)--(v4) (v1)--(v3) (v1)--(v5) (v1)--(v6) (v1)--(v7) (v1)--(v8); \draw [loosely dotted,thick] (v3) --(v4) (v5)--(v6) (v7)--(v8); \draw[->] (3,0)--(4,0); \draw (8,0.5) ellipse (0.5 and 0.9); \draw (7.4,-1) ellipse (0.9 and 0.6); \draw (5.5,-1) ellipse( 0.9 and 0.6); \end{tikzpicture} \caption{$\eta$-transformation} \label{fig:pepper} \end{figure} \begin{lemma}\label{lemma4.2} For $G\in\mathcal{U}_{n,l}^3$, if $G^\ast$ is obtained from $G$ by $\eta$-transformation, then $G^\ast\preceq G$, i.e., $c_{k}(G^\ast)\leq c_{k}(G)$ with equality if and only if $k\in \{0,1,n-1,n\}$. \end{lemma} \begin{proof} Clearly, when $k\in\{0,1,n-1,n\}$, $c_k(G)=c_k(G^\ast)$. For $2\leq k\leq n-2$, let $\mathcal{F}^\prime$ (resp. $\mathcal{F}$) be the set of all spanning forests of $G^\ast$ (resp. $G$) with exactly $n-k$ components. Let $\mathcal{F}^\prime=\mathcal{F}^{\prime(1)}\cup \mathcal{F}^{\prime(2)}\cup\mathcal{F}^{\prime(3)}$, where $\mathcal{F}^{\prime(j)}, j=1,2,3$ is the set of all spanning forests of $G^\ast$ in which $u_1, u_2, u_3$ belong to exactly $j$ different components. Similarly, $\mathcal{F}^{(j)}, j=1,2,3$ can be defined. Let $f: \mathcal{F}^\prime\rightarrow \mathcal{F}$ with $F^\prime \rightarrow F=f(F^\prime)$, where $V(F)=V(F^\prime)$ and \begin{align*} E(F)=&E(F^\prime)\\ &-\{u_1x| x\in N_{R^\prime}(u_1)\cap N_G(u_2)\setminus \{u_3\}\} -\{u_1x| x\in N_{R^\prime}(u_1)\cap N_G(u_3)\setminus \{u_2\}\}\\ &+\{u_2x| x\in N_{R^\prime}(u_1)\cap N_G(u_2)\setminus \{u_3\}\} +\{u_3x| x\in N_{R^\prime}(u_1)\cap N_G(u_3)\setminus \{u_2\}\}, \end{align*} for $R^\prime$ being a component of $F^\prime$ containing $u_1$. Clearly $f$ is injective and $f(\mathcal{F}^{\prime(j)})\subseteq \mathcal{F}^{(j)}$ for $j=1,2,3$. Denote $|V(T_1)\cap V(R^\prime)\setminus\{u_1\}|=a, |V(T_2)\cap V(R^\prime)\setminus\{u_2\}|=b, |V(T_3)\cap V(R^\prime)\setminus\{u_3\}|=c$, where $a,b,c\geq 0$. Let $N$ be the product of the orders of all components of $F^\prime$ containing no $\{u_1, u_2, u_3\}$. Now we distinguish the proof into the following three cases. {\bf Case 1:} $F^\prime\in \mathcal{F}^{\prime(1)}$, $u_1, u_2, u_3$ belong to one component, then $\gamma(F)=\gamma(F^\prime)$. Thus $$\sum_{F^\prime\in \mathcal{F}^{\prime(1)}}[\gamma(F)-\gamma(F^\prime)]=0.$$ {\bf Case 2:} $F^\prime\in \mathcal{F}^{\prime(2)}$, $u_1, u_2, u_3$ are in two components, then $\gamma(F)-\gamma(F^\prime)=[(a+b+2)(c+1)+(a+c+2)(b+1)+(b+c+2)(a+1)-2(a+b+c+1)-2(a+b+c+2)]N=[(a+b)c+(a+c)b+(b+c)a]N\geq 0$. Thus $$\sum_{F^\prime\in \mathcal{F}^{\prime(2)}}[\gamma(F)-\gamma(F^\prime)]\geq 0.$$ {\bf Case 3:} $F^\prime\in \mathcal{F}^{\prime(3)}$, $u_1, u_2, u_3$ are in three components, then $\gamma(F)-\gamma(F^\prime)=[(a+1)(b+1)(c+1)-(a+b+c+1)]N=(abc+ab+ac+bc)N\geq 0$. Thus $$\sum_{F^\prime\in \mathcal{F}^{\prime(3)}}[\gamma(F)-\gamma(F^\prime)]\geq 0.$$ Now the inequality $c_{k}(G^\ast)<c_{k}(G), k=2,3,\cdots, n-2$ holds from Theorem~\ref{theorem1.1} by summing over all possible subsets $\mathcal{F}^\prime$ of spanning forests $F^\prime$ of $G^\ast$ with $n-k$ components. \end{proof} \begin{theorem}\label{theorem4.3} There are exactly $p+1$ minimal elements $U_{n,l}^{3,0}, U_{n,l}^{3,1} \cdots, U_{n,l}^{3,p}$ in the poset $(\mathcal{U}_{n,l}^{3},\preceq)$, where $p= \lfloor\tfrac{n-3-2l}{l+1}\rfloor$. \end{theorem} \begin{proof} The assertion follows from Lemma~\ref{lemma4.2} and Theorems~\ref{theorem3.3} and \ref{theorem3.4}. \end{proof} \begin{definition}\label{definition4.4} For $ G\in\mathcal{U}_{n,l}^{4}$, let $G^\star$ be the graph obtained from $G$ by changing all the edges (except $E(C_4)$) incident with $u_2, u_3, u_4$ into new edges between $u_1$ and $\cup_{i=2}^4 N_G(u_i)\setminus V(C_4)$. In other words, \begin{eqnarray*} G^\star&=&G-\{u_2x| x\in N_G(u_2)\setminus V(C_4)\}-\{u_3x| x\in N_G(u_3)\setminus V(C_4)\}\\ &&-\{u_4x| x\in N_G(u_4)\setminus V(C_4)\}+\{u_1x| x\in \cup_{i=2}^4 N_G(u_i)\setminus V(C_4)\}. \end{eqnarray*} We say $G^\star$ is a $\kappa$-transformation of $G$. \end{definition} Since for bipartite graphs, the Laplacian coefficients and the signless Laplacian coefficients are equal, from Lemma 3.1 in \cite{zhang2012}, we have the following lemma: \begin{lemma}\label{lemma4.5} For $G\in\mathcal{U}_{n,l}^{4}$, if $G^\star$ is obtained from $G$ by $\kappa$-transformation, then $G^\star\preceq G$, i.e., $c_{k}(G^\star)\leq c_{k}(G)$ with equality if and only if $k\in \{0,1,n-1,n\}$. \end{lemma} By combining Lemma~\ref{lemma4.5} and Theorem~\ref{theorem3.3}, we have the following theorem. \begin{theorem}\label{theorem4.6} There are exactly $p+1$ minimal elements $U_{n,l}^{4,0}, U_{n,l}^{4,1} \cdots, U_{n,l}^{4,p}$ in the poset $(\mathcal{U}_{n,l}^{4},\preceq)$, where $p= \lfloor\tfrac{n-4-3l}{l+1}\rfloor$. \end{theorem} From Theorem~\ref{theorem1.2}, we have the following corollary: \begin{corollary}\label{corollary4.7} (1). Let $G=C_{T_1,T_2,T_3}\in \mathcal{U}_{n,l}^3$. Then for $p= \lfloor\tfrac{n-3-2l}{l+1}\rfloor$, $$LEL(G)\geq \mbox{min}\{U_{n,l}^{3,0}, U_{n,l}^{3,1} \cdots, U_{n,l}^{3,p}\}.$$ (2). Let $G=C_{T_1,T_2,T_3,T_4}\in \mathcal{U}_{n,l}^4$. Then for $p= \lfloor\tfrac{n-4-3l}{l+1}\rfloor$, $$LEL(G)\geq \mbox{min}\{U_{n,l}^{4,0}, U_{n,l}^{4,1} \cdots, U_{n,l}^{4,p}\}.$$ \end{corollary} \section{Remarks} Although Ili\'{c} and Ili\'{c}'s conjecture is false, we may modify the condition or result such that the conjecture is still true. In fact, if there are at least two vertices in the cycle with degrees at least 3, then the conjecture is true for $g=3$ and $g=4 $. Moreover, we checked that the conjecture is still true for all unicyclic graphs on $\leq 30$ vertices with fixed $l, g$ and having at least three vertices in the cycle with degrees at least 3. Hence their conjecture can be modified as follows: \begin{conjecture} (1). For $G\in \mathcal{U}_{n,l}^g$, if there are more than two vertices in the cycle having degrees $\geq 3$, then $U^0\preceq G$, with equality if and only if $G\cong U^0$. (2). There are exactly $p+1$ minimal elements $U_{n,l}^{3,0}, U_{n,l}^{3,1} \cdots, U_{n,l}^{3,p}$ in the poset $(\mathcal{U}_{n,l},\preceq)$, where $p= \lfloor\tfrac{n-3-2l}{l+1}\rfloor$. \end{conjecture} \frenchspacing
\section{Background} \label{sec:background} In this section, we review three-point branched covers of Riemann surfaces. For a reference, see the introductory texts of Miranda \cite{Miranda} and Girondo--Gonz\'{a}lez-Diez \cite{GirondoGonzalezDiez}. \subsection*{Triangle groups and Bely\u{\i}\ maps} By the Riemann existence theorem, Riemann surfaces are the same as (algebraic) curves over $\mathbb C$; we will pass between these categories without mention. A \defi{Bely\u{\i}\ map} over $\mathbb C$ is a morphism $\phi:X \to \mathbb P_\mathbb C^1$ of curves over $\mathbb C$ that is unramified outside $\{0,1,\infty\}$. Two Bely\u{\i}\ maps $\phi,\phi':X,X' \to \mathbb P^1$ are \defi{isomorphic} if there is an isomorphism $\iota:X \to X'$ such that $\phi = \phi' \circ \iota$. Bely\u{\i}\ \cite{Belyi,Belyi2} proved that a curve $X$ over $\mathbb C$ can be defined over $\overline{\mathbb Q}$, an algebraic closure of $\mathbb Q$, if and only if $X$ admits a Bely\u{\i}\ map, so we may work equivalently with curves defined over $\overline{\mathbb Q}$ or over $\mathbb C$. A Bely\u{\i}\ map $\phi:X \to \mathbb P^1$ admits a short combinatorial description as a \defi{transitive permutation triple} $\sigma = (\sigma_0, \sigma_1, \sigma_{\infty}) \in S_d^3$ such that $\sigma_{\infty} \sigma_1 \sigma_{0} = 1$ and $\langle \sigma_0,\sigma_1, \sigma_{\infty} \rangle \leq S_d$ is a transitive subgroup, according to the following lemma. \begin{lem} \label{lem:biject} There is a bijection between transitive permutation triples up to simultaneous conjugacy and isomorphism classes of Bely\u{\i}\ maps over $\mathbb C$ (or $\overline{\mathbb Q}$). \end{lem} By the phrase \defi{up to simultaneous conjugation}, we mean that we consider two triples $\sigma,\sigma' \in S_d^3$ to be equivalent if there exists $\rho \in S_d$ such that $\sigma=\rho\sigma\rho^{-1}$, i.e., \[ (\sigma_0',\sigma_1',\sigma_\infty') = \rho(\sigma_0,\sigma_1,\sigma_{\infty})\rho^{-1} = (\rho\sigma_0\rho^{-1}, \rho\sigma_1\rho^{-1}, \rho\sigma_{\infty}\rho^{-1}). \] In this correspondence, the cycles of the permutation $\sigma_0,\sigma_1,\sigma_\infty$ correspond to the points of $X$ above $0,1,\infty$, respectively, and the length of the cycle corresponds to its multiplicity. Note in particular that there are only finitely many $\overline{\mathbb Q}$-isomorphism classes of curves $X$ with a Bely\u{\i}\ map of bounded degree. The proof of Lemma \ref{lem:biject} has many strands and is a piece of Grothendieck's theory of dessin d'enfants. The bijection can be realized explicitly using the theory of triangle groups and this constructive proof was the starting point of this paper and is the subject of this section. For an introduction to triangle groups, including their relationship to Bely\u{\i}\ maps and dessins d'enfants, see the surveys of Wolfart \cite{WolfartDessins,WolfartObvious}, and for a more classical treatment see Magnus \cite{Magnus}. Let $a,b,c \in \mathbb Z_{\geq 2} \cup \{\infty\}$ with $a \leq b \leq c$. We will call the triple $(a,b,c)$ \defi{spherical}, \defi{Euclidean}, or \defi{hyperbolic} according to whether the value $$ \chi(a,b,c) = 1 - \frac{1}{a} - \frac{1}{b} - \frac{1}{c} $$ is respectively negative, zero, or positive, and we associate the geometry \[ H= \begin{cases} \text{sphere $\mathbb P^1_{\mathbb C}$}, & \text{ if $\chi(a,b,c)<0$}; \\ \text{plane $\mathbb C$}, & \text{ if $\chi(a,b,c)=0$}; \\ \text{upper half-plane $\mathcal{H}$}, & \text{ if $\chi(a,b,c)>0$}. \end{cases} \] accordingly. The spherical triples are $(2,3,3)$, $(2,3,4)$, $(2,3,5)$, and $(2,2,c)$ for $c \in \mathbb Z_{\geq 2}$. The Euclidean triples are $(2,2,\infty)$, $(2,3,6)$, $(2,4,4),$ and $(3,3,3)$. All other triples are hyperbolic. For each triple, we have the associated \defi{triangle group} \begin{equation} \label{eqn:Deltaabcpres} \Delta(a,b,c) = \langle \delta_a, \delta_b, \delta_c \mid \delta_a^a = \delta_b^b = \delta_c^c = \delta_c \delta_b \delta_a =1 \rangle \end{equation} analogously classified as spherical, Euclidean, or hyperbolic. The triangle group $\Delta(a,b,c)$ has a simple geometric description. Let $T$ be a triangle with angles $\pi/a, \pi/b, \pi/c$ on $H$. Let $\tau_a, \tau_b,$ and $\tau_c$ be the reflections in the three sides of $T$. The group generated by $\tau_a, \tau_b,$ and $\tau_c$ is a discrete group with fundamental domain $T$. The subgroup of orientation-preserving isometries in this group is generated by \[ \delta_a = \tau_c \tau_b,\ \delta_b = \tau_a \tau_c,\text{ and }\delta_c = \tau_b \tau_a \] and has fundamental domain $D_\Delta$ a copy of $T$ and its mirror. The above relations can be seen in Figure \ref{fig:goabc}. For instance, note that $\delta_b(D_\Delta) = \tau_a(D_\Delta)$, the only difference being that the shaded and unshaded triangles are reversed. Precomposing $\tau_a$ by $\tau_c$ corrects this reversal, hence $\delta_b = \tau_a \tau_c$. The other relations follow similarly. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-01-edit.pdf} \begin{comment} \begin{equation} \psset{unit=.75cm} \begin{pspicture}(-5,-5)(5,5) \def\Reflections{% \psline[linewidth=.5pt]{<->}(-3,0)(3,0) \psline[linewidth=.5pt]{<->}(0,-1)(0,5) \uput*{5pt}[0]{0}(0.67,2.65){$\tau_a$} \uput*{5pt}[0]{0}(0.01,1.4){$\tau_b$} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psline(0,1)(0,3.5122186) \psarc(3.512218,0){4.96702}{135}{170.44683} \psarcn(-0.57735,0){1.15470}{134.4468}{60} \closepath} \pscustom[linewidth=.5pt]{ \psline(0,1)(0,3.5122) \psarcn(-3.5122,0){4.96702}{45}{9.55316} \psarc(0.577350,0){1.15470}{45.55316}{120} \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn(3.5122,0){4.9670}{135}{51.9346} \psarc(6.433661,0){3.9131}{87.934}{151.78545} \psarc(0,0){3.5122}{31.7854597}{90} \closepath} \pscustom[linewidth=.5pt]{ \psarcn(0,0){3.51221}{90}{31.78545} \psarc(3.04314,0){1.85092}{91.78545}{153.5531} \psarc(-3.51221,0){4.967027}{9.553164}{45} \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn(-0.577350,0){1.15470}{60.0000}{18.6699} \psarc(0.33365,0){0.41242}{63.66997}{144.00} \psline(0,0.24241)(0,1) \closepath} \pscustom[linewidth=.5pt]{ \psarcn(0.577350,0){1.15470}{120}{45.5531} \psarc(1.26350,0){0.83338}{81.55316}{153.6699} \psarc(-0.577350,0){1.15470}{18.66997}{60.0} \closepath} \uput*{5pt}[225]{0}(0,1){$z_a$} \uput*{5pt}[-45]{0}(1.38592,0.8243){$z_c$} \uput*{5pt}[135]{0}(0,3.5122){$z_b$} \uput*{5pt}[225]{0}(-1.3859,0.82434){$-\overline{z_c}$} \psdots*(0,1)(0,3.5122)(1.385926,0.82434)(-1.3859,0.82434) \psdots*(2.985478,1.8500)(6.574682,3.910)(0.51658,0.369638)(0,0.24241) \psarcn{->}(1.2,1.8){0.5}{155}{80} \psarcn{->}(0.25,1.25){0.5}{10}{-50} } \def\Rotations{% \psline[linewidth=.5pt]{<->}(-3,0)(3,0) \psline[linewidth=.5pt]{<->}(0,-1)(0,5) \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psline(0,1)(0,3.512218) \psarc(3.512218,0){4.96702}{135}{170.4468} \psarcn(-0.57735,0){1.15470}{134.4468}{60} \closepath} \pscustom[linewidth=.5pt]{ \psline(0,1)(0,3.5122) \psarcn(-3.5122,0){4.96702}{45}{9.553164} \psarc(0.577350,0){1.154700}{45.5531648}{120} \closepath} \pscustom[linewidth=.5pt]{ \psarcn(3.512218,0){4.96702}{135}{51.9346} \psarc(6.433661,0){3.91312}{87.934}{151.78545} \psarc(0,0){3.5122}{31.7854597}{90} \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn(0,0){3.51221}{90}{31.78545} \psarc(3.043148,0){1.85092}{91.78545}{153.5531} \psarc(-3.512218,0){4.96702}{9.553164}{45} \closepath} \pscustom[linewidth=.5pt]{ \psarc(-0.577350,0){1.15470}{60}{134.4468} \psarcn(-1.26350,0){0.83338}{98.446}{26.33002} \psarcn(0.57735,0){1.1547}{161.3300}{120} \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psarc(0.5773502,0){1.15470}{120}{161.3300} \psarcn(-0.33365,0){0.4124262}{116.33002}{36} \psline(0,0.2424)(0,1) \closepath} \uput*{7pt}[-30]{0}(0,1){$z_a$} \uput*{5pt}[-45]{0}(1.38592,0.824341){$z_c$} \uput*{5pt}[135]{0}(0,3.5122){$z_b$} \uput*{5pt}[225]{0}(-1.385926,0.82434){$-\overline{z_c}$} \psarc[linewidth=.5pt]{->}(0,1){0.3}{90}{218} \psarc[linewidth=.5pt]{->}(0,3.5122){0.3}{-90}{-2} \uput{15pt}[120]{0}(0,1){$\delta_a$} \uput{14pt}[-65]{0}(0,3.5122){$\delta_b$} \psdots*(0,1)(0,3.5122)(1.385926,0.82434)(-1.38592,0.824341) \psdots*(2.985478,1.85002)(6.57468,3.9105)(-0.516587,0.36963)(0,0.2424) } \rput{0}(-5,0){\Reflections} \rput{0}(5,0){\Rotations} \end{pspicture} \end{equation} \end{comment} \caption{A normalized triangle and its images under the reflections $\tau_a$ and $\tau_b$, and under the rotations $\delta_a = \tau_c \tau_b$ and $\delta_b = \tau_a \tau_c$.} \label{fig:goabc} \end{figure} From these gluing relations, we see that $\Delta \backslash H \cong \mathbb P^1 \setminus S$ as Riemann surfaces, where $S$ consists of a set of points indexed by indices with $a,b,c=\infty$. So if $a,b,c<\infty$, the case of interest here, then we simply have $\Delta \backslash H \cong \mathbb P^1$. From now on we restrict to the case where $a,b,c<\infty$; for the remaining cases, one has a different approach not pursued here. \begin{rmk} Mapping elements to their inverses gives an isomorphism from the group $\Delta(a,b,c)$ to the group with the relation with $\delta_c \delta_b \delta_a = 1$ replaced by $\delta_a \delta_b \delta_c = 1$ with the same fundamental domain. The choice corresponds to an orientation of the Riemann surface $\Delta(a,b,c) \backslash H$; we have chosen the counterclockwise orientation, so that $\delta_s$ acts by counterclockwise rotation by $2\pi/s$ for $s=a,b,c$; and with this choice, we obtain the presentation \eqref{eqn:Deltaabcpres}. \end{rmk} \subsection*{Conventions} First, a convention on permutations (following a convention in computational group theory): we compose permutations from left to right. This is contrary to the usual functional composition, so to avoid confusion we will write it in the exponentiated form: for example, if $\sigma=(1\ 2)$ and $\tau=(2\ 3)$, then we write $1^{\sigma\tau} = (1^\sigma)^\tau = 2^\tau = 3$, so $\sigma\tau = (1\ 3\ 2)$. In this way, the action of $S_d$ on $\{1,\dots,d\}$ is on the right. There is an injective map which, given a transitive permutation triple $\sigma=(\sigma_0,\sigma_1,\sigma_\infty) \in S_d^3$ with orders $a,b,c \in \mathbb Z_{\geq 2}$ up to simultaneous conjugation, associates a subgroup $\Gamma \leq \Delta=\Delta(a,b,c)$ of index $[\Delta:\Gamma]=d$ up to conjugation. Explicitly, to the permutation triple $\sigma$ we associate \begin{equation} \label{eqn:gammaperm} \Gamma = \pi^{-1}\Stab(1) = \{ \delta \in \Delta : 1^{\pi(\delta)} = 1\}, \end{equation} where \begin{equation} \label{eqn:piperm} \begin{aligned} \pi : \Delta &\to S_d \\ \delta_a, \delta_b, \delta_c &\mapsto \sigma_0, \sigma_1, \sigma_\infty \end{aligned} \end{equation} is the permutation representation induced by $\sigma$. This map has a right inverse: given $\Gamma \leq \Delta$ of index $d$, we define a permutation triple as follows. We choose cosets $\Gamma\alpha_1,\dots,\Gamma\alpha_d$ of $\Gamma$, and define a permutation representation $\pi:\Delta \to S_d$ by \begin{equation} \label{eqn:deltaaction} i^{\pi(\delta)}=j \quad \text{if} \quad \Gamma \alpha_i \delta = \Gamma \alpha_j, \end{equation} for $\delta \in \Delta$ and $i,j \in \{1,\dots,d\}$. We then let $\sigma_0=\pi(\delta_a)$, $\sigma_1=\pi(\delta_b)$, and $\sigma_\infty=\pi(\delta_c)$. This defines a group homomorphism, but it changes the left action of $\Delta$ on $\mathcal{H}$ into a right action of $\Delta$ on $\{1, \dots, d\}$, according to our convention. Indeed, if $\Gamma \alpha_i \delta = \Gamma \alpha_j$ and $\Gamma \alpha_j \delta' = \Gamma \alpha_k$, so that $i^{\pi(\delta)}=j$ and $j^{\pi(\delta')}=k$, then \[ \Gamma \alpha_i \delta\delta' = \Gamma \alpha_j \delta' = \Gamma \alpha_k \] so \[ i^{\pi(\delta\delta')} = k = j^{\pi(\delta')} = (i^{\pi(\delta)})^{\pi(\delta')} = i^{\pi(\delta)\pi(\delta')}. \] If we begin with the permutation triple $\sigma$ and associate the group $\Gamma$ as in (\ref{eqn:gammaperm}) and choose cosets $\Gamma\alpha_i$ with the property that \begin{equation} \label{eqn:permtocoset} 1^{\pi(\alpha_i)} = i \end{equation} (well-defined, since $1^{\pi(\gamma\alpha_i)}=1^{\pi(\gamma)\pi(\alpha_i)}=1^{\pi(\alpha_i)}=i$ for all $\gamma \in \Gamma$ by definition). Since $i^{\sigma_0} = j = i^{\pi(\delta_a)}$ if and only if \[ \Gamma \alpha_i \delta_a = \Gamma \alpha_j \] and similarly for $\sigma_1$ (resp.~$\sigma_\infty$) and $\delta_b$ (resp.~$\delta_c$), we recover $\sigma$ as claimed. \begin{rmk} These maps can be made into bijections by artificially restricting the image to those subgroups such that $\sigma_0,\sigma_1,\sigma_{\infty}$ have orders $a,b,c$, i.e., eliminating those subgroups that ``come from a smaller triple''. \end{rmk} \subsection*{Correspondence} We now prove Lemma \ref{lem:biject}, as sketched in the introduction. As mentioned above, $\Delta$ (and also $\Gamma$) act on $H$ on the left and the quotient map $\phi: X = \Gamma \backslash H \to \Delta \backslash H \cong \mathbb P^1$ is then a Bely\u{\i}\ map. We first show that this association is well-defined on simultaneous conjugacy classes of permutation triples. Suppose $\sigma,\sigma' \in S_d^3$ are simultaneously conjugate with $\sigma'=\rho\sigma\rho^{-1}$ for some $\rho \in S_d$. The permutations of these triples have the same respective orders, hence the same triangle group $\Delta=\Delta(a,b,c)$ is associated to both triples. Let $\pi,\Gamma$ be as in \eqref{eqn:gammaperm}, \eqref{eqn:piperm}. Then $\Gamma' = \pi^{-1} \Stab(k)$ where $k=1^\rho$: if a word $\upsilon'$ written in $\sigma_0', \sigma_1', \sigma_\infty'$ fixes 1, then $\upsilon' = \rho \upsilon \rho^{-1}$, where $\upsilon$ is the corresponding word written in $\sigma_0, \sigma_1, \sigma_\infty$, so $\upsilon$ fixes $1^\rho = k$. Since $\langle \sigma_0, \sigma_1, \sigma_\infty \rangle \leq S_d$ is transitive, then there exists $\mu \in \langle \sigma_0, \sigma_1, \sigma_\infty \rangle$ with $1^\mu = k$. Then $$ \Gamma' = (\pi')^{-1} \Stab(1) = \pi^{-1} \Stab(k) = \pi^{-1}(\mu \Stab(1) \mu^{-1}) \, . $$ Choosing $\delta \in \pi^{-1}(\mu)$, then $$ \Gamma' = \pi^{-1}(\mu \Stab(1) \mu^{-1}) = \delta \pi^{-1}(\Stab(1)) \delta^{-1} = \delta \Gamma \delta^{-1} $$ so these subgroups are conjugate; and the map \begin{align*} \iota : \Gamma \backslash H &\mapsto \Gamma' \backslash H\\ z &\mapsto \delta z \end{align*} is an isomorphism of Riemann surfaces with $\phi' \circ \iota = \phi$, where $\phi : \Gamma \backslash H \to \Delta \backslash H$ and $\phi' : \Gamma' \backslash H \to \Delta \backslash H$ are the quotient maps described above. We now construct the inverse map to the correspondence in Lemma \ref{lem:biject}. Let $\phi: X \to \mathbb P^1$ be a Bely\u{\i}\ map of degree $d$, let $Y = X \setminus \phi^{-1}(\{0, 1, \infty\})$ and let $U = \mathbb P^1 \setminus \{0, 1, \infty\}$. Then the induced map $\phi|_Y : Y \to U$ is an unramified covering map. Let $* \in U$ be a basepoint. We have a presentation of the fundamental group as \[ \pi_1(U,*)=\langle \eta_0, \eta_1, \eta_\infty \mid \eta_\infty \eta_1 \eta_0 = 1 \rangle \] where $\eta_0,\eta_1,\eta_\infty$ are homotopy classes represented by loops around 0, 1, and $\infty$, respectively. For each $x \in \phi^{-1}(*)$, a path $\eta$ in $U$ with initial point $*$ can be lifted to a unique path $\widetilde{\eta}$ in $Y$ with initial point $x$ such that $f \circ \widetilde{\eta} = \eta$. Thus the terminal point of $\widetilde{\eta}$ will be a unique $x' \in \phi^{-1}(*)$, and this induces a right action of $\pi_1(U,*)$ on $\phi^{-1}(*)$ by $x^\eta = \widetilde{\eta}(1)$. If we label the $d$ points in $\phi^{-1}(*)$ with $\{1,\dots,d\}$, then the action yields a permutation representation $\sigma : \pi_1(U, *) \to S_d$. Letting $\sigma_0 = \sigma(\eta_0)$, $\sigma_1 = \sigma(\eta_1)$, and $\sigma_\infty = \sigma(\eta_\infty)$ yields a permutation triple. Moreover, since $Y$ is path-connected, the group $\langle \sigma_0, \sigma_1, \sigma_\infty \rangle \leq S_d$ is transitive. \begin{rmk} Instead of considering $U=\mathbb P^1 \setminus \{0,1,\infty\}$, one can also consider the orbifold fundamental group of $X(\Delta)=\Delta \backslash H$, which is exactly $\pi_1(X(\Delta),*) \cong \Delta$ (with a choice of basepoint $*$). Then the map $X(\Gamma) \to X(\Delta)$ is a cover in the orbifold category, and the argument in the previous paragraph becomes a tautology. \end{rmk} We now show that this map is well-defined on the set of isomorphism classes of Bely\u{\i}\ maps. Let $\phi : X \to \mathbb P^1$ and $\phi' : X' \to \mathbb P^1$ be isomorphic Bely\u{\i}\ maps. Then there is an isomorphism $\iota:X \to X'$ of Riemann surfaces such that $\phi = \phi' \circ \iota$. The restriction \[ \iota|_{\phi^{-1}(*)} : \phi^{-1}(*)=\{x_1,\dots,x_d\} \to \{x_1',\dots,x_d'\}=(\phi')^{-1}(*) \] is a bijection and so if these sets are labelled, we obtain a relabeling $\rho \in S_d$ with $\iota(x_i)=x_j'$ if and only if $i^\rho=j$. Let $\sigma,\sigma'$ be the permutation triples associated to $\phi,\phi'$ above. Then \[ (\sigma_0, \sigma_1, \sigma_\infty) = (\rho {\sigma_0}' \rho^{-1}, \rho {\sigma_1}' \rho^{-1}, \rho {\sigma_\infty}' \rho^{-1}) \] so the triples are simultaneously conjugate, as desired. To complete the proof of Lemma \ref{lem:biject}, we check directly that these maps are inverses; an explicit version of this is the subject of section \ref{sec:cosets}. \begin{rmk} The triple $(\sigma_0, \sigma_1, \sigma_\infty)$ encodes several pieces of information about the associated ramified covering. The number of disjoint cycles in $\sigma_0$ (resp., $\sigma_1$, $\sigma_\infty$) is the number of distinct points in the fiber above $0$ (resp., $1$, $\infty$). The ramification indices of the points in the fiber are given by the lengths of these cycles. \end{rmk} \subsection*{Genus, passports, and moduli} The genus of a Bely\u{\i}\ map is given by the Riemann-Hurwitz formula. If we define the \defi{excess} $e(\tau)$ of a cycle $\tau \in S_n$ to be its length minus one, and the excess $e(\sigma)$ of a permutation to be the sum of the excesses of its disjoint cycles (also known as the \defi{index} of the permutation, equal to $d$ minus the number of orbits), then the genus of a Bely\u{\i}\ map of degree $d$ with monodromy $\sigma$ is \begin{equation} \label{eqn:RH} g = 1-d + \frac{e(\sigma_0)+e(\sigma_1)+e(\sigma_\infty)}{2}. \end{equation} A \defi{refined passport} consists of the data of a transitive subgroup $G \leq S_d$ and three conjugacy classes $C_0,C_1,C_\infty$ in $G$. The refined passport of a triple $\sigma$ is given by the group $G= \langle \sigma \rangle$ and the conjugacy classes of the elements of $\sigma$. The \defi{size} of a refined passport is the cardinality of the set \[ \{\sigma \in S_d : \sigma_i \in C_i \text{ for $i=0,1,\infty$ and $\langle \sigma \rangle = G$}\}/\!\sim \] where $\sim$ denotes simultaneous conjugation in $S_d$. If a refined passport has size $1$, then we call the refined passport (and any associated triple) \defi{rigid}. For more on rigid triples, with applications to the inverse Galois problem, see e.g.\ Serre \cite{Serre}. For more on issues related to passports and moduli, see Sijsling--Voight \cite{SijslingVoight}. Let $X$ be a curve defined over $\mathbb C$. The \defi{field of moduli} $M(X)$ of $X$ is the fixed field of the group $\{\tau \in \Aut(\mathbb C) : X^{\tau} \cong X\}$, where $X^{\tau}$ is the base change of $X$ by the automorphism $\tau \in \Aut(\mathbb C)$ obtained by applying $\tau$ to any set of defining equations for $X$. If $F$ is a field of definition for $X$ then clearly $F \supseteq M(X)$. If $X$ has a minimal field of definition $F$ then necessarily $F=M(X)$. However, not all curves can be defined over their field of moduli. For more information, see work of Coombes and Harbater \cite{CoombesHarbater}, D\`ebes and Emsalem \cite{DebesEmsalem}, and K\"ock \cite{Kock}. Similarly, we define the field of moduli $M(\phi)$ of a Bely\u{\i}\ map $\phi$ to be the fixed field of $\{ \tau \in \Aut(C) : \phi^{\tau} \cong \phi\}$. The \defi{field of rationality} of a refined passport is the field $\mathbb Q(\chi(C_i))$ obtained by adjoining the character values on the conjugacy classes $C_0,C_1,C_\infty$ in $G$. In general, the degree of the field of moduli of a Bely\u{\i}\ map over its \defi{field of rationality} is bounded above by the size of its refined passport. Again, a field of definition of $\phi$ may be bigger than the field of moduli (itself bigger than the field of rationality), and these issues quickly become quite delicate: see Sijsling--Voight \cite[Section 7]{SijslingVoight}. \subsection*{Enumerating triples} To use this bijection to provide tables of Bely\u{\i}\ maps, we will need to enumerate triples: it is straightforward, using computational group theory techniques, to enumerate all transitive permutation triples up to simultaneous conjugacy. Indeed, for $d \leq 30$, the set of transitive groups of degree $d \leq 30$ have been classified \cite{Hulpke}, and these are available in \textsf{Magma} \cite{Magma} and given a unique identifier. However, the number of such triples grows rapidly with $d$, and so computing a complete database of triples is only practical for small degree. A simple way to compute these triples using double cosets is as follows. \begin{lem} \label{lem:c0c1} Let $G$ be a group, let $C_0,C_1$ be conjugacy classes in $G$ represented by $\tau_0,\tau_1$. Then the map \begin{align*} C_G(\tau_0) \backslash G / C_G(\tau_1) &\to \{(\sigma_0,\sigma_1) : \sigma_0 \in C_0, \sigma_1 \in C_1 \}/\!\sim_{G} \\ C_G(\tau_0) g C_G(\tau_1) &\mapsto (\tau_0, g \tau_1 g^{-1}) \end{align*} is a bijection, where the latter set is taken up to simultaneous conjugation in $G$. \end{lem} \noindent (Here $C_G(\tau_i)$ denotes the centralizer of $\tau_i$ in $G$.) One completes the triples arising from Lemma \ref{lem:c0c1} by taking $\sigma_\infty=(\sigma_1\sigma_0)^{-1}$; enumerating over all conjugacy classes of subgroups $G \leq S_d$ and all conjugacy classes $C_0,C_1$ in $G$ then lists all triples up to simultaneous conjugation. To avoid overcounting, one will want to restrict to those triples $\sigma$ obtained in Lemma \ref{lem:c0c1} which generate $G$ (and not a proper subgroup). \begin{proof} First, surjectivity. Let $\sigma_0 \in C_0$ and $\sigma_1 \in C_1$. Conjugating in $G$, we may assume that $\sigma_0=\tau_0$, and then indeed $\sigma_1=g \tau_1 g^{-1}$ for some $g \in G$. Next, injectivity: if $(\tau_0,g\tau_1 g^{-1})$ and $(\tau_0, g'\tau_1 (g')^{-1})$ are simultaneously conjugate in $G$ then there exists $c \in C_G(\tau_0)$ such that $cg\tau_1 (cg)^{-1} = g' \tau_1 (g')^{-1}$ so $(g')^{-1} cg \in C_G(\tau_1)$ and $C_G(\tau_0) g C_G(\tau_1) = C_G(\tau_0) g' C_G(\tau_1)$ as claimed. \end{proof} There are efficient algorithms for computing double cosets \cite{Linton}, so Lemma \ref{lem:c0c1} has an advantage over more naive enumeration. \subsection*{Spherical and Euclidean Bely\u{\i}\ maps} To conclude this section, we describe the calculation of the spherical and Euclidean Bely\u{\i}\ maps. The spherical triangle groups $\Delta(a,b,c)$ are finite groups acting on the sphere $\mathbb P^1$, and the corresponding quotients correspond to the classically known finite subgroups of $\PSL_2(\mathbb C)$. The case $(2,2,c)$ corresponds to dihedral groups and they are given explicitly by Chebyshev polynomials. This leaves the three cases $\Delta(2,3,3)=A_4$, $\Delta(2,3,4)=S_4$, and $\Delta(2,3,5)=A_5$ and these are referred to as the tetrahedral, octahedral, and icosahedral groups respectively because they are the groups of rigid motions of the corresponding Platonic solids. We refer the reader to Magnus \cite{Magnus} and the references therein for further discussion. The Bely\u{\i}\ maps for these cases are also classical: see Magot--Zvonkin \cite{MagotZvonkin} for references and a discussion of maps arising from Archimedean solids. (One can also view the spherical dessins using hyperbolic triples, including $\infty$: see recent work of He--Read \cite{HeRead} for a complete description.) The Euclidean groups $\Delta(a,b,c)$ are infinite groups associated to flat tori. First, we have that $\Delta(3,3,3) \leq \Delta(2,3,6)$ with index $2$, so it suffices to consider only the triples $(2,3,6)$ and $(2,4,4)$, associated to classical triangulations of the plane. First consider the triple $(2,3,6)$, to which we associate the elliptic curve $E(2,3,6) : y^2 = x^3 - 1$ with CM by $\mathbb Z[(-1+\sqrt{-3})/2]=\mathbb Z[\omega]$. The quotient map $x^3: E(2,3,6) \to \mathbb P^1$ of $E$ by its automorphism group (as an elliptic curve) $\Aut(E(2,3,6)) = \langle -\omega \rangle \cong \mathbb Z/6\mathbb Z$ gives a Galois Bely\u{\i}\ map of degree $6$ corresponding to the permutation triple \[ \sigma_0=(1\ 4)(2\ 5)(3\ 6), \quad \sigma_1=(1\ 3\ 5)(2\ 4\ 6),\quad \sigma_{\infty}=(1\ 2\ 3\ 4\ 5\ 6) \] and subgroup $\Gamma(E) \leq \Delta(2,3,6)$ of index $6$. For any other finite index subgroup $\Gamma$, the intersection $\Gamma \cap \Gamma(E)$ corresponds to an elliptic curve $X$ equipped with a isogeny $X \to E(2,3,6)$ induced by the inclusion $\Gamma \cap \Gamma(E) \hookrightarrow \Gamma(E)$. Similar statements hold for $\Delta(2,4,4)$, with the elliptic curve $E(2,4,4): y^2 = x^3 - x$ with CM by $\mathbb Z[i]$ and Bely\u{\i}\ map $x^2: E(2,4,4) \to \mathbb P^1$ of degree $4$. In other words, all Bely\u{\i}\ maps with these indices can be written down explicitly via CM theory of these two elliptic curves \cite{Silverman,Silverman2}, and so other methods are easier to apply. See also work of Singerman and Sydall \cite{SingermanSydall}. \section{Embedding Fuchsian triangle groups} \label{sec:embedtriangle} Let $(a,b,c)$ be a hyperbolic triple. In this section, we explicitly construct a hyperbolic triangle $T$ in the hyperbolic plane $H=\mathcal{H}$ with angles $\pi / a$, $\pi / b$, and $\pi / c$; see also Petersson \cite{Petersson}. We assume that one side of the triangle is on the imaginary axis with one vertex at $i$. We label the vertices $z_a, z_b,$ and $z_c$ corresponding to the angle at each vertex. Recall that the hyperbolic metric on $\mathcal{H}$ is given by \begin{equation} \label{eqn:hyperbolicmetric} \begin{aligned} d : \mathcal{H} \times \mathcal{H} &\to \mathbb R_{\geq 0} \\ \cosh d(x_1+iy_1, x_2+iy_2) &= 1 + \frac{(x_2-x_1)^2 + (y_2-y_1)^2}{2 y_1 y_2}. \end{aligned} \end{equation} Since $\PSL_2(\mathbb R)$ acts transitively on $\mathcal{H}$ by orientation-preserving isometries, we may take $z_a=i$; further composing by a local rotation around $i$, we may assume that $z_b=\mu i$ for some $\mu \in \mathbb R_{\geq 1}$. Finally, reflecting in the imaginary axis, we may assume that $\Real(z_c) > 0$. The triangle $T(a,b,c)$ is then unique, and in particular the positions of $z_b$ and $z_c$ are determined by insisting that $T(a,b,c)$ have angles $\pi / a,\pi / b,\pi / c$ at $z_a, z_b,z_c$, as in Figure \ref{fig:tabc}. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-02.pdf} \begin{comment} \begin{equation} \label{fig:tabc} \notag \psset{unit=.6cm} \begin{pspicture} \psline[linewidth=.5pt]{<->}(-4,0)(4,0) \psline[linewidth=.5pt]{<->}(0,-1)(0,14) \pscustom[linewidth=.5pt]{ \psarcn(1.732050){2}{150}{27.74601} \psarc(-22.74459){26.2631}{2.0317}{30} \psline(0,13.1315)(0,1) \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psarc(-1.73205){2}{30}{152.2539} \psarcn(22.74459){26.26319}{177.97}{150} \psline(0,13.1315)(0,1) \closepath} \uput{5pt}[-45]{0}(0,1){$z_a = i$} \uput{5pt}[-45]{0}(3.50209, 0.931105){$z_c$} \uput{5pt}[45]{0}(0,13.1315){$z_b = \mu i$} \uput{5pt}[225]{0}(-3.502090, 0.931105){$-\overline{z_c}$} \psarcn[linewidth=.5pt](0,1){0.7}{90}{51} \psarc[linewidth=.5pt](0,13.13159){0.9}{-90}{-60} \psarc[linewidth=.5pt](3.502090, 0.9311058){0.7}{93}{126} \uput{18pt}[67]{0}(0,1){$\pi/a$} \uput{20pt}[114]{0}(3.50209, 0.93110){$\pi/c$} \uput{35pt}[-77]{0}(0,13.1315){$\pi/b$} \psdots*(0,1)(0, 13.1315)(3.50209, 0.931105)(-3.50209, 0.93110) \end{pspicture} \end{equation} \end{comment} \caption{Normalized triangle with angles $\pi/a,\pi/b,\pi/c$ and its mirror, comprising $D_\Delta$.} \label{fig:tabc} \end{figure} We can solve for $\mu$ by recalling the law of cosines from hyperbolic trigonometry (see Ratcliffe \cite[Theorem 3.5.4]{Ratcliffe}). \begin{prop} \label{prop:lawcosines} If $\alpha, \beta, \gamma$ are the angles of a hyperbolic triangle and $C$ is the (hyperbolic) side length of the side opposite $\gamma$, then \begin{equation*} \cosh(C) = \frac{\cos \alpha \cos \beta + \cos \gamma}{\sin \alpha \sin \beta}. \end{equation*} \end{prop} Using (\ref{eqn:hyperbolicmetric}) and Proposition \ref{prop:lawcosines} we have $$ 1 + \frac{(\mu-1)^2}{2\mu} = \frac{\cos \pi/a \cos \pi/b + \cos \pi/c}{\sin \pi/a \sin \pi/b} $$ so writing $$ \lambda = \frac{\cos\pi/a\cos\pi/b+\cos\pi/c}{2\sin\pi/a\sin\pi/b} $$ we have $$ \mu = \lambda + \sqrt{\lambda^2 -1}. $$ Note that we are justified in taking the positive square root since we assume that $\mu > 1$. Having found $z_b$ we now find equations for the geodesics through $z_a$ and $z_b$ that intersect the imaginary axis with the angles $\pi /a$ and $\pi /b$ respectively. These geodesics are given by the intersection of $\mathcal{H}$ with the set of points $x+iy \in \mathbb C$ satisfying the equations \begin{equation} \label{eqn:circles} \begin{aligned} (x- \cot \pi/a)^2 + y^2 &= \csc^2 \pi/a \\ (x + \mu \cot \pi /b)^2 + y^2 &= \mu^2 \csc^2 \pi/b. \end{aligned} \end{equation} We then find the unique point of intersection of the two circles \eqref{eqn:circles} in $\mathcal{H}$ as \begin{equation} \label{eqn:zc} z_c = \frac{\mu^2-1}{2(\cot \pi/a + \mu \cot \pi/b)} + i\sqrt{\csc^2 \pi/a - \left( \frac{\mu ^2-1}{2(\cot \pi/a + \mu \cot \pi/b)} - \cot \pi/a \right ) ^2}. \end{equation} Note that $z_c \in \overline{\mathbb Q}$, in fact, $z_c$ lies in an at most quadratic extension of $\mathbb Q(\zeta_{2a},\zeta_{2b},\zeta_{2c})$, where $\zeta_s=\exp(2\pi i/s)$. We now find the elements in $\PSL_2(\mathbb R)$ that realize $\delta_a$ and $\delta_b$ geometrically, yielding an explicit embedding of $\Delta(a,b,c) \hookrightarrow \PSL_2(\mathbb R)$. Recall from section \ref{sec:background} that the image of $\delta_a$ in $\PSL_2(\mathbb R)$ is equal to $\tau_c \tau_b$. \begin{comment} \begin{equation} \label{fig:deltaa_deltab} \notag \psset{unit=1.75cm} \begin{pspicture} \psline[linewidth=.5pt]{<->}(-3,0)(3,0) \psline[linewidth=.5pt]{->}(0,0)(0,5) \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psline(0,1)(0,3.51221866389091100300080386381) \psarc(3.51221866389091100300080386381,0){4.96702726849443742278534092833}{135}{170.446835130473672273980047616} \psarcn(-0.577350269189625764509148780504,0){1.15470053837925152901829756101}{134.446835130473672273980047616}{60} \closepath} \pscustom[linewidth=.5pt]{ \psline(0,1)(0,3.51221866389091100300080386381) \psarcn(-3.51221866389091100300080386381,0){4.96702726849443742278534092833}{45}{9.55316486952632772601995238394} \psarc(0.577350269189625764509148780504,0){1.15470053837925152901829756101}{45.5531648695263277260199523841}{120} \closepath} \uput*{5pt}[-45]{0}(0,1){$z_a$} \uput*{5pt}[-45]{0}(1.38592601426635740621012014225,0.824341311476588330530384628265){$z_c$} \uput*{5pt}[45]{0}(0,3.51221866389091100300080386381){$z_b$} \uput*{5pt}[225]{0}(-1.38592601426635740621012014225,0.824341311476588330530384628265){$-\overline{z_c}$} \psarc[linewidth=.5pt]{->}(0,1){0.2}{26}{154} \psarc[linewidth=.5pt]{->}(0,3.51221866389091100300080386381){0.2}{-135}{-45} \uput{10pt}[55]{0}(0,1){$\delta_a$} \uput{11pt}[-65]{0}(0,3.51221866389091100300080386381){$\delta_b$} \psdots*(0,1)(0,3.51221866389091100300080386381)(1.38592601426635740621012014225,0.824341311476588330530384628265)(-1.38592601426635740621012014225,0.824341311476588330530384628265) \end{pspicture} \end{equation} \begin{center} \textbf{Figure \ref{fig:deltaa_deltab}}: The rotations $\delta_a = \tau_b\tau_c$ and $\delta_b = \tau_c\tau_a$ \\ \vspace{2ex} \end{center} \addtocounter{equation}{1} \end{comment} Therefore, the element of $\PSL_2(\mathbb R)$ corresponding to $\delta_a$ will fix $z_a$ and $\overline{z_a}$, while sending the vertex $z_c$ to $-\overline{z_c}$. Similarly, the element of $\PSL_2(\mathbb R)$ corresponding to $\delta_b$ will fix $z_b$ and $\overline{z_b}$, while sending the vertex $-\overline{z_c}$ to $z_c$. Since a linear fractional transformation is uniquely determined by its action on three points, these conditions uniquely determine the elements of $\PSL_2(\mathbb R)$ corresponding to $\delta_a$ and $\delta_b$. The element $\delta_a$ acts by rotation around $i$ by the angle $2\pi/a$, and conjugating by the matrix $\begin{pmatrix} \sqrt{\mu} & 0 \\ 0 & 1/\sqrt{\mu} \end{pmatrix}$ we obtain a matrix for $\delta_b$, as follows (cf.\ Petersson \cite{Petersson}). \begin{comment} Let \begin{align*} f : \mathbb C &\to\mathbb C \\ z &\mapsto \frac{s_{11} z+s_{12}}{s_{21} z+s_{22}} \end{align*} for $s_{11},s_{12},s_{21},s_{22} \in \mathbb R$ with $s_{11} s_{22}- s_{12} s_{21} \neq 0$, such that $f( \pm i \mu_0) = \pm i \mu_0 $ and $f(z_0) = -\overline{z_0}$ for $\mu_0 \in \mathbb R^\times$ and $z_0 \in \mathbb C$. Then \begin{align*} s_{21} z^2 + (s_{22}-s_{11})z - s_{12} &= s_{21}(z+i \mu_0)(z- i \mu_0) \\ &= s_{21}(z^2 + \mu_0^2), \end{align*} and since $f$ is only defined up to scaling $s_{11},s_{12},s_{21},s_{22}$ simultaneously by an element in $\mathbb R^\times$ we have $s_{11}=s_{22}$, $s_{21}=1$, and $s_{12} = -\mu_0^2$. Since $f(z_0) = -\overline{z_0}$, we may take $$ s_{11} = \frac{\mu_0^2 - z_0 \overline{z_0}}{z_0 + \overline{z_0}}. $$ Taking $\mu_0 = 1$ and $z_0 = z_c$ we obtain the matrix corresponding to the linear fractional transformation given by $$ \begin{pmatrix} \cot(\pi / a) & -1 \\ 1 & \cot(\pi/a) \end{pmatrix} $$ so by scaling through by $\sin(\pi /a)$ we obtain a matrix in $\SL_2(\mathbb R)$ corresponding to $\gamma_a$. Similarly, we find such a matix for $\gamma_b$ by taking $\mu_0 = \mu$, $z_0 = -\overline{z_c}$ and scaling the resulting matrix by $-\sin(\pi /b) / \mu$. \end{comment} \begin{prop} \label{prop:triangleembed} We have an embedding \begin{align*} \Delta(a,b,c) &\hookrightarrow \PSL_2(\mathbb R) \\ \delta_a &\mapsto \begin{pmatrix} \cos(\pi /a) & \sin(\pi /a) \\ -\sin(\pi /a) & \cos(\pi /a) \end{pmatrix} \\ \delta_b &\mapsto \begin{pmatrix} \cos(\pi / b) & \mu\sin(\pi / b) \\ -\sin(\pi / b )/\mu & \cos(\pi/b) \end{pmatrix} \end{align*} where \[ \mu = \lambda + \sqrt{\lambda^2 -1} \quad \text{and} \quad \lambda = \frac{\cos\pi/a\cos\pi/b+\cos\pi/c}{2\sin\pi/a\sin\pi/b}. \] In this embedding, the fixed points of $\delta_a,\delta_b,\delta_c$ are $z_a=i$, $z_b=\mu i$, and \[ z_c = \frac{\mu^2-1}{2(\cot \pi/a + \mu \cot \pi/b)} + i\sqrt{\csc^2 \pi/a - \left( \frac{\mu^2-1}{2(\cot \pi/a + \mu \cot \pi/b)} - \cot \pi/a \right ) ^2} \in \overline{\mathbb Q}. \] \end{prop} By conformally mapping $\mathcal{H}$ to the unit disc $\mathcal{D}$ by the map \begin{align*} w_p : \mathcal{H} &\to \mathcal{D} \\ z &\mapsto w_p(z) = \frac{z-p}{z-\overline{p}} \end{align*} for a point $p \in \mathcal{H}$, we can extend the action of $\Delta$ to $\mathcal{D}$ by the action \begin{align*} \Delta \times \mathcal{D} &\to \mathcal{D} \\ (\delta, w) &\mapsto w_p(\delta w_p^{-1}(w)). \end{align*} With the identification map $w_{z_a} = w_a : \mathcal{H} \to \mathcal{D}$ the element $\delta_a$ then acts on $\mathcal{D}$ by rotation about $2\pi / a$ about the origin. To avoid excessive notation, when the identification $w_p : \mathcal{H} \to \mathcal{D}$ is clear, we will just identify $D_\Delta$ with its image $w_p(D_\Delta) \subset \mathcal{D}$. \section{Coset enumeration and reduction theory} \label{sec:cosets} In this section, we exhibit an algorithm that takes as input a permutation triple $\sigma$ and gives as output the cosets of the point stabilizer subgroup $\Gamma \leq \Delta(a,b,c)$; this algorithm also yields a reduction algorithm and a fundamental domain for the action of $\Gamma$ on $\mathcal{H}$. Some similar methods were considered in the context of curves with signature $(1;e)$ by Sijsling \cite{Sijsling}. Throughout, we will use implicitly the conventions of computational group theory for representing finitely presented groups in bits as well as basic algorithms for manipulating them: for a reference, see Holt \cite{Holt} or Johnson \cite{Johnson}. Throughout, let $\Delta=\Delta(a,b,c)$ be a triangle group with indices $a,b,c \in \mathbb Z_{\geq 2}$ and let $\Gamma \leq \Delta$ be a subgroup of finite index $d$. \subsection*{Coset graph} The set $\{\delta_a^{\pm 1}, \delta_b^{\pm 1}\}$ of the generators of $\Delta(a,b,c)$ and their inverses acts on the cosets of $\Gamma$. We can represent this action as a directed, vertex- and edge-labelled graph with multiple edges allowed. This multidigraph is similar to the Cayley graph of a group; however, since $\Gamma$ is not assumed to be normal the coset space does not necessarily form a group. Our algorithm is similar to the classical Todd-Coxeter algorithm (see e.g.\ Rotman \cite{Rotman}) for enumerating the cosets of a subgroup of a group given by a presentation, but it is tailored for our specific application. The \defi{(symmetric) Cayley graph} for $\Delta$ on the generators $\delta_a,\delta_b$ is a symmetric, directed $4$-regular graph (i.e., every vertex has indegree and outdegree equal to 4) with vertex set $\Delta$ and edge set \[ \{\delta \xrightarrow{\epsilon} \delta \epsilon: \delta \in \Delta,\epsilon = \delta_a^{\pm 1}, \delta_b^{\pm 1}\}. \] Consider the quotient of the Cayley graph by the group $\Gamma$, where $\Gamma$ acts on the vertices on the left, identifying $w \sim \gamma w$ for $\gamma \in \Gamma$. In the quotient, edges map to edges, and so we obtain again a symmetric directed $4$-regular graph with edges labelled by $\delta_a^{\pm 1},\delta_b^{\pm 1}$ and now with vertex set equal to $\Gamma \backslash \Delta$. \begin{exm} Let $\Gamma \leq \Delta = \Delta(5,6,4)$ correspond to the permutation triple $\sigma = (\sigma_0,\sigma_1,\sigma_\infty)$ with \[ \sigma_0 = (1\;5\;4\;3\;2),\ \sigma_1 = (1\;6\;4\;2\;3\;5),\ \sigma_\infty = (1\;4\;3\;6). \] In the top left of Figure \ref{fig:QuotientGraph} is the Cayley graph for $\Delta$ which we quotient by $\Gamma$ to get the quotient graph in the top right. The coset graphs (bottom) exactly resemble the quotient graph. The only difference is that in the quotient graph the vertices are the cosets, whereas in the coset graphs we choose a particular coset representative to label each vertex. In this way there are many different coset graphs for the same quotient graph. \end{exm} \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-03.pdf} \begin{comment} \begin{equation} \label{fig:QuotientGraph} \psset{unit=2cm} \begin{pspicture}(-5,-5)(5,5) \def\CayleyGraph{% \cnode(0,0){3pt}{id} \cnode(1,0){3pt}{A} \cnode(-1,0){3pt}{Ainverse} \cnode(0,1){3pt}{B} \cnode(0,-1){3pt}{Binverse} \psnode(.5,1){BA}{$\cdots$} \psnode(0,1.5){BB}{$\vdots$} \psnode(-.5,1){BAinverse}{$\cdots$} \psnode(-1,.5){AinverseB}{$\vdots$} \psnode(-1.5,0){AinverseAinverse}{$\cdots$} \psnode(-1,-.5){AinverseBinverse}{$\vdots$} \psnode(-.5,-1){BinverseAinverse}{$\cdots$} \psnode(0,-1.5){BinverseBinverse}{$\vdots$} \psnode(.5,-1){BinverseA}{$\cdots$} \psnode(1,-.5){ABinverse}{$\vdots$} \psnode(1.5,0){AA}{$\cdots$} \psnode(1,.5){AB}{$\vdots$} \ncline{->}{id}{A}\naput{$\delta_a$} \ncline{->}{Ainverse}{id} \ncline{->}{Binverse}{id} \ncline{->}{id}{B}\naput{$\delta_b$} \ncline{->}{B}{BA} \ncline{->}{B}{BB} \ncline{->}{BAinverse}{B} \ncline{->}{Ainverse}{AinverseB} \ncline{->}{AinverseAinverse}{Ainverse} \ncline{->}{AinverseBinverse}{Ainverse} \ncline{->}{BinverseAinverse}{Binverse} \ncline{->}{BinverseBinverse}{Binverse} \ncline{->}{Binverse}{BinverseA} \ncline{->}{ABinverse}{A} \ncline{->}{A}{AA} \ncline{->}{A}{AB} } \def\Quotient{% \psset{unit=4cm} \cnodeput(0.40453,0.00000){1}{$\Gamma$} \cnodeput(0.12501,0.38473){5}{$\delta_a\Gamma$} \cnodeput(-0.32727,0.23778){4}{$\delta_a^2\Gamma$} \cnodeput(-0.32727,-0.23778){3}{$\delta_a^{-2}\Gamma$} \cnodeput(0.12501,-0.38473){2}{$\delta_a^{-1}\Gamma$} \cnodeput(0.77987,-0.16657){6}{$\delta_b\Gamma$} \ncline{->}{1}{5}\naput{$\delta_a$} \ncline{->}{5}{4}\naput{$\delta_a$} \ncline{->}{4}{3}\naput{$\delta_a$} \ncline{->}{3}{2}\nbput{$\delta_a$} \ncline{->}{2}{1}\naput{$\delta_a$} \ncline{->}{1}{6}\naput{$\delta_b$} \ncarc[arcangle=30]{->}{5}{1}\naput{$\delta_b$} \ncarc[arcangle=-30]{->}{2}{3}\nbput{$\delta_b$} \ncangles[angleA=180,angleB=140,armA=.75cm,linearc=.15]{->}{3}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{4}{2}\naput{$\delta_b$} \ncangles[angleA=90,angleB=90,armA=1.5cm,linearc=.15]{<-}{4}{6}\nbput{$\delta_b$} \nccircle[angleA=-90]{<-}{6}{.7cm}\naput{$\delta_a$} } \def\CosetGraphOne{% \psset{unit=4cm} \cnodeput(0.40453,0.00000){1}{$1$} \cnodeput(0.12501,0.38473){5}{$\delta_a$} \cnodeput(-0.32727,0.23778){4}{$\delta_a^2$} \cnodeput(-0.32727,-0.23778){3}{$\delta_a^{-2}$} \cnodeput(0.12501,-0.38473){2}{$\delta_a^{-1}$} \cnodeput(0.77987,-0.16657){6}{$\delta_b$} \ncline{->}{1}{5}\naput{$\delta_a$} \ncline{->}{5}{4}\naput{$\delta_a$} \ncline{->}{4}{3}\naput{$\delta_a$} \ncline{->}{3}{2}\nbput{$\delta_a$} \ncline{->}{2}{1}\naput{$\delta_a$} \ncline{->}{1}{6}\naput{$\delta_b$} \ncarc[arcangle=30]{->}{5}{1}\naput{$\delta_b$} \ncarc[arcangle=-30]{->}{2}{3}\nbput{$\delta_b$} \ncangles[angleA=180,angleB=180,armA=.75cm,linearc=.15]{->}{3}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{4}{2}\naput{$\delta_b$} \ncangles[angleA=90,angleB=90,armA=1.5cm,linearc=.15]{<-}{4}{6}\nbput{$\delta_b$} \nccircle[angleA=-90]{<-}{6}{.7cm}\naput{$\delta_a$} } \def\CosetGraphTwo{% \psset{unit=5cm} \cnodeput(0.40453,0.00000){1}{$\delta_b\delta_a\delta_b^{-1}$} \cnodeput(0.12501,0.38473){5}{$\delta_a\delta_b^{-1}\delta_a^{-1}$} \cnodeput(-0.32727,0.23778){4}{$\delta_a^2\delta_b^{-1}\delta_a^{-1}$} \cnodeput(-0.32727,-0.23778){3}{$\delta_a^{-2}\delta_b^{-1}\delta_a^{-1}$} \cnodeput(0.12501,-0.38473){2}{$\delta_a^{-1}\delta_b^{-1}\delta_a^{-1}$} \cnodeput(0.77987,-0.16657){6}{$\delta_b\delta_a^2\delta_b ^{-2}$} \ncline{->}{1}{5}\naput{$\delta_a$} \ncline{->}{5}{4}\naput{$\delta_a$} \ncline{->}{4}{3}\naput{$\delta_a$} \ncline{->}{3}{2}\nbput{$\delta_a$} \ncline{->}{2}{1}\naput{$\delta_a$} \ncline{->}{1}{6}\naput{$\delta_b$} \ncarc[arcangle=30]{->}{5}{1}\naput{$\delta_b$} \ncarc[arcangle=-30]{->}{2}{3}\nbput{$\delta_b$} \ncangles[angleA=180,angleB=140,armA=1cm,linearc=.15]{->}{3}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{4}{2}\naput{$\delta_b$} \ncangles[angleA=90,angleB=90,armA=1.5cm,linearc=.15]{<-}{4}{6}\nbput{$\delta_b$} \nccircle[angleA=-90]{<-}{6}{.8cm}\naput{$\delta_a$} } \rput{0}(-4,2){\CayleyGraph} \rput{0}(-0.25,2){\Quotient} \rput{0}(-4.5,-1.5){\CosetGraphOne} \rput{0}(-0.2,-1.5){\CosetGraphTwo} \end{pspicture} \end{equation} \end{comment} \caption{Quotient and coset graphs for $\Gamma$ the subgroup of $\Delta(5,6,4)$ corresponding to the permutation triple $(\sigma_0,\sigma_1,\sigma_{\infty})$ with $\sigma_0 = (1\;5\;4\;3\;2)$, $\sigma_1 = (1\;6\;4\;2\;3\;5)$, and $\sigma_\infty = (1\;4\;3\;6)$.} \label{fig:QuotientGraph} \end{figure} We will need to choose representatives for the cosets (vertex set) of the aforementioned quotient graph. So we define a \defi{coset graph} $G=G(\Gamma \backslash \Delta)$ for a permutation representation (\ref{eqn:piperm}) to be the quotient graph of the Cayley graph for $\Delta$ by the group $\Gamma$, as above, but with vertices labelled $\alpha_i$ for $i=1,\dots,d$ representing the cosets $\Gamma \alpha_i$ such that equation (\ref{eqn:permtocoset}) is satisfied. In both the Cayley graph of $\Delta$ and in its quotient by $\Gamma$ with labels $\delta_a$ or $\delta_b$, we have the corresponding reverse edge labeled with $\delta_a^{-1}$ or $\delta_b^{-1}$ respectively. Let $G$ be a coset graph for $\Gamma$ with vertex labels $\alpha_i$. Then we have a corresponding fundamental domain $D_\Gamma = D_\Gamma(G)$ for $\Gamma$ given by \begin{equation} \label{eqn:fundGamma} D_\Gamma = \bigcup_{i=1}^d \alpha_i D_\Delta. \end{equation} where $D_\Delta$ is as in Proposition \ref{prop:triangleembed}. The connectedness of the corresponding fundamental domain depends on how we label the vertices in the coset graph. If $e:v(\alpha_i) \xrightarrow{\epsilon} v(\alpha_j)$ is an edge of $G$, then we obtain a geodesic segment \[ s(e)=\alpha_i D_\Delta \cap \alpha_i\epsilon D_\Delta \subseteq D_\Gamma. \] Those edges that correspond to geodesics on the boundary of $D_\Gamma$ we call the \defi{sides} of $G$. Not all edges of $G$ are sides: those edges that are not sides correspond to geodesics in the interior of $D_\Gamma$. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-04-edit.pdf} \begin{comment} \begin{equation} \label{fig:EdgesSides} \notag \psset{unit=4.5cm} \begin{pspicture}(-5,-5)(5,5) \def\FundamentalDomain{% \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.80906,0.00000) \psarc[linewidth=.1pt] (1.0225,-0.36977) {0.42697} {120.00} {189.00} \psline[linewidth=.1pt] (0.60084,-0.43653) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.80906,0.00000) \psarcn[linewidth=.1pt] (1.0225,0.36977) {0.42697} {240.00} {171.00} \psline[linewidth=.1pt] (0.60084,0.43653) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.25001,-0.76946) \psarc[linewidth=.1pt] (-0.035688,-1.0867) {0.42696} {47.998} {117.00} \psline[linewidth=.1pt] (-0.22950,-0.70633) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.25001,-0.76946) \psarcn[linewidth=.1pt] (0.66766,-0.85822) {0.42697} {168.00} {99.004} \psline[linewidth=.1pt] (0.60084,-0.43653) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.65454,-0.47555) \psarc[linewidth=.1pt] (-1.0446,-0.30191) {0.42694} {336.00} {45.002} \psline[linewidth=.1pt] (-0.74268,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.65454,-0.47555) \psarcn[linewidth=.1pt] (-0.60992,-0.90016) {0.42695} {95.999} {27.000} \psline[linewidth=.1pt] (-0.22950,-0.70633) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.65454,0.47555) \psarc[linewidth=.1pt] (-0.60992,0.90016) {0.42695} {264.00} {333.00} \psline[linewidth=.1pt] (-0.22950,0.70633) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.65454,0.47555) \psarcn[linewidth=.1pt] (-1.0446,0.30191) {0.42694} {23.998} {315.00} \psline[linewidth=.1pt] (-0.74268,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.25001,0.76946) \psarc[linewidth=.1pt] (0.66766,0.85822) {0.42697} {192.00} {261.00} \psline[linewidth=.1pt] (0.60084,0.43653) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.25001,0.76946) \psarcn[linewidth=.1pt] (-0.035688,1.0867) {0.42696} {312.00} {243.00} \psline[linewidth=.1pt] (-0.22950,0.70633) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.0225,-0.12325) {0.24650} {230.24} {150.00} \psarc[linewidth=.1pt] (1.0225,0.00000) {0.21346} {180.00} {244.22} \psarc[linewidth=.1pt] (0.96302,-0.28785) {0.10126} {109.22} {194.23} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.0225,-0.12325) {0.24650} {230.24} {150.00} \psarc[linewidth=.1pt] (1.0225,-0.36977) {0.42697} {120.00} {189.00} \psarcn[linewidth=.1pt] (0.84509,-0.61399) {0.30191} {144.00} {86.242} \closepath} \rput(0.71692,-0.34063) {\scalebox{0.82101} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}}} \rput(0.87383,-0.23987) {\scalebox{0.82101} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}}} \rput(0.63716,0.18595) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(0.37377,0.54851) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.020055,0.66344) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(-0.62477,-0.22409) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(0.020055,-0.66345) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(-0.40620,-0.52495) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(0.37372,-0.54852) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}} \rput(-0.40618,0.52496) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}} \rput(-0.62479,0.22410) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{6}$}} \rput(0.84171,-0.11348) {\scalebox{0.82101} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{6}$}}} \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.80906,0.00000) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.80906,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.60084,-0.43653) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.25001,-0.76946) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.25001,-0.76946) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.22950,-0.70633) \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.65454,-0.47555) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.65454,-0.47555) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.74268,0.00000) \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.65454,0.47555) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.65454,0.47555) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.22950,0.70633) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.25001,0.76946) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.25001,0.76946) \psdots*[dotstyle=x,dotsize=7.0000pt](0.60084,0.43653) \psarcn[linewidth=1.2315pt] (1.0225,-0.12325) {0.24650} {230.24} {150.00} \psdots*[dotstyle=o,dotsize=5.7471pt](0.86488,-0.31274) \psdots*[dotstyle=*,dotsize=5.7471pt](0.80906,0.00000) \psdots*[dotstyle=x,dotsize=5.7471pt](0.92969,-0.19223) \rput(0.40453,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.12501,-0.38473) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.32727,-0.23778) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(-0.32727,0.23778) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(0.12501,0.38473) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \rput(0.77987,-0.16657) {\scalebox{0.82101} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}}} } \def\CosetGraph{% \cnodeput(0.40453,0.00000){1}{$1$} \cnodeput(0.12501,0.38473){5}{$\delta_a$} \cnodeput(-0.32727,0.23778){4}{$\delta_a^2$} \cnodeput(-0.32727,-0.23778){3}{$\delta_a^{-2}$} \cnodeput(0.12501,-0.38473){2}{$\delta_a^{-1}$} \cnodeput(0.77987,-0.16657){6}{$\delta_b$} \ncline[doubleline=true]{->}{1}{5}\naput{$\delta_a$} \ncline[doubleline=true]{->}{5}{4}\naput{$\delta_a$} \ncline[doubleline=true]{->}{4}{3}\naput{$\delta_a$} \ncline[doubleline=true]{->}{3}{2}\nbput{$\delta_a$} \ncline[doubleline=true]{->}{2}{1}\naput{$\delta_a$} \ncline[doubleline=true]{->}{1}{6}\naput{$\delta_b$} \ncarc[arcangle=30]{->}{5}{1}\naput{$\delta_b$} \ncarc[arcangle=-30]{->}{2}{3}\nbput{$\delta_b$} \ncangles[angleA=180,angleB=180,armA=.75cm,linearc=.15]{->}{3}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{4}{2}\naput{$\delta_b$} \ncangles[angleA=90,angleB=90,armA=1.5cm,linearc=.15]{<-}{4}{6}\nbput{$\delta_b$} \nccircle[angleA=-90]{<-}{6}{.7cm}\naput{$\delta_a$} } \rput{0}(-2.25,0){\CosetGraph} \rput{0}(-0.15,0){\FundamentalDomain} \end{pspicture} \end{equation} \end{comment} \caption{The two coset graphs in Figure \ref{fig:QuotientGraph} and their corresponding fundamental domains.} \label{fig:EdgesSides} \end{figure} \begin{exm} In Figure \ref{fig:EdgesSides}, the fundamental domain corresponding to the coset graph on the top is connected whereas the fundamental domain corresponding to the coset graph on the bottom is not. The double arrows are the edges of $G$ that are not sides, and the single arrows are the sides of $G$. In the bottom left of Figure \ref{fig:EdgesSides} is another choice of coset graph for $\Gamma$ as in Figure \ref{fig:QuotientGraph} with the corresponding fundamental domain $D_\Gamma$ on the right. \end{exm} Let $S$ be the set of sides of $D_\Gamma$. A \defi{side pairing} of $G$ is a set of elements \begin{equation} \label{eqn:SG} S(G) = \{ (\gamma, s, s^*) \in \Gamma \times S \times S : \gamma s^* = s \} \end{equation} that defines a labeled equivalence relation on $S$ that induces a partition of $S$ into pairs. The elements $\gamma$ in (\ref{eqn:SG}) are called the \defi{side pairing elements} of the side pairing $S(G)$. In bits, we specify a side $s\in S$ as a pair $(i,\epsilon)$ denoting the unique edge $e:v(\alpha_i) \xrightarrow{\epsilon} v(\alpha_j)$ in $G$ corresponding to $s$. A side pairing always exists: we give a constructive proof of this fact in the next subsection. \subsection*{Coset algorithm} In Algorithm~\ref{alg:cosetalg}, we present an algorithm that to compute a coset graph. The intuition behind this algorithm is as follows. We start in the domain $D_\Delta$, corresponding to the identity coset with representative $\alpha_1=1$. We then consider those translates of the domain $D_\Delta$ which are adjacent to $D_\Delta$ (in the sense that they share a side): with normalizations as in Proposition \ref{prop:triangleembed}, they are $\delta_a^{\pm 1} D_\Delta$ and $\delta_b^{\pm 1} D_\Delta$. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-05.pdf} \begin{comment} \begin{equation} \label{fig:translates} \notag \psset{unit=7cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.66874,0.00000) \psarc[linewidth=.001pt] (1.0820,-0.41330) {0.58449} {135.00} {189.00} \psline[linewidth=.001pt] (0.50474,-0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.66874,0.00000) \psarcn[linewidth=.001pt] (1.0820,0.41330) {0.58449} {225.00} {171.00} \psline[linewidth=.001pt] (0.50474,0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.00000,0.66874) \psarc[linewidth=.001pt] (0.41331,1.0820) {0.58450} {225.00} {279.00} \psline[linewidth=.001pt] (0.50474,0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.00000,0.66874) \psarcn[linewidth=.001pt] (-0.41331,1.0820) {0.58450} {315.00} {261.00} \psline[linewidth=.001pt] (-0.50474,0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.00000,-0.66874) \psarc[linewidth=.001pt] (-0.41331,-1.0820) {0.58450} {45.000} {98.998} \psline[linewidth=.001pt] (-0.50474,-0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.00000,-0.66874) \psarcn[linewidth=.001pt] (0.41331,-1.0820) {0.58450} {135.00} {81.000} \psline[linewidth=.001pt] (0.50474,-0.50474) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psarcn[linewidth=.001pt] (1.0820,0.00000) {0.41331} {228.19} {180.00} \psarc[linewidth=.001pt] (1.0820,0.41327) {0.58448} {225.00} {253.33} \psarc[linewidth=.001pt] (0.96780,-0.29910) {0.16157} {109.32} {183.18} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.001pt] (1.0820,0.00000) {0.41331} {228.19} {180.00} \psarc[linewidth=.001pt] (1.0820,-0.41330) {0.58449} {135.00} {189.00} \psarcn[linewidth=.001pt] (0.82661,-0.66875) {0.36125} {153.00} {93.188} \closepath} \pscustom[linewidth=.001pt]{ \psarc[linewidth=.001pt] (1.0820,0.00000) {0.41331} {131.81} {180.00} \psarcn[linewidth=.001pt] (1.0820,0.41330) {0.58449} {225.00} {171.00} \psarc[linewidth=.001pt] (0.82661,0.66875) {0.36125} {207.00} {266.81} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.001pt] (1.0820,0.00000) {0.41331} {131.81} {180.00} \psarcn[linewidth=.001pt] (1.0820,-0.41327) {0.58448} {135.00} {106.67} \psarcn[linewidth=.001pt] (0.96780,0.29910) {0.16157} {250.68} {176.82} \closepath} \pscustom[linewidth=.001pt]{ \psarcn[linewidth=.001pt] (0.00000,1.0820) {0.41331} {318.19} {270.00} \psarc[linewidth=.001pt] (-0.41327,1.0820) {0.58448} {315.00} {343.32} \psarc[linewidth=.001pt] (0.29910,0.96780) {0.16157} {199.32} {273.18} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.001pt] (0.00000,1.0820) {0.41331} {318.19} {270.00} \psarc[linewidth=.001pt] (0.41331,1.0820) {0.58450} {225.00} {279.00} \psarcn[linewidth=.001pt] (0.66874,0.82660) {0.36124} {243.00} {183.19} \closepath} \pscustom[linewidth=.001pt]{ \psarc[linewidth=.001pt] (0.92420,-0.41327) {0.15786} {138.20} {206.57} \psarc[linewidth=.001pt] (0.66872,-0.82662) {0.36126} {71.562} {117.00} \psarcn[linewidth=.001pt] (0.82661,-0.66875) {0.36125} {153.00} {93.188} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.001pt] (0.92420,-0.41327) {0.15786} {138.20} {206.57} \psarcn[linewidth=.001pt] (0.86395,-0.51075) {0.085304} {161.66} {87.638} \psarcn[linewidth=.001pt] (0.96780,-0.29910) {0.16152} {231.59} {183.18} \closepath} \psdots*(0,0)(0.668740305001847445964813232422,0) \uput{5pt}[180]{0}(0,0){$w_a$} \uput{5pt}[0]{0}(0.668740305001847445964813232422,0){$w_b$} \psarc{->}(0,0){0.1}{0}{90} \psarc{->}(0.668740305001847445964813232422,0){0.1}{180}{278} \rput(0.33437,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.00000,0.33437) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_a$}} \rput(0.00000,-0.33437) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_a^{-1}$}} \rput(0.70475,0.16873) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_b^{-1}$}} \rput(0.16874,0.70476) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_a \delta_b$}} \rput(0.70475,-0.16873) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_b$}} \rput(0.76773,-0.39236) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$\delta_b \delta_a$}} \end{pspicture} \end{equation} \end{comment} \caption{The translates $\delta_a^{\pm 1}D_\Delta, \delta_b^{\pm 1}D_\Delta, \delta_a\delta_bD_\Delta, \delta_b\delta_aD_\Delta$.} \label{fig:translates} \end{figure} We check each of these cosets if they are old (in the start, equivalent to the identity): those that are not, we add to the \defi{frontier}, labelling them in the graph; those that are, we find an element in $\Gamma$ matching it with an existing coset. We continue in this manner until the frontier is exhausted. \begin{alg} \label{alg:cosetalg} Let $\sigma \in S_d^3$ be a transitive permutation triple. This algorithm returns a coset graph $G=G(\Gamma \backslash \Delta)$ for the group $\Gamma \leq \Delta$ associated to $\sigma$ and a side pairing for $G$. \begin{enumalg} \item Initialize a graph $G$ with $d$ vertices $v_1,\dots,v_d$, label $v_1$ with $1 \in \Delta$. Initialize the empty side pairing $S(G)$. \item Let $v_j$ be the first vertex with no out edges. If no such vertex exists, return $G,S(G)$. \item For $\epsilon \in \{\delta_a,\delta_a^{-1},\delta_b,\delta_b^{-1}\}$: \begin{enumalgalph} \item Compute $i :=1^{\pi(\alpha_j\epsilon)}$. Draw a directed edge from $v_j$ to $v_i$ with label $\epsilon$. \item If vertex $v_i$ is labelled, let $\gamma := \alpha_j\epsilon\alpha_i^{-1}$. If further $\gamma \neq 1$, then add $(\gamma,(j,\epsilon),(i,\epsilon^{-1}))$ to $S(G)$. \item If vertex $v_i$ is not labelled, label $v_i$ with $\alpha_i := \alpha_j \epsilon$. \end{enumalgalph} \item Return to Step 2. \end{enumalg} \end{alg} In Step 3b, we only add the side pairing if $\gamma \neq 1$; this removes spurious (non-boundary) sides that are contained inside the fundamental domain $D_\Gamma$. We can check if $\gamma=1$ exactly by computing the linear transformation corresponding to $\gamma$ as in Proposition \ref{prop:triangleembed} over the number field $K$ such that $\Delta(a,b,c) \hookrightarrow \PSL_2(K)$. \begin{rmk} In fact, the $\mathbb Q$-algebra generated by any lift of the matrices $\delta_a,\delta_b,\delta_c$ to $\SL_2(\mathbb R)$ is a quaternion algebra over a smaller totally real number field \cite{ClarkVoight}, so the complexity of the check that $\gamma=1$ can be reduced somewhat. Alternatively, since the group is discrete, we can check if $\gamma z = z$ for some point $z$ in the interior of $D_\Delta$ up to some precision depending on the group. \end{rmk} \begin{proof}[Proof of correctness of Algorithm \ref{alg:cosetalg}] First, we show that the output graph $G$ of the algorithm is indeed a coset graph for $\Gamma$. Let $k \in \{1,...,d\}$ and consider the vertex $v_k$ of $G$. Since $\sigma$ is a transitive permutation triple, there exists $\delta \in \Delta$ such that $1^{\pi(\delta)} = k$. By writing $\delta$ as a word in $\delta_a$ and $\delta_b$, we find that there will be a directed path from the vertex labeled $1$ to the vertex $v_k$. Specifically, this directed path is given by following the edges labeled by elements $\delta_a$ and $\delta_b$ in the word equaling $\delta$, read from left to right. Since an in-edge for $v_k$ is constructed by the algorithm, Step 2 of the algorithm will label the vertex $v_k$ with a specific coset representative. We have therefore shown that every vertex in $G$ is labeled with a coset representative and all of the cosets of $\Gamma$ in $\Delta$ are represented by exactly one vertex in $G$. It remains to show that the edges in the graph $G$ that is returned are such that $G$ is a coset graph. Consider the map from $G$ to the quotient graph of the Cayley graph of $\Delta$ acted on by $\Gamma$ that sends each vertex to the coset that the vertices label representatives and sends each labeled edge in $G$ to the edge with the same label in the quotient graph. As each vertex is considered in Step 3 of the algorithm, all of the incident edges of that vertex are constructed such that the aforementioned map between graphs is a graph isomorphism. Therefore the returned graph $G$ is in fact a coset graph. We now show that the set $S(G)$ that is returned by the algorithm is in fact a side pairing for $G$. First, assume that $(\gamma, (j, \epsilon), (i, \epsilon^{-1})) \in S(G)$ where $\gamma = \alpha_j \epsilon \alpha_i^{-1}$ is added in Step 3b of the algorithm. Then $1^{\pi(\alpha_i \epsilon)} = i = 1^{\pi(\alpha_j)}$ so we have \[ (1^{\pi(\alpha_j \epsilon)})^{\pi(\alpha_i^{-1})} = 1^{\pi(\alpha_j \epsilon \alpha_i^{-1})} = 1\] and therefore $\alpha_j \epsilon \alpha_i^{-1} = \gamma \in \Gamma$. Consider such an element $\gamma \in \Gamma$. We show that the edges $(j,\epsilon)$ and $(i, \epsilon^{-1})$ are sides of $D_\Gamma$ if and only if $\gamma \neq 1$. Indeed if $\gamma \neq 1$ then the two edges are distinct subsets of $D_\Gamma$ that are in the same $\Gamma$-orbit. Since $D_\Gamma$ is a fundamental domain for $\Gamma$, these two edges are on the boundary of $D_\Gamma$ and therefore are sides of $D_\Gamma$. Conversely, if two distinct sides of $D_\Gamma$ are paired by an element in $\gamma \in \Gamma$, then we must have $\gamma \neq 1$. Lastly, we show that $S(G)$ induces a partition of $S$ into pairs. Since the sides of $G$ are in bijection with a subset of the pairs $(j, \epsilon)$ with $j \in \{1,...,d \}$ and $\epsilon \in \{ \delta_a^{\pm 1}, \delta_b^{\pm 1} \}$ and all such pairs are considered in Step 3 of the algorithm, we know that each side will appear in a side pairing in the output $S(G)$. Furthermore, each side can be paired with at most one other side in the partition of $S$ that is induced by $S(G)$, since the value of $1^{\pi(\alpha_j \epsilon)}$ computed in Step 3a of the algorithm is unique. \end{proof} \begin{lem}\label{lem:connected}\notag Let $D_\Delta$ be a connected fundamental domain for $\Delta$ and let $D_\Gamma$ be the fundamental domain corresponding to a coset graph $G$ computed in Algorithm \ref{alg:cosetalg}. Then $D_\Gamma$ is connected. \end{lem} The conclusion of this lemma is not true in general for an arbitrarily chosen coset graph (cf. Figure \ref{fig:EdgesSides}). \begin{proof} Let $\epsilon \in \{\delta_a^{\pm 1}, \delta_b^{\pm 1} \}$. For any element $\delta \in \Delta$, note that $\delta(D_\Delta)$ and $\delta \epsilon (D_\Delta)$ share an edge and therefore $\delta(D_\Delta) \cup \delta \epsilon(D_\Delta)$ is connected. For example, if $\epsilon = \delta_a$, then the side between $\delta(w_a)$ and $\delta(\overline{w_c})$ of $\delta(D_\Delta)$ is connected to the side between $\delta \epsilon (w_a)$ and $\delta \epsilon(w_c)$ of $\delta \epsilon(D_\Delta)$. Note that for each adjacent pair of vertices in $G$, their representatives differ by multiplication on the left by some $\epsilon \in \{ \delta_a^{\pm 1}, \delta_b^{\pm 1} \}$. Therefore, the translates of $D_\Delta$ by the labels of adjacent vertices of $G$ are connected to one another. For a chosen coset representative $\alpha_i$ that is a label of a vertex of $G$, we showed that there is a directed path from the vertex labeled $1$ to the vertex labeled $\alpha_i$ in $G$. By following along this path we are writing $\alpha_i$ as a word in $\delta_a$ and $\delta_b$ from left to right, and all of the translates of $D_\Delta$ by the coset representatives labeling the vertices along this path are a part of $D_\Gamma$. Therefore, the union of the translates of $D_\Delta$ by all of the labels of the vertices in this path is a connected subset of $D_\Gamma$, that contains $\alpha_i(D_\Delta)$ and $D_\Delta$. Hence, all $D_\Delta$ translates by the elements that are the labels of our coset graph $G$ are in the connected component of $D_\Delta$ in $D_\Gamma$. Thus $D_\Gamma$ is connected. \end{proof} We propose also three potential variations on Algorithm \ref{alg:cosetalg}; these do not affect its correctness and can be used to obtain fundamental domains that are better (in some way). First, in Step 2, we can instead choose a vertex $v_j$ with label $\alpha_j$ such that $d(\alpha_j0,0)$ is minimal, where $d$ is hyperbolic metric. This can potentially yield a fundamental domain $D_\Gamma$ that is contained in a circle of smaller radius. Second, we can use a \defi{petalling} approach in which we first attempt to extend the fundamental domain by translates consisting of powers of $\delta_a$, before considering those with powers of $\delta_b$. This results in the following modification of Algorithm \ref{alg:cosetalg}. For $n \in \mathbb Z$, we define $\sgn(n)=+1,0,-1$ according as $n>0,n=0,n<0$. \begin{alg} \label{alg:petal} Let $\sigma \in S_d^3$ be a transitive permutation triple. This algorithm returns a coset graph $G=G(\Gamma \backslash \Delta)$ for the group $\Gamma \leq \Delta$ associated to $\sigma$ and a side pairing for $G$. \begin{enumalg} \item Initialize a graph $G$ with $d$ vertices $v_1,\dots,v_d$, label $v_1$ with $1 \in \Delta$. Initialize the empty side pairing $S(G)$. \item Let $v_j$ be the first vertex with fewer than $4$ out edges. If no such vertex exists, return $G,S(G)$. \item For $k=1,-1,2,-2,\dots$ until both $+1$ and $-1$ cases are done, do: \begin{enumalgalph} \item If the $\sgn(k)$ case is done, return to Step 3. \item Let $\epsilon := \delta_a^{\sgn(k)}$ and let $j' := j^{\sigma_a^{k-\sgn(k)}}$. \item Let $i := 1^{\pi(\alpha_{j'} \epsilon)}$. Draw a directed edge from $v_{j'}$ to $v_i$ with label $\epsilon$. \item If vertex $v_i$ is labelled, let $\gamma := \alpha_{j'}\epsilon\alpha_i^{-1}$, and declare that the $\sgn(k)$ case is done. If $\gamma \neq 1$, then add $(\gamma,(j',\epsilon),(i,\epsilon^{-1}))$ to $S(G)$. \item If vertex $v_i$ is not labelled, label $v_i$ with $\alpha_i := \alpha_{j'} \epsilon$. \end{enumalgalph} \item For $\epsilon\in\{\delta_b,\delta_b^{-1}\}$, repeat Steps 3c--3e with $j' := j$. \item Return to Step 2. \end{enumalg} \end{alg} In Step 3, we alternate signs in order to make the nicest pictures; it is equivalent to simply take $k=1,2,\dots$. Since the coset representatives labelling adjacent vertices in the coset graph produced by this approach still differ by multiplication on the right by $\epsilon\in\{\delta_a^{\pm 1},\delta_b^{\pm 1}\}$, the argument in the proof of Algorithm \ref{alg:cosetalg} applies here. By the same reasoning, the translates of $D_\Delta$ by the labels of adjacent vertices of $G$ are connected and Lemma \ref{lem:connected} also applies. \begin{exm} In this example, we compare the results of Algorithms (\ref{alg:cosetalg}) and (\ref{alg:petal}). Let $\Gamma\leq \Delta(6,6,7)$ corresponding to the permutations \[ \sigma_0 = (1\;3\;2\;6\;4\;5),\ \sigma_1 = (1\;4\;5\;3\;7\;6),\ \sigma_\infty = (1\;4\;7\;3\;6\;2\;5) . \] The advantage to preferring powers of $\delta_a$ is the resulting ``petalling" effect illustrated in Figure \ref{fig:CompareAlg6,6,7}. \end{exm} \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-06.pdf} \begin{comment} \begin{equation} \label{fig:CompareAlg6,6,7} \notag \psset{unit=3.25cm} \begin{pspicture}(-5,-5)(5,5) \def\Full{% \pscircle[linewidth=.1pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.85847,0.00000) \psarc[linewidth=.1pt] (1.0117,-0.26534) {0.30638} {120.00} {214.29} \psline[linewidth=.1pt] (0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.85847,0.00000) \psarcn[linewidth=.1pt] (1.0117,0.26534) {0.30638} {240.00} {145.71} \psline[linewidth=.1pt] (0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,0.74346) \psarc[linewidth=.1pt] (-0.27604,1.0088) {0.30639} {240.00} {334.29} \psline[linewidth=.1pt] (0.00000,0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,0.74346) \psarcn[linewidth=.1pt] (-0.73563,0.74346) {0.30639} {360.00} {265.71} \psline[linewidth=.1pt] (-0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,0.74346) \psarc[linewidth=.1pt] (0.73563,0.74346) {0.30639} {180.00} {274.29} \psline[linewidth=.1pt] (0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,0.74346) \psarcn[linewidth=.1pt] (0.27604,1.0088) {0.30639} {300.00} {205.72} \psline[linewidth=.1pt] (0.00000,0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.0117,-0.088472) {0.17686} {240.61} {149.99} \psarc[linewidth=.1pt] (1.0117,0.00000) {0.15318} {180.00} {253.88} \psarc[linewidth=.1pt] (0.97976,-0.21011) {0.063831} {99.583} {210.59} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.0117,-0.088472) {0.17686} {240.61} {149.99} \psarc[linewidth=.1pt] (1.0117,-0.26534) {0.30638} {120.00} {214.29} \psarcn[linewidth=.1pt] (0.92665,-0.41258) {0.17003} {188.58} {90.606} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,-0.74346) \psarc[linewidth=.1pt] (0.27604,-1.0088) {0.30639} {60.000} {154.28} \psline[linewidth=.1pt] (0.00000,-0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,-0.74346) \psarcn[linewidth=.1pt] (0.73563,-0.74346) {0.30639} {180.00} {85.714} \psline[linewidth=.1pt] (0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.0117,0.088472) {0.17686} {119.39} {210.01} \psarcn[linewidth=.1pt] (1.0117,0.26534) {0.30638} {240.00} {145.71} \psarc[linewidth=.1pt] (0.92665,0.41258) {0.17003} {171.42} {269.39} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.0117,0.088472) {0.17686} {119.39} {210.01} \psarcn[linewidth=.1pt] (1.0117,0.00000) {0.15318} {180.00} {106.12} \psarcn[linewidth=.1pt] (0.97976,0.21011) {0.063831} {260.42} {149.41} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.58244,0.83189) {0.17688} {300.61} {210.00} \psarc[linewidth=.1pt] (0.50585,0.87611) {0.15319} {239.99} {313.88} \psarc[linewidth=.1pt] (0.67185,0.74342) {0.063852} {159.57} {270.58} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.58244,0.83189) {0.17688} {300.61} {210.00} \psarc[linewidth=.1pt] (0.73563,0.74346) {0.30639} {180.00} {274.29} \psarcn[linewidth=.1pt] (0.82063,0.59622) {0.17004} {248.58} {150.61} \closepath} \rput{0}(0,-1.1){\text{Basic algorithm}} } \def\Petal{% \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.85847,0.00000) \psarc[linewidth=.1pt] (1.0117,-0.26534) {0.30638} {120.00} {214.29} \psline[linewidth=.1pt] (0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.85847,0.00000) \psarcn[linewidth=.1pt] (1.0117,0.26534) {0.30638} {240.00} {145.71} \psline[linewidth=.1pt] (0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,0.74346) \psarc[linewidth=.1pt] (-0.27604,1.0088) {0.30639} {240.00} {334.29} \psline[linewidth=.1pt] (0.00000,0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,0.74346) \psarcn[linewidth=.1pt] (-0.73563,0.74346) {0.30639} {360.00} {265.71} \psline[linewidth=.1pt] (-0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,0.74346) \psarc[linewidth=.1pt] (0.73563,0.74346) {0.30639} {180.00} {274.29} \psline[linewidth=.1pt] (0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,0.74346) \psarcn[linewidth=.1pt] (0.27604,1.0088) {0.30639} {300.00} {205.72} \psline[linewidth=.1pt] (0.00000,0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,-0.74346) \psarc[linewidth=.1pt] (-0.73563,-0.74346) {0.30639} {360.00} {94.285} \psline[linewidth=.1pt] (-0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.42924,-0.74346) \psarcn[linewidth=.1pt] (-0.27604,-1.0088) {0.30639} {120.00} {25.716} \psline[linewidth=.1pt] (0.00000,-0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,-0.74346) \psarc[linewidth=.1pt] (0.27604,-1.0088) {0.30639} {60.000} {154.28} \psline[linewidth=.1pt] (0.00000,-0.87586) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.42924,-0.74346) \psarcn[linewidth=.1pt] (0.73563,-0.74346) {0.30639} {180.00} {85.714} \psline[linewidth=.1pt] (0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.85847,0.00000) \psarc[linewidth=.1pt] (-1.0117,0.26534) {0.30638} {300.00} {34.287} \psline[linewidth=.1pt] (-0.75852,0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.85847,0.00000) \psarcn[linewidth=.1pt] (-1.0117,-0.26534) {0.30638} {60.001} {325.71} \psline[linewidth=.1pt] (-0.75852,-0.43793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.58244,0.83189) {0.17688} {300.61} {210.00} \psarc[linewidth=.1pt] (0.50585,0.87611) {0.15319} {239.99} {313.88} \psarc[linewidth=.1pt] (0.67185,0.74342) {0.063852} {159.57} {270.58} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.58244,0.83189) {0.17688} {300.61} {210.00} \psarc[linewidth=.1pt] (0.73563,0.74346) {0.30639} {180.00} {274.29} \psarcn[linewidth=.1pt] (0.82063,0.59622) {0.17004} {248.58} {150.61} \closepath} \rput{0}(0,-1.1){\text{Petalling approach}} } \rput{0}(-1.15,1.15){\Full} \rput{0}(1.15,1.15){\Petal} \end{pspicture} \end{equation} \end{comment} \caption{Comparing the petalling approach} \label{fig:CompareAlg6,6,7} \end{figure} A third possible variation is to find the ``smallest'' cosets in each case. More precisely, once we have computed a set of generators for $\Gamma$ coming from side pairing elements, we utilize the algorithm of Voight \cite{Voightfd} (the original idea is due to Ford) to compute the \defi{Dirichlet domain} for $\Gamma$, the set \[ \{w \in \mathcal{D} : d(0,\gamma w) \geq d(0,w) \text{ for all $\gamma \in \Gamma$}\} \] that picks out in each $\Gamma$ orbit the points closest to the origin. From there, we choose a point $p$ in the fundamental triangle $D_\Delta$ and, for each coset of $\Gamma$ in $\Delta$, we choose a representative $\alpha$ such that $\alpha p$ lies in the Dirichlet domain. In this way, we can compute a fundamental domain which has the property that it lies in a circle of radius $\rho$ with $\rho$ minimal among all possible fundamental domains that are triangulated by $D_\Delta$. But we choose not to employ this modification in practice for the following reasons. First, we find the petalling approach to be more natural for our calculations. Below, we compute power series centered at the elements of the orbit of $0$ and so it is enough to minimize the distance between an arbitrary in the fundamental domain to one of these centers. Second, we found in practice that the petalling approach produced more aesthetically pleasing dessins. Finally, it can be a bit expensive to compute the Dirichlet domain. That being said, it is possible that the Dirichlet domain may be useful in other variants of our method, so we mention it here. \begin{exm} Below in Figure \ref{fig:CompareAlgorithms} is a simple example comparing the variations of Algorithm \ref{alg:cosetalg}. The basic algorithm and petalling approach yield a fundamental domain inside the solid blue circle. The Dirichlet domain approach yields a fundamental domain inside the dashed red circle. When we compare these circles we see that the radius of the dashed red circle is smaller than the radius of the solid blue circle. In Figure \ref{fig:CompareAlgorithms}, we have three fundamental domains for $\Gamma \leq \Delta(3,5,3)$ corresponding to the permutations \[ \sigma_0 = (1\;3\;4)(2\;6\;5),\ \sigma_1 = (1\;4\;2\;5\;6),\ \sigma_\infty = (3\;6\;4). \] \end{exm} \begin{figure}[h] \includegraphics[scale=1]{belyi-triangle-pics/belyi-triangle-pics-07.pdf} \begin{comment} \begin{equation} \label{fig:CompareAlgorithms} \notag \psset{unit=3.25cm} \begin{pspicture}(-5,-5)(5,5) \def\Full{% \pscircle[linewidth=.1pt](0,0){1} \pscircle[linewidth=0.1pt,linecolor=blue](0,0){0.874903546969607} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarc[linewidth=.1pt] (1.2098,-0.93707) {1.1583} {126.00} {150.00} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (-0.79646,-0.93709) {0.71587} {74.478} {42.000} \psarc[linewidth=.1pt] (-0.41330,-1.1583) {0.71588} {78.000} {105.52} \psarc[linewidth=.1pt] (-1.0332,-0.35797) {0.44241} {345.53} {14.482} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (-0.79646,-0.93709) {0.71587} {74.478} {42.000} \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {330.00} {314.48} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarc[linewidth=.1pt] (0.20664,1.5162) {1.1583} {246.00} {270.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {354.00} {330.00} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarcn[linewidth=.1pt] (0.20664,-1.5162) {1.1583} {114.00} {89.999} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarc[linewidth=.1pt] (0.97298,0.35791) {0.27347} {152.84} {194.47} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarcn[linewidth=.1pt] (0.76114,0.67809) {0.19787} {260.85} {210.01} \psarc[linewidth=.1pt] (1.2098,0.22116) {0.71592} {150.00} {165.52} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,-0.22120) {0.71587} {162.00} {134.48} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \rput{0}(0,-1.1){\text{Basic algorithm}} } \def\Petal{% \pscircle[linewidth=.1pt](0,0){1} \pscircle[linewidth=0.1pt,linecolor=blue](0,0){0.874903546969607} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarc[linewidth=.1pt] (1.2098,-0.93707) {1.1583} {126.00} {150.00} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.82663,0.71587) {0.44244} {225.52} {186.00} \psarc[linewidth=.1pt] (0.58981,0.85258) {0.27344} {222.00} {270.00} \psarc[linewidth=.1pt] (1.2098,0.22116) {0.71592} {150.00} {165.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.82663,0.71587) {0.44244} {225.52} {186.00} \psline[linewidth=.1pt] (0.38660,0.66962) (0.20665,0.35793) \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarc[linewidth=.1pt] (0.20664,1.5162) {1.1583} {246.00} {270.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {354.00} {330.00} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarcn[linewidth=.1pt] (0.20664,-1.5162) {1.1583} {114.00} {89.999} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarc[linewidth=.1pt] (0.97298,0.35791) {0.27347} {152.84} {194.47} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarcn[linewidth=.1pt] (0.76114,0.67809) {0.19787} {260.85} {210.01} \psarc[linewidth=.1pt] (1.2098,0.22116) {0.71592} {150.00} {165.52} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,-0.22120) {0.71587} {162.00} {134.48} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \rput{0}(0,-1.1){\text{Petalling approach}} } \def\FSC{% \pscircle[linewidth=.1pt](0,0){1} \pscircle[linewidth=.1pt,linecolor=red,linestyle=dashed](0,0){0.826611605834324} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarc[linewidth=.1pt] (1.2098,-0.93707) {1.1583} {126.00} {150.00} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (-0.79646,-0.93709) {0.71587} {74.478} {42.000} \psarc[linewidth=.1pt] (-0.41330,-1.1583) {0.71588} {78.000} {105.52} \psarc[linewidth=.1pt] (-1.0332,-0.35797) {0.44241} {345.53} {14.482} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (-0.79646,-0.93709) {0.71587} {74.478} {42.000} \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {330.00} {314.48} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarc[linewidth=.1pt] (0.20664,1.5162) {1.1583} {246.00} {270.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {354.00} {330.00} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarcn[linewidth=.1pt] (0.20664,-1.5162) {1.1583} {114.00} {89.999} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.20670,-1.5163) {1.1584} {129.28} {114.00} \psline[linewidth=.1pt] (-0.26447,-0.45807) (-0.41330,-0.71587) \psarc[linewidth=.1pt] (-0.60488,-0.82646) {0.22121} {29.997} {69.273} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.20670,-1.5163) {1.1584} {129.28} {114.00} \psarc[linewidth=.1pt] (-0.41330,-1.1583) {0.71588} {78.000} {105.52} \psarcn[linewidth=.1pt] (-0.79649,-0.66361) {0.27343} {45.514} {9.2686} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,-0.22120) {0.71587} {162.00} {134.48} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \rput{0}(0,-1.1){\text{Dirichlet domain approach}} } \def\CompareRadii{% \pscircle[linewidth=.1pt](0,0){1} \pscircle[linewidth=.1pt,linecolor=blue](0,0){0.874903546969607} \pscircle[linewidth=.1pt,linecolor=red,linestyle=dashed](0,0){0.826611605834324} \rput{0}(0,-1.1){\text{Comparing radii}} } \rput{0}(-1.15,1.15){\Full} \rput{0}(1.15,1.15){\Petal} \rput{0}(-1.15,-1.15){\FSC} \rput{0}(1.15,-1.15){\CompareRadii} \end{pspicture} \end{equation} \end{comment} \caption{Comparison of variations of Algorithm \ref{alg:cosetalg}.} \label{fig:CompareAlgorithms} \end{figure} \subsection*{Drawing dessins}\label{subsec:drawdessin} We now show how the preceding methods fit together in our first main result, to conformally draw dessins. Compare the work of Schneps \cite[\S III.1]{Schneps} for genus zero dessins (given by an explicit rational function). Rather than getting bogged down in algorithmic generalities, it is clearer instead to illustrate with an extended example. Let \[ \sigma_0 =(1\;3\;5\;4)(2\;6),\ \sigma_1 =(1\;5\;2\;3\;6\;4),\ \sigma_\infty = (1\;6\;5\;2\;3\;4) \] and let $\Gamma \leq \Delta(4,6,6)$ be the corresponding subgroup. Let $D_\Delta$ be the fundamental domain for $\Delta$ as in Proposition \ref{prop:triangleembed}. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-08.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic1} \notag \psset{unit=4cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.78359,0.00000) \psarc[linewidth=.001pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psline[linewidth=.001pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.78359,0.00000) \psarcn[linewidth=.001pt] (1.0299,0.42658) {0.49257} {240.00} {165.00} \psline[linewidth=.001pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \psdots*(0,0)(0.554082307686481868582393694073,-0.554082307686481868582393694073) (0.783590714201204868242763136647,0)(0.554082307686481868582393694073,0.554082307686481868582393694073) \uput{5pt}[180]{0}(0,0){$w_a$} \uput{5pt}[270]{0}(0.554082307686481868582393694073,-0.554082307686481868582393694073){$w_c$} \uput{5pt}[0]{0}(0.783590714201204868242763136647,0){$w_b$} \uput{5pt}[90]{0}(0.554082307686481868582393694073,0.554082307686481868582393694073){$\overline{w_c}$} \end{pspicture} \end{equation} \end{comment} \caption{$D_\Delta$ for $\Delta = \Delta(4,6,6)$ as in Proposition \ref{prop:triangleembed}.} \label{fig:dessinpic1} \end{figure} We first compute a coset graph $G = G(\Gamma \backslash \Delta)$ and a side pairing $S = S(G)$ for $\Gamma$ using Algorithm \ref{alg:cosetalg}. In the first iteration of Step $3$ we have $j=1$ and $\alpha_1=1$. After computing $1^{\pi(\alpha_1\epsilon)}$ for $\epsilon\in\{\delta_a,\delta_a^{-1},\delta_b,\delta_b^{-1}\}$ we have the partial coset graph and side pairing as shown in Figure \ref{fig:dessinpic2a} at the end of the first iteration of Step $3$. Explicitly, since $1^{\pi(\alpha_1\delta_a)}= 3$ we add an edge from $v_1$ to $v_3$ labeled $\delta_a$ and we label $v_3$ with $\alpha_1\delta_a = \delta_a$. Similarly, since $1^{\pi(\alpha_1\delta_a^{-1})}=4$ and $1^{\pi(\alpha_1\delta_b)}=5$, we add edges from $v_1$ to $v_4$ and $v_1$ to $v_5$ labeled $\delta_a^{-1}$ and $\delta_b$ respectively, and we label $v_4$ and $v_5$ with $\delta_a^{-1}$ and $\delta_b$ respectively. Now for $\epsilon = \delta_b^{-1}$ we see that $1^{\pi(\alpha_1\delta_b^{-1})}= 4$. Since $v_4$ is already labeled $\delta_a^{-1}$, we add the edge from $v_1$ to $v_4$ labeled $\delta_b^{-1}$, compute $\gamma = \alpha_1\epsilon\alpha_4^{-1} = \delta_b^{-1}\delta_a$, verify that the linear transformation corresponding to $\delta_b^{-1}\delta_a$ as in Proposition \ref{prop:triangleembed} is not the identity matrix, and append $(\gamma,(j,\epsilon),(i,\epsilon^{-1})) = (\delta_b^{-1}\delta_a,(1,\delta_b^{-1}),(4,\delta_b))$ to the side pairing $S$. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-09.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic2a} \notag \psset{unit=7cm} \begin{pspicture}(-1,-1)(1,1) \cnodeput(0.39180,0.00000){1}{$1$} \cnodeput(0.74794,-0.17933){5}{$\delta_b$} \cnodeput(0.00000,-0.39180){4}{$\delta_a^{-1}$} \cnodeput(0.00000,0.39180){3}{$\delta_a$} \cnodeput(-0.17930,0.74792){2}{$v_2$} \cnodeput(0.17935,0.74794){6}{$v_6$} \ncline{->}{1}{3}\naput{$\delta_a$} \ncline{->}{1}{4}\nbput{$\delta_a^{-1}$} \ncline{->}{1}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{1}{4}\naput{$\delta_b^{-1}$} \end{pspicture} \end{equation} \begin{align*} S = \{(\delta_b^{-1}\delta_a,(1,\delta_b^{-1}),(4,\delta_b)) \} \end{align*} \end{comment} \caption{A partial coset graph and side pairing at the end of the first iteration of Step $3$ in Algorithm \ref{alg:cosetalg} for our example.} \label{fig:dessinpic2a} \end{figure} \FloatBarrier \begin{comment} After the first iteration of Step $3$ is complete we return to Step $2$ and compare the hyperbolic distances $d(\alpha_30,0)=d(\delta_a 0,0), d(\alpha_40,0)=d(\delta_a^{-1}0,0),d(\alpha_50,0)=d(\delta_b0,0)$. Since we are choosing this distance to be minimal, and $d(\delta_a0,0)=d(\delta_a^{-1}0,0)=0$, we need to choose between $\delta_a$ and $\delta_a^{-1}$ for the next iteration of Step $3$. \end{comment} By convention, we prefer earlier labels appearing in the coset graph and therefore choose $\delta_a$ in the next iteration of Step $3$, and take $j=3$ and $\alpha_3 = \delta_a$. After computing $1^{\pi(\alpha_3\epsilon)}$ for $\epsilon\in\{\delta_a,\delta_a^{-1},\delta_b,\delta_b^{-1}\}$ we have the following partial coset graph and side pairing in Figure \ref{fig:dessinpic2b} at the end of the second iteration of Step $3$. Explicitly, since $1^{\pi(\delta_a\delta_b)}=6$ and $1^{\pi(\delta_a\delta_b^{-1})}=2$, we label $v_6$ and $v_2$ with $\delta_a\delta_b$ and $\delta_a\delta_b^{-1}$ respectively. Since $1^{\pi(\delta_a^2)}=5$ we append $(\delta_a^2\delta_b^{-1},(3,\delta_a),(5,\delta_a^{-1}))$ to $S$. For $\epsilon=\delta_a^{-1}$ we find that the resulting side pairing element $\gamma$ corresponds to the identity matrix and therefore we do not add this side pairing element to $S$. The new edges added in this iteration are colored blue. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-10.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic2b} \notag \psset{unit=7cm} \begin{pspicture}(-1,-1)(1,1) \cnodeput(0.39180,0.00000){1}{$1$} \cnodeput(0.74794,-0.17933){5}{$\delta_b$} \cnodeput(0.00000,-0.39180){4}{$\delta_a^{-1}$} \cnodeput(0.00000,0.39180){3}{$\delta_a$} \cnodeput(-0.17930,0.74792){2}{$\delta_a\delta_b^{-1}$} \cnodeput(0.17935,0.74794){6}{$\delta_a\delta_b$} \ncline{->}{1}{3}\naput{$\delta_a$} \ncline{->}{1}{4}\nbput{$\delta_a^{-1}$} \ncline{->}{1}{5}\nbput{$\delta_b$} \ncarc[arcangle=30]{->}{1}{4}\naput{$\delta_b^{-1}$} \ncarc[linecolor=blue,arcangle=15]{->}{3}{5}\naput{$\delta_a$} \ncline[linecolor=blue]{->}{3}{6}\naput{$\delta_b$} \ncline[linecolor=blue]{->}{3}{2}\naput{$\delta_b^{-1}$} \ncarc[linecolor=blue,arcangle=-50]{->}{3}{1}\nbput{$\delta_a^{-1}$} \end{pspicture} \end{equation} \end{comment} \begin{align*} S = \{(\delta_b^{-1}\delta_a,(1,\delta_b^{-1}),(4,\delta_b)),(\delta_a^2,(3,\delta_a),(5,\delta_a^{-1})) \} \end{align*} \caption{A partial coset graph and side pairing at the end of the second iteration of Step $3$ in Algorithm \ref{alg:cosetalg} for our example.} \label{fig:dessinpic2b} \end{figure} \FloatBarrier We continue according to Algorithm \ref{alg:cosetalg} taking the labels $\delta_a^{-1}, \delta_b, \delta_a\delta_b,\delta_a\delta_b^{-1}$ in order for the four remaining iterations of Step $3$. At this point the algorithm terminates and returns $G$ and $S$ shown in Figure \ref{fig:dessinpic2c}. In Figure \ref{fig:dessinpic2c} we have the output of Algorithm \ref{alg:cosetalg}. We have only included the edges of the coset graph labeled $\delta_a$ and $\delta_b$ since for every edge labeled $\delta_a$ or $\delta_b$ there is a corresponding reverse edge labeled $\delta_a^{-1}$ or $\delta_b^{-1}$ respectively. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-11.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic2c} \notag \psset{unit=7cm} \begin{pspicture}(-1,-1)(1,1) \cnodeput(0.39180,0.00000){1}{$1$} \cnodeput(0.74794,-0.17933){5}{$\delta_b$} \cnodeput(0.00000,-0.39180){4}{$\delta_a^{-1}$} \cnodeput(0.00000,0.39180){3}{$\delta_a$} \cnodeput(-0.17930,0.74792){2}{$\delta_a\delta_b^{-1}$} \cnodeput(0.17935,0.74794){6}{$\delta_a\delta_b$} \ncline{->}{1}{3}\naput{$\delta_a$} \ncline{->}{4}{1}\naput{$\delta_a$} \ncline{->}{1}{5}\naput{$\delta_b$} \ncline{->}{3}{6}\nbput{$\delta_b$} \ncline{->}{2}{3}\naput{$\delta_b$} \ncarc[arcangle=50]{->}{4}{1}\naput{$\delta_b$} \ncarc[arcangle=30]{->}{3}{5}\nbput{$\delta_a$} \ncline{->}{5}{4}\nbput{$\delta_a$} \ncangles[angleA=0,angleB=0,armA=4.5cm,linearc=.15]{->}{6}{4}\nbput{$\delta_b$} \ncangles[angleA=270,angleB=270,armA=2cm,linearc=.15]{->}{5}{2}\nbput{$\delta_b$} \ncline{->}{6}{2}\naput{$\delta_a$} \ncarc[arcangle=60]{->}{2}{6}\nbput{$\delta_a$} \end{pspicture} \end{equation} \end{comment} \[ S = \{(\delta_b^{-1}\delta_a,(1,\delta_b^{-1}),(4,\delta_b)),(\delta_a^2\delta_b^{-1},(3,\delta_a),(5,\delta_a^{-1})),\dots,(\delta_a\delta_b^2\delta_a,(6,\delta_b),(4,\delta_b^{-1}))\} \] \begin{comment} S = \{&(\delta_b^{-1}\delta_a,(1,\delta_b^{-1}),(4,\delta_b)),(\delta_a^2\delta_b^{-1},(3,\delta_a),(5,\delta_a^{-1})),(\delta_a^{-2}\delta_b^{-1},(4,\delta_a^{-1}),(5,\delta_a)),\\ &(\delta_a^{-1}\delta_b,(4,\delta_b),(1,\delta_b^{-1})),(\delta_a^{-1}\delta_b^{-2}\delta_a^{-1},(4,\delta_b^{-1}),(6,\delta_b)),(\delta_b\delta_a^2,(5,\delta_a),(4,\delta_a^{-1})),\\ &(\delta_b\delta_a^{-2},(5,\delta_a^{-1}),(3,\delta_a)),(\delta_b^3\delta_a^{-1},(5,\delta_b),(2,\delta_b^{-1})),(\delta_a\delta_b\delta_a\delta_b\delta_a^{-1},(6,\delta_a),(2,\delta_a^{-1})),\\ &(\delta_a\delta_b\delta_a^{-1}\delta_b\delta_a^{-1},(6,\delta_a^{-1}),(2,\delta_a)),(\delta_a\delta_b^2\delta_a,(6,\delta_b),(4,\delta_b^{-1})),(\delta_a\delta_b^{-3},(2,\delta_b^{-1}),(5,\delta_b)),\\ &(\delta_a\delta_b^{-1}\delta_a\delta_b^{-1}\delta_a^{-1},(2,\delta_a),(6,\delta_a^{-1})),(\delta_a\delta_b^{-1}\delta_a^{-1},\delta_b^{-1}\delta_a^{-1},(2,\delta_a^{-1}),(6,\delta_a))\} \end{comment} \caption{Output of Algorithm \ref{alg:cosetalg}.} \label{fig:dessinpic2c} \end{figure} Next we draw all of the translates of $D_\Delta$ by the coset representatives of $\Gamma \backslash \Delta$ given by the vertices of $G$ as in \eqref{eqn:fundGamma}. For example, for the vertex of $G$ labeled $\delta = \delta_a\delta_b^{-1}$ we compute the matrix corresponding to $\delta$ as a matrix as in Proposition \ref{prop:triangleembed}, namely $$ \delta = \begin{pmatrix} 0.6552\dots & -2.3015\dots\\ -0.5694\dots & 3.5262\dots \end{pmatrix}. $$ Next we compute $\delta(w_a), \delta(w_b), \delta(w_c), \delta(\overline{w_c})$ and draw in the corresponding translates of the geodesics in $D_\Delta$ between $w_a,w_b,w_c,\overline{w_c}$ as in Figure \ref{fig:dessinpic1}. By doing this for all of the vertices of $G$ we obtain a fundamental domain for $\Gamma$. Given a vertex of $G$ labeled $\delta$ such that $1^{\pi(\delta)} = j$ we label the translate of $D_\Delta$ with $j$ and explicitly give the element $\delta$ as a word in $\delta_a^{\pm 1},\delta_b^{\pm 1}$. For example, again taking $\delta = \delta_a\delta_b^{-1}$ we have $1^{\pi(\delta)} = 2$ so we label $\delta D_\Delta$ with $2$. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-12.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic3} \notag \psset{unit=5cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.78359,0.00000) \psarc[linewidth=.001pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psline[linewidth=.001pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.78359,0.00000) \psarcn[linewidth=.001pt] (1.0299,0.42658) {0.49257} {240.00} {165.00} \psline[linewidth=.001pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psarc[linewidth=.001pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.001pt] (-0.42659,1.0299) {0.49258} {330.00} {255.00} \psarc[linewidth=.001pt] (-0.62769,0.82880) {0.28440} {285.00} {360.00} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.001pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.001pt] (0.00000,1.0299) {0.24628} {270.00} {203.23} \psarcn[linewidth=.001pt] (-0.34331,0.94659) {0.11781} {353.24} {270.01} \closepath} \rput(0.39180,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(-0.17930,0.74792) {\scalebox{0.93886} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}}} \uput*{5pt}[180](0,0){$w_a$} \psdots*(0,0) \uput*{5pt}[270](0.554082307686481868582393694073,-0.554082307686481868582393694073){$w_c$} \psdots*(0.554082307686481868582393694073,-0.554082307686481868582393694073) \uput*{5pt}[90](0.554082307686481868582393694073,0.554082307686481868582393694073){$\overline{w_c}$} \psdots*(0.554082307686481868582393694073,0.554082307686481868582393694073) \uput*{5pt}[0](0.783590714201204868242763136647,0){$w_b$} \psdots*(0.783590714201204868242763136647,0) \uput*{5pt}[180](-0.343294523984519625971712875866,0.828786295691843046255433503358){$\delta(w_a)$} \psdots*(-0.343294523984519625971712875866,0.828786295691843046255433503358) \uput*{5pt}[135](-0.554082307686481868582393694075,0.554082307686481868582393694075){$\delta(w_c)$} \psdots*(-0.554082307686481868582393694075,0.554082307686481868582393694075) \uput*{5pt}[45](0,0.783590714201204868242763136647){$\delta(w_b)$} \psdots*(0,0.783590714201204868242763136647) \uput*{5pt}[90](-0.226314211160785481389272710181,0.932713362493597915794980941618){$\delta(\overline{w_c})$} \psdots*(-0.226314211160785481389272710181,0.932713362493597915794980941618) \end{pspicture} \end{equation} \begin{center} \end{comment} \caption{$D_\Delta$ and the translate $\delta D_\Delta$ where $\delta = \delta_a\delta_b^{-1}$.} \label{fig:dessinpic3} \end{figure} \FloatBarrier Next we must consider the side pairings of $G$ and make them visible on $D_\Gamma$. For each $(\gamma, (i, \epsilon_1), (j, \epsilon_2)) \in S$ we locate the sides $(i, \epsilon_1)$ and $(j, \epsilon_2)$ of $D_\Gamma$ and give them the same label. We signify that $\gamma$ maps the side $(j, \epsilon_2)$ to the side $(i, \epsilon_1)$ by coloring the label for side $(j,\epsilon_2)$ red and the label for side $(i,\epsilon_1)$ blue. For example, in Figure \ref{fig:dessinpic2c} we see that $(\delta_a^2\delta_b^{-1},(3,\delta_a),(5,\delta_a^{-1})) \in S$ so the side $(3,\delta_a)$ which is $\delta_aD_\Delta \cap \delta_a^2D_\Delta$ is identified with the side $(5,\delta_a^{-1})$ which is $\delta_bD_\Delta \cap \delta_b\delta_aD_\Delta$ by the element $\delta_a^2\delta_b^{-1}$. In Figure \ref{fig:dessinpic4} we describe our convention for labeling the elements of $S$. For example, the element $(\delta_a^2\delta_b^{-1},(3,\delta_a),(5,\delta_a^{-1}))$ in $S$ tells us that $\gamma=\delta_a^2\delta_b^{-1}$ maps the side $(5,\delta_a^{-1})$ to the side $(3,\delta_a)$. This is indicated in the picture by labeling the two sides $s_2$ (subscript $2$ since we are considering the second element in $S$) and coloring the side $(5,\delta_a^{-1})$ red and the side $(3,\delta_a)$ light blue, i.e.: \begin{center} \includegraphics{belyi-triangle-pics/belyi-triangle-pics-13.pdf} \begin{comment} \begin{equation}\label{fig:redblue} \begin{pspicture} \psline{->}(-1.5,0)(1.22,0) \rput(-1.5,0){\psframebox*[linewidth=.0001,fillcolor=red!80]{$s_{2}$}} \rput(1.5,0){\psframebox*[linewidth=.0001,fillcolor=blue!10]{$s_{2}$}} \rput(0,.25){$\gamma$} \end{pspicture} \end{equation} \end{comment} \end{center} Also note that $S$ contains the element $(\delta_b\delta_a^{-2},(5,\delta_a^{-1}),(3,\delta_a))$ which is just the same side pairing in the opposite direction. \begin{figure}[h] \includegraphics[scale=.8]{belyi-triangle-pics/belyi-triangle-pics-14.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic4} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.78359,0.00000) \psarc[linewidth=.1pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psline[linewidth=.1pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.78359,0.00000) \psarcn[linewidth=.1pt] (1.0299,0.42658) {0.49257} {240.00} {165.00} \psline[linewidth=.1pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.1pt] (-0.42659,1.0299) {0.49258} {330.00} {255.00} \psarc[linewidth=.1pt] (-0.62769,0.82880) {0.28440} {285.00} {360.00} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.1pt] (0.00000,1.0299) {0.24628} {270.00} {203.23} \psarcn[linewidth=.1pt] (-0.34331,0.94659) {0.11781} {353.24} {270.01} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,0.78359) \psarc[linewidth=.1pt] (0.42659,1.0299) {0.49258} {210.00} {285.00} \psline[linewidth=.1pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,0.78359) \psarcn[linewidth=.1pt] (-0.42659,1.0299) {0.49258} {330.00} {255.00} \psline[linewidth=.1pt] (-0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,-0.78359) \psarc[linewidth=.1pt] (-0.42659,-1.0299) {0.49258} {29.999} {105.00} \psline[linewidth=.1pt] (-0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,-0.78359) \psarcn[linewidth=.1pt] (0.42659,-1.0299) {0.49258} {150.00} {74.998} \psline[linewidth=.1pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.0299,-0.14222) {0.28439} {225.00} {150.00} \psarc[linewidth=.1pt] (1.0299,0.00000) {0.24628} {180.00} {246.77} \psarc[linewidth=.1pt] (0.94658,-0.34331) {0.11781} {96.757} {179.99} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.0299,-0.14222) {0.28439} {225.00} {150.00} \psarc[linewidth=.1pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psarcn[linewidth=.1pt] (0.82879,-0.62768) {0.28439} {165.00} {90.000} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.14222,1.0299) {0.28437} {315.00} {240.00} \psarc[linewidth=.1pt] (0.00000,1.0299) {0.24628} {270.00} {336.77} \psarc[linewidth=.1pt] (0.34331,0.94659) {0.11781} {186.76} {270.00} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.14222,1.0299) {0.28437} {315.00} {240.00} \psarc[linewidth=.1pt] (0.42659,1.0299) {0.49258} {210.00} {285.00} \psarcn[linewidth=.1pt] (0.62769,0.82880) {0.28440} {255.00} {180.00} \closepath} \rput(-0.27704,0.27704) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(0.85855,-0.26511) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}}} \rput(0.00000,0.39180) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(0.74794,-0.17933) {\scalebox{0.93886} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}}} \end{pspicture} \end{equation} \end{comment} \caption{An illustration of our convention for labeling the elements of $S$.} \label{fig:dessinpic4} \end{figure} \FloatBarrier Finally, for each translate $\delta D_\Delta$ we draw a white dot at $\delta(w_a)$, a black dot at $\delta(w_b)$, an $\times$ at $\delta(w_c)$, and bold the geodesic from $\delta(w_a)$ to $\delta(w_b)$. The white and black dots are the vertices of the dessin, and the bold geodesics are the edges of the dessin. By identifying the sides that are paired together we form a quotient Riemann surface with the dessin conformally embedded. In Figure \ref{fig:dessinpic5} is the final fundamental domain for $\Gamma$ which, after identifying the sides, yields a Riemann surface with the dessin conformally embedded. Notice that some of the white and black vertices are identified in the quotient and we are left with $2$ white vertices and $1$ black vertex. From this we can read off each cycle in $\sigma_0$ and $\sigma_1$ by rotating counterclockwise around each of the white dots and black dots respectively. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-15.pdf} \begin{comment} \begin{equation} \label{fig:dessinpic5} \notag \psset{unit=6.5cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.78359,0.00000) \psarc[linewidth=.1pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psline[linewidth=.1pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.78359,0.00000) \psarcn[linewidth=.1pt] (1.0299,0.42658) {0.49257} {240.00} {165.00} \psline[linewidth=.1pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.1pt] (-0.42659,1.0299) {0.49258} {330.00} {255.00} \psarc[linewidth=.1pt] (-0.62769,0.82880) {0.28440} {285.00} {360.00} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psarcn[linewidth=.1pt] (0.00000,1.0299) {0.24628} {270.00} {203.23} \psarcn[linewidth=.1pt] (-0.34331,0.94659) {0.11781} {353.24} {270.01} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,0.78359) \psarc[linewidth=.1pt] (0.42659,1.0299) {0.49258} {210.00} {285.00} \psline[linewidth=.1pt] (0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,0.78359) \psarcn[linewidth=.1pt] (-0.42659,1.0299) {0.49258} {330.00} {255.00} \psline[linewidth=.1pt] (-0.55408,0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,-0.78359) \psarc[linewidth=.1pt] (-0.42659,-1.0299) {0.49258} {29.999} {105.00} \psline[linewidth=.1pt] (-0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.00000,-0.78359) \psarcn[linewidth=.1pt] (0.42659,-1.0299) {0.49258} {150.00} {74.998} \psline[linewidth=.1pt] (0.55408,-0.55408) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.0299,-0.14222) {0.28439} {225.00} {150.00} \psarc[linewidth=.1pt] (1.0299,0.00000) {0.24628} {180.00} {246.77} \psarc[linewidth=.1pt] (0.94658,-0.34331) {0.11781} {96.757} {179.99} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.0299,-0.14222) {0.28439} {225.00} {150.00} \psarc[linewidth=.1pt] (1.0299,-0.42658) {0.49257} {120.00} {195.00} \psarcn[linewidth=.1pt] (0.82879,-0.62768) {0.28439} {165.00} {90.000} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.14222,1.0299) {0.28437} {315.00} {240.00} \psarc[linewidth=.1pt] (0.00000,1.0299) {0.24628} {270.00} {336.77} \psarc[linewidth=.1pt] (0.34331,0.94659) {0.11781} {186.76} {270.00} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.14222,1.0299) {0.28437} {315.00} {240.00} \psarc[linewidth=.1pt] (0.42659,1.0299) {0.49258} {210.00} {285.00} \psarcn[linewidth=.1pt] (0.62769,0.82880) {0.28440} {255.00} {180.00} \closepath} \rput(0.57481,0.23806) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(-0.23809,-0.57479) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(-0.27704,0.27704) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(0.85855,-0.26511) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}}} \rput(-0.27704,-0.27704) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(0.65565,-0.40206) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}}} \rput(0.23808,-0.57480) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.13552,0.82422) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}}} \rput(0.82423,-0.13552) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}}} \rput(-0.13551,0.82423) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}}} \rput(0.40207,0.65565) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{6}$}}} \rput(-0.40207,0.65569) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{6}$}}} \rput(0.26511,0.85855) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{7}$}}} \rput(-0.26505,0.85852) {\scalebox{0.93886} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{7}$}}} \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.78359,0.00000) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.78359,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.55408,-0.55408) \psarc[linewidth=1.4083pt] (-0.14222,1.0299) {0.28437} {225.00} {300.00} \psdots*[dotstyle=o,dotsize=6.5720pt](-0.34330,0.82879) \psdots*[dotstyle=*,dotsize=6.5720pt](0.00000,0.78359) \psdots*[dotstyle=x,dotsize=6.5720pt](-0.55408,0.55408) \psdots*[dotstyle=x,dotsize=6.5720pt](-0.22631,0.93272) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.00000,0.78359) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.00000,0.78359) \psdots*[dotstyle=x,dotsize=7.0000pt](0.55408,0.55408) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.00000,-0.78359) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.00000,-0.78359) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.55408,-0.55408) \psarcn[linewidth=1.4083pt] (1.0299,-0.14222) {0.28439} {225.00} {150.00} \psdots*[dotstyle=o,dotsize=6.5720pt](0.82879,-0.34330) \psdots*[dotstyle=*,dotsize=6.5720pt](0.78359,0.00000) \psdots*[dotstyle=x,dotsize=6.5720pt](0.93272,-0.22631) \psarcn[linewidth=1.4083pt] (0.14222,1.0299) {0.28437} {315.00} {240.00} \psdots*[dotstyle=o,dotsize=6.5720pt](0.34330,0.82879) \psdots*[dotstyle=*,dotsize=6.5720pt](0.00000,0.78359) \psdots*[dotstyle=x,dotsize=6.5720pt](0.22631,0.93272) \rput(0.39180,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(-0.17930,0.74792) {\scalebox{0.93886} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}}} \rput(0.00000,0.39180) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(0.00000,-0.39180) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(0.74794,-0.17933) {\scalebox{0.93886} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}}} \rput(0.17935,0.74794) {\scalebox{0.93886} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}}} \end{pspicture} \end{equation} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_a^{}\delta_b^{-1}$ \\ $3$ & $\delta_a^{}$ \\ $4$ & $\delta_a^{-1}$ \\ $5$ & $\delta_b^{}$ \\ $6$ & $\delta_a^{}\delta_b^{}$ \\ \bottomrule \end{tabular} \;\;\;\;\;\;\;\;\;\; \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_b^{-1}\delta_a^{}$ \\ $s_{2}$ & $\delta_a^{2}\delta_b^{-1}$ \\ $s_{3}$ & $\delta_a^{-2}\delta_b^{-1}$ \\ $s_{4}$ & $\delta_a^{-1}\delta_b^{-2}\delta_a^{-1}$ \\ $s_{5}$ & $\delta_b^{3}\delta_a^{-1}$ \\ $s_{6}$ & $\delta_a^{}\delta_b^{}\delta_a^{}\delta_b^{}\delta_a^{-1}$ \\ $s_{7}$ & $\delta_a^{}\delta_b^{}\delta_a^{-1}\delta_b^{}\delta_a^{-1}$ \\ \bottomrule \end{tabular} \end{center} \end{comment} \caption{A fundamental domain for $\Gamma$ with the embedded dessin.} \label{fig:dessinpic5} \end{figure} \FloatBarrier \subsection*{Reduction algorithm} We now describe the reduction algorithm for $\Delta$ that takes a point $w \in \mathcal{D}$ and returns a point $w' = \delta w \in D_\Delta$ and an element $\delta \in \Delta$. This algorithm \cite{VoightThesis} generalizes the classical reduction theory for $\SL_2(\mathbb Z)$ (see e.g.\ Serre \cite[Theorem 7.1]{SerreCourse}); it is a special (and very efficient) case of a full algorithmic theory of Dirichlet fundamental domains for Fuchsian groups whose quotients have finite area \cite{Voightfd}. \begin{alg} \label{alg:fdreddelta} Let $z \in \mathcal{H}$ and let $\Delta=\Delta(a,b,c)$ be a triangle group with fundamental domain $D_\Delta$ given as in Proposition \ref{prop:triangleembed}. This algorithm returns an element $\delta \in \Delta$ as a word in $\delta_a,\delta_b,\delta_c$ such that $z'=\delta z \in D_\Delta$. \begin{enumalg} \item Initialize $\delta := 1$, $w := w_{z_a}(z)$. \item Let \[ \alpha :=\arg(w) \in [0, 2 \pi). \] Let \[ i := - \left \lfloor \frac{a \alpha}{2 \pi} + \frac{1}{2} \right \rfloor. \] Let $w := \delta_a^i w$ and $\delta := \delta_a^i \delta$. \item Let \begin{equation} \label{Amat} A = \begin{pmatrix} \mu+1 & -\mu+1 \\ -\mu+1 & \mu+1 \end{pmatrix} \end{equation} with $\mu$ as in Proposition \ref{prop:triangleembed}. Let \[ \beta :=\arg(-A w_{a}) \in [0, 2 \pi). \] Let \[ j := -\left \lfloor \frac{b \beta}{2 \pi} + \frac{1}{2} \right \rfloor. \] Let $w := \delta_b^jw$ and $\delta := \delta_b^j \delta$. If $j = 0$, then return $\delta$, otherwise return to Step 2. \end{enumalg} \end{alg} \begin{proof} Let $w = w_{z_a}(z)$. By identifying $\mathcal{H}$ and $\mathcal{D}$ with the map $w_{z_a}$, the element $\delta_a$ acts on $\mathcal{D}$ by a rotation by $2\pi/a$ about the origin. We can therefore rotate $w$ by a unique element $\delta_a^i$ with $-a < i \leq 0$ so that it is in the region of $\mathcal{D}$ bounded by the images of the geodesics $z_a z_c$ and $z_a z_c'$. In fact $i = - \left \lfloor \frac{a \alpha}{2 \pi} + \frac{1}{2} \right \rfloor$ is the unique such integer. We now show that by applying Step $3$ to the resulting point, when nontrivial we get a point with smaller absolute value. Since the orbits of the action of $\Delta$ on $\mathcal{H}$ are discrete, the algorithm will terminate after finitely many steps with a point in $D_\Delta$. Let $w$ be a point in the region bounded by the images of the geodesics $z_a z_c$ and $z_a z_c'$. There exists a unique isometry of the unit disc that maps $w_{z_a}(z_b)$ to $0$ and $w_{z_a}(z_a)$ to the negative real axis: this isometry is given by the linear fraction transformation corresponding to $A$ in \eqref{Amat}. We can therefore rotate $w$ by a unique element $\delta_b^j$ with $0 \leq j < b$ so that it is in the region of $\mathcal{D}$ bounded by the goedesics $z_b z_c$ and $z_b z_c'$. In fact $j = \left \lfloor \frac{b \beta}{2 \pi} + \frac{1}{2} \right \rfloor$ is the unique such integer. Then $\delta_b^j Aw $ will be closer to $A w_{z_a}(z_a)$ than $A w$. Then, $\delta_b^j w$ has smaller absolute value than $w$ when $j \neq 0$, and the proof is complete. \end{proof} \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-16.pdf} \begin{comment} \begin{equation} \label{fig:reduction2} \notag \psset{unit=1.5cm} \begin{pspicture}(-5,-5)(5,5) \def\mathbb H{% \psline[linewidth=.5pt]{<->}(-3,0)(3,0) \psline[linewidth=.5pt]{<->}(0,-1)(0,5) \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psline(0,1)(0,3.51221866) \psarc(3.51221866,0){4.9670272}{135}{170.446835} \psarcn(-0.577350,0){1.1547001}{134.44683}{60} \closepath} \pscustom[linewidth=.5pt]{ \psline(0,1)(0,3.512) \psarcn(-3.5122186,0){4.967027268}{45}{9.5531648} \psarc(0.5773502,0){1.154700}{45.553164}{120} \closepath} \uput{5pt}[-45]{0}(0,1){$z_a$} \uput{5pt}[-45]{0}(1.38592,0.82434131){$z_c$} \uput{5pt}[45]{0}(0,3.51221){$z_b$} \uput{5pt}[225]{0}(-1.385926,0.824341){$-\overline{z_c}$} \psdots*(0,1)(0,3.51221)(1.3859260,0.824341311)(-1.3859260142,0.82434131) \psdots*[dotstyle=*](1,.3) \uput{5pt}[0]{0}(1,.3){$z$} } \def\DDa{% \pscircle[linewidth=.1pt](0,0){1} \pscustom[linewidth=.01pt]{ \psline[linewidth=.01pt](0,0)(0.5567590,0) \psarc[linewidth=.01pt](1.176434,-0.6196749){0.8763527}{135}{174} \psline[linewidth=.01pt](0.3048820,-0.5280711)(0,0) \closepath} \pscustom[linewidth=.01pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.01pt](0,0)(0.5567590,0) \psarcn[linewidth=.01pt](1.1764340,0.61967498){0.8763527}{225}{186} \psline[linewidth=.01pt](0.3048820,0.5280711)(0,0) \closepath} \pscustom[linewidth=.01pt,fillstyle=solid, fillcolor = red,opacity=0.25]{ \psarc[linewidth=.01pt](1.176434,-0.6196749){0.8763527}{103.452219}{200.99263} \psarc[linewidth=.01pt](0,0){1}{-69.0073688}{13.452219} \closepath} \pscustom[linewidth=.01pt,fillstyle=solid,fillcolor=blue,opacity=0.25]{ \psarc[linewidth=.01pt](1.1764340, 0.6196749){0.8763527}{159.00736}{256.547780} \psarc[linewidth=.1pt](0,0){1}{-13.45221}{69.0073688} \closepath} \psdots*[dotstyle=*,dotsize=5pt](0,0) \psdots*[dotstyle=*,dotsize=5pt](0.5567590,0) \uput{5pt}[180]{0}(0,0){$w_a$} \uput{8pt}[205]{0}(0.55675906){$w_b$} \psdots*[dotstyle=*,dotsize=5pt](0.033457, - 0.743494423) \uput{5pt}[260]{0}(0.0334572, - 0.7434944){$w = w_{z_a}(z)$} \psdots*[dotstyle=*,dotsize=5pt](0.627156434, 0.4007220) \uput{5pt}[0]{0}(0.6271564, 0.400722){$\delta_a^i w$} } \def\DDb{% \pscircle[linewidth=.1pt](0,0){1} \pscustom[linewidth=.01pt]{ \psline[linewidth=.01pt](0,0)(-0.5567590,0) \psarcn[linewidth=.01pt](-1.176434, - 0.3577695){0.715539041}{30}{-9} \psline[linewidth=.01pt](-0.4697044, - 0.4697044)(0,0) \closepath} \pscustom[linewidth=.01pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.01pt](0,0)(-0.55675 ,0) \psarc[linewidth=.01pt](-1.176434, 0.357769){0.71553909}{-30}{9} \psline[linewidth=.01pt](-0.4697044, 0.469704)(0,0) \closepath} \pscustom[linewidth=.01pt,fillstyle=solid,fillcolor=blue,opacity=0.25]{ \psarcn[linewidth=.01pt](-1.17643405, - 0.35776952){0.7155390}{71.3300267}{-37.49957} \psarcn[linewidth=.1pt](0,0){1}{232.5004}{161.330026} \closepath} \pscustom[linewidth=.01pt,fillstyle=solid,fillcolor=red,opacity=0.25]{ \psarcn[linewidth=.01pt](-1.1764349, 0.3577695){0.71553904}{37.49957}{-71.33003} \psarcn[linewidth=.1pt](0,0){1}{198.669}{127.49957} \closepath} \psdots*[dotstyle=*,dotsize=5pt](0,0) \psdots*[dotstyle=*,dotsize=5pt](-0.556759,0) \uput{5pt}[30]{0}(0,0){$A w_b$} \uput{3.5pt}[180]{0}(-0.556759,0){$A w_a$} \psdots*[dotstyle=*,dotsize=5pt](-0.72398, - 0.452226) \uput{5pt}[-45]{0}(-0.72398, - 0.452226){$A\delta_a^i w$} \psdots*[dotstyle=*,dotsize=5pt](-0.030048, - 0.13874) \uput{5pt}[-30]{0}(-0.0300480, - 0.138747290) {$\delta_b^j A \delta_a^i w$} } \rput{0}(0,3){\mathbb H} \psset{unit=3cm} \rput{0}(-1.5,-0.5){\DDa} \rput{0}(1.5,-0.5){\DDb} \psline[linewidth=1pt]{->}(-.75,1.25)(-1.2,.55) \uput{5pt}[135]{0}(-1,.9){$w_{z_a}$} \psline[linewidth=1pt]{->}(.75,1.25)(1.2,.55) \uput{5pt}[45]{0}(1,.9){$w_{z_b}$} \psline[linewidth=1pt]{->}(-.4,-.5)(.4,-.5) \uput{5pt}[90]{0}(0,-.5){$A$} \end{pspicture} \end{equation} \end{comment} \caption{An example of the reduction algorithm. Points in the blue region are moved closer to $D_\Delta$ by $\delta_b$, while points in the red region are moved closer to $D_\Delta$ by $\delta_b^{-1}$.} \label{fig:reduction2} \end{figure} \begin{rmk} Due to roundoff issues, for points near the boundary of the fundamental triangle, one either needs to pick a preferred side among the possible two paired sides or to slightly thicken the fundamental domain (dependent on the precision) to avoid bouncing between points equivalent under $\Delta$. For the purposes of our numerical algorithms, any such choice works as well as any other. \end{rmk} We now describe how a reduction algorithm for $\Gamma$ is obtained from the reduction algorithm for $\Delta$ together with Algorithm \ref{alg:cosetalg}. \begin{alg} \label{alg:fdredgamma} Let $\Gamma \leq \Delta(a,b,c)$ be a subgroup of index $d=[\Delta:\Gamma]$. Let $\Gamma \alpha_1 , \dots , \Gamma \alpha_d$ be a set of right cosets for $\Gamma$ in $\Delta$ and let $D_\Gamma = \alpha_1 D_\Delta \cup \cdots \cup \alpha_d D_\Delta$. This algorithm returns an element $\gamma \in \Gamma$ such that $z'=\gamma z \in D_\Gamma$. \begin{enumalg} \item Using Algorithm \ref{alg:fdreddelta}, let $\delta \in \Delta$ be such that $\delta z \in D_\Delta$. \item Compute $i := 1^{\pi(\delta^{-1})}$ and return $\gamma=\alpha_i \delta$. \end{enumalg} \end{alg} \begin{proof}[Proof of correctness] We have $1^{\pi(\delta^{-1})}=i$ if and only if $\Gamma \delta^{-1} = \Gamma \alpha_i$, so we have $\gamma=\alpha_i \delta \in \Gamma$. But since $\delta z \in D_\Delta$ we have $\gamma z = \alpha_i (\delta z) \in \alpha_i D_\Delta \in D_\Gamma$, as claimed. \end{proof} \begin{rmk} The reason we need to take an inverse in Step 2 of Algorithm \ref{alg:fdredgamma} is because the cosets are labelled by element moving from the center $0$ to the translated piece, but we want to move into the fundamental domain, so we need an inverse. If the tessellates are labelled by the element that maps to the fundamental triangle, the inverse would be replaced somewhere else (adjacent tessellates would be inverted). \end{rmk} \begin{exm} We illustrate the reduction algorithms (Algorithms \ref{alg:fdreddelta} and \ref{alg:fdredgamma}) with an example. Let $\Delta = \Delta(3,4,5)$ and define $\pi : \Delta \to S_6$ by $\delta_a, \delta_b, \delta_c \mapsto \sigma_0, \sigma_1, \sigma_\infty$, where \[ \sigma_0 = (1\;2\;3),\ \sigma_1 = (1\;2)(3\;4\;5\;6),\ \sigma_\infty = (1\;6\;5\;4\;3). \] Given the example point $w = 0.34992\ldots + 0.82246\ldots i$ , the reduction algorithm for $\Delta$ returns the point $\delta w = 0.08700\ldots + 0.09353\ldots i \in D_\Delta$, where $\delta = \delta_a^{-1} \delta_b^{-1} \delta_a^{-1} \delta_b^{-1} \delta_a^{-1}$. This $\delta$ corresponds to the linear fractional transformation represented by the matrix $$ \begin{pmatrix} 1.9532\dots & 2.6172\dots\\ 1.0755\dots & 1.9531\dots \end{pmatrix}. $$ (See Figure \ref{fig:reduction3}.) In the reduction algorithm for $\Gamma$, we compute $\pi(\delta^{-1}) = (1\;2\;5)(3\;4\;6)$, so $1^{\pi(\delta^{-1})} = 2$. This means that $\gamma w$ will be in the coset labeled 2. From the table in Figure \ref{fig:reduction3}, we see that $\alpha_2 = \delta_a$, i.e.~$\delta_a$ is a representative for the coset labeled 2. Then we have: \begin{align*} \gamma = \alpha_2 \delta = \delta_a \delta_a^{-1} \delta_b^{-1} \delta_a^{-1} \delta_b^{-1} \delta_a^{-1} = (\delta_b^{-1} \delta_a^{-1})^2 \end{align*} This $\gamma$ corresponds to the linear fractional transformation represented by the matrix $$ \begin{pmatrix} 1.9080\dots & 3.0001\dots\\ -1.1537\dots & -1.2900\dots \end{pmatrix}. $$ Applying $\gamma$ to $w$ yields the point $\gamma w = -0.12450\dots + 0.02857\dots i \in \alpha_2 D_\Delta \in D_\Gamma$. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-17.pdf} \begin{comment} \begin{equation} \label{fig:reduction3} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.55676,0.00000) \psarc[linewidth=.001pt] (1.1764,-0.61968) {0.87636} {135.00} {174.00} \psline[linewidth=.001pt] (0.30488,-0.52807) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.55676,0.00000) \psarcn[linewidth=.001pt] (1.1764,0.61968) {0.87636} {225.00} {186.00} \psline[linewidth=.001pt] (0.30488,0.52807) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (-0.27838,0.48217) \psarc[linewidth=.001pt] (-0.051568,1.3287) {0.87637} {255.00} {294.00} \psline[linewidth=.001pt] (0.30488,0.52807) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (-0.27838,0.48217) \psarcn[linewidth=.001pt] (-1.1249,0.70898) {0.87634} {345.00} {306.00} \psline[linewidth=.001pt] (-0.60976,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (-0.27838,-0.48217) \psarc[linewidth=.001pt] (-1.1249,-0.70898) {0.87634} {15.000} {54.000} \psline[linewidth=.001pt] (-0.60976,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (-0.27838,-0.48217) \psarcn[linewidth=.001pt] (-0.051568,-1.3287) {0.87637} {105.00} {66.000} \psline[linewidth=.001pt] (0.30488,-0.52807) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt]{ \psarcn[linewidth=.001pt] (-0.58823,-1.0188) {0.61967} {94.442} {60.000} \psarc[linewidth=.001pt] (-0.051529,-1.3287) {0.87641} {105.00} {128.42} \psarc[linewidth=.001pt] (-0.87458,-0.56447) {0.28902} {344.42} {34.442} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.001pt] (-0.58823,-1.0188) {0.61967} {94.442} {60.000} \psarc[linewidth=.001pt] (-1.1249,-0.70898) {0.87634} {15.000} {54.000} \psarcn[linewidth=.001pt] (-1.1249,-0.16738) {0.54163} {18.000} {334.45} \closepath} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (-0.42501,-0.73615) (-0.27838,-0.48217) \psarcn[linewidth=.001pt] (-1.1249,-0.70897) {0.87635} {14.999} {351.57} \psarc[linewidth=.001pt] (-0.42507,-0.92457) {0.18849} {27.561} {89.982} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (-0.42501,-0.73615) (-0.27838,-0.48217) \psarc[linewidth=.001pt] (-0.051529,-1.3287) {0.87641} {105.00} {128.42} \psarcn[linewidth=.001pt] (-0.58819,-0.83039) {0.18847} {92.431} {30.010} \closepath} \pscustom[linewidth=.001pt]{ \psarc[linewidth=.001pt] (-0.58822,-1.0188) {0.61968} {25.556} {60.000} \psarcn[linewidth=.001pt] (-0.051568,-1.3287) {0.87637} {105.00} {66.000} \psarc[linewidth=.001pt] (0.41747,-1.0578) {0.54161} {102.00} {145.55} \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.001pt] (-0.58822,-1.0188) {0.61968} {25.556} {60.000} \psarcn[linewidth=.001pt] (-1.1249,-0.70897) {0.87635} {14.999} {351.57} \psarcn[linewidth=.001pt] (-0.051496,-1.0397) {0.28907} {135.59} {85.568} \closepath} \rput(0.38544,-0.24242) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(-0.40263,0.21265) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.38545,0.24239) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(0.017208,0.45500) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(-0.58443,-0.20306) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(-0.58949,-0.51711) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(-0.15310,-0.76904) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.11636,-0.60767) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(-0.49745,-0.66520) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}} \rput(-0.32736,-0.76343) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}} \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.55676,0.00000) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.55676,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.30488,-0.52807) \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.27838,0.48217) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.27838,0.48217) \psdots*[dotstyle=x,dotsize=7.0000pt](0.30488,0.52807) \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.27838,-0.48217) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.27838,-0.48217) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.60976,0.00000) \psarcn[linewidth=1.5000pt] (-0.58823,-1.0188) {0.61967} {94.442} {60.000} \psdots*[dotstyle=o,dotsize=7.0000pt](-0.63624,-0.40101) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.27838,-0.48217) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.59618,-0.64209) \psline[linewidth=1.5000pt] (-0.42501,-0.73615) (-0.27838,-0.48217) \psdots*[dotstyle=o,dotsize=7.0000pt](-0.42501,-0.73615) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.27838,-0.48217) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.25797,-0.83736) \psarc[linewidth=1.5000pt] (-0.58822,-1.0188) {0.61968} {25.556} {60.000} \psdots*[dotstyle=o,dotsize=7.0000pt](-0.029165,-0.75150) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.27838,-0.48217) \psdots*[dotstyle=x,dotsize=7.0000pt](0.30488,-0.52807) \rput(0.27838,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(-0.13919,0.24108) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.13919,-0.24108) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(-0.45116,-0.41451) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(-0.35170,-0.60915) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \rput(-0.13338,-0.59796) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}} \psdot*[dotsize=6pt,linecolor=blue](0.349927106111882585456754867815, 0.822467033424113218236207583323) {\color{blue}\uput{5pt}[0]{0}(0.349927106111882585456754867815, 0.822467033424113218236207583323){$w$}} \psdots*[dotsize=6pt,linecolor=blue](0.0870031338945347927969723613417, 0.0935374900506349823790715163708) {\color{blue}\uput{5pt}[0]{0}(0.0870031338945347927969723613417, 0.0935374900506349823790715163708){$\delta w$}} \psdots*[dotsize=6pt,linecolor=blue](-0.124507409537351469536214814904, 0.0285781791362085828564156171686) {\color{blue}\uput{5pt}[180]{0}(-0.124507409537351469536214814904, 0.0285781791362085828564156171686){$\gamma w$}} \end{pspicture} \end{equation} \end{comment} \begin{center} \begin{tabular}{ll} \toprule Coset Number & Label\\ \midrule $1$ & 1 \\ $2$ & $\delta_a$ \\ $3$ & $\delta_a^{-1}$ \\ $4$ & $\delta_a^{-1 } \delta_b$ \\ $5$ & $\delta_a^{-1 } \delta_b^{-2}$ \\ $6$ & $\delta_a^{-1 } \delta_b^{-1}$ \\ \bottomrule \end{tabular} \hspace{2cm} \begin{tabular}{ll} \toprule Side Pairing & Label\\ \midrule $s_{1}$ & $\delta_b \delta_a^{-1}$ \\ $s_{2}$ & $\delta_b^{-1 } \delta_a^{-1}$ \\ $s_{3}$ & $\delta_a^{-1 } \delta_b^{-1 } \delta_a \delta_b \delta_a$ \\ $s_{4}$ & $\delta_a^{-1 } \delta_b \delta_a \delta_b^{-1 } \delta_a$ \\ $s_{5}$ & $\delta_a^{-1 } \delta_b^{-2 } \delta_a \delta_b^{2 } \delta_a$ \\ \bottomrule \end{tabular} \end{center} \bigskip \caption{For the example point $w = 0.34992\ldots+0.82246\ldots i$ in the unit disc, the reduction algorithm for $\Gamma$ yields the point $\gamma w = -0.12450\ldots+0.02857\ldots i$ in $D_\Gamma$. The reduction algorithm for $\Delta$ yields the point $\delta w = 0.08700\ldots+0.09353\ldots i$ in $D_\Delta$.} \label{fig:reduction3} \end{figure} \end{exm} \subsection*{Algorithm to compute a group presentation} We conclude this section by showing how to determine a presentation for the group $\Gamma$ from the coset graph. Although this will not figure in our method to compute Bely\u{\i}\ maps, it can be computed easily in our setup and may be useful for other applications. Let $D_\Gamma$ be a fundamental domain for $\Gamma$ as in \ref{eqn:fundGamma} and let $S$ be the associated side pairing. Having computed a side pairing for $\Gamma$, all of the relevant algorithms for obtaining a presentation for $\Gamma$ appear in work of Voight \cite{Voightfd}. \begin{rmk} The assumption that the fundamental domains of the Fuchsian groups under examination are hyperbolically convex is verified \cite{Voightfd} in order to apply the Poincar\'{e} polygon theorem. However, this condition is extraneous; the needed general version of the Poincar\'{e} polygon theorem was proven by Maskit \cite{Maskit}. In particular, even when the fundamental domain is disconnected, the Poincar\'{e} polygon theorem can still be applied: see Epstein and Petronio \cite[Theorem 8.1]{EP} (which also includes the colorful history of this theorem). \end{rmk} The \defi{vertices} of $D_\Gamma$ are the intersections of the sides of $D_\Gamma$ in $S$ (i.e. the vertices of $D_\Gamma$ as a hyperbolic polygon). A \defi{pairing cycle} for $D_\Gamma$ is a sequence $v_1,\dots,v_n$ of vertices of $D_\Gamma$ which is the (ordered) intersection of the $\Gamma$-orbit of of $v_1$ with $D_\Gamma$. To each cycle associate a word $g = g_n g_{n-1} \cdots g_2 g_1$ where $g_i(v_i) = v_{i+1}$. A cycle is \defi{minimal} if $v_i \neq v_j$ for all $i \neq j$. Any side pairing element $\gamma \in \Gamma$ of $S$ will appear at most once in any word associated to a minimal cycle. A set of minimal cycles is \defi{complete} if every side pairing element in $S$ occurs in (a necessarily unique) one of the cycles in the set. Given $S$ there is an algorithm to compute a complete set of minimal cycles for $D_\Gamma$ (see \cite{Voightfd} Algorithm 5.2). Given such a complete set of minimal cycles of $D_\Gamma$ there is an algorithm to compute a minimal set of generators and relations for $\Gamma$ \cite[Algorithm 5.7]{Voightfd}. We illustrate these methods with an example. \begin{exm} Let $\Gamma \leq \Delta(3,5,3)$ correspond to the permutation triple \[ \sigma_0 = (1\;3\;4)(2\;6\;5),\ \sigma_1 = (1\;4\;2\;5\;6),\ \sigma_\infty = (3\;6\;4). \] We first obtain a fundamental domain for $\Gamma$ using the petalling approach. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-18.pdf} \begin{comment} \begin{equation} \label{fig:presentation} \notag \psset{unit=6.5cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarc[linewidth=.1pt] (1.2098,-0.93707) {1.1583} {126.00} {150.00} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.52894,0.00000) \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.82663,0.71587) {0.44244} {225.52} {186.00} \psarc[linewidth=.1pt] (0.58981,0.85258) {0.27344} {222.00} {270.00} \psarc[linewidth=.1pt] (1.2098,0.22116) {0.71592} {150.00} {165.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.82663,0.71587) {0.44244} {225.52} {186.00} \psline[linewidth=.1pt] (0.38660,0.66962) (0.20665,0.35793) \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarc[linewidth=.1pt] (0.20664,1.5162) {1.1583} {246.00} {270.00} \psline[linewidth=.1pt] (0.20665,0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,0.45807) \psarcn[linewidth=.1pt] (-1.4164,0.57915) {1.1583} {354.00} {330.00} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarc[linewidth=.1pt] (-1.4164,-0.57915) {1.1583} {5.9999} {29.999} \psline[linewidth=.1pt] (-0.41330,0.00000) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (-0.26447,-0.45807) \psarcn[linewidth=.1pt] (0.20664,-1.5162) {1.1583} {114.00} {89.999} \psline[linewidth=.1pt] (0.20665,-0.35793) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarc[linewidth=.1pt] (0.97298,0.35791) {0.27347} {152.84} {194.47} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psarcn[linewidth=.1pt] (0.76114,0.67809) {0.19787} {260.85} {210.01} \psarc[linewidth=.1pt] (1.2098,0.22116) {0.71592} {150.00} {165.52} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,0.93707) {1.1583} {234.00} {210.00} \psarc[linewidth=.1pt] (0.20667,1.5162) {1.1583} {270.00} {285.52} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psarcn[linewidth=.1pt] (1.2098,-0.22120) {0.71587} {162.00} {134.48} \psarcn[linewidth=.1pt] (0.82662,0.71586) {0.44243} {254.48} {225.52} \closepath} \rput(0.34897,-0.16203) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(-0.034159,-0.38326) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(-0.034132,0.38324) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(-0.31480,0.22124) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(-0.31482,-0.22121) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(0.29663,0.51378) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(0.60110,0.15561) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.70120,0.38812) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(0.47860,0.60277) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}} \rput(0.64886,0.51516) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}} \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.52894,0.00000) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.52894,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.20665,-0.35793) \uput{10pt}[-45]{0}(0.52894,0.00000){$a$} \uput{10pt}[-45]{0}(0.20665,-0.35793){$j$} \psarcn[linewidth=1.5000pt] (0.82663,0.71587) {0.44244} {225.52} {186.00} \psdots*[dotstyle=o,dotsize=7.0000pt](0.51663,0.40018) \psdots*[dotstyle=*,dotsize=7.0000pt](0.38660,0.66962) \psdots*[dotstyle=x,dotsize=7.0000pt](0.58981,0.57915) \uput{10pt}[90]{0}(0.38660,0.66962){$e$} \uput{10pt}[45]{0}(0.58981,0.57915){$d$} \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.26447,0.45807) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.26447,0.45807) \psdots*[dotstyle=x,dotsize=7.0000pt](0.20665,0.35793) \uput{10pt}[135]{0}(-0.26447,0.45807){$g$} \uput{10pt}[135]{0}(0.20665,0.35793){$f$} \psline[linewidth=1.5000pt] (0.00000,0.00000) (-0.26447,-0.45807) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](-0.26447,-0.45807) \psdots*[dotstyle=x,dotsize=7.0000pt](-0.41330,0.00000) \uput{10pt}[-135]{0}(-0.26447,-0.45807){$i$} \uput{10pt}[180]{0}(-0.41330,0.00000){$h$} \psarc[linewidth=1.5000pt] (0.20662,1.5163) {1.1583} {285.52} {296.84} \psdots*[dotstyle=o,dotsize=7.0000pt](0.51663,0.40018) \psdots*[dotstyle=*,dotsize=7.0000pt](0.72967,0.48274) \psdots*[dotstyle=x,dotsize=7.0000pt](0.70821,0.28957) \uput{10pt}[45]{0}(0.72967,0.48274){$c$} \uput{10pt}[0]{0}(0.70821,0.28957){$b$} \psarc[linewidth=1.5000pt] (1.2098,0.22123) {0.71586} {165.52} {198.00} \psdots*[dotstyle=o,dotsize=7.0000pt](0.51663,0.40018) \psdots*[dotstyle=*,dotsize=7.0000pt](0.52894,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.20665,0.35793) \rput(0.26447,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.42817,0.52356) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.13223,0.22904) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(-0.13223,-0.22904) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(0.62524,0.43622) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \rput(0.49424,0.19921) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}} \end{pspicture} \end{equation} \end{comment} \begin{comment} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_b^{-1}\delta_a^{-1}$ \\ $3$ & $\delta_a^{}$ \\ $4$ & $\delta_a^{-1}$ \\ $5$ & $\delta_b^{-1}\delta_a^{}$ \\ $6$ & $\delta_b^{-1}$ \\ \bottomrule \end{tabular} \hspace{2cm} \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_b^{}\delta_a^{}$ \\ $s_{2}$ & $\delta_a^{}\delta_b^{}\delta_a^{-1}$ \\ $s_{3}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{}\delta_b^{}$ \\ $s_{4}$ & $\delta_b^{-2}\delta_a^{-1}\delta_b^{}$ \\ $s_{5}$ & $\delta_b^{-1}\delta_a^{-1}\delta_b^{}\delta_a^{-1}\delta_b^{}$ \\ \bottomrule \end{tabular} \end{center} \bigskip \end{comment} \caption{A labelled fundamental domain for $\Gamma$.} \label{fig:presentation} \end{figure} Recall that the side pairings follow the convention that we move from red labels to blue labels. We now compute the pairing cycles for the preimages of $1$ (black dots). These pairing cycles are \begin{align*} &a\stackrel{s_1^{-1}}{\longrightarrow} i \stackrel{s_3^{-1}}{\longrightarrow}e\stackrel{s_5^{-1}}{\longrightarrow}c\stackrel{s_4}{\longrightarrow}a\\ &g\stackrel{s_2}{\longrightarrow}g \end{align*} which correspond to the words $\gamma_4\gamma_5^{-1}\gamma_3^{-1}\gamma_1^{-1}$ and $\gamma_2$ respectively. Some power of each word will be a relation in our presentation where this power is the order of the stabilizer of the associated point. This can be observed by dividing the order of $\sigma_1$ by the number of triangles incident to this point. Since the order of $\sigma_1$ is $5$ and there are $5$ triangles incident to $a$, we obtain the relation $\gamma_4\gamma_5^{-1}\gamma_3^{-1}\gamma_1^{-1}$. Similarly, since only one triangle is incident to $g$, we obtain the relation $\gamma_2^{5}$. Now we compute the pairing cycles for the preimages of $\infty$ (crosses). These pairing cycles are \begin{align*} &b\stackrel{s_4}{\longrightarrow} b\\ &d\stackrel{s_5}{\longrightarrow}d\\ &f\stackrel{s_3}{\longrightarrow} h\stackrel{s_2}{\longrightarrow} f\\ &j\stackrel{s_1}{\longrightarrow} j \end{align*} which correspond to the words $\gamma_4$, $\gamma_5$, $\gamma_2\gamma_3$, $\gamma_1$ respectively. Again, since the order of $\sigma_{\infty}$ is $3$, by considering the stabilizers we find the relations $\gamma_4^3$, $\gamma_5^3$, $\gamma_2\gamma_3$, $\gamma_1^3$. Thus $$ \langle \gamma_1,\gamma_2,\gamma_3,\gamma_4,\gamma_5\mid \gamma_2^5=\gamma_4\gamma_5^{-1}\gamma_3^{-1}\gamma_1^{-1}=\gamma_4^3=\gamma_5^3=\gamma_2\gamma_3=\gamma_1^3=1\rangle $$ is a presentation for $\Gamma$. This presentation simplifies to $$ \langle \gamma_1,\gamma_2,\gamma_3\mid \gamma_1^3=\gamma_3^3=\gamma_2^5=(\gamma_3\gamma_1^{-1}\gamma_2)^3=1\rangle $$ the recognizable $2$-orbifold group of signature $(0;3,3,3,5;0)$ associated to a surface with genus zero and $3$ (resp.~$1$) orbifold points of orders $3$ (resp.~$5$), respectively. \end{exm} \section{Power series expansions of differentials} \label{sec:powser} In this section, we discuss power series expansions of differentials on Riemann $2$-orbifolds in the language of modular forms on cocompact Fuchsian groups and we describe a method for computing these expansions following Voight--Willis \cite{VoightWillis}; the key ingredients are the tools developed in the previous section involving reduction to a fundamental domain. \subsection*{Computing power series expansions} Let $\Gamma < \PSL_2(\mathbb R)$ be a cocompact Fuchsian group. A \defi{modular form} for $\Gamma$ of \defi{weight} $k \in 2 \mathbb Z$ is a holomorphic function $f : \mathcal{H} \to \mathbb C$ such that \begin{equation} \label{eqn:fgamma} f(\gamma z) = j(\gamma,z)^k f(z) \text{ for all $z \in \mathcal{H}$ and all $\gamma \in \Gamma$} \end{equation} where $j(\gamma ,z) = c z+ d$ for $\gamma = \pm \begin{pmatrix} a & b \\ c & d \end{pmatrix} \in \PSL_2(\mathbb R)$. Equation (\ref{eqn:fgamma}) is equivalent to \begin{equation} \label{eqn:dgammaz} f(\gamma z)\, d(\gamma z)^{\otimes k/2} = f(z)\, dz^{\otimes k/2} \end{equation} so that a modular form $f$ of weight $k$ can be thought of as a holomorphic differential $k/2$-form on the quotient $X=\Gamma \backslash \mathcal{H}$. Since $X=\Gamma \backslash \mathcal{H}$ has no cusps, a modular form for $\Gamma$ will also sometimes be called a \defi{cusp form} (extending the classical terminology). Let $S_k(\Gamma)$ be the $\mathbb C$-vector space of cusp forms of weight $k$ for $\Gamma$. Let $f \in S_k(\Gamma)$ be a modular form of weight $k$ for $\Gamma$. Since $\Gamma$ is cocompact, we do not have $q$-expansions available; however, $f$ is holomorphic and therefore has a Taylor series expansion. Let $p \in \mathcal{H}$. By conformally mapping $\mathcal{H}$ to $\mathcal{D}$ with the map \begin{align*} w_p : \mathcal{H} & \to \mathcal{D} \\ z &\mapsto w_p=w = \frac{z-p}{z-\overline{p}} \end{align*} we express $f$ as a Taylor series in $w$ that will be valid for all $z \in \mathcal{H}$. So we define a \defi{power series expansion} for $f$ centered at $p$ to be an expression of the form \begin{equation} \label{eqn:fbnpow} f(z) = (1-w)^k \sum_{n=0}^\infty b_n w^n \end{equation} where $b_n \in \mathbb C$ and $w = w_p(z)$. The factor $(1-w)^k$ is the automorphy factor corresponding to the linear fractional transformation $w$ and is included for arithmetic reasons \cite{VoightWillis}. The basic idea of the algorithm is as follows: we use the relation of modularity (\ref{eqn:fgamma}) and reduction to the fundamental domain to solve for the unknown coefficients $b_n$. Specifically, let $D_\Gamma$ be a fundamental domain for $\Gamma$ acting on $\mathcal{D}$ as in (\ref{eqn:fundGamma}), and let $\rho > 0$ be the radius of a circle centered at the origin containing $D_\Gamma$, possible since $\Gamma$ is cocompact. To compute the power series expansion for $f$ centered at $p \in \mathcal{H}$ to some precision $\epsilon > 0$, we consider a truncation \begin{equation} \label{eqn:fNz} f(z) \approx f_N(z) = (1-w)^k \sum_{n=0}^N b_n w^n \end{equation} valid for all $|w| \leq \rho$ and some $N \in \mathbb Z_{\geq 0}$. For a point $w = w_p(z) \in \mathcal{D}$ on the circle of radius $\rho$ with $w \notin D_\Gamma$, we use Algorithm \ref{alg:fdreddelta} to find $\gamma \in \Gamma$ such that $w' = \gamma w \in D_\Gamma$. Letting $z' = w_p^{-1}(w')$, by the modularity of $f$ we have \[ (1-w')^k \sum_{n=0}^N b_n (w')^n \approx f_N(z') = j(\gamma,z)^kf(z) \approx j(\gamma,z)^k (1-w)^k \sum_{n=0}^N b_n w^n \] valid to the precision $\epsilon >0$. For each such point $w$, we obtain such a relation; therefore by taking enough such points we can obtain enough nontrivial linear relations on the unknown coefficients $b_n$ to determine them. An alternative method for obtaining linear relations amongst the coefficients $b_n$ which we found to have greater numerical stability uses Cauchy's integral formula. We identify $\mathcal{H}$ with $\mathcal{D}$ via the map $w$, we abuse notation slightly and write simply $f(w) = f(w(z))$. From the expansion (\ref{eqn:fbnpow}), we apply Cauchy's integral formula to $f(w)/(1-w)^k$: for $n \ge 0$, we have $$ b_n = \frac{1}{2\pi i} \int_C \frac{f(w)}{w^{n+1}(1-w)^k}\,dw $$ where $C$ is a simple contour around $0$. We take $C(\theta) = \rho e^{i \theta}$ for $0 \le \theta \le 2 \pi$ and thus obtain $$ b_n = \frac{1}{2\pi} \int_0^{2\pi} \frac{f(\rho e^{i\theta})}{(\rho e^{i\theta})^{n}(1-\rho e^{i\theta})^{k}}\,d\theta . $$ Breaking up $[0, 2 \pi]$ into $Q \in \mathbb Z_{\ge0}$ intervals and letting $w_m = \rho \exp(2 \pi m i/Q)$ for $m=1,\dots,Q$, we obtain the approximation (by Riemann summation) $$ b_n \approx \frac{1}{Q} \sum_{m=1}^{Q} \frac{f(z_m)}{w_m^n(1-w_m)^k} $$ valid to precision $\epsilon$ for sufficiently large $Q$. For each $m$, let $\gamma_m \in \Gamma$ be such that $w_m'=\gamma w_m \in D$, and let $z_m=\phi^{-1}(w_m)$ and $z_m'=\phi^{-1}(w_m')$. Then by the modularity of $f$, we have $$ b_n \approx \frac{1}{Q} \sum_{m=1}^{Q} \frac{f(z_m)}{w_m^n(1-w_m)^k} \approx \frac{1}{Q} \sum_{m=1}^{Q} \frac{f_N(z_m') j(\gamma_m, z_m)^{-k}}{w_m^n(1-w_m)^k}. $$ Expanding $f_N(z)$ as in (\ref{eqn:fNz}) and substituting, we obtain an approximate linear equation involving the coefficients $b_n$. Specifically, we have \begin{equation} \label{eqn:bigdealrelation} b_n \approx \sum_{r=0}^N a_{nr} b_r \end{equation} where \begin{equation} \label{eqn:jwmfact} a_{nr} = \frac{1}{Q} \sum_{m=1}^Q j(\gamma_m,z)^{-k}\frac{(w_m')^r (1-w_m')^k}{w_m^n(1-w_m)^k} . \end{equation} Let $A$ be the matrix with entries $a_{nr}$, rows indexed by $n = 0, \dots, N$, and columns indexed by $r = 0, \dots, N$. If we write $b$ as the column vector with entries $b_n$ for $n = 0, \dots, N$, we have that \begin{equation} \label{eqn:yayA} Ab \approx b \end{equation} and therefore $b$ is approximately in the eigenspace for the eigenvalue $1$ of the matrix $A$. In this way, computing the approximate eigenspace for $1$ of the matrix $K$ yields a basis for the space $S_k(\Gamma)$. In the next section, we discuss how to take this algorithm to furnish equations for Bely\u{\i}\ maps. \subsection*{Algorithmic improvements} We now discuss a few improvements on this algorithm which extend the range of practical computation. First, from (\ref{eqn:jwmfact}), conveniently the matrix $K$ factors as \begin{equation} \label{matrixmult} QK = JW' \end{equation} where $J$ is the matrix with entries $$ J_{nm}=\frac{j(\gamma_m,z_m)^{-k}}{w_m^n(1-w_m)^k} $$ with $0 \leq n \leq N$ and $1 \leq m \leq Q$ and $W'$ is the Vandermonde-like matrix with entries $$ W'_{mr} = (w_m')^r (1-w_m')^k $$ with $1 \leq m \leq Q$ and $0 \leq r \leq N$. (The matrix $W'$ further factors as the product of an honest Vandermonde matrix and a diagonal matrix.) Since $J$ and $W'$ are both fast to compute, the computation of $K$ requires just a single matrix multiplication. Taking $Q=O(N)$, one can compute the matrix $K$ using $O(N^3)$ ring operations in $\mathbb C$ using storage space $O(N^2)$; however, when $N$ is large, one can do better by finding the needed eigenspace for $K$ without ever computing the matrix $K$ directly. Second, we apply a federalist approach: we consider power series expansions at several points $p_1,\dots,p_s$, and express $f$ as piecewise-defined power series expansions in the associated Voronoi regions (evaluating in the neighborhood of the closest point). As the number of points grows, the expansion degree $N$ drops as points can be taken in smaller neighborhoods (smaller radius); in general, the geometry of the fundamental domain can lead to several optimal configurations of points. In our context, where $\Gamma \leq \Delta$ is a subgroup of finite index with cosets $\Gamma \alpha_i$ for $i=1,\dots,d$ and fundamental domain $D_\Gamma = \bigcup_{i=1}^d \alpha_i D_\Delta$ as in (\ref{eqn:fundGamma}), it is simpler to take expansions at the set of points $\{\alpha_i z_a : i=1,\dots,d\}$ as the output, since the reduction algorithm (Algorithm \ref{alg:fdredgamma}) exactly produces a point in a coset translate $\alpha_i D_\Delta$. In this way, the expansion degree $N$ depends only on that needed for convergence inside $D_\Delta$. In this approach, we repeat the application of Cauchy's theorem around each center $p_i$, so the resulting matrix is nearly block diagonal, but with some points evaluated in other neighborhoods: examples of what this approach looks like in practice can be found in section \ref{sec:examples}. Finally, we use the method of Krylov subspaces and Arnoldi iteration, a standard technique in numerical linear algebra \cite{GolubvanLoan}, to compute the eigenspace for eigenvalue $1$. Given an initial random vector $v$ and $n \in \mathbb Z_{\geq 0}$, if the sequence $v, Av, A^2 v, \dots$ converges, it converges to an eigenvector for $A$ with eigenvalue of largest absolute value; this is called the \defi{power method}. To get finer information, we define the \defi{Krylov subspace} $K_n$ for a matrix $A$ to be the span of the vectors $v, Av, A^2 v, \dots, A^n v$. The approximate eigenspaces for $A$, when intersected with $K_n$, converge to the full approximate eigenspaces, so we can expect that the intersections are good approximations to the full eigenspaces for the large eigenvalues of $A$ when $n$ is not too large; then another technique (Gram-Schmidt orthogonalization, LU decomposition, SVD decomposition, etc.)\ can be applied to the subspace more efficiently, as this subspace is of much smaller dimension. However, this method is unstable as described, and \defi{Arnoldi iteration} uses a stable version of the Gram-Schmidt process to find a sequence of orthonormal vectors spanning the Krylov subspace $K_n$. Many implementations of Arnoldi iteration typically restart after some number of iterations (such as the Implicitly Restarted Arnoldi method, IRAM). More experiments are needed to fully optimize these techniques in our setting, to select the right variant of these methods and to choose optimal parameters; we leave these experiments for future work. \subsection*{Hypergeometric series} To conclude this section, we discuss the connection between hypergeometric series and an explicit power series expansion for a uniformizer $\phi$ for $\Delta \backslash \mathcal{H}$ where $\Delta$ is a triangle group; this provides an analytic expansion for the Bely\u{\i}\ map for a subgroup $\Gamma \leq \Delta$ of finite index. Let $\Delta=\Delta(a,b,c)$ be a triangle group. Then $X(\Delta) = \Delta \backslash \mathcal{H}$ naturally possesses the structure of a Riemann $2$-orbifold with three elliptic points of orders $a,b,c$. Further, it can be given the structure of a Riemann surface of genus zero, and consequently there is an isomorphism $\phi:X(\Delta) \xrightarrow{\sim} \mathbb P^1(\mathbb C)$ of Riemann surfaces that is unique if we insist that the elliptic points map to the points $0,1,\infty$. The function $\phi$ is not holomorphic (it has a single simple pole at the elliptic point of order $c$), but it is still well-defined as a function on $X(\Delta)$, so that $\phi(\delta z) = \phi(z)$ for all $\delta \in \Delta$, and so we say that $\phi$ is a \defi{meromorphic modular form} for $\Gamma$ of weight $0$. We consider a power series expansion for $\phi$ in a neighborhood of $p=z_a$ of the form \[ \phi(w) = \phi(w(z)) = \sum_{n=0}^{\infty} b_n w^{n}. \] But since $\delta_a$ fixes $z_a \in \mathcal{H}$ and so acts by rotation by $\zeta_a=\exp(2\pi i/a)$ in $\mathcal{D}$ centered at $p$, we have \[ \phi(\delta_a w) = \phi(\zeta_a w)=\phi(w) \] so in fact \[ \phi(w) = \sum_{n=0}^{\infty} b_{an} w^{an}. \] In particular, $\phi:\mathcal{D} \to \mathbb P^1(\mathbb C)$ is an $a$-to-1 map at $w_a=0 \in \mathcal{D}$. We now give an explicit expression for $\phi$ and its coefficients $b_n$. We use the classical fact that the functional inverse of $\phi$ can be expressed by the quotient of two holomorphic functions that form a basis for the vector space of solutions to a second order linear differential equation {\cite[Theorem 15, \S 44]{Ford}}. For the triangle group $\Delta(a,b,c)$, this differential equation has $3$ regular singular points \cite[\S 109]{Ford} corresponding to the elliptic points. After a bit of simplification \cite[\S 2.3]{AndrewsAskey} we recognize this differential equation as the \defi{hypergeometric ${}_2F_1$ differential equation} \begin{equation} \label{eqn:hypergeom} z(1-z)\frac{d^2y}{dz^2} + \left(\gamma-(A+B+1)y\right)\frac{dy}{dz} - AB y =0 \end{equation} where \begin{equation} \label{eqn:alphabetagamma} A=\frac{1}{2}\left(1+\frac{1}{a}-\frac{1}{b}-\frac{1}{c}\right),\quad B=\frac{1}{2}\left(1+\frac{1}{a}-\frac{1}{b}+\frac{1}{c}\right),\quad C=1+\frac{1}{a}. \end{equation} (For other Fuchsian groups with at least 4 elliptic points or genus $>0$, one must instead look at the relevant Schwarzian differential equation; see Elkies \cite[p.~8]{ElkiesSCC} for a discussion, and Ihara \cite{Ihara} for the general case.) The solutions to the equation (\ref{eqn:hypergeom}) are well-studied. For a moment, let $A,B,C \in \mathbb R$ be arbitrary real numbers with $C \not\in \mathbb Z_{\leq 0}$. Define the \defi{hypergeometric series} \begin{equation} \label{eqn:Fabc} F(A,B,C;t)= \sum_{n=0}^{\infty} \frac{(A)_n(B)_n}{(C)_n}\frac{t^n}{n!} \in \mathbb R[[t]] \end{equation} where $(x)_n = x(x+1) \cdots (x+n-1)$ is the \defi{Pochhammer symbol}. This series is also known as the \defi{${}_2F_1$ series} or the \defi{Gaussian hypergeometric function} (see Slater \cite[\S 1.1]{Slater} for a historical introduction). The hypergeometric series is convergent for all $t \in \mathbb C$ with $|t|<1$ by the ratio test and is a solution to the hypergeometric differential equation by direct substitution. A basis of solutions in a neighborhood of $t=0$, convergent for $|t|<1$, is given by \cite[1.3.6]{Slater} \begin{align*} F_1(t) &=F(A,B,C;t) \\ F_2(t) &= t^{1-C}F(1+A-C,1+B-C,2-C;t). \end{align*} To obtain locally an inverse to $\phi$ as above, since $1-C=-1/a$, we have \[ F_1(t)/F_2(t) \in t^{1/a} \mathbb Q[[t]] \] which uniquely defines this as the correct ratio up to a scalar $\kappa \in \mathbb C^\times$. The value of $\kappa$ can be recovered by plugging in $t=1$: when $C-A-B>0$, we have \cite[1.7.6]{Slater} \begin{equation} \label{eqn:FABC1} F(A,B,C;1)=\frac{\Gamma(C)\Gamma(C-A-B)}{\Gamma(C-A)\Gamma(C-B)} \end{equation} where $\Gamma$ is the complex $\Gamma$-function. Since \[ 2-C-(1+A-C)=C-A-B =\left(1+\frac{1}{a}\right)-\left(1+\frac{1}{a}+\frac{1}{b}\right)=\frac{1}{c}>0 \] we can use (\ref{eqn:FABC1}) to recover $\kappa$ by solving \[\kappa \frac{F_1(1)}{F_2(1)}=w_b=\frac{z_b-z_a}{z_b-\overline{z_a}} \] where $z_a=i$ and $z_b=\mu i$ are as in Proposition \ref{prop:triangleembed}. We obtain \begin{equation} \label{eqn:kdef} \kappa=\left(\frac{\mu-1}{\mu+1}\right) \frac{\Gamma(2-C)\Gamma(C-A)\Gamma(C-B)}{\Gamma(1-A)\Gamma(1-B)\Gamma(C)} \in \mathbb R. \end{equation} Therefore, the function \begin{equation} \label{eqn:psit} \psi(t) = \kappa\frac{F_1(t)}{F_2(t)} \end{equation} maps the neighborhood of $t=0$ with $|t|<1$ to a neighborhood of $w=0 \in \mathcal{D}$ (intersected with $D_\Delta$) and $\phi$ is an inverse for the map $\psi$. \begin{rmk} If we make a cut in the $t$-plane along the real axis from $t=1$ to $t=\infty$, that is, we insist that $-\pi < \arg(-t) \leq \pi$ for $|t| \geq 1$, then the function $F(A,B,C;t)$ defined by the hypergeometric series can be analytically continued to a holomorphic function in the cut plane. Similar formulas then hold for explicitly given Puiseux series expansions around the two other elliptic points \cite[\S 5.2]{VoightThesis}. \end{rmk} \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-19.pdf} \begin{comment} \begin{equation} \label{fig:puiseux} \notag \psset{unit=.75cm} \begin{pspicture}(-5,-5)(5,5) \def\mathbb H{% \psline[linewidth=.5pt]{<->}(-3,0)(3,0) \psline[linewidth=.5pt]{<->}(0,-1)(0,8) \uput{5pt}[0]{0}(3,0){coordinate $z$} \pscustom[linewidth=.5pt]{ \psline(0,1)(0,7.32763187) \psarcn(-12.69184,0){14.6552637}{30}{4.18839} \psarc(1,0){1.414213}{49.18839}{135} \closepath} \pscustom[linewidth=.5pt,fillstyle=solid,fillcolor=lightgray]{ \psline(0,1)(0,7.327631) \psarc(12.6918307,0){14.65526}{150}{175.81160} \psarcn(-1,0){1.41421}{130.8116016}{45} \closepath} \psdots*[dotstyle=o,dotsize=4.5000pt](0,1) \psdots*[dotstyle=*,dotsize=4.5000pt](0,7.327631879) \uput{5pt}[-45]{0}(0,1){$z_a = i$} \uput{5pt}[30]{0}(0,7.327631){$z_b$} } \def\DD{% \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.001pt]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.75983,0.00000) \psarc[linewidth=.001pt] (1.0379,-0.48172) {0.55624} {120.00} {180.00} \psline[linewidth=.001pt] (0.48172,-0.48172) (0.00000,0.00000) \closepath} \pscustom[linewidth=.001pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.001pt] (0.00000,0.00000) (0.75983,0.00000) \psarcn[linewidth=.001pt] (1.0379,0.48172) {0.55624} {240.00} {180.00} \psline[linewidth=.001pt] (0.48172,0.48172) (0.00000,0.00000) \closepath} \psdots*[dotstyle=o,dotsize=4.5pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=4.5pt](0.75983,0.00000) \uput{5pt}[180]{0}(0,0){$w_a$} \uput{5pt}[0]{0}(0.75983,0){$w_b$} \pscircle[linewidth=.001pt, linestyle = dashed](0,0){0.75983} \uput{3.5pt}[-90]{0}(0,-0.75983){$|w| < |w_b|$} \uput{3cm}[-90]{0}(0,0){coordinate $w$} } \def\CC{% \newgray{lightestgray}{.93} \psframe[linecolor=black, linewidth=.5pt, fillcolor=white, fillstyle=solid](-0.75,-0.75)(1,1) \pscircle[linewidth=.001pt, linestyle = dashed](0,0){0.5} \psdots*[dotsize=4.5000pt](0,0) \uput{7pt}[-90]{0}(0,0){$\phi(0) = 0$} \uput{1.5cm}[45]{0}(0,0){$|t| < 1$} \uput{3cm}[-90]{0}(0,0){coordinate $t$} } \rput{0}(-8,4){\mathbb H} \psset{unit=2.4cm} \rput{0}(-2.25,-1){\DD} \rput{0}(2.15,-1){\CC} \rput{0}(-3.4,0){ \rput{0}(0,1.1){$\mathcal{H}$} \rput{0}(0,-1){$\mathcal{D}$} \pcline[linewidth=1pt, nodesep=8pt]{->}(0,1.1)(0,-1) \uput{5pt}[135]{0}(0,.05){$w_{z_a}$} } \rput{0}(-.8,-1){$\mathcal{D}$} \rput{0}(.8,-1){$\mathbb C$} \pcline[linewidth=1pt, nodesep=8pt]{->}(-.8,-1)(.8,-1) \uput{5pt}[90]{0}(0,-1){$\phi$} \uput{25pt}[90]{0}(0,-1){$\displaystyle \phi(w) = \left ( \frac{w}{\kappa}\right )^a + \ldots \in \mathbb Q[[(w/\kappa)^a]]$} \pscurve[]{->}(.7,-1.05)(.2,-1.2)(-.2,-1.2)(-.7,-1.05) \uput{20pt}[-90]{0}(0,-1){$\psi$} \uput{80pt}[-90]{0}(0,-1){$\displaystyle \psi(t) = \kappa \frac{F_1(t)}{F_2(t)} = \kappa t^{1/a} + \cdots \in \kappa t^{1/a} \mathbb Q[[t]]$} \rput{0}(0,-2.95){$\displaystyle \phi(\psi(t)) = \frac{1}{\kappa^a}\left(\kappa t^{1/a} + \cdots \right)^a + \cdots = t$} \end{pspicture} \end{equation} \end{comment} \caption{Hyperbolic uniformization of the triangle quotient via hypergeometric series.} \label{fig:puiseux} \end{figure} Computing the reversion of the Puiseux series $\phi(t)=w$ then gives a power series expansion for $\psi(w)=t$. We have \begin{equation} \label{eqn:expandpuiseux} \phi(w) = \left(\frac{w}{\kappa}\right)^a + \frac{a(C(1+A-C)(1+B-C)-AB(2-C))}{C(2-C)}\left(\frac{w}{\kappa}\right)^{2a} + \dots \in \mathbb Q\left[\!\left[\left(\frac{w}{\kappa}\right)^a\right]\!\right]. \end{equation} Note in particular that the expansion (\ref{eqn:expandpuiseux}) has rational coefficients when considered as a series in $w/\kappa$. \section{Computing equations, and examples} \label{sec:examples} In this section, we compute equations for Bely\u{\i}\ maps combining the algorithms of the past two sections. Along the way, we present several explicit examples in full. Throughout, we retain the notation introduced in section 3: in particular, let $\Gamma \leq \Delta=\Delta(a,b,c)$ be a subgroup of index $d$ corresponding to a transitive permutation triple $\sigma \in S_d^3$. We compute defining equations for the Bely\u{\i}\ map \[ \phi : X(\Gamma)=\Gamma \backslash \mathcal{H} \to X(\Delta) = \Delta \backslash \mathcal{H} \] defined over a number field. Our description and method depend on the genus, with low genus cases handled separately. \subsection*{Genus zero} First, suppose that $X(\Gamma)$ has genus zero. Then we have an analytic isomorphism $x:X(\Gamma) \cong \mathbb P_\mathbb C^1$ of Riemann surfaces and the Bely\u{\i}\ map $\phi:X(\Gamma) \to X(\Delta)$ corresponds to a rational function $\phi(x) \in \mathbb C(x)$ of degree $d$. The isomorphism $x$ is not uniquely specified. One way to fix this is to specify the images of three points on $X(\Gamma)$, as was done for the function $\phi$; but this may extend the field of definition. Instead, let $e$ be the order of the stabilizer of $w=0$, so that $x \in \mathbb C[[w^e]]$; we require that $x$ has Laurent series expansion \begin{equation} \label{eqn:xw} x(w) = (\Theta w)^e + O(w^{3e}) \end{equation} for some $\Theta \in \mathbb C^\times$. For a given choice of $\Theta$, this normalization specifies $x$ uniquely. Experimentally, we observe that a choice \begin{equation} \label{eqn:Theta0} \Theta = \sqrt[a]{\alpha} \left(\frac{1}{\kappa}\right) \end{equation} for some $\alpha \in \overline{\mathbb Q}$ yields a series \[ x(w) = \sum_{\substack{n=e \\ e \mid n}}^{\infty} \frac{c_n}{n!} (\Theta w)^n \] (weight $k=0$ so no factor $(1-w)^k$ appears) with Taylor coefficients $c_n$ that belong to a number field $K$ (in fact, with small denominators). \begin{rmk} \label{rmk:Shimura} The algebraicity of the coefficients is similar in spirit to results of Shimura \cite{Shimura1,Shimura2,Shimura3}, who proved algebraicity of $c_n$ when the group $\Gamma$ is \defi{arithmetic}, commensurable with the units in the maximal order of a quaternion algebra defined over a totally real field and split at a unique real place. There are exactly $85$ triples $(a,b,c)$ with $a \leq b \leq c$ that are arithmetic, by a theorem of Takeuchi \cite{Takeuchi,Takeuchi2}; so in the infinitely many remaining cases, one does not know that the coefficients are algebraic, and we merely report the observation that the coefficients seem to be algebraic (and, in fact, nearly integral after suitable normalization). \end{rmk} We compute an expansion for $x$ as follows. First, we find $k \in 2\mathbb Z_{\geq 0}$ such that $\dim_\mathbb C S_k(\Gamma) \geq 2$. The dimension of $S_k(\Gamma)$ is given by Riemann-Roch \cite{Shimura}: we have $\dim_\mathbb C S_2(\Gamma)=g$ and \begin{equation} \label{eqn:dimformula} \dim_\mathbb C S_k(\Gamma) = (k-1)(g-1) + \displaystyle{\sum_{i=1}^r \left\lfloor \frac{k(e_i-1)}{2e_i} \right\rfloor}, \quad \text{ if $k \geq 4$}. \end{equation} Next, we compute an \defi{echelonized basis} \cite[\S 4.4]{Klug} of functions with respect to the coefficients $b_n$. Let \[ s \equiv \frac{k}{2}(e-1) \pmod e \] with $0 \leq s < e$. Then the expansions in $S_k(\Gamma)$ belong to $w^s \mathbb C[[w^e]]$. We find functions \begin{equation} \label{eqn:gandh} g(w)=w^{m} + O(w^{m+2e})\quad \text{and} \quad h(w)=w^{m+e} + O(w^{m+2e}) \end{equation} with $m \equiv s \pmod{e}$ maximal with this property and then take \[ x(w)=\frac{\Theta^e h(w)}{g(w) + ch(w)} \] where $c$ is the coefficient of $w^{m+2e}$ in $h$ to obtain an expansion as in \eqref{eqn:xw}. The fact that the function $x(w)$ has degree $1$ and thereby gives an isomorphism of Riemann surfaces $x:X=\Gamma \backslash \mathcal{H} \to \mathbb P_\mathbb C^1$ follows from Riemann-Roch: if $K$ is a canonical divisor for orbifold $X$ and $m=\dim H^0(X,dK)$ with $d=k/2$, then $g,h$ span $H^0(X,dK-(m-1)P)$ where $P$ is the point $w=0$. Finally, using the expansion for $\phi$ in (\ref{eqn:expandpuiseux}), we use linear algebra to find the rational function $\phi(x)$ of degree $d$. \begin{rmk} We could instead compute the ring of modular forms for $\Gamma$, making no choices at all: the structure of this ring is described by Voight--Zureick-Brown \cite{JVDZB} and contains complete information about differentials on the orbifold $X(\Gamma)$. This remark holds in all genera below as well. \end{rmk} These calculations are all performed over the complex numbers. The Bely\u{\i}\ map $\phi(x)$ will be defined over a number field of degree at most equal to the size of the corresponding refined passport, and so one can use the LLL algorithm \cite{LLL} to find putative polynomials that these coefficients satisfy. Alternatively, one can compute the maps $\phi(x)$ for all Galois conjugates (possibly belonging to different refined passports), and then, using continued fractions, recognize the symmetric functions in these conjugates to obtain the minimal polynomial for each coefficient. \begin{exm} The smallest degree $d$ for which there exists a hyperbolic, transitive permutation triple $\sigma$ with genus $g=0$ is $d=5$. There are nine refined passports and each is rigid, having a unique triple up to simultaneous conjugation. Two triples generate $G=A_5\cong \PSL_2(\mathbb F_5)$ and seven generate $G=S_5 \cong \PGL_2(\mathbb F_5)$. We consider the first two triples in detail; two of the resulting maps from the latter collection are merely reported below. First, consider the triples with $G=A_5$: they both have $(a,b,c)=(5,3,3)$, up to reordering. The conjugacy class of an element of order $3$ is unique, but there are two conjugacy classes in $A_5$ of elements of order $5$, and the two triples correspond to these two choices. The two conjugacy classes of order $5$ are interchanged by the outer automorphism of $A_5$ obtained from conjugating by an element in $\PGL_2(\mathbb F_5) \setminus \PSL_2(\mathbb F_5)$. We will see (in several ways) below that this automorphism corresponds to complex conjugation on the associated Bely\u{\i}\ maps, and that the Bely\u{\i}\ map is defined over $\mathbb Q$ (and hence over $\mathbb R$). We begin with the first triple \begin{equation} \label{eqn:sigmaexmgenus0} \sigma_0=(1\ 5\ 4\ 3\ 2), \quad \sigma_1 = (1\ 2\ 3), \quad \sigma_\infty=(3\ 4\ 5). \end{equation} Let $\Gamma \leq \Delta(5,3,3)$ be the group of index $5$ associated to this triple as in (\ref{eqn:gammaperm}). The ramification diagram then looks as in Figure \ref{fig:ramification}. The points are labelled with the order of their stabilizer group (no label indicates trivial stabilizer). \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-20.pdf} \begin{comment} \begin{equation} \label{fig:ramification} \notag \psset{unit=2cm} \begin{pspicture} \psdots*(0,0)(2,0)(4,0) \psdots*(0,2)(1.25,2)(2,2)(2.75,2)(3.25,2)(4,2)(4.75,2) \rput(6,0){$X(\Delta)$} \rput(6,2){$X(\Gamma)$} \uput{10pt}[-90]{0}(0,0){$\phi = 0$} \uput{10pt}[-90]{0}(2,0){$\phi = 1$} \uput{10pt}[-90]{0}(4,0){$\phi = \infty$} \uput{10pt}[0]{0}(0,0){\psscalebox{.8 .8}{$5$}} \uput{10pt}[0]{0}(2,0){\psscalebox{.8 .8}{$3$}} \uput{10pt}[0]{0}(4,0){\psscalebox{.8 .8}{$3$}} \uput{10pt}[90]{0}(2,2){\psscalebox{.8 .8}{$3$}} \uput{10pt}[90]{0}(2.75,2){\psscalebox{.8 .8}{$3$}} \uput{10pt}[90]{0}(4,2){\psscalebox{.8 .8}{$3$}} \uput{10pt}[90]{0}(4.75,2){\psscalebox{.8 .8}{$3$}} \uput{10pt}[180]{0}(0,2){$x = \infty$} \multirput(-.06,0)(.03, 0){5}{\pcline[nodesep = 3pt](0,0)(0,2)} \pcline[nodesep = 3pt](2,0)(2,2) \pcline[nodesep = 3pt](2,0)(2.75,2) \multirput(-.04,0)(.03, 0){3}{\pcline[nodesep = 3pt](2,0)(1.25,2)} \pcline[nodesep = 3pt](4,0)(4,2) \pcline[nodesep = 3pt](4,0)(4.75,2) \multirput(-.04,0)(.03, 0){3}{\pcline[nodesep = 3pt](4,0)(3.25,2)} \end{pspicture} \end{equation} \end{comment} \caption{The ramification diagram for the triple (\ref{eqn:sigmaexmgenus0}).} \label{fig:ramification} \end{figure} From the Riemann-Hurwitz formula (\ref{eqn:RH}) we find that $g(X(\Gamma))=0$. We have $e=1$ so $s=0$ and our power series belong to ring $\mathbb C[[w]]$ in all weights $k$. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-21.pdf} \begin{comment} \begin{equation} \label{fig:dessin-5,3,3} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0,0) (0.5289355,0) \psarc[linewidth=.1pt] (1.20976257,-0.393075688) {0.786151377} {150} {174} \psline[linewidth=.1pt] (0.42791781,-0.310900493) (0,0) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0,0) (0.52893551,0) \psarcn[linewidth=.1pt] (1.20976257,0.393075688) {0.786151377} {210} {186} \psline[linewidth=.1pt] (0.427917812,0.310900493) (0,0) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0,0) (0.16345008,-0.503047567) \psarc[linewidth=.1pt] (0,-1.27201964) {0.786151358} {78.0000000} {102.000000} \psline[linewidth=.1pt] (-0.163450188,-0.503047567) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (0.163450088,-0.503047547) \psarcn[linewidth=.1pt] (0.747674611,-1.02901364) {0.786151377} {138.000000} {114.000000} \psline[linewidth=.1pt] (0.427917812,-0.310900466) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (-0.427917272,-0.310900466) \psarc[linewidth=.1pt] (-1.20976653,-0.393078878) {0.786157757} {5.99999994} {29.9999999} \psline[linewidth=.1pt] (-0.528935168,0.000000000) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (-0.427918272,-0.310993466) \psarcn[linewidth=.1pt] (-0.747674390,-1.02908551) {0.786151377} {66.0000001} {42.0000001} \psline[linewidth=.1pt] (-0.163450068,-0.503047567) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000000,0.0000000000) (-0.427917818,0.310900466) \psarc[linewidth=.1pt] (-0.747674390,1.02908564) {0.786151757} {294.000000} {318.000000} \psline[linewidth=.1pt] (-0.163450062,0.503047565) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (-0.427917272,0.310900496) \psarcn[linewidth=.1pt] (-1.20976257,0.393075678) {0.786151377} {354.000000} {330.000000} \psline[linewidth=.1pt] (-0.528935568,0.000000000) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (0.163450062,0.503047547) \psarc[linewidth=.1pt] (0.747674391,1.02908551) {0.786151759} {222.000000} {246.000000} \psline[linewidth=.1pt] (0.427917818,0.310900496) (0.000000000,0.000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000,0.000000000) (0.163450068,0.50304757) \psarcn[linewidth=.1pt] (0.000000000,1.27201961) {0.786151378} {282.0000000} {258.000000} \psline[linewidth=.1pt] (-0.163450068,0.503047567) (0.000000000,0.000000000) \closepath} \rput(0.462088184,-0.150141554) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(0.285586204,-0.393075689) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.462088185,0.150141553) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(-0.462088915,-0.150141550) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.285586204,0.393075689) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(0,0.485868271) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(0,-0.485868271) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(-0.285586206,-0.393075688) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(-0.285586294,0.393075688) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}} \rput(-0.462088185,0.150141000) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}} \psline[linewidth=1.50000000pt] (0.000000000,0.000000000) (0.528935512,0.000000000) \psdots*[dotstyle=o,dotsize=7.00000000pt](0.000000000,0.000000000) \psdots*[dotstyle=*,dotsize=7.00000000pt](0.528935518,0.000000000) \psdots*[dotstyle=x,dotsize=7.00000000pt](0.427917872,-0.310900496) \psline[linewidth=1.50000000pt] (0.000000000,0.000000000) (0.163450068,-0.503047565) \psdots*[dotstyle=o,dotsize=7.00000000pt](0.000000000,0.000000000) \psdots*[dotstyle=*,dotsize=7.00000000pt](0.163450068,-0.503047547) \psdots*[dotstyle=x,dotsize=7.00000000pt](-0.163450068,-0.503047547) \psline[linewidth=1.50000000pt] (0.000000000,0.000000000) (-0.427917818,-0.310900496) \psdots*[dotstyle=o,dotsize=7.00000000pt](0.000000000,0.000000000) \psdots*[dotstyle=*,dotsize=7.00000000pt](-0.427917872,-0.310900466) \psdots*[dotstyle=x,dotsize=7.00000000pt](-0.528935518,0.00000000) \psline[linewidth=1.5000000pt] (0.000000000,0.000000000) (-0.427917818,0.310900493) \psdots*[dotstyle=o,dotsize=7.00000000pt](0.000000000,0.000000000) \psdots*[dotstyle=*,dotsize=7.00000000pt](-0.427917812,0.310903466) \psdots*[dotstyle=x,dotsize=7.00000000pt](-0.163450068,0.503047547) \psline[linewidth=1.50000000pt] (0.000000000,0.000000000) (0.163450088,0.503047547) \psdots*[dotstyle=o,dotsize=7.00000000pt](0.000000000,0.000000000) \psdots*[dotstyle=*,dotsize=7.00000000pt](0.163450068,0.503047547) \psdots*[dotstyle=x,dotsize=7.00000000pt](0.427917872,0.310900466) \rput(0.264467784,0.000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.0817250312,-0.251523782) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.213958906,-0.155450246) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(-0.213958936,0.155450243) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(0.0817250312,0.251523783) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \end{pspicture} \end{equation} \end{comment} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_a^{-1}$ \\ $3$ & $\delta_a^{-2}$ \\ $4$ & $\delta_a^{2}$ \\ $5$ & $\delta_a^{}$ \\ \bottomrule \end{tabular} \;\;\;\;\;\;\;\;\;\; \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_b^{}\delta_a^{}$ \\ $s_{2}$ & $\delta_b^{-1}\delta_a^{2}$ \\ $s_{3}$ & $\delta_a^{}\delta_b^{}\delta_a^{-1}$ \\ $s_{4}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{2}$ \\ $s_{5}$ & $\delta_a^{2}\delta_b^{}\delta_a^{-2}$ \\ \bottomrule \end{tabular} \end{center} \bigskip \caption{A fundamental domain and conformally drawn dessin for $\Gamma$ corresponding to the permutation triple \eqref{eqn:sigmaexmgenus0} using the methods in section \ref{sec:cosets}.} \label{fig:dessin-5,3,3} \end{figure} From the dimension formula (\ref{eqn:dimformula}), we compute that \[ \dim_\mathbb C S_k(\Gamma) = 0,1,3 \quad \text{ for $k=2,4,6$} \] so we look at the space of modular forms of weight $6$ for $\Gamma$. We work in precision $\varepsilon=10^{-30}$; the fundamental domain is contained in a circle of radius $\rho=0.528935$, and so (assuming $b_n=O(1)$) we take $N= \lceil \log \varepsilon/\log \rho \rceil = 109$ and $Q=2N$. We compute the matrix $A$ (\ref{eqn:yayA}) and the singular value decomposition of the matrix $A-1$ and find it has three small singular values, the largest of which is $< 8 \cdot 10^{-29}$, yielding an approximate eigenspace for $A$ with eigenvalue $1$ of dimension $3$. Computing as in (\ref{eqn:gandh}) and taking \begin{equation} \label{eqn:exmgen0Theta} \Theta = 0.3917053\ldots + 1.205545\ldots i= \sqrt[5]{\frac{81}{2}}\exp(2\pi i/5)\left(\frac{1}{\kappa}\right) \end{equation} where $\kappa$ is as in (\ref{eqn:kdef}), we find the (approximate) basis \begin{align*} f(w)&=(1-w)^6\left(1 - \frac{81}{4!}(\Theta w)^4 - \frac{6075}{6!} (\Theta w)^6 + \frac{382725}{2\cdot 8!} (\Theta w)^8 \right. \\ & \qquad\qquad\qquad\qquad \left. - \frac{24111675}{10!} (\Theta w)^{10} + O(w^{14})\right) \\ g(w)&=(1-w)^6\left((\Theta w) + \frac{9}{3!}(\Theta w)^3 - \frac{675}{2 \cdot 5!}(\Theta w)^5 - \frac{1148175}{2\cdot 9!} (\Theta w)^9 \right. \\ &\qquad\qquad\qquad\qquad \left. - \frac{1148175}{11!}(\Theta w)^{11} + \frac{27807650325}{4\cdot 13!}(\Theta w)^{13} + O(w^{15})\right) \\ h(w) &= (1-w)^6\left((\Theta w)^2 + \frac{881798400}{12!} (\Theta w)^{12} + O(w^{22})\right) \end{align*} which satisfy \[ g^2-fh - 3h^2 = 0 \] to the precision computed. This gives \begin{equation} \label{eqn:exmgen0x} \begin{aligned} x(w) = \frac{h(w)}{g(w)} &= (\Theta w) - \frac{9}{3!}(\Theta w)^3 + \frac{1215}{2\cdot 5!}(\Theta w)^5 - \frac{-59535}{7!}(\Theta w)^7 \\ &\qquad\quad + \frac{12170655}{9!}(\Theta w)^9 - \frac{-6708786525}{2\cdot 11!}(\Theta w)^{11} + O(w^{13}). \end{aligned} \end{equation} From (\ref{eqn:psit}) we obtain \[ \frac{w}{\kappa}=\frac{\psi(t)}{\kappa} = t^{1/5} \left(1 + \frac{1}{10} t + \frac{3943}{89100}t^2 + \frac{2161}{81000}t^3 + \frac{23027911}{1246590000}t^4 + \dots \right) \] so \[ \phi(w)=\psi^{-1}(t) = \left(\frac{w}{\kappa}\right)^5 - \frac{1}{2}\left(\frac{w}{\kappa}\right)^{10} + \frac{637}{3564}\left(\frac{w}{\kappa}\right)^{15} - \frac{383}{7128}\left(\frac{w}{\kappa}\right)^{20}+O(w^{25}) \] so with $\Theta$ as in (\ref{eqn:exmgen0Theta}) we have \[ \phi(w)=\psi^{-1}(t) = \frac{81}{2}(\Theta w)^5 - \frac{6561}{8}(\Theta w)^{10} + \frac{4179357}{352} (\Theta w)^{15} +O(w^{20}). \] From the ramification data, we have $\phi(x)$ is a rational function in $x$ of degree $5$ with numerator $x^5$; comparing power series, we then compute \begin{equation} \label{eqn:phix0} \phi(x) = \frac{648x^5}{324x^5+405x^4-120x^2+16} = \frac{648x^5}{(3x+2)^3(12x^2-9x+2)} \end{equation} and \begin{equation} \label{eqn:phix1} \phi(x)-1 = \frac{(3x-2)^3(12x^2+9x+2)}{(3x+2)^3(12x^2-9x+2)}. \end{equation} We also verify this numerically: we find that \[ \{x(\alpha_i \cdot z_c)\}_i = \{0.666\ldots, -0.375000\ldots \pm 0.16137\ldots \sqrt{-1}\} \] and \[ \{x(\alpha_i \cdot z_b)\}_i = \{-0.666\ldots, 0.375000\ldots \pm 0.16137\ldots \sqrt{-1}\} \] which agree with the roots of the numerator and denominator in (\ref{eqn:phix1}). The expression for $\phi(x)$ in (\ref{eqn:phix0}) is not the simplest one possible, allowing for linear fractional transformations in the $x$ and $\phi$: for example, we have \[ \frac{1}{\phi(1/3x)} = 6x^5 - 5x^3 + 15/8 x + 1/2 \] and further substituting $x \leftarrow x-1/2$ we obtain the nicer form \begin{equation} \label{eqn:phiop} \phi_{\text{op}}(x) = 6x^5 - 15x^4 + 10x^3 = x^3(6x^2 - 15x + 10) = 1 + (x-1)^3(6x^2 + 3x + 1). \end{equation} We can verify that the computed map is correct by computing it using a direct method: beginning with \[ x^3(a_2x^2 + a_1x+a_0) = 1 + (x-1)^3(b_2 x^2 + b_1x + b_0) \] we find the system of equations \[ b_0 -1 = 3b_0 - b_1 = 3b_0-3b_1+b_2 = a_0-b_0+3b_1-3b_2 = a_1-b_1+3b_2=a_2-b_2 = 0\] which has the unique solution $(a_0,a_1,a_2,b_0,b_1,b_2)=(10,-15,6,1,3,6)$ as in (\ref{eqn:phiop}). (For more on the direct method and its extensions, see Sijsling--Voight \cite{SijslingVoight}.) The other triple, companion to \eqref{eqn:sigmaexmgenus0}, is \begin{equation} \label{eqn:othertriple} \sigma_0=(1\ 3\ 2\ 4\ 5), \quad \sigma_1 = (1\ 2\ 3), \quad \sigma_\infty=(1\ 5\ 4). \end{equation} If we run the above procedure again on this triple, we find the same expression for $x(w)$ as in (\ref{eqn:exmgen0x}) except now with $\Theta=0.3917053\ldots - 1.205545\ldots i$, i.e., the complex conjugate of the previous value (\ref{eqn:exmgen0Theta}). In other words, the antiholomorphic map given by complex conjugation identifies these two Bely\u{\i}\ maps. The federalist approach in this example does not give anything new, since the origin $0 \in \mathcal{D}$ does not move $\{\alpha_i \cdot 0 : i=0,\dots,5\}=\{0\}$ under the coset representatives $\alpha_i$. Using our implementation, the whole calculation takes a few seconds with a standard desktop CPU. (We will see the benefits of using improved numerical linear algebra techniques in the next example.) There are seven other refined passports of genus zero with $d=5$, all with monodromy group $G=S_5$. The tables below list representative permutation triples for each refined passport, as well as polynomials $\phi_0$, $\phi_1$, $\phi_\infty$ such that \begin{align*} \phi(x) = \frac{\phi_0(x)}{\phi_\infty(x)} = 1 + \frac{\phi_1(x)}{\phi_\infty(x)} \, . \end{align*} \begin{comment} \begin{itemize} \item $(a,b,c) = (5,2,4)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3\ 4\ 5) & 128x^5\\ 1 & (1\ 2) & (3x-2)^2 (14x^3 + 17 x^2 + 12 x + 4)\\ \infty & (1\ 5\ 4\ 3) & (x+2)^4(2x-1) \end{array} $$ \item $(a,b,c) = (5,2,6)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3\ 4\ 5) & 108 x^5\\ 1 & (1\ 3) & (x+1)^2 (36 x^3 - 12 x^2 - 2 x + 1)\\ \infty & (1\ 5\ 4)(2\ 3) & (2x - 1)^3 (3x + 1)^2 \end{array} $$ \item $(a,b,c) = (4,2,6)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3\ 4) & 84375 x^4 (15x + 16)\\ 1 & (1\ 3)(4\ 5) & (3x^2 + 512x +512)^2 (17x - 16)\\ \infty & (1\ 5\ 4)(2\ 3) & (2x - 1)^3 (3x + 1)^2 \end{array} $$ \item $(a,b,c) = (6,2,6)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3)(4\ 5) & -6250x^3 (5x + 1)^2\\ 1 & (1\ 5)(3\ 4) & -(113x^2 + 23x + 2)^2 (13x - 1)\\ \infty & (1\ 4)(2\ 5\ 3) & (19x + 2)^2 (3x - 1)^3 \end{array} $$ \item $(a,b,c) = (4,3,4)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3\ 4) & -84375 x^4 (5x - 28)\\ 1 & (1\ 4\ 5) & -(11x - 28)^3 (321 x^2 + 784 x + 784)\\ \infty & (2\ 5\ 4\ 3) & 1792 (x+7)^4 (3x - 4) \end{array} $$ \item $(a,b,c) = (4,3,6)$ $$ \begin{array}{c|c|c|c} i & \multicolumn{2}{c|}{\sigma_i} & \phi_i\\ \hline 0 & (1\ 2\ 3\ 4) & (1\ 2\ 3\ 4) & 6(26496 + 10719\sqrt{6}) x^4 ((8 - 3\sqrt{6})x -1)) \\ 1 & (2\ 4\ 5) & (3\ 4\ 5) & -6(2244584 + 99851\sqrt{6})((830 - 339\sqrt{6})x^2 - x - 1)((3-\sqrt{6})x - 1)^3\\ \infty & (1\ 2)(3\ 5\ 4) & (1\ 3\ 2)(4\ 5) & (10x - (7 + 3\sqrt{6}))^2 (6x + (12 + 5\sqrt{6}))^3 \end{array} $$ \item $(a,b,c) = (3,6,6)$ $$ \begin{array}{c|c|c} i & \sigma_i & \phi_i\\ \hline 0 & (1\ 2\ 3) & 2x^3 (9x^2 - 5) \\ 1 & (1\ 2\ 5)(3\ 4) & (x - 1)^3 (3x + 2)^2\\ \infty & (1\ 4\ 3)(2\ 5) & (x + 1)^3 (3x - 2)^2 \end{array} $$ \end{itemize} \end{comment} \begin{landscape} $$ \begin{array}{c|c|c|c} a,b,c & i & \sigma_i & \phi_i\\ \hline 5 & 0 & (1\ 2\ 3\ 4\ 5) & 128x^5\\ 2 & 1 & (1\ 2) & (3x-2)^2 (14x^3 + 17 x^2 + 12 x + 4)\\ 4 & \infty & (1\ 5\ 4\ 3) & (x+2)^4(2x-1) \end{array} $$ $$ \begin{array}{c|c|c|c} a,b,c, & i & \sigma_i & \phi_i\\ \hline 5 & 0 & (1\ 2\ 3\ 4\ 5) & 108 x^5\\ 2 & 1 & (1\ 3) & (x+1)^2 (36 x^3 - 12 x^2 - 2 x + 1)\\ 6 & \infty & (1\ 5\ 4)(2\ 3) & (2x - 1)^3 (3x + 1)^2 \end{array} $$ $$ \begin{array}{c|c|c|c} a,b,c & i & \sigma_i & \phi_i\\ \hline 4 & 0 & (1\ 2\ 3\ 4) & 84375 x^4 (15x + 16)\\ 2 & 1 & (1\ 3)(4\ 5) & (3x^2 + 512x +512)^2 (17x - 16)\\ 6 & \infty & (1\ 5\ 4)(2\ 3) & (2x - 1)^3 (3x + 1)^2 \end{array} $$ $$ \begin{array}{c|c|c|c} a,b,c & i & \sigma_i & \phi_i\\ \hline 6 & 0 & (1\ 2\ 3)(4\ 5) & -6250x^3 (5x + 1)^2\\ 2 & 1 & (1\ 5)(3\ 4) & -(113x^2 + 23x + 2)^2 (13x - 1)\\ 6 & \infty & (1\ 4)(2\ 5\ 3) & (19x + 2)^2 (3x - 1)^3 \end{array} $$ $$ \begin{array}{c|c|c|c} a,b,c & i & \sigma_i & \phi_i\\ \hline 4 & 0 & (1\ 2\ 3\ 4) & -84375 x^4 (5x - 28)\\ 3 & 1 & (1\ 4\ 5) & -(11x - 28)^3 (321 x^2 + 784 x + 784)\\ 4 & \infty & (2\ 5\ 4\ 3) & 1792 (x+7)^4 (3x - 4) \end{array} $$ $$ \begin{array}{c|c|c|c|c} a,b,c & i & \multicolumn{2}{c|}{\sigma_i} & \phi_i\\ \hline 4 & 0 & (1\ 2\ 3\ 4) & (1\ 2\ 3\ 4) & 6(26496 + 10719\sqrt{6}) x^4 ((8 - 3\sqrt{6})x -1)) \\ 3 & 1 & (2\ 4\ 5) & (3\ 4\ 5) & -6(2244584 + 99851\sqrt{6})((830 - 339\sqrt{6})x^2 - x - 1)((3-\sqrt{6})x - 1)^3\\ 6 & \infty & (1\ 2)(3\ 5\ 4) & (1\ 3\ 2)(4\ 5) & (10x - (7 + 3\sqrt{6}))^2 (6x + (12 + 5\sqrt{6}))^3 \end{array} $$ $$ \begin{array}{c|c|c|c} a,b,c & i & \sigma_i & \phi_i\\ \hline 3 & 0 & (1\ 2\ 3) & 2x^3 (9x^2 - 5) \\ 6 & 1 & (1\ 2\ 5)(3\ 4) & (x - 1)^3 (3x + 2)^2\\ 6 & \infty & (1\ 4\ 3)(2\ 5) & (x + 1)^3 (3x - 2)^2 \end{array} $$ \end{landscape} Note in the example where $(a,b,c) = (4,3,6)$ that the size of the refined passport is 2 and correspondingly the field of definition of $\phi$ is the quadratic extension $\mathbb Q(\sqrt{6})$. \end{exm} \begin{exm} \label{exm:genus0nonewt} We now consider a more complicated example. Consider the triple \begin{equation}\label{eqn:dessin-12,2,5} \sigma_0=(1\ 2\ 3\ 4)(5\ 6\ 7), \quad \sigma_1=(2\ 3)(4\ 5)(6\ 7), \quad \sigma_\infty=(1\ 5\ 6\ 4\ 2) \end{equation} which generates $S_7$. The size of the refined passport for this triple is $2$; the other triple (up to uniform conjugation) is \[ \sigma_0=(1\ 2\ 3\ 4)(5\ 6\ 7), \quad \sigma_1=(1\ 2)(4\ 5)(6\ 7), \quad \sigma_\infty=(1\ 5\ 6\ 4\ 3). \] We find $g(X(\Gamma))=0$ and $\dim_\mathbb C S_4(\Gamma) =2$ so we take forms in weight $k=4$. We work in precision $\varepsilon=10^{-100}$. The federalist approach has an advantage here: expanding around the vertices $\alpha_i\cdot 0$ for $i=1,5$, we can take the radius of expansion to be $\rho=0.821032\ldots$ and degree $N=1168$; if we did not, we would need to take radius $0.966411$ and $N=6740$. Our implementation uses 71 Arnoldi iterations, each requiring about $2$ seconds, to find a basis element, and the total calculation time was about 6 minutes; computing a full numerical kernel instead took over 15 minutes. We find with \begin{equation} \Theta = 1.3259366\ldots + 0.63528305\ldots \sqrt{-1} = \sqrt[12]{\frac{1632000-961536\sqrt{-5}}{4117715}}\left(\frac{1}{\kappa}\right) \end{equation} that \[ x = (\Theta w)^3 + \frac{-537128+59872\sqrt{-5}}{194145\cdot 9!}(\Theta w)^9 + \frac{37661184-1520640\sqrt{-5}}{3895843\cdot 12!}(\Theta w)^{12} + O(w^{15}) \] and \[ \phi(x) = \frac{p(x)}{q(x)} = 1+\frac{r(x)}{q(x)} \] where \begin{align*} p(x) &= (379-2102\sqrt{-5})x^4\left((4-\sqrt{-5})x+(-5+\sqrt{-5})\right)^3 \\ q(x) &= 3\left((-2-\sqrt{-5})x+(2+2\sqrt{-5})\right)^5\left((-57+20\sqrt{-5})x^2+(25-19\sqrt{-5})x+60\right) \\ r(x) &= -2(3x-(1+\sqrt{-5}))\left((259+5\sqrt{-5})x^3 + (-356+43\sqrt{-5})x^2\right. \\ &\qquad\qquad\qquad\qquad\qquad\qquad \left. - (144+144\sqrt{-5})x + (240+96\sqrt{-5})\right)^2. \end{align*} See below (Example \ref{exm:genus0newt}) for the use of Newton's method on this example. Finally, we repeat this calculation with the other triple (\ref{eqn:othertriple}) and we find that all coefficients are (to numerical precision) the complex conjugates of the first triples, thus confirming that the Galois action is as expected on this refined passport. \end{exm} \begin{figure}[h] \includegraphics[scale=1]{belyi-triangle-pics/belyi-triangle-pics-22-1.pdf} \begin{comment} \begin{equation} \label{fig:dessin-12,2,5} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.71780,0.00000) \psarc[linewidth=.1pt] (1.0555,0.00000) {0.33768} {180.00} {219.00} \psline[linewidth=.1pt] (0.79305,-0.21250) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.71780,0.00000) \psarcn[linewidth=.1pt] (1.0555,0.00000) {0.33768} {180.00} {141.00} \psline[linewidth=.1pt] (0.79305,0.21250) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.62164,0.35890) \psarc[linewidth=.1pt] (0.91406,0.52776) {0.33768} {210.01} {249.00} \psline[linewidth=.1pt] (0.79305,0.21250) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.62164,0.35890) \psarcn[linewidth=.1pt] (0.91407,0.52773) {0.33767} {210.00} {171.00} \psline[linewidth=.1pt] (0.58056,0.58056) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.35890,0.62164) \psarc[linewidth=.1pt] (0.52774,0.91409) {0.33769} {240.00} {279.00} \psline[linewidth=.1pt] (0.58056,0.58056) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.35890,0.62164) \psarcn[linewidth=.1pt] (0.52777,0.91405) {0.33768} {239.99} {201.00} \psline[linewidth=.1pt] (0.21250,0.79305) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.62164,-0.35890) \psarc[linewidth=.1pt] (0.91407,-0.52773) {0.33767} {150.00} {189.00} \psline[linewidth=.1pt] (0.58056,-0.58056) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.00000,0.00000) (0.62164,-0.35890) \psarcn[linewidth=.1pt] (0.91406,-0.52776) {0.33768} {149.99} {111.00} \psline[linewidth=.1pt] (0.79305,-0.21250) (0.00000,0.00000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.82051,-0.47372) (0.62164,-0.35890) \psarcn[linewidth=.1pt] (0.91406,-0.52776) {0.33768} {149.99} {111.00} \psarc[linewidth=.1pt] (0.96806,-0.32618) {0.20869} {146.99} {225.00} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.82051,-0.47372) (0.62164,-0.35890) \psarc[linewidth=.1pt] (0.91407,-0.52773) {0.33767} {150.00} {189.00} \psarcn[linewidth=.1pt] (0.76651,-0.67528) {0.20868} {153.01} {75.001} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.82057,-0.58166) {0.10793} {90.032} {169.25} \psarc[linewidth=.1pt] (0.67520,-0.76660) {0.20881} {79.179} {116.96} \psarcn[linewidth=.1pt] (0.76651,-0.67528) {0.20868} {153.01} {75.001} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.82057,-0.58166) {0.10793} {90.032} {169.25} \psarcn[linewidth=.1pt] (0.67555,-0.76614) {0.20829} {79.250} {62.448} \psarcn[linewidth=.1pt] (0.84028,-0.54750) {0.076384} {206.44} {105.00} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.91404,-0.41977) {0.10797} {209.98} {130.77} \psarc[linewidth=.1pt] (1.0014,-0.20172) {0.20857} {220.78} {237.56} \psarc[linewidth=.1pt] (0.89429,-0.45398) {0.076407} {93.561} {194.99} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.91404,-0.41977) {0.10797} {209.98} {130.77} \psarcn[linewidth=.1pt] (1.0015,-0.20157) {0.20870} {220.80} {183.00} \psarc[linewidth=.1pt] (0.96806,-0.32618) {0.20869} {146.99} {225.00} \closepath} \rput(0.29028,-0.29028) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(0.10625,0.39653) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.73717,-0.11273) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(0.73717,0.11271) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.69476,0.27096) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(0.27099,0.69476) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(0.58206,0.46619) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.46622,0.58205) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(0.77065,-0.51609) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}}} \rput(0.83231,-0.40942) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}}} \rput(0.64590,-0.55991) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{6}$}}} \rput(0.80783,-0.27942) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{6}$}}} \rput(0.74388,-0.56938) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{7}$}}} \rput(0.86507,-0.35953) {\scalebox{0.36517} {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{7}$}}} \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.71780,0.00000) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.71780,0.00000) \psdots*[dotstyle=x,dotsize=7.0000pt](0.79305,-0.21250) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.62164,0.35890) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.62164,0.35890) \psdots*[dotstyle=x,dotsize=7.0000pt](0.79305,0.21250) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.35890,0.62164) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.35890,0.62164) \psdots*[dotstyle=x,dotsize=7.0000pt](0.58056,0.58056) \psline[linewidth=1.5000pt] (0.00000,0.00000) (0.62164,-0.35890) \psdots*[dotstyle=o,dotsize=7.0000pt](0.00000,0.00000) \psdots*[dotstyle=*,dotsize=7.0000pt](0.62164,-0.35890) \psdots*[dotstyle=x,dotsize=7.0000pt](0.58056,-0.58056) \psarc[linewidth=0.54774pt] (0.82057,-0.58166) {0.10793} {90.032} {169.25} \psdots*[dotstyle=o,dotsize=2.5562pt](0.82051,-0.47372) \psdots*[dotstyle=*,dotsize=2.5562pt](0.71440,-0.56151) \psdots*[dotstyle=x,dotsize=2.5562pt](0.58056,-0.58056) \psarcn[linewidth=0.54774pt] (0.91404,-0.41977) {0.10797} {209.98} {130.77} \psdots*[dotstyle=o,dotsize=2.5562pt](0.82051,-0.47372) \psdots*[dotstyle=*,dotsize=2.5562pt](0.84348,-0.33794) \psdots*[dotstyle=x,dotsize=2.5562pt](0.88954,-0.37772) \psline[linewidth=1.5000pt] (0.82051,-0.47372) (0.62164,-0.35890) \psdots*[dotstyle=o,dotsize=7.0000pt](0.82051,-0.47372) \psdots*[dotstyle=*,dotsize=7.0000pt](0.62164,-0.35890) \psdots*[dotstyle=x,dotsize=7.0000pt](0.79305,-0.21250) \rput(0.35890,0.00000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.31082,0.17945) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(0.17945,0.31082) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(0.31082,-0.17945) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(0.72107,-0.41631) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \rput(0.75171,-0.49855) {\scalebox{0.36517} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}}} \rput(0.80759,-0.40172) {\scalebox{0.36517} {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$7$}}} \end{pspicture} \end{center} \end{comment} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_a^{}$ \\ $3$ & $\delta_a^{2}$ \\ $4$ & $\delta_a^{-1}$ \\ $5$ & $\delta_a^{-1}\delta_b^{}$ \\ $6$ & $\delta_a^{-1}\delta_b^{}\delta_a^{}$ \\ $7$ & $\delta_a^{-1}\delta_b^{}\delta_a^{-1}$ \\ \bottomrule \end{tabular} \hfill \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_a^{-4}$ \\ $s_{2}$ & $\delta_b^{}$ \\ $s_{3}$ & $\delta_a^{}\delta_b^{}\delta_a^{-2}$ \\ $s_{4}$ & $\delta_a^{}\delta_b^{-1}\delta_a^{-2}$ \\ $s_{5}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{3}\delta_b^{-1}\delta_a^{}$ \\ $s_{6}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{}\delta_b^{}\delta_a^{}\delta_b^{-1}\delta_a^{}$ \\ $s_{7}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{}\delta_b^{-1}\delta_a^{}\delta_b^{-1}\delta_a^{}$ \\ \bottomrule \end{tabular} \end{center} \bigskip \caption{A fundamental domain and conformally drawn dessin for $\Gamma$ corresponding to the permutation triple \ref{eqn:dessin-12,2,5} using the methods in section \ref{sec:cosets}.} \label{fig:dessin-12,2,5} \end{figure} \FloatBarrier \subsection*{Genus one} Now suppose that $X(\Gamma)$ has genus $g=1$. Then the space $S_2(\Gamma)$ of modular forms of weight $2$ has dimension $\dim S_2(\Gamma)=1$ and so is spanned by a form \[ f(z)=(1-w)^2\sum_{n=0}^{\infty} b_n w^n = (1-w)^2 \sum_{n=0}^{\infty} \frac{c_n}{n!} (\Theta w)^n \] where as usual $w=w(z)$ and $\Theta$ is chosen as in (\ref{eqn:Theta0}) to (experimentally) obtain coefficients $c_n$ that are algebraic integers in a number field. Then $f(z)\,dz$ is the unique (nonzero) holomorphic differential $1$-form on the Riemann surface $X(\Gamma)$ up to scaling by $\mathbb C^\times$. To compute an equation for $X(\Gamma)$, we compute the Abel-Jacobi map, computing its lattice of periods, and then use the Weierstrass uniformization to obtain a model, as follows. (For a reference, see Silverman \cite[Chapter VI]{Silverman} or Cremona \cite[Chapters II--III]{Cremona}.) Let $\{\gamma_i\}_i$ be the side pairing elements arising from the coset graph and fundamental domain $D_\Gamma$ for $\Gamma$ as computed in Algorithm \ref{alg:cosetalg}. Suppose that $\gamma_i$ pairs the vertex $v_i$ with $v_i'=\gamma_i v_i$ in $D_\Gamma$. In particular, these vertices are within the domain of convergence of $D_\Gamma$. Then we compute the set of normalized periods by \begin{equation} \label{eqn:varpi} \varpi_i = \int_{z(v_i)}^{z(v_i')} \Theta f(z)\,dz = \int_{v_i}^{v_i'} f(w) \frac{d(\Theta w)}{(1-w)^2} \approx \sum_{n=0}^N \frac{c_n}{(n+1)!} (\Theta w)^{n+1} \bigg|_{v_i}^{v_i'}. \end{equation} The span \[ \Lambda = \sum_i \mathbb Z \varpi_i = \mathbb Z \omega_1 + \mathbb Z\omega_2 \subset \mathbb C \] is a (full) lattice in $\mathbb C$, and the Abel-Jacobi map \begin{equation} \label{eqn:AJ} \begin{aligned} X(\Gamma)=\Gamma \backslash \mathcal{H} &\xrightarrow{\sim} \mathbb C/\Lambda \\ z &\mapsto u(z) = \int_{0}^{w(z)} f(w) \frac{d(\Theta w)}{(1-w)^2} \end{aligned} \end{equation} is an isomorphism of Riemann surfaces. We may assume without loss of generality, interchanging $\omega_1,\omega_2$ if necessary, that $\tau=\omega_1/\omega_2 \in \mathcal{H}$. (Further, by application of the Euclidean algorithm, i.e., fundamental domain reduction for $\SL_2(\mathbb Z)$, we may assume $|\repart \tau\,| \leq 1/2$ and $|\tau| \geq 1$, so that $\impart \tau \geq \sqrt{3}/2$.) Let $q=\exp(2\pi i\tau)$. Then $X(\Gamma)$ has $j$-invariant $j(\Lambda)=j(\tau)$ \[ j(q) = \frac{1}{q} + 744 + 196884q + 21493760q^2 + \dots, \] where $j(q)$ is the modular elliptic $j$-invariant (not to be confused with the automorphy factor by the same name!), and we have a uniformization \begin{equation} \label{eqn:AJtoE} \begin{aligned} \mathbb C/\Lambda &\xrightarrow{\sim} E(\Gamma) : y^2 = x^3 - 27c_4(\Lambda)x - 54c_6(\Lambda) \\ u &\mapsto \left(\wp(\Lambda;u),\frac{\wp'(\Lambda;u)}{2}\right) \end{aligned} \end{equation} where \[ \wp(\Lambda;u) = \frac{1}{u^2} + \sum_{k=2}^{\infty} (2k-1)\frac{2\zeta(2k)}{\omega_2^{2k}}E_{2k}(\tau) u^{2k-2} \] is the Weierstrass $\wp$-function, \[ E_{2k}(\tau) = 1+(-1)^k\frac{4k}{B_{2k}}\sum_{n=0}^{\infty} \sigma_{2k-1}(n) q^n \] is the normalized Eisenstein series with $B_{2k}$ the Bernoulli numbers, so that \begin{equation} \label{eqn:c4c6} \begin{aligned} E_4(\tau) &= 1 + 240\sum_{n=1}^{\infty} \sigma_3(n)q^n= 1 + 240q + 2160q^2 + 6720q^3 + \dots \\ E_6(\tau) &= 1 - 504\sum_{n=1}^{\infty} \sigma_5(n)q^n = 1 - 504q - 16632q^2 - 122976q^3 + \dots, \end{aligned} \end{equation} and finally \[ c_{2k}(\Lambda) = \left(\frac{2\pi}{6\omega_2}\right)^{2k} E_{2k}(\tau). \] Implicitly, we are taking the point $w=0$ to be the origin of the group law, so we always get an elliptic curve (instead of a genus one curve). The Bely\u{\i}\ map is computed from the explicit power series $\phi(w) \in \mathbb Q[[(w/\kappa)^a]]$ as in (\ref{eqn:expandpuiseux}). A general function on $E(\Gamma)$ is of the form \[ \phi(x,y) = \frac{p(x) + q(x) y}{r(x)} \] Since $x(w)=\wp(\Lambda;u(w))$ and $y(w)=\wp'(\Lambda;u(w))/2$ are also computed power series in $w$, we can solve for the coefficients of $p(x),q(x),r(x)$ using linear algebra. \begin{exm} The smallest degree $d$ for which there exists a hyperbolic refined passport of genus one is $d=4$, and there is a unique such refined passport with $(a,b,c)=(4,3,4)$ and representative triple \[ \sigma_0 = (1\ 2\ 3\ 4), \quad \sigma_1=(1\ 2\ 3), \quad \sigma_\infty = (1\ 4\ 2\ 3). \] \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-22.pdf} \begin{comment} \begin{equation} \label{fig:dessin-4,3,4} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psarc[linewidth=.1pt] (1.22474487139,-0.408248290461) {0.816496580928} {150.000000000} {180.000000000} \psline[linewidth=.1pt] (0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psarcn[linewidth=.1pt] (1.22474487139,0.408248290461) {0.816496580928} {210.000000000} {180.000000000} \psline[linewidth=.1pt] (0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psarc[linewidth=.1pt] (0.408248290461,1.22474487139) {0.816496580928} {240.000000000} {270.000000000} \psline[linewidth=.1pt] (0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psarcn[linewidth=.1pt] (-0.408248290461,1.22474487139) {0.816496580928} {300.000000000} {270.000000000} \psline[linewidth=.1pt] (-0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psarc[linewidth=.1pt] (-1.22474487139,0.408248290461) {0.816496580928} {330.000000000} {1.91465129622E-10} \psline[linewidth=.1pt] (-0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psarcn[linewidth=.1pt] (-1.22474487139,-0.408248290461) {0.816496580928} {29.9999999998} {360.000000000} \psline[linewidth=.1pt] (-0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psarc[linewidth=.1pt] (-0.408248290461,-1.22474487139) {0.816496580928} {60.0000000003} {90.0000000002} \psline[linewidth=.1pt] (-0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psarcn[linewidth=.1pt] (0.408248290461,-1.22474487139) {0.816496580928} {120.000000000} {89.9999999999} \psline[linewidth=.1pt] (0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \rput(0.436069736795,-0.196923425061) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(-0.196923425054,0.436069736798) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.436069736797,0.196923425061) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(-0.436069736796,0.196923425056) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.196923425053,0.436069736797) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(-0.436069736798,-0.196923425057) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(-0.196923425060,-0.436069736795) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.196923425059,-0.436069736797) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.517638090205,0.000000000000) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.408248290464,-0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.000000000000,0.517638090205) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.408248290464,0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.517638090205,0.000000000000) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.408248290464,0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.000000000000,-0.517638090205) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.408248290464,-0.408248290464) \rput(0.258819045102,0.000000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.000000000000,0.258819045102) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.258819045102,0.000000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(0.000000000000,-0.258819045102) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \end{pspicture} \end{equation} \end{comment} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_a^{}$ \\ $3$ & $\delta_a^{2}$ \\ $4$ & $\delta_a^{-1}$ \\ \bottomrule \end{tabular} \;\;\;\;\;\;\;\;\;\; \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_b^{}\delta_a^{-1}$ \\ $s_{2}$ & $\delta_b^{-1}\delta_a^{-2}$ \\ $s_{3}$ & $\delta_a^{}\delta_b^{}\delta_a^{-2}$ \\ $s_{4}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{}$ \\ \bottomrule \end{tabular} \end{center} \bigskip \caption{A fundamental domain and conformally drawn dessin for $\Gamma$ corresponding to the permutation triple $$ \sigma_0 = (1\;2\;3\;4),\quad\sigma_1=(1\;2\;3),\quad\sigma_\infty = (1\;4\;2\;3) $$ using the methods in \ref{sec:cosets}.} \label{fig:dessin-4,3,4} \end{figure} In under a second, we compute in precision $\varepsilon=10^{-30}$ the normalized differential \[ f(z) = 1 + (\Theta w) - \frac{1}{2!}(\Theta w)^2 + \frac{6}{4!}(\Theta w)^4 - \frac{6}{5!}(\Theta w)^5 + \frac{126}{6!}(\Theta w)^6 + O(w^8) \] with \[ \Theta = -1.786853\ldots i = \sqrt[4]{\frac{27}{4}} i \left(\frac{1}{\kappa}\right). \] We find two classes which span the homology of $X(\Gamma)$, as in Figure \ref{fig:homology}. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-23.pdf} \begin{comment} \begin{equation} \label{fig:homology} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psarc[linewidth=.1pt] (1.22474487139,-0.408248290461) {0.816496580928} {150.000000000} {180.000000000} \psline[linewidth=.1pt] (0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psarcn[linewidth=.1pt] (1.22474487139,0.408248290461) {0.816496580928} {210.000000000} {180.000000000} \psline[linewidth=.1pt] (0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psarc[linewidth=.1pt] (0.408248290461,1.22474487139) {0.816496580928} {240.000000000} {270.000000000} \psline[linewidth=.1pt] (0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psarcn[linewidth=.1pt] (-0.408248290461,1.22474487139) {0.816496580928} {300.000000000} {270.000000000} \psline[linewidth=.1pt] (-0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psarc[linewidth=.1pt] (-1.22474487139,0.408248290461) {0.816496580928} {330.000000000} {1.91465129622E-10} \psline[linewidth=.1pt] (-0.408248290464,0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psarcn[linewidth=.1pt] (-1.22474487139,-0.408248290461) {0.816496580928} {29.9999999998} {360.000000000} \psline[linewidth=.1pt] (-0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psarc[linewidth=.1pt] (-0.408248290461,-1.22474487139) {0.816496580928} {60.0000000003} {90.0000000002} \psline[linewidth=.1pt] (-0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psarcn[linewidth=.1pt] (0.408248290461,-1.22474487139) {0.816496580928} {120.000000000} {89.9999999999} \psline[linewidth=.1pt] (0.408248290464,-0.408248290464) (0.000000000000,0.000000000000) \closepath} \rput(0.436069736795,-0.196923425061) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(-0.196923425054,0.436069736798) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.436069736797,0.196923425061) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(-0.436069736796,0.196923425056) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.196923425053,0.436069736797) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(-0.436069736798,-0.196923425057) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(-0.196923425060,-0.436069736795) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(0.196923425059,-0.436069736797) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.517638090205,0.000000000000) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.517638090205,0.000000000000) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.408248290464,-0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.000000000000,0.517638090205) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.000000000000,0.517638090205) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.408248290464,0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.517638090205,0.000000000000) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.517638090205,0.000000000000) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.408248290464,0.408248290464) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.000000000000,-0.517638090205) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.000000000000,-0.517638090205) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.408248290464,-0.408248290464) \pcarc[linewidth = 1.5pt,arrowsize=6pt,ArrowInside=->,ArrowInsidePos=0.65](0.517638090205,0.000000000000)(0.000000000000,0.517638090205) \pcarc[linewidth = 1.5pt,arrowsize=6pt,ArrowInside=->,ArrowInsidePos=0.65](0.000000000000,0.517638090205)(-0.517638090205,0.000000000000) \uput{10pt}[0]{0}(0.517638090205,0.000000000000){$\gamma_1$ (via $s_1$)} \uput{10pt}[90]{0}(0.000000000000,0.517638090205){$\gamma_2$ (via $s_3$)} \rput(0.258819045102,0.000000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.000000000000,0.258819045102) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.258819045102,0.000000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(0.000000000000,-0.258819045102) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \end{pspicture} \end{equation} \end{comment} \caption{Representatives $\gamma_1, \gamma_2$ of the homology classes of $X(\Gamma)$.} \label{fig:homology} \end{figure} We integrate as in (\ref{eqn:varpi}): we find the two periods \[ \omega_1,\omega_2=\pm 1.68575035481259604287120365776 - 1.07825782374982161771933749941 i \] and numerically we verify that all other periods lie in the lattice spanned by these two periods. We find \[ \tau=\omega_1/\omega_2 = -0.4193\ldots + 0.9078\ldots\sqrt{-1} \] and $\tau$ belongs to the fundamental domain for $\SL_2(\mathbb Z)$. We then compute \[ j(q)=31.648529\ldots=\frac{207646}{6561} \] and from (\ref{eqn:c4c6}) that \[ c_4=-0.03626\ldots = -\frac{47}{1296}, \qquad c_6=-0.05056\ldots = -\frac{2359}{46656} \] so $X(\Gamma)$ has equation \begin{equation} \label{eqn:y2x3} y^2 = x^3 + \frac{47}{48} x + \frac{2359}{864} \end{equation} with minimal model \begin{equation} \label{genus1X} y^2 = x^3 + x^2 + 16x + 180, \end{equation} (via the map $(x,y) \mapsto (4x-1/3, 8y)$), the elliptic curve with conductor $48$ with Cremona label \textsf{48a6}. We compute the Bely\u{\i}\ map by consideration of power series. From (\ref{eqn:AJ}) we have \[ u(w) = (\Theta w) + \frac{1}{2}(\Theta w)^2 - \frac{1}{6}(\Theta w)^3 + O(w^5) \] and from (\ref{eqn:AJtoE}), with \[ \wp(\Lambda;u) = \frac{1}{u^2} - \frac{47}{240}u^2 - \frac{337}{864} u^4 + \frac{2209}{172800} u^6 + O(u^8) \] we compute that \begin{align*} x(w) &= \frac{1}{(\Theta w)^2} - \frac{1}{\Theta w} + \frac{13}{12} - (\Theta w) + \frac{3}{5} (\Theta w)^2 + O(w^3) \\ y(w) &= -\frac{1}{(\Theta w)^3} + \frac{3}{2} \frac{1}{(\Theta w)^2} - 2\frac{1}{(\Theta w)} + \frac{9}{4} - \frac{12}{5}(\Theta w) + O(w^2) \end{align*} and confirm again that $x(w),y(w)$ satisfy \eqref{eqn:y2x3} to the precision computed. Then from (\ref{eqn:expandpuiseux}) we find \[ \phi(w) = \frac{27}{4}(\Theta w)^4 - \frac{108}{5}(\Theta w)^8 + \frac{8181}{175}(\Theta w)^{12} + O(w^{16}). \] Then we compute a linear relation among the functions \[ \phi,\phi x, \phi y, \phi x^2, 1\] and we find the Bely\u{\i}\ map \[ \phi(x,y) = \frac{139968x^2 - 23328x + 490860 + 279936y}{(12x-13)^4} \] of degree $4$ with divisor $4(13/12,9/4)-4\infty$ and $\phi-1$ with divisor \[ \ddiv(\phi-1) = 3(-5/12,-3/2)+(67/12,27/2)-4\infty \] as desired. \end{exm} \begin{comment} > phiE; 279936/(20736*xE^4 - 89856*xE^3 + 146016*xE^2 - 105456*xE + 28561)*yE + (139968*xE^2 - 23328*xE + 490860)/(20736*xE^4 - 89856*xE^3 + 146016*xE^2 - 105456*xE + 28561) > Support(Divisor(phiE)); [ Place at (0 : 1 : 0), Place at (13/12 : 9/4 : 1) ] [ 4, -4 ] > Support(Divisor(phiE-1)); [ Place at (67/12 : 27/2 : 1), Place at (-5/12 : -3/2 : 1), Place at (13/12 : 9/4 : 1) ] [ 1, 3, -4 ] \end{comment} \subsection*{Hyperelliptic} Now suppose that $X(\Gamma)$ has genus $g=g(X(\Gamma)) \geq 2$. Then we define the \emph{canonical map} \begin{equation} \label{eqn:canhyp} \begin{aligned} X(\Gamma) &\xrightarrow{\sim} \mathbb P^{g-1} \\ z &\mapsto (f_1(w) : \dots : f_g(w)) \end{aligned} \end{equation} where $f_1,\dots,f_g$ is a basis for $S_2(\Gamma)$. We have two possibilities. We suppose first, in this subsection, that $X(\Gamma)$ is \defi{hyperelliptic}, meaning that there exists a (nonconstant) degree $2$ map $X(\Gamma) \to \mathbb P^1$. We test if $X$ is hyperelliptic as follows: we compute the function $x(w)=g(w)/h(w)$ where $g,h \in S_2(\Gamma)$ are as in (\ref{eqn:gandh}), compute $y=(dx/dw)/h(w)$, and test if there is an equation of the form \[ y^2 + u(x)y = v(x) \] with $\deg u(x) \leq g+1$ and $\deg v(x) \leq 2g+2$; if so, we have found our embedding as a hyperelliptic curve. One can use Riemann--Roch to verify that this test is correct (see e.g.\ \cite[\S 4]{Galbraith}). Analogous to the case in genus $g=1$, a general function on $X$ is of the form \[ \phi(x,y) = \frac{p(x)+q(x) y}{r(x)} \] and linear algebra yields an expression for $\phi$. \begin{exm} Consider the triple \[ \sigma_0=\sigma_1=(1\ 2\ 3 \ 4\ 5), \quad \sigma_{\infty}=(1\ 4\ 2\ 5\ 3) \] generating $\mathbb Z/5\mathbb Z$. Then $g=2$ so $X$ is hyperelliptic. We compute a basis for $S_2(\Gamma)$ to precision $10^{-30}$ in $15$ seconds. With \[ \Theta = \sqrt[5]{\frac{1}{24}} \left(\frac{1}{\kappa}\right) \] we find the following forms in $S_2(\Gamma)$: \[ \begin{aligned} \frac{f(w)}{(1-w)^2} &= 1 + \frac{120}{5!}(\Theta w)^5 + \frac{21}{10!}(\Theta w)^{10} - \frac{45864}{15!}(\Theta w)^{15} + \frac{237414996}{20!}(\Theta w)^{20} + O(w^{25}) \\ \frac{g(w)}{(1-w)^2} &= (\Theta w) - \frac{3}{6!}(\Theta w)^6 - \frac{126}{11!}(\Theta w)^{11} - \frac{366912}{16!}(\Theta w)^{16} +O(w^{21}) \end{aligned} \] We then compute \[ x(w) = \frac{f(w)}{g(w)} = (\Theta w)^{-1} + \frac{3}{10\cdot 4!}(\Theta w)^4 + \frac{1218}{55 \cdot 9!}(\Theta w)^9 + O(w^{14}) \] and $y(w)=x'(w)/g(w)$ where $x'$ denotes the derivative with respect to $\Theta w$. Then $x$ has a pole of order $1$ at $w=0$ and $y$ has a pole of order $3$, therefore we obtain the equation \[ y^2 = x^6 - \frac{1}{6} x. \] We then compute using (\ref{eqn:expandpuiseux}) the expansion \[ \phi(w) = \frac{1}{24}(\Theta w)^5 - \frac{1}{1152}(\Theta w)^{10} + \frac{23}{1824768}(\Theta w)^{15} + O(w^{20}) \] and hence \[ \phi(x,y) = \frac{y+1}{2x^3}. \] This dessin corresponds to a Galois Bely\u{\i}\ map and so it can be constructed using other means; however, it makes for a nice example and our method did not use this property. \end{exm} \subsection*{General case} Now suppose still that $g \geq 2$ but now that $X(\Gamma)$ is not hyperelliptic; then the canonical map (\ref{eqn:canhyp}) is a closed embedding. If $g=3$ then $f_1,f_2,f_3$ satisfy an equation of degree $4$; otherwise, the ideal of equations vanishing on the image is generated in degree $2$ and $3$: in fact, in degree $2$ as long as $X$ is neither trigonal nor isomorphic to a plane quintic. So again using linear algebra and the power series expansions for $f_1,\dots,f_g \in S_2(\Gamma)$, we can compute a generating set for the ideal of quadrics and cubics vanishing on the image to obtain equations for $X(\Gamma)$. Once equations for $X(\Gamma)$ are obtained, we again write down using Riemann-Roch and the explicit representation of the function field of $X(\Gamma)$ a general function of degree $d$ and use linear algebra using explicit power series to compute an equation for the Bely\u{\i}\ map $\phi$; alternatively, we compute the ramification values and use Riemann-Roch to find a function with specified divisor. \begin{exm} Consider the rigid permutation triple \[ \sigma_0 = (1\ 2\ 3\ 4\ 5\ 6\ 7), \quad \sigma_1=(1\ 6\ 2\ 5\ 7\ 3\ 4), \quad \sigma_{\infty}=(1\ 5\ 3\ 6\ 2\ 4\ 7), \] a $(7,7,7)$-triple generating $\PSL_2(\mathbb F_7)\cong \GL_3(\mathbb F_2)$ and with refined passport of genus $g=3$. With \[ \Theta = 0.9280540\ldots + 0.4469272\ldots\sqrt{-1} = \sqrt[7]{-16}\left(\frac{1}{\kappa}\right) \] we find a basis of $3$ forms $f,g,h \in S_2(\Gamma)$ with coefficients in the field $\mathbb Q(\nu)$ where $\nu=(-1+\sqrt{-7})/2$. The canonical embedding of a genus three curve is a quartic in $\mathbb P^2$, which we compute in the echelonized basis to be \begin{align*} &11f^3h - 11f^2g^2 + (5\nu + 2)f^2h^2 + (-10\nu + 29)fg^2h + (-25\nu - 10)fh^3 \\ &\qquad + (5\nu + 2)g^4 + (-8\nu - 12)g^2h^2 + (-21\nu + 18)h^4 = 0 \end{align*} valid to the precision computed. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-24.pdf} \begin{comment} \begin{equation} \label{fig:dessin-7,7,7} \notag \psset{unit=6cm} \begin{pspicture}(-1,-1)(1,1) \pscircle[linewidth=.001pt](0,0){1} \newgray{lightestgray}{.93} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.895509763099,0.000000000000) \psarc[linewidth=.1pt] (1.00609608630,-0.229634866309) {0.254875472937} {115.714285714} {218.571428572} \psline[linewidth=.1pt] (0.806826417454,-0.388547124429) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.895509763099,0.000000000000) \psarcn[linewidth=.1pt] (1.00609608630,0.229634866309) {0.254875472937} {244.285714286} {141.428571428} \psline[linewidth=.1pt] (0.806826417454,0.388547124429) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.558341204757,0.700137725648) \psarc[linewidth=.1pt] (0.806826417455,0.643422597365) {0.254875472939} {167.142857142} {270.000000000} \psline[linewidth=.1pt] (0.806826417454,0.388547124429) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.558341204757,0.700137725648) \psarcn[linewidth=.1pt] (0.447754881546,0.929772591958) {0.254875472940} {295.714285715} {192.857142857} \psline[linewidth=.1pt] (0.199269668852,0.873057463677) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.199269668852,0.873057463677) \psarc[linewidth=.1pt] (0.000000000000,1.03196972180) {0.254875472939} {218.571428572} {321.428571428} \psline[linewidth=.1pt] (0.199269668852,0.873057463677) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.199269668852,0.873057463677) \psarcn[linewidth=.1pt] (-0.447754881547,0.929772591958) {0.254875472935} {347.142857143} {244.285714285} \psline[linewidth=.1pt] (-0.558341204757,0.700137725648) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.806826417454,0.388547124429) \psarc[linewidth=.1pt] (-0.806826417455,0.643422597366) {0.254875472937} {270.000000000} {12.8571428574} \psline[linewidth=.1pt] (-0.558341204757,0.700137725648) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.806826417454,0.388547124429) \psarcn[linewidth=.1pt] (-1.00609608630,0.229634866309) {0.254875472936} {38.5714285716} {295.714285714} \psline[linewidth=.1pt] (-0.895509763099,0.000000000000) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.806826417454,-0.388547124429) \psarc[linewidth=.1pt] (-1.00609608630,-0.229634866309) {0.254875472936} {321.428571428} {64.2857142861} \psline[linewidth=.1pt] (-0.895509763099,0.000000000000) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.806826417454,-0.388547124429) \psarcn[linewidth=.1pt] (-0.806826417455,-0.643422597366) {0.254875472937} {89.9999999999} {347.142857143} \psline[linewidth=.1pt] (-0.558341204757,-0.700137725648) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.199269668852,-0.873057463677) \psarc[linewidth=.1pt] (-0.447754881547,-0.929772591958) {0.254875472935} {12.8571428574} {115.714285715} \psline[linewidth=.1pt] (-0.558341204757,-0.700137725648) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (-0.199269668852,-0.873057463677) \psarcn[linewidth=.1pt] (0.000000000000,-1.03196972180) {0.254875472939} {141.428571428} {38.5714285717} \psline[linewidth=.1pt] (0.199269668852,-0.873057463677) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.558341204757,-0.700137725648) \psarc[linewidth=.1pt] (0.447754881546,-0.929772591958) {0.254875472940} {64.2857142851} {167.142857143} \psline[linewidth=.1pt] (0.199269668852,-0.873057463677) (0.000000000000,0.000000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.000000000000,0.000000000000) (0.558341204757,-0.700137725648) \psarcn[linewidth=.1pt] (0.806826417455,-0.643422597365) {0.254875472939} {192.857142858} {90.0000000002} \psline[linewidth=.1pt] (0.806826417454,-0.388547124429) (0.000000000000,0.000000000000) \closepath} \rput(0.757610873608,-0.172919738028) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{1}$}} \rput(0,-0.777094248860) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{1}$}} \rput(0.757610873608,0.172919738028) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{2}$}} \rput(-0.607556748602,0.484510339245) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{2}$}} \rput(0.607556748604,0.484510339246) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{3}$}} \rput(-0.607556748604,-0.484510339247) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{3}$}} \rput(0.337168558339,0.700137725647) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{4}$}} \rput(-0.337168558340,-0.700137725647) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{4}$}} \rput(0.337168558342,-0.700137725651) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{5}$}} \rput(-0.337168558343,0.700137725650) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{5}$}} \rput(0.607556748602,-0.484510339245) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{6}$}} \rput(-0.757610873608,-0.172919738029) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{6}$}} \rput(0,0.777094248860) {\psframebox*[linewidth=.0001,fillcolor=blue!10] {$s_{7}$}} \rput(-0.757610873608,0.172919738028) {\psframebox*[linewidth=.0001,fillcolor=red!80] {$s_{7}$}} \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.895509763099,0.000000000000) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.895509763099,0.000000000000) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.806826417454,-0.388547124429) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.558341204757,0.700137725648) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.558341204757,0.700137725648) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.806826417454,0.388547124429) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.199269668852,0.873057463677) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.199269668852,0.873057463677) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.199269668852,0.873057463677) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.806826417454,0.388547124429) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.806826417454,0.388547124429) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.558341204757,0.700137725648) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.806826417454,-0.388547124429) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.806826417454,-0.388547124429) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.895509763099,0.000000000000) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (-0.199269668852,-0.873057463677) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](-0.199269668852,-0.873057463677) \psdots*[dotstyle=x,dotsize=7.00000000000pt](-0.558341204757,-0.700137725648) \psline[linewidth=1.50000000000pt] (0.000000000000,0.000000000000) (0.558341204757,-0.700137725648) \psdots*[dotstyle=o,dotsize=7.00000000000pt](0.000000000000,0.000000000000) \psdots*[dotstyle=*,dotsize=7.00000000000pt](0.558341204757,-0.700137725648) \psdots*[dotstyle=x,dotsize=7.00000000000pt](0.199269668852,-0.873057463677) \rput(0.447754881549,0.000000000000) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$1$}} \rput(0.279170602379,0.350068862824) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$2$}} \rput(-0.0996348344258,0.436528731838) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$3$}} \rput(-0.403413208727,0.194273562215) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$4$}} \rput(-0.403413208727,-0.194273562215) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$5$}} \rput(-0.0996348344258,-0.436528731838) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$6$}} \rput(0.279170602379,-0.350068862824) {\psshadowbox[linewidth=.0001,fillcolor=lightestgray] {$7$}} \end{pspicture} \end{equation} \end{comment} \begin{center} \begin{tabular}{ll} \toprule Label & Coset Representative\\ \midrule $1$ & 1 \\ $2$ & $\delta_a^{}$ \\ $3$ & $\delta_a^{2}$ \\ $4$ & $\delta_a^{3}$ \\ $5$ & $\delta_a^{-3}$ \\ $6$ & $\delta_a^{-2}$ \\ $7$ & $\delta_a^{-1}$ \\ \bottomrule \end{tabular} \;\;\;\;\;\;\;\;\;\; \begin{tabular}{ll} \toprule Label & Side Pairing Element\\ \midrule $s_{1}$ & $\delta_b^{}\delta_a^{2}$ \\ $s_{2}$ & $\delta_b^{-1}\delta_a^{-3}$ \\ $s_{3}$ & $\delta_a^{}\delta_b^{}\delta_a^{3}$ \\ $s_{4}$ & $\delta_a^{}\delta_b^{-1}\delta_a^{2}$ \\ $s_{5}$ & $\delta_a^{-1}\delta_b^{}\delta_a^{-2}$ \\ $s_{6}$ & $\delta_a^{-1}\delta_b^{-1}\delta_a^{3}$ \\ $s_{7}$ & $\delta_a^{2}\delta_b^{}\delta_a^{-3}$ \\ \bottomrule \end{tabular} \end{center} \bigskip \caption{A fundamental domain and conformally drawn dessin for $\Gamma$ corresponding to the permutation triple $$ \sigma_0 = (1\;2\;3\;4),\quad\sigma_1=(1\;2\;3),\quad\sigma_\infty = (1\;4\;2\;3) $$ using the methods in Section \ref{sec:cosets}.} \label{fig:dessin-7,7,7} \end{figure} We find by evaluating the normalized power series the values $\phi^{-1}(0)=\{(1:0:0)\}$, $\phi^{-1}(1)=\{(0:-\nu:1)\}$, and $\phi^{-1}(\infty)=\{(0:\nu:1)\}$. Computing with Riemann-Roch (linear algebra), we find a generator for the $1$-dimensional space of functions with divisor $7(1:0:0)-7(0:\nu:1)$, containing $\phi$; computing similarly with $7(0:-\nu:1)-7(0:\nu:1)$, containing $\phi-1$, we find that \[ \phi(x,y) = \frac{p_2(y)x^2 + p_1(y)x + p_0(y)}{44(y-\nu)^7} \] where \begin{align*} p_2(y) &= (-55\nu - 154)y^4 + (-77\nu - 110)y^3 + (187\nu - 638)y^2 + (385\nu + 550)y \\ p_1(y) &= (55\nu + 154)y^6 + (-44\nu + 440)y^4 + (202\nu + 380)y^3 \\ &\qquad + (-715\nu + 990)y^2 + (-570\nu - 668)y \end{align*} and \begin{align*} p_0(y) &= (77\nu + 110)y^7 + (-308\nu - 264)y^6 + (-1049\nu - 1590)y^5 \\ &\qquad + (2112\nu - 3520)y^4 + (3299\nu + 3634)y^3 \\ &\qquad + (1980\nu + 4840)y^2 + (-2535\nu + 1846)y - 616\nu - 880 \end{align*} where $x=f/h$ and $y=g/h$. \end{exm} \subsection*{Refinement using Newton's method} We have shown how to obtain equations for the Bely\u{\i}\ map $\phi:X(\Gamma) \to X(\Delta)$ defined over a number field. As the complexity of the examples increases, it is easier to use Newton's method to refine the solution: we set up a system of equations with rational coefficients that define the Bely\u{\i}\ map and then refine the approximate solution. \begin{exm} \label{exm:genus0newt} We return to Example \ref{exm:genus0nonewt}. Computing initially in precision $\varepsilon=10^{-30}$, we find an approximation to $\phi(x)$ \[ \frac{(8.265 - 2.345\sqrt{-1})x^4(x-1.190 - 0.1064\sqrt{-1})^3}{(x-1.555- 0.4969\sqrt{-1})^5(x^2 - (0.6334-0.2483\sqrt{-1})x-0.6515-0.5111\sqrt{-1})} \] in about $10$ seconds. We then set up the equations that describe rational functions with the same ramification pattern \cite{SijslingVoight}: we have \[ \phi(x) = \frac{a_9x^4(x+a_1)^3}{(x+a_6)^5(x^2+a_8x+a_7)} = 1+(a_9-1)\frac{(x+a_5)(x^3+a_4x^2+a_3x+a_2)^2}{(x+a_6)^5(x^2+a_8x+a_7)}. \] Expanding and setting the coefficients equal to zero, we obtain a set of $7$ equations in $9$ unknowns. By scaling $x$ we may assume $a_2=a_3$. The condition that $x$ is normalized as in (\ref{eqn:xw}) implies the further condition that \[ a_1a_6a_8 + 5a_1a_7 - 3a_6a_7 = 0, \] and this now gives in total $9$ equations in $9$ unknowns. We apply Newton's method to the approximate solution above: the solution is correct to error $10^{-20}$, so Newton iteration converges after $20$ iterations in about $4$ seconds to a solution which is correct to $10^{-500}$. Now the coefficients of $\phi$ are very easy to recognize! \end{exm} \begin{exm} \label{exm:PSU3(5)} We now consider a much larger example, for which the combined efficiency gains introduced in this paper are essential to make the calculation practical. We consider the exceptional permutation representation of $G=\PSU_3(\mathbb F_5) \hookrightarrow S_{50}$ arising from the action on the cosets of the subgroup $A_7 \leq G$ of index $50$. This exceptional representation can be seen on the Hoffman-Singleton graph \cite{Hafner}, a 7-regular undirected graph with 50 vertices and 175 edges with automorphism group equal to the semidirect product $\PSU_3(\mathbb F_5):2$ (extension by Frobenius). We look for rigid refined passports and refined passports of genus zero. There are $6$ rigid refined passports, all with orders $(5,2,7)$: there are $2$ of genus zero and $4$ of genus two. There are $5$ refined passports of genus zero: the $2$ rigid $(5,2,7)$-refined passports, along with $2$ $(5,2,8)$-triples with refined passport of size 2, and one $(4,2,10)$-triple with refined passport of size 6. We first consider the rigid $(5,2,7)$-refined passports. They arise from the two conjugacy classes of order $7$ in $G$, interchanged by an automorphism of $G$ (also an automorphism of the associated dessin), so as above it suffices to consider just one of these triples: \begin{equation}\label{eqn:sigmasPSU3(5)} \begin{aligned} \sigma_0 &= (2\ 44\ 11\ 7\ 28)(3\ 42\ 4\ 46\ 17)(5\ 32\ 6\ 34\ 16)(8\ 12\ 43\ 33\ 25) \\ &\qquad (10\ 38\ 37\ 31\ 50) (13\ 49\ 40\ 22\ 23)(15\ 48\ 24\ 29\ 26) \\ &\qquad (18\ 21\ 27\ 30\ 45)(19\ 20\ 39\ 41\ 36) \\ \sigma_1 &= (1\ 5)(2\ 25)(4\ 20)(6\ 33)(7\ 47)(8\ 17)(9\ 42)(10\ 41)(11\ 35) \\ &\qquad (13\ 30)(14\ 18)(16\ 37)(19\ 40)(22\ 27)(24\ 29)(26\ 46) \\ &\qquad (28\ 39)(31\ 49)(34\ 48)(38\ 44) \\ \sigma_{\infty} &= (1\ 5\ 37\ 44\ 25\ 6\ 32)(2\ 39\ 4\ 9\ 42\ 3\ 8)(7\ 35\ 11\ 38\ 41\ 28\ 47) \\ &\qquad (10\ 50\ 49\ 30\ 22\ 19\ 36)(12\ 17\ 26\ 24\ 34\ 33\ 43) \\ &\qquad (13\ 23\ 27\ 21\ 14\ 18\ 45)(15\ 46\ 20\ 40\ 31\ 16\ 48) \\ \end{aligned} \end{equation} In Figure \ref{fig:dessin-PSU3(5)}, we display the fundamental domain, using the methods in Section \ref{sec:cosets}. The picture is in the unit disc (centered at $i$) and with the left most point of the fundamental domain at the center. The labels are omitted because of the small triangles close to the boundary. \begin{figure}[h] \includegraphics{belyi-triangle-pics/belyi-triangle-pics-25.pdf} \begin{comment} \begin{equation} \label{fig:dessin-PSU3(5)} \notag \psset{unit=16.5cm} \begin{pspicture}(-1,-1)(1,1) \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.0000000000,0.0000000000) (0.5493821703,0.0000000000) \psarc[linewidth=.1pt] (1.184804349,0.0000000000) {0.6354221785} {180.0000000} {208.2857143} \psline[linewidth=.1pt] (0.6252544176,-0.3011066578) (0.0000000000,0.0000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.0000000000,0.0000000000) (0.5493821703,0.0000000000) \psarcn[linewidth=.1pt] (1.184804349,0.0000000000) {0.6354221785} {180.0000000} {151.7142857} \psline[linewidth=.1pt] (0.6252544176,0.3011066578) (0.0000000000,0.0000000000) \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.014412779,0.1358827424) {0.2179389974} {218.5714286} {164.5318834} \psarc[linewidth=.1pt] (0.9091049986,0.5724955983) {0.3927125012} {254.5318832} {270.2767345} \psarc[linewidth=.1pt] (1.014412779,0.03889076404) {0.1747735059} {126.2767346} {192.8571429} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.014412779,0.1358827424) {0.2179389974} {218.5714286} {164.5318834} \psarcn[linewidth=.1pt] (0.9091049979,0.5724956007) {0.3927125034} {254.5318834} {223.7142857} \psarc[linewidth=.1pt] (1.014412779,0.3538217396) {0.3927125035} {187.7142857} {244.2857143} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9960963062,-0.09784819315) {0.04221516457} {229.0267645} {252.0543854} \psarc[linewidth=.1pt] (0.9901378382,-0.1402923933) {0.007409071865} {162.0544773} {213.9664053} \psarcn[linewidth=.1pt] (0.9932609989,-0.1190147655) {0.02705413710} {249.9663337} {203.3124697} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9960963062,-0.09784819315) {0.04221516457} {229.0267645} {252.0543854} \psarcn[linewidth=.1pt] (0.9901378341,-0.1402924233) {0.007409077170} {162.0542469} {116.9046445} \psarcn[linewidth=.1pt] (1.014412772,0.03889071110) {0.1747734529} {260.9047240} {254.7410475} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9572089103,-0.2895172846) {0.008316007691} {181.2302763} {228.9613976} \psarc[linewidth=.1pt] (0.9545120121,-0.2981949672) {0.003663300784} {138.9615475} {193.2624289} \psarcn[linewidth=.1pt] (0.9561462420,-0.2929980405) {0.007967914241} {229.2623259} {155.5160856} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9572089103,-0.2895172846) {0.008316007691} {181.2302763} {228.9613976} \psarcn[linewidth=.1pt] (0.9545120065,-0.2981949852) {0.003663308353} {138.9612770} {100.0511443} \psarcn[linewidth=.1pt] (0.9586716234,-0.2847261874) {0.01096737560} {244.0512329} {206.9446895} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9960963062,0.09784819315) {0.04221516457} {130.9732355} {107.9456146} \psarc[linewidth=.1pt] (0.9901378341,0.1402924233) {0.007409077170} {197.9457531} {243.0953555} \psarc[linewidth=.1pt] (1.014412772,-0.03889071110) {0.1747734529} {99.09527597} {105.2589525} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9960963062,0.09784819315) {0.04221516457} {130.9732355} {107.9456146} \psarcn[linewidth=.1pt] (0.9901378382,0.1402923933) {0.007409071865} {197.9455227} {146.0335947} \psarc[linewidth=.1pt] (0.9932609989,0.1190147655) {0.02705413710} {110.0336663} {156.6875303} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9318217455,-0.3726726023) {0.08471501574} {128.6858626} {91.05700503} \psarc[linewidth=.1pt] (0.9582013479,-0.2874564625) {0.02794711034} {181.0569952} {221.8768672} \psarc[linewidth=.1pt] (0.9091049980,-0.4378018930) {0.1346937080} {77.87687223} {102.9715756} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9318217455,-0.3726726023) {0.08471501574} {128.6858626} {91.05700503} \psarcn[linewidth=.1pt] (0.9582013498,-0.2874564563) {0.02794711232} {181.0570078} {128.5548118} \psarc[linewidth=.1pt] (0.9440022176,-0.3377509875) {0.07222130003} {92.55480939} {154.4001479} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9458616219,-0.3252438131) {0.02092237691} {80.13668444} {74.52999990} \psarc[linewidth=.1pt] (0.9522552119,-0.3053043119) {0.0008434131268} {164.5376077} {212.9521677} \psarcn[linewidth=.1pt] (0.9549395677,-0.2969504861) {0.009442947702} {248.9473176} {234.4222274} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9458616219,-0.3252438131) {0.02092237691} {80.13668444} {74.52999990} \psarcn[linewidth=.1pt] (0.9522551450,-0.3053045203) {0.0008434042592} {164.5227505} {114.5060898} \psarc[linewidth=.1pt] (0.9508679810,-0.3096404423) {0.005207775775} {78.51088354} {105.8505230} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9882015265,-0.1567855182) {0.03352545021} {126.1696220} {188.6436086} \psarc[linewidth=.1pt] (0.9703271462,-0.2622769828) {0.1016070200} {98.64360681} {125.7232644} \psarcn[linewidth=.1pt] (0.9819802436,-0.2032295978) {0.07474936882} {161.7232642} {100.4553351} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9882015265,-0.1567855182) {0.03352545021} {126.1696220} {188.6436086} \psarcn[linewidth=.1pt] (0.9703271472,-0.2622769770) {0.1016070144} {98.64360783} {87.46025330} \psarcn[linewidth=.1pt] (0.9902631543,-0.1413947366) {0.02477067757} {231.4602533} {151.8839069} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9931040509,-0.1172724830) {0.002913957238} {167.1427068} {112.6908455} \psarc[linewidth=.1pt] (0.9934873969,-0.1139537715) {0.001633896515} {202.6908510} {238.8422117} \psarc[linewidth=.1pt] (0.9929299558,-0.1187507367) {0.003410940555} {94.84232622} {141.4289057} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9931040509,-0.1172724830) {0.002913957238} {167.1427068} {112.6908455} \psarcn[linewidth=.1pt] (0.9934873948,-0.1139537899) {0.001633887486} {202.6902843} {148.2109490} \psarc[linewidth=.1pt] (0.9932610014,-0.1159398051) {0.003074926827} {112.2117406} {192.8574147} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9549395747,-0.2969504642) {0.009442969122} {234.4222701} {195.8590104} \psarcn[linewidth=.1pt] (0.9497649549,-0.3132908304) {0.01430432967} {105.8587974} {85.26231539} \psarc[linewidth=.1pt] (0.9531212497,-0.3026176566) {0.004190787898} {121.2623797} {208.7079655} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9549395747,-0.2969504642) {0.009442969122} {234.4222701} {195.8590104} \psarc[linewidth=.1pt] (0.9497649608,-0.3132908119) {0.01430431347} {105.8588404} {149.8768768} \psarcn[linewidth=.1pt] (0.9458614650,-0.3252443023) {0.02092288461} {113.8768641} {80.13649049} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9318217407,-0.3726726174) {0.08471503056} {78.37323967} {91.05700161} \psarcn[linewidth=.1pt] (0.9582013498,-0.2874564563) {0.02794711232} {181.0570078} {128.5548118} \psarc[linewidth=.1pt] (0.9626645980,-0.2716472530) {0.02270149850} {164.5548146} {232.6589566} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9318217407,-0.3726726174) {0.08471503056} {78.37323967} {91.05700161} \psarc[linewidth=.1pt] (0.9582013479,-0.2874564625) {0.02794711034} {181.0569952} {221.8768672} \psarcn[linewidth=.1pt] (0.9526211187,-0.3045445340) {0.01530912085} {185.8768744} {104.0875273} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9916960326,0.1306977927) {0.02330095014} {182.4018064} {128.3927811} \psarc[linewidth=.1pt] (0.9876527123,0.1572231869) {0.01330453039} {218.3927952} {254.0336706} \psarc[linewidth=.1pt] (0.9932609989,0.1190147655) {0.02705413710} {110.0336663} {156.6875303} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9916960326,0.1306977927) {0.02330095014} {182.4018064} {128.3927811} \psarcn[linewidth=.1pt] (0.9876527134,0.1572231799) {0.01330452685} {218.3927687} {164.5397335} \psarc[linewidth=.1pt] (0.9902631543,0.1413947366) {0.02477067757} {128.5397467} {208.1160931} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9531212370,-0.3026176961) {0.004190764344} {208.7075760} {238.9132676} \psarc[linewidth=.1pt] (0.9518016821,-0.3067157154) {0.0009859016876} {148.9094155} {201.8872654} \psarcn[linewidth=.1pt] (0.9525270691,-0.3044695456) {0.003085682434} {237.8903486} {182.9934452} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9531212370,-0.3026176961) {0.004190764344} {208.7075760} {238.9132676} \psarcn[linewidth=.1pt] (0.9518017178,-0.3067156040) {0.0009858748139} {148.9160284} {104.9487821} \psarcn[linewidth=.1pt] (0.9549395677,-0.2969504861) {0.009442947702} {248.9473176} {234.4222274} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9586716217,-0.2847261927) {0.01096737014} {206.9446687} {142.4175537} \psarc[linewidth=.1pt] (0.9671642204,-0.2557092058) {0.02817493754} {232.4175018} {245.6198548} \psarc[linewidth=.1pt] (0.9572089164,-0.2895172643) {0.008316008612} {101.6198909} {181.2304149} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9586716217,-0.2847261927) {0.01096737014} {206.9446687} {142.4175537} \psarcn[linewidth=.1pt] (0.9671642424,-0.2557091302) {0.02817501093} {232.4175602} {200.5548216} \psarc[linewidth=.1pt] (0.9626645980,-0.2716472530) {0.02270149850} {164.5548146} {232.6589566} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.014412779,-0.03889076311) {0.1747735054} {105.2589502} {122.5769067} \psarcn[linewidth=.1pt] (0.9916960324,0.1539987445) {0.08471501741} {212.5769074} {162.2767353} \psarc[linewidth=.1pt] (0.9819802436,0.2032295978) {0.07474936882} {198.2767358} {259.5446649} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.014412779,-0.03889076311) {0.1747735054} {105.2589502} {122.5769067} \psarc[linewidth=.1pt] (0.9916960328,0.1539987415) {0.08471501612} {212.5769056} {248.3413338} \psarcn[linewidth=.1pt] (0.9960963060,0.09784819479) {0.04221516454} {212.3413357} {130.9732367} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9902631612,-0.1413946793) {0.02477062026} {90.00001588} {105.9973864} \psarcn[linewidth=.1pt] (0.9934535014,-0.1147114967) {0.01042050378} {195.9973613} {142.1125299} \psarc[linewidth=.1pt] (0.9941247946,-0.1086052762) {0.008900171490} {178.1125429} {244.2857266} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9902631612,-0.1413946793) {0.02477062026} {90.00001588} {105.9973864} \psarc[linewidth=.1pt] (0.9934534996,-0.1147115117) {0.01042049792} {195.9972848} {235.1905012} \psarcn[linewidth=.1pt] (0.9926071761,-0.1214914824) {0.005402417173} {199.1904325} {115.7142298} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.014412780,-0.1358827416) {0.2179389981} {231.5430043} {195.4681168} \psarcn[linewidth=.1pt] (0.9091049986,-0.5724955983) {0.3927125012} {105.4681168} {89.72326551} \psarc[linewidth=.1pt] (0.9703271468,-0.2622769797) {0.1016070171} {125.7232657} {205.8287188} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.014412780,-0.1358827416) {0.2179389981} {231.5430043} {195.4681168} \psarc[linewidth=.1pt] (0.9091049979,-0.5724956007) {0.3927125034} {105.4681166} {136.2857143} \psarcn[linewidth=.1pt] (0.7387134287,-0.9263173402) {0.6354221783} {100.2857143} {77.25728995} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.014412779,-0.1358827424) {0.2179389974} {141.4285714} {195.4681166} \psarc[linewidth=.1pt] (0.9091049979,-0.5724956007) {0.3927125034} {105.4681166} {136.2857143} \psarcn[linewidth=.1pt] (1.014412779,-0.3538217396) {0.3927125035} {172.2857143} {115.7142857} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.014412779,-0.1358827424) {0.2179389974} {141.4285714} {195.4681166} \psarcn[linewidth=.1pt] (0.9091049986,-0.5724955983) {0.3927125012} {105.4681168} {89.72326551} \psarcn[linewidth=.1pt] (1.014412779,-0.03889076404) {0.1747735059} {233.7232654} {167.1428571} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9549395608,-0.2969505080) {0.009442943026} {129.8017515} {195.8587776} \psarc[linewidth=.1pt] (0.9497649608,-0.3132908119) {0.01430431347} {105.8588404} {149.8768768} \psarcn[linewidth=.1pt] (0.9526211187,-0.3045445340) {0.01530912085} {185.8768744} {104.0875273} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9549395608,-0.2969505080) {0.009442943026} {129.8017515} {195.8587776} \psarcn[linewidth=.1pt] (0.9497649549,-0.3132908304) {0.01430432967} {105.8587974} {85.26231539} \psarcn[linewidth=.1pt] (0.9561462420,-0.2929980405) {0.007967914241} {229.2623259} {155.5160856} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9549395647,-0.3063934320) {0.07606909759} {180.1144338} {119.8110419} \psarc[linewidth=.1pt] (0.9819802436,-0.2032295969) {0.07474936801} {209.8110408} {236.5548089} \psarc[linewidth=.1pt] (0.9440022176,-0.3377509875) {0.07222130003} {92.55480939} {154.4001479} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9549395647,-0.3063934320) {0.07606909759} {180.1144338} {119.8110419} \psarcn[linewidth=.1pt] (0.9819802439,-0.2032295957) {0.07474936883} {209.8110414} {161.7232658} \psarc[linewidth=.1pt] (0.9703271468,-0.2622769797) {0.1016070171} {125.7232657} {205.8287188} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9525270771,-0.3044695206) {0.003085698463} {182.9939010} {118.1037561} \psarc[linewidth=.1pt] (0.9538633234,-0.3002578274) {0.003162698666} {208.1032151} {235.1743744} \psarc[linewidth=.1pt] (0.9521165317,-0.3057490719) {0.002895618255} {91.17466186} {157.2797873} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9525270771,-0.3044695206) {0.003085698463} {182.9939010} {118.1037561} \psarcn[linewidth=.1pt] (0.9538633276,-0.3002578144) {0.003162708439} {208.1033877} {157.2620965} \psarc[linewidth=.1pt] (0.9531212497,-0.3026176566) {0.004190787898} {121.2623797} {208.7079655} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.014412779,0.3538217405) {0.3927125042} {244.2857143} {260.4388943} \psarc[linewidth=.1pt] (1.000467104,-0.04207377268) {0.05200604367} {170.4388948} {219.6586661} \psarcn[linewidth=.1pt] (1.014412779,0.1358827429) {0.2179389980} {255.6586657} {218.5714287} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.014412779,0.3538217405) {0.3927125042} {244.2857143} {260.4388943} \psarcn[linewidth=.1pt] (1.000467104,-0.04207377446) {0.05200604385} {170.4388929} {125.9999991} \psline[linewidth=.1pt] (0.9698987185,0.0000000000) (0.8440212099,0.0000000000) \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9440022171,-0.3377509886) {0.07222130086} {154.4001469} {211.0298703} \psarc[linewidth=.1pt] (0.9091049975,-0.4198444162) {0.05235676052} {121.0298692} {172.2857129} \psarcn[linewidth=.1pt] (0.9318217448,-0.3726726044) {0.08471501695} {208.2857143} {128.6858614} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9440022171,-0.3377509886) {0.07222130086} {154.4001469} {211.0298703} \psarcn[linewidth=.1pt] (0.9091050000,-0.4198444101) {0.05235675665} {121.0298750} {89.43434320} \psarcn[linewidth=.1pt] (0.9549395642,-0.3063934338) {0.07606909670} {233.4343410} {180.1144325} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9929299544,-0.1187507493) {0.003410957913} {141.4287253} {207.8393129} \psarc[linewidth=.1pt] (0.9922081503,-0.1246881113) {0.004913098469} {117.8392822} {163.1907277} \psarcn[linewidth=.1pt] (0.9926071761,-0.1214914824) {0.005402417173} {199.1904325} {115.7142298} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9929299544,-0.1187507493) {0.003410957913} {141.4287253} {207.8393129} \psarcn[linewidth=.1pt] (0.9922081333,-0.1246882509) {0.004913213959} {117.8383468} {96.36692729} \psarcn[linewidth=.1pt] (0.9931040531,-0.1172724646) {0.002913966355} {240.3673373} {167.1430682} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9932610030,-0.1159397917) {0.003074950084} {192.8576497} {239.9814767} \psarc[linewidth=.1pt] (0.9928637471,-0.1192615471) {0.001317838137} {149.9832281} {204.3690806} \psarcn[linewidth=.1pt] (0.9931040531,-0.1172724646) {0.002913966355} {240.3673373} {167.1430682} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9932610030,-0.1159397917) {0.003074950084} {192.8576497} {239.9814767} \psarcn[linewidth=.1pt] (0.9928637280,-0.1192617071) {0.001317901625} {149.9767921} {110.7171349} \psarcn[linewidth=.1pt] (0.9934828011,-0.1140564895) {0.004118070532} {254.7193494} {218.5712872} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.014412779,0.03889076311) {0.1747735054} {254.7410498} {237.4230933} \psarcn[linewidth=.1pt] (0.9916960328,-0.1539987415) {0.08471501612} {147.4230944} {111.6586662} \psarc[linewidth=.1pt] (0.9960963060,-0.09784819479) {0.04221516454} {147.6586643} {229.0267633} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.014412779,0.03889076311) {0.1747735054} {254.7410498} {237.4230933} \psarc[linewidth=.1pt] (0.9916960324,-0.1539987445) {0.08471501741} {147.4230926} {197.7232647} \psarcn[linewidth=.1pt] (0.9819802436,-0.2032295978) {0.07474936882} {161.7232642} {100.4553351} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9902631544,-0.1413947358) {0.02477067722} {151.8839086} {105.9973352} \psarc[linewidth=.1pt] (0.9934534996,-0.1147115117) {0.01042049792} {195.9972848} {235.1905012} \psarc[linewidth=.1pt] (0.9882015274,-0.1567855118) {0.03352544368} {91.19052856} {126.1696297} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9902631544,-0.1413947358) {0.02477067722} {151.8839086} {105.9973352} \psarcn[linewidth=.1pt] (0.9934535014,-0.1147114967) {0.01042050378} {195.9973613} {142.1125299} \psarc[linewidth=.1pt] (0.9916960329,-0.1306977906) {0.02330094788} {106.1125073} {177.5981987} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.7387134323,-0.9263173300) {0.6354221674} {77.25729007} {72.87041935} \psarc[linewidth=.1pt] (0.9458614670,-0.3252442960) {0.02092288137} {162.8704251} {210.5980361} \psarcn[linewidth=.1pt] (1.014412778,-0.1358827450) {0.2179389946} {246.5980323} {231.5430040} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.7387134323,-0.9263173300) {0.6354221674} {77.25729007} {72.87041935} \psarcn[linewidth=.1pt] (0.9458614662,-0.3252442988) {0.02092288135} {162.8704171} {113.8768709} \psarc[linewidth=.1pt] (0.9091049980,-0.4378018930) {0.1346937080} {77.87687223} {102.9715756} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9932609997,-0.1190147603) {0.02705414201} {203.3124792} {138.4080925} \psarc[linewidth=.1pt] (0.9980546135,-0.07285883468) {0.03770174048} {228.4081037} {250.1125100} \psarc[linewidth=.1pt] (0.9916960329,-0.1306977906) {0.02330094788} {106.1125073} {177.5981987} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9932609997,-0.1190147603) {0.02705414201} {203.3124792} {138.4080925} \psarcn[linewidth=.1pt] (0.9980546119,-0.07285885043) {0.03770172766} {228.4080896} {183.6586624} \psarc[linewidth=.1pt] (0.9960963060,-0.09784819479) {0.04221516454} {147.6586643} {229.0267633} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9526211250,-0.3045445133) {0.01530910158} {104.0875689} {82.37969553} \psarc[linewidth=.1pt] (0.9571204387,-0.2897010043) {0.002491215690} {172.3788870} {218.0099146} \psarc[linewidth=.1pt] (0.9318218176,-0.3726723656) {0.08471476732} {74.01038089} {78.37325627} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9526211250,-0.3045445133) {0.01530910158} {104.0875689} {82.37969553} \psarcn[linewidth=.1pt] (0.9571204501,-0.2897009668) {0.002491221966} {172.3797768} {120.4097652} \psarc[linewidth=.1pt] (0.9549395636,-0.2969504992) {0.009442939159} {84.40969203} {129.8017988} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (1.014412779,0.03889076396) {0.1747735062} {192.8571429} {237.4230934} \psarc[linewidth=.1pt] (0.9916960324,-0.1539987445) {0.08471501741} {147.4230926} {197.7232647} \psarcn[linewidth=.1pt] (1.014412779,-0.03889076404) {0.1747735059} {233.7232654} {167.1428571} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (1.014412779,0.03889076396) {0.1747735062} {192.8571429} {237.4230934} \psarcn[linewidth=.1pt] (0.9916960328,-0.1539987415) {0.08471501612} {147.4230944} {111.6586662} \psarcn[linewidth=.1pt] (1.014412779,0.1358827429) {0.2179389980} {255.6586657} {218.5714287} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9819802454,-0.2032295841) {0.07474935686} {100.4553383} {87.86316294} \psarc[linewidth=.1pt] (0.9916960347,-0.1287907148) {0.006933490838} {177.8632877} {224.9048008} \psarcn[linewidth=.1pt] (1.014412772,0.03889071110) {0.1747734529} {260.9047240} {254.7410475} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9819802454,-0.2032295841) {0.07474935686} {100.4553383} {87.86316294} \psarcn[linewidth=.1pt] (0.9916960316,-0.1287907390) {0.006933488600} {177.8630867} {127.1904743} \psarc[linewidth=.1pt] (0.9882015274,-0.1567855118) {0.03352544368} {91.19052856} {126.1696297} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9626645937,-0.2716472663) {0.02270148716} {232.6589447} {246.9683109} \psarc[linewidth=.1pt] (0.9559713749,-0.2934695649) {0.002378067063} {156.9687354} {208.0515272} \psarcn[linewidth=.1pt] (0.9586716234,-0.2847261874) {0.01096737560} {244.0512329} {206.9446895} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9626645937,-0.2716472663) {0.02270148716} {232.6589447} {246.9683109} \psarcn[linewidth=.1pt] (0.9559713691,-0.2934695840) {0.002378069164} {156.9682578} {110.0103444} \psarc[linewidth=.1pt] (0.9318218176,-0.3726723656) {0.08471476732} {74.01038089} {78.37325627} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9941248152,-0.1086051011) {0.008900343247} {244.2860961} {257.6859801} \psarc[linewidth=.1pt] (0.9930749379,-0.1174858353) {0.0008682661867} {167.6877410} {218.7207048} \psarcn[linewidth=.1pt] (0.9934828011,-0.1140564895) {0.004118070532} {254.7193494} {218.5712872} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9941248152,-0.1086051011) {0.008900343247} {244.2860961} {257.6859801} \psarcn[linewidth=.1pt] (0.9930749423,-0.1174857988) {0.0008682626734} {167.6901544} {120.5099523} \psarc[linewidth=.1pt] (0.9902631256,-0.1413949810) {0.02477092355} {84.50733028} {89.99993349} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9561462460,-0.2929980271) {0.007967914125} {155.5161857} {101.9772780} \psarc[linewidth=.1pt] (0.9587447774,-0.2843015053) {0.004346696224} {191.9777562} {228.4099443} \psarc[linewidth=.1pt] (0.9549395636,-0.2969504992) {0.009442939159} {84.40969203} {129.8017988} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9561462460,-0.2929980271) {0.007967914125} {155.5161857} {101.9772780} \psarcn[linewidth=.1pt] (0.9587447670,-0.2843015398) {0.004346678943} {191.9773406} {137.6199841} \psarc[linewidth=.1pt] (0.9572089164,-0.2895172643) {0.008316008612} {101.6198909} {181.2304149} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9916960326,-0.1306977927) {0.02330095014} {177.5981936} {231.6072189} \psarc[linewidth=.1pt] (0.9876527134,-0.1572231799) {0.01330452685} {141.6072313} {195.4602665} \psarcn[linewidth=.1pt] (0.9902631543,-0.1413947366) {0.02477067757} {231.4602533} {151.8839069} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9916960326,-0.1306977927) {0.02330095014} {177.5981936} {231.6072189} \psarcn[linewidth=.1pt] (0.9876527123,-0.1572231869) {0.01330453039} {141.6072048} {105.9663294} \psarcn[linewidth=.1pt] (0.9932609989,-0.1190147655) {0.02705413710} {249.9663337} {203.3124697} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9091049980,-0.4378018928) {0.1346937078} {102.9715756} {154.2857143} \psline[linewidth=.1pt] (0.7877501606,-0.3793604833) (0.6252544176,-0.3011066578) \psarcn[linewidth=.1pt] (0.7387134287,-0.9263173402) {0.6354221783} {100.2857143} {77.25728995} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9091049980,-0.4378018928) {0.1346937078} {102.9715756} {154.2857143} \psline[linewidth=.1pt] (0.7877501606,-0.3793604833) (0.8572220794,-0.4128163961) \psarcn[linewidth=.1pt] (0.9318217448,-0.3726726044) {0.08471501695} {208.2857143} {128.6858614} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9926071625,-0.1214915972) {0.005402527392} {115.7135717} {92.99613231} \psarc[linewidth=.1pt] (0.9932439937,-0.1160483356) {0.0009204720154} {182.9954855} {228.5068387} \psarc[linewidth=.1pt] (0.9902631256,-0.1413949810) {0.02477092355} {84.50733028} {89.99993349} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9926071625,-0.1214915972) {0.005402527392} {115.7135717} {92.99613231} \psarcn[linewidth=.1pt] (0.9932440047,-0.1160482423) {0.0009204878811} {183.0012483} {130.8459092} \psarc[linewidth=.1pt] (0.9929299558,-0.1187507367) {0.003410940555} {94.84232622} {141.4289057} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9932609997,0.1190147603) {0.02705414201} {156.6875208} {221.5919075} \psarc[linewidth=.1pt] (0.9980546119,0.07285885043) {0.03770172766} {131.5919104} {176.3413376} \psarcn[linewidth=.1pt] (0.9960963060,0.09784819479) {0.04221516454} {212.3413357} {130.9732367} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9932609997,0.1190147603) {0.02705414201} {156.6875208} {221.5919075} \psarcn[linewidth=.1pt] (0.9980546135,0.07285883468) {0.03770174048} {131.5918963} {109.8874900} \psarcn[linewidth=.1pt] (0.9916960329,0.1306977906) {0.02330094788} {253.8874927} {182.4018013} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9508679893,-0.3096404164) {0.005207746351} {105.8506883} {164.1651576} \psarc[linewidth=.1pt] (0.9091050167,-0.4378018357) {0.1346936477} {74.16531970} {77.87687489} \psarcn[linewidth=.1pt] (0.9458614650,-0.3252443023) {0.02092288461} {113.8768641} {80.13649049} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9508679893,-0.3096404164) {0.005207746351} {105.8506883} {164.1651576} \psarcn[linewidth=.1pt] (0.9091048109,-0.4378024676) {0.1346943119} {74.16530879} {72.77793843} \psarcn[linewidth=.1pt] (0.9516703794,-0.3071395991) {0.003353263736} {216.7779920} {131.5651369} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.014412779,-0.3538217405) {0.3927125042} {115.7142857} {99.56110572} \psarc[linewidth=.1pt] (1.000467104,0.04207377446) {0.05200604385} {189.5611071} {234.0000009} \psline[linewidth=.1pt] (0.9698987185,0.0000000000) (0.8440212099,0.0000000000) \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.014412779,-0.3538217405) {0.3927125042} {115.7142857} {99.56110572} \psarcn[linewidth=.1pt] (1.000467104,0.04207377268) {0.05200604367} {189.5611052} {140.3413339} \psarc[linewidth=.1pt] (1.014412779,-0.1358827429) {0.2179389980} {104.3413343} {141.4285713} \closepath} \pscustom[linewidth=.1pt]{ \psline[linewidth=.1pt] (0.8440212099,0.0000000000) (0.5493821703,0.0000000000) \psarcn[linewidth=.1pt] (1.184804349,0.0000000000) {0.6354221785} {180.0000000} {151.7142857} \psarc[linewidth=.1pt] (1.014412779,0.3538217396) {0.3927125035} {187.7142857} {244.2857143} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psline[linewidth=.1pt] (0.8440212099,0.0000000000) (0.5493821703,0.0000000000) \psarc[linewidth=.1pt] (1.184804349,0.0000000000) {0.6354221785} {180.0000000} {208.2857143} \psarcn[linewidth=.1pt] (1.014412779,-0.3538217396) {0.3927125035} {172.2857143} {115.7142857} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9516703918,-0.3071395604) {0.003353236413} {131.5657342} {91.11606470} \psarc[linewidth=.1pt] (0.9527477885,-0.3037646419) {0.001142928239} {181.1198952} {222.5130791} \psarc[linewidth=.1pt] (0.9508679810,-0.3096404423) {0.005207775775} {78.51088354} {105.8505230} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9516703918,-0.3071395604) {0.003353236413} {131.5657342} {91.11606470} \psarcn[linewidth=.1pt] (0.9527477687,-0.3037647042) {0.001142907177} {181.1167901} {127.1753625} \psarc[linewidth=.1pt] (0.9521165317,-0.3057490719) {0.002895618255} {91.17466186} {157.2797873} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9703271464,-0.2622769811) {0.1016070167} {205.8287182} {235.5375961} \psarc[linewidth=.1pt] (0.9332546214,-0.3600686019) {0.02477067862} {145.5375860} {197.4343330} \psarcn[linewidth=.1pt] (0.9549395642,-0.3063934338) {0.07606909670} {233.4343410} {180.1144325} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9703271464,-0.2622769811) {0.1016070167} {205.8287182} {235.5375961} \psarcn[linewidth=.1pt] (0.9332546243,-0.3600685942) {0.02477067665} {145.5376045} {102.5980370} \psarcn[linewidth=.1pt] (1.014412778,-0.1358827450) {0.2179389946} {246.5980323} {231.5430040} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9882015265,0.1567855182) {0.03352545021} {233.8303780} {171.3563914} \psarc[linewidth=.1pt] (0.9703271472,0.2622769770) {0.1016070144} {261.3563922} {272.5397467} \psarc[linewidth=.1pt] (0.9902631543,0.1413947366) {0.02477067757} {128.5397467} {208.1160931} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9882015265,0.1567855182) {0.03352545021} {233.8303780} {171.3563914} \psarcn[linewidth=.1pt] (0.9703271462,0.2622769828) {0.1016070200} {261.3563932} {234.2767356} \psarc[linewidth=.1pt] (0.9819802436,0.2032295978) {0.07474936882} {198.2767358} {259.5446649} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9902631544,0.1413947358) {0.02477067722} {208.1160914} {254.0026648} \psarc[linewidth=.1pt] (0.9934535014,0.1147114967) {0.01042050378} {164.0026387} {217.8874701} \psarcn[linewidth=.1pt] (0.9916960329,0.1306977906) {0.02330094788} {253.8874927} {182.4018013} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9902631544,0.1413947358) {0.02477067722} {208.1160914} {254.0026648} \psarcn[linewidth=.1pt] (0.9934534996,0.1147115117) {0.01042049792} {164.0027152} {124.8094988} \psarcn[linewidth=.1pt] (0.9882015274,0.1567855118) {0.03352544368} {268.8094714} {233.8303703} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (1.014412779,-0.03889076396) {0.1747735062} {167.1428571} {122.5769066} \psarc[linewidth=.1pt] (0.9916960328,0.1539987415) {0.08471501612} {212.5769056} {248.3413338} \psarc[linewidth=.1pt] (1.014412779,-0.1358827429) {0.2179389980} {104.3413343} {141.4285713} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (1.014412779,-0.03889076396) {0.1747735062} {167.1428571} {122.5769066} \psarcn[linewidth=.1pt] (0.9916960324,0.1539987445) {0.08471501741} {212.5769074} {162.2767353} \psarc[linewidth=.1pt] (1.014412779,0.03889076404) {0.1747735059} {126.2767346} {192.8571429} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9819802454,0.2032295841) {0.07474935686} {259.5446617} {272.1368371} \psarc[linewidth=.1pt] (0.9916960316,0.1287907390) {0.006933488600} {182.1369133} {232.8095257} \psarcn[linewidth=.1pt] (0.9882015274,0.1567855118) {0.03352544368} {268.8094714} {233.8303703} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9819802454,0.2032295841) {0.07474935686} {259.5446617} {272.1368371} \psarcn[linewidth=.1pt] (0.9916960347,0.1287907148) {0.006933490838} {182.1367123} {135.0951992} \psarc[linewidth=.1pt] (1.014412772,-0.03889071110) {0.1747734529} {99.09527597} {105.2589525} \closepath} \pscustom[linewidth=.1pt]{ \psarcn[linewidth=.1pt] (0.9934828009,-0.1140564901) {0.004118074041} {218.5712816} {154.0623498} \psarc[linewidth=.1pt] (0.9948753149,-0.1017783860) {0.01165041310} {244.0624382} {256.2115603} \psarc[linewidth=.1pt] (0.9932610014,-0.1159398051) {0.003074926827} {112.2117406} {192.8574147} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarcn[linewidth=.1pt] (0.9934828009,-0.1140564901) {0.004118074041} {218.5712816} {154.0623498} \psarcn[linewidth=.1pt] (0.9948753237,-0.1017783082) {0.01165048692} {244.0625670} {214.1125522} \psarc[linewidth=.1pt] (0.9941247946,-0.1086052762) {0.008900171490} {178.1125429} {244.2857266} \closepath} \pscustom[linewidth=.1pt]{ \psarc[linewidth=.1pt] (0.9521165168,-0.3057491177) {0.002895641413} {157.2788368} {216.8823995} \psarc[linewidth=.1pt] (0.9510253741,-0.3091195618) {0.002041037870} {126.8821737} {180.7773379} \psarcn[linewidth=.1pt] (0.9516703794,-0.3071395991) {0.003353263736} {216.7779920} {131.5651369} \closepath} \pscustom[linewidth=.1pt,fillstyle=solid,fillcolor=lightgray]{ \psarc[linewidth=.1pt] (0.9521165168,-0.3057491177) {0.002895641413} {157.2788368} {216.8823995} \psarcn[linewidth=.1pt] (0.9510254070,-0.3091194605) {0.002040976575} {126.8846191} {93.89127357} \psarcn[linewidth=.1pt] (0.9525270691,-0.3044695456) {0.003085682434} {237.8903486} {182.9934452} \closepath} \end{pspicture} \end{equation} \end{comment} \caption{Fundamental domain for $\Gamma$ corresponding to the permutation triple \ref{eqn:sigmasPSU3(5)}.} \label{fig:dessin-PSU3(5)} \end{figure} We work in precision $\varepsilon=10^{-60}$, and it takes about $20$ minutes (with twice $172$ Arnoldi iterations) for each triple followed by another $15$ minutes of Newton iteration to obtain $1000$ digits of precision. Initially, we find a Bely\u{\i}\ map defined over $\mathbb Q(\zeta_7)^+$, the cubic totally real subfield of the cyclotomic field $\mathbb Q(\zeta_7)$; this is also the field of definition of the normalized power series we computed. After some creative simplification, involving a change of coordinates as in (\ref{eqn:phiop}), we descend to the Bely\u{\i}\ map \[ \phi(x) = \frac{p(x)}{q(x)} = 1+ \frac{r(x)}{q(x)} \] where \begin{align*} p(x) &= 2^6 (x^4 + 11x^3 - 29x^2 + 11x + 1)^5 (64x^5 - 100x^4 + 150x^3 - 25x^2 + 5x + 1)^5 \\ &\qquad \cdot (196x^5 - 430x^4 + 485x^3 - 235x^2 + 30x + 4) \\ q(x) &= 5^{10} x^7 (x+1)^7 (2x^2 - 3x + 2)^7 (8x^3 - 32x^2 + 10x + 1)^7 \\ \end{align*} and \begin{align*} r(x) &= (28672x^{20} - 2114560x^{19} + 13722240x^{18} - 65614080x^{17} + 245351840x^{16} \\ &\qquad\qquad - 660267008x^{15} + 1248458280x^{14} - 1700835920x^{13} + 1704958640x^{12} \\ &\qquad\qquad - 1267574420x^{11} + 690436992x^{10} - 257110380x^9 + 52736995x^8 - 948040x^7 \\ &\qquad\qquad - 1171555x^6 - 246148x^5 + 86660x^4 + 11060x^3 - 1520x^2 - 240x - 8)^2 \\ &\qquad\cdot (16384x^{10} - 34960x^9 - 160960x^8 + 620820x^7 - 792960x^6 + 416087x^5 \\ &\qquad\qquad + 57435x^4 + 935x^3 + 705x^2 + 110x + 4). \end{align*} Remarkably, the Galois group of the degree $20$ polynomial $r(x)$ is a group of order $240$, an extension of the normal subgroup $A_5$ by the Klein 4-group $V_4$, and the associated number field has discriminant $2^{14} 5^{26} 7^{15}$. Similarly, the other factors have non-generic Galois group with small ramification. Computing monodromy (as in the next subsection), we verify that the polynomial $f(x) = p(x) - tq(x) \in \mathbb C[x;t]$ indeed has Galois group $\PSU_3(\mathbb F_5)$ and this descends to a $\PSU_3(\mathbb F_5)$-extension of the field of rationality $K=\mathbb Q(\sqrt{-7})$. As a Galois extension of $\mathbb Q(t)$, therefore, we find the group $\PSU_3(\mathbb F_5):2$; the additional descent to $\mathbb Q$ follows from the fact that the dessin has an automorphism of order $2$ identifying it with its complex conjugate, which allows us to descend the cover defined by $\phi$ to $\mathbb Q$ as a \defi{mere cover} (see D\`ebes and Emsalem \cite{DebesEmsalem} for the definition), but the full Galois group is only defined over $K$. The existence of the rational function $f(x)$ was assured by the general theory of rigidity \cite{MalleMatzat}, but the explicit polynomial itself is new. Independently of this, we verify that the Galois group $G$ over $K=\mathbb Q(\sqrt{-7})$ is indeed $\PSU_3(\mathbb F_5)$ following a suggestion of Kl\"uners. First, we prove that $G$ is primitive. The Galois group of a specialization is a subgroup of $G$. Specializing at $t=2$ and reducing modulo a prime of norm $29$, we find an element $\sigma$ of order $7$ with cycle type $7^7 1$. Then $\sigma \in G$ has a unique fixed point and so fixes any block that contains it. Then the cardinality $s$ of this block has $s \mid 50$ and $s \equiv 1 \pmod{7}$, so $s=1,50$, hence $G$ is primitive. (See more generally Kl\"uners \cite[\S 3.3]{KluenersSubfields}.) Alternatively, we check that the field extension $K(f)/K(t)$ has no proper subfields. Now there are $9$ primitive permutation groups of degree $50$. We eliminate the smaller groups $H$ by finding a conjugacy class that belongs to $\PSU_3(\mathbb F_5)$ that does not belong to $G$; alternately, using the algorithm of Fieker--Kl\"uners \cite{FiekerKlueners}, we compute that the Galois group of the specialization at $t=2$ is $\PSU_3(\mathbb F_5)$. Using the classification of primitive subgroups of $S_{50}$, this leaves $4$ possibilities: $\PSU_3(\mathbb F_5)$, $\PSU_3(\mathbb F_5):2$, $A_{50}$, or $S_{50}$. The discriminant of $f(t)$ is \[ \disc(f(t)) = \frac{5^{560} 7^{1092}}{2^{1918}} t^{36}(t-1)^{20}, \] which is a square, so immediately $G \leq A_{50}$. To rule out $A_{50}$ (and again $S_{50}$ at the same time), we show that $G \leq S_{50}$ is not $2$-transitive: the polynomial of degree $50\cdot 49$ with roots $x_i-x_j$ with $i \neq j$, where $x_i$ are the roots of $f(x)$---computed symbolically using resolvents---factors over $\mathbb Q(t)$. Thus $G \leq \PSU_3(\mathbb F_5):2$. (Since the prime $p=53 \nmid \#S_{50}$ is a prime of good reduction for the cover by a theorem of Beckmann \cite{Beckmann}, it is enough to compute this factorization over $\mathbb F_p$.) Computing another relative resolvent to distinguish between $\PSU_3(\mathbb F_5)$ and $\PSU_3(\mathbb F_5):2$, we verify that $G = \PSU_3(\mathbb F_5)$; alternatively, we construct the fixed field of $\PSU_3(\mathbb F_5)$ in the Galois extension over $\mathbb Q$ and find that it is equal to $K$. To conclude with the remaining examples at the other extreme, for the refined passport of size $6$ with orders $(4,2,10)$, we repeat these steps with all $6$ triples, and in about $3$ CPU hours we have computed computed the result to $1000$ digits of precision. Computing the minimal polynomial of each coefficient, we find a Bely\u{\i}\ map defined over the number field $L$ with defining polynomial \[ x^{12} - x^{11} - 6x^{10} - 10x^9 - 15x^8 - 16x^7 - 19x^6 - 24x^5 - 35x^4 - 30x^3 - 6x^2 - 9x - 9. \] The field $L$ is a quadratic extension of the field $K=\mathbb Q(\sqrt{5},\sqrt[3]{10})$ and $L$ (like $K$) is ramified only at $2,3,5$. We were not able to find a descent of this map to $K$, and we guess that there might be an obstruction of some kind, for example, the curve $X(\Gamma)$ may be a conic defined over $K$ that is not $K$-isomorphic to $\mathbb P^1$. For more on the issue of descent, see Sijsling--Voight \cite[\S 7]{SijslingVoight} and the references therein. \end{exm} \subsection*{Verification} Once a Bely\u{\i}\ map has been computed, it remains to verify that it is in fact correct. There are a number of methods to achieve this: see the survey by Sijsling--Voight \cite{SijslingVoight}. In our situation, it is enough to verify that the cover computed is a three-point cover; this can be checked by a discriminant calculation. Once this has been checked, the way that the cover was constructed guarantees that it has the correct monodromy, computed as they were to sufficient precision to separate triangles. This verifies that the cover is correct over $\mathbb C(t)$ (and thus over $\overline{\mathbb Q}(t)$ and identifies the embedding of the number field yielding the specific monodromy triple). Example \ref{exm:PSU3(5)} was also independently verified over $\mathbb C(t)$ using code due to Bartholdi, Buff, Graf von Bothmer, and Kr\"oker \cite{Bartholdi}. \begin{rmk} \label{rmk:makerigorous} At the present time, our method is not rigorous, and we have not analyzed its running time. However, our computations are quite successful in practice, and we believe that it may be possible to furnish this rigorous analysis. A bound on the height of a Bely\u{\i}\ map, in the context of Arakelov geometry, has been established by Javanpeykar \cite{Javanpeykar}, and so it would suffice to establish that the linear system yielding power series expansions of modular forms is well-conditioned. In this vein, it would also be helpful to establish the integrality of coefficients of the power series expansions under a suitable choice of $\Theta$, as indicated in Remark \ref{rmk:Shimura}. We leave these investigations for future work. \end{rmk}
\section{Introduction} \label{s:intro} The sheer volume of networked information, in conjunction with the complexity and polymorphous nature of modern cyberthreats, calls for scalable, accurate, and, {\em at the same time}, flexible and programmable monitoring systems \cite{deri10, dain12}. The challenge is to {\em promptly} react to fastly mutating needs by deploying {\em custom traffic analyses}, capable of tracking event chains and multi-stage attacks, and efficiently handle the many heterogeneous features, events, and conditions which characterize an operational failure, a network application mis-behavior, an anomaly or an incoming attack. Such needed level of flexibility and programmability should address scalability {\em by design}, through systematic exploitation of stream-based analysis techniques. And, even more challenging, traffic analyses and mitigation primitives should be ideally brought {\em inside} the monitoring probes themselves at data plane, so as to avoid exporting traffic data to central analysis points, an hardly adequate way to cope with the traffic scale and the strict (ideally real time) mitigation delay requirements. \begin{figure*}[t] \centering \includegraphics[width=17cm, height=3cm]{figs/abstraction.pdf} \caption{StreaMon data plane identification/measurement/decision abstraction, and its mapping to implementation-specific workflow tasks} \vspace*{-.5cm} \label{fig:logic-wf} \end{figure*} To face this rapidly evolving scenario, in this paper we propose StreaMon, a data-plane programming abstraction for stream-based monitoring tasks directly running over network probes. StreaMon is devised as a pragmatic tradeoff between full programmability and vendors' need to keep their platforms {\em closed}. StreaMon's strategy closely resembles that pioneered by Openflow \cite{OF08} in the abstraction of networking functionalities, thus paving the road towards software-defined networking\footnote{ And in fact, even if the focus of this paper is (as a first step) only on the data plane interface itself, we believe that StreaMon could eventually candidate to play the role of a possible southbound data-plane interface for Software-Defined Monitoring frameworks, devised to translate high level monitoring tasks into a sequence of low-level traffic analysis programs distributed into network-wide StreaMon probes.}. However, the analogy with Openflow limits to the strategic level; in its technical design, StreaMon significantly departs from Openflow for the very simple reason that (as discussed in Section \ref{s:abstraction}) the data-plane programmability of monitoring tasks exhibits very different requirements with respect to the data-plane programmability of networking functionalities, and thus mandate for different programming abstractions. \\ The actual contribution of this paper is threefold. {\bf (1)} We identify (and design an execution platform for) an extremely simple abstraction which appears capable of supporting a wide range of monitoring application requirements. The proposed API decouples the monitoring application ``logic'', externally provided by a third party programmer via (easy to code) eXtended Finite State Machines (XFSM), from the actual ``primitives'', namely configurable sketch-based measurement modules, d-left hash tables, state management primitives, and export/mitigation actions, hard-coded in the device. While our handling of sketch-based measurement primitives may recall \cite{opensk}, we radically differentiate from any other work we are aware of, in our proposal to inject monitoring logic in the form of XFSM devised to locally orchestrate and run-time adapt measurements to the tracked state for each monitored entity (flows, hosts, DNS names, etc), with no need to resort to an external controlling device. {\bf (2)} We implement two StreaMon platform prototypes, a full SW and a FPGA/SW integrated implementation. We functionally validate them with five use case examples (P2P traffic classification, Conficker botnet detection, packet entropy HW analysis, DDos detection, Port Knocking), not meant as stand-alone contributions, but rather selected to showcase the StreaMon's adaptability to different application requirements. {\bf (3)} We assess the performance of the proposed approach: even if the current prototype implementation is not primarily designed with performance requirements in mind, we show that it can already sustain traffic in the multi-gbps range even with several instantiated metrics (for example 2.315 Gbps of real world replayed traffic with 16 metrics, see section \ref{s:perf}); moreover, we show that scalability can be easily enhanced by offloading the SW implementation with HW accelerated metrics; in essence, it seems fair to say that scalability appears to be an {\em architectural} property of our proposed API, rather than a side effect of an efficient implementation. \section{StreaMon abstraction} \label{s:abstraction} Our strategy in devising an abstraction for deploying stream-based monitoring tasks over a (general-purpose) network monitoring probe is similar in spirit to that brought about by the designers of the Openflow \cite{OF08} match/action abstraction, for programming networking functionalities over a switching fabric. Indeed, we also aim at identifying a compromise between full programming flexibility, so as to adapt to the very diverse needs of monitoring application developers and permit a broad range of innovation, and consistency with the vendors' need for closed platforms. However, the requirements of monitoring applications appear largely different from that of a networking functionality, and this naturally drives towards a {\em different} pragmatic abstraction with respect to a match/action table. At least three important differences do emerge. First, the ``entity'' being monitored is not consistently associated with the same field (or set of fields) in the packet header. For instance, if the target is to detect whether an IP address (monitored entity) is a bot, and the chosen mechanism is to analyze if the percentage of DNS NXDomain replies (feature) is greater than a given threshold (condition), the flow key to use for accounting is the source IP address when the arriving packet is a DNS query (event), but becomes the destination IP address when the packet is a DNS response (a different event). Second, the type of analysis (and possibly the monitoring entity target) entailed by a monitoring application may change over time, dynamically adapting to the knowledge gathered so far. For instance, if an IP address exhibits a critical percentage of DNS NXDomain replies, hence a bot suspect, we may further track its TCP SYNACK/SYN ratio and determine whether horizontal network scans occur, so as to reinforce our suspicion. And we might then follow up by deriving even more in-depth features, e.g., based on deep packet inspection. But at the same time, we would like to avoid tracking {\em all} features for {\em all} the possible flows, as this would unnecessarily drain computational resources. Finally, activities associated to a monitoring task are not all associated to a matching functionality: only measurement and accounting tasks are. Rather, less frequent, but crucial, activities (such as forging and exporting alerts, changing states, setting a mitigation filtering rule, and so on) are based on {\em decisions} taken on what we learned so far, and which are hence triggered by {\em conditions} applied to the gathered features. Our proposed StreaMon abstraction, illustrated in figure \ref{fig:logic-wf}, appears capable to cope with such requirements (as more extensively shown with the use cases presented in section \ref{ss:usecases}). It comprises of three ``stages'', programmable by the monitoring application developer via external means (i.e. not accessing the internal probe platform implementation). {\bf (1)} The {\bf Identification} stage permits the programmer to specify what is the monitored entity (more precisely, deriving its primary key, i.e., a combination of packet fields that identify the entity) associated to an event triggered by the actual packet under arrival, as well as retrieve an eventually associated state\footnote{ Our implementation uses d-left hashes for O(1) complexity in managing states. Also, memory consumption can be greatly reduced by {\em opportunistically} storing only non-default (e.g. anomalous) states. See the illustrative experiment in Fig. \ref{fig:offloading} for a rough idea of the attainable saving in the Conficker detection example, use case \ref{ss:use-conficker}.}. {\bf (2)} The {\bf Measurement} stage permits the programmer to configure which information should be accounted. It integrates {\em hard-coded and efficiently implemented} hash-based measurement primitives (metric modules), fed by configurable packet/protocol fields, with externally programmed {\em features}, expressed as arbitrary arithmetic operations on the metric modules' output. {\bf (3)} The {\bf Decision} stage is the most novel aspect of our abstraction. It permits to take decisions programmed in the form of eXtended Finite State Machines (XFSM), i.e. check conditions associated to the current state and tested over the currently computed features, and trigger associated actions and/or state transitions. \begin{figure}[!t] \centering \includegraphics[width=.45\textwidth]{figs/offloading.pdf} \caption{Experimental analysis of number of NXDomain DNS responses received (a feature commonly used in Botnet detection, e.g., \cite{confickerc}). Data obtained using a trace of 155 minutes with 955 different hosts belonging to 36 different /24 subnetworks. (a): Number $F_1$ of DNS NXDomain responses received per host (430 IPs performing at least 1 DNS query, 401 of them don't receive any NXDomain response); (b): number of tcp packet per seconds for all the flows (pkt1) and only for the flows for which $F_1 >= 1$ (pkt2), 15 minutes sample; analysis time for whole pkt1 trace: 1179 seconds, versus 68 seconds for pkt2.} \vspace*{-.5cm} \label{fig:offloading} \end{figure} The obvious compromise in our proposed approach is that (as per the match/action Openflow primitives) new metrics or actions can only be added by directly implementing them over the monitoring devices, thus extending the device capabilities. However, even restricting to our actual StreaMon prototype, the subset of metrics we implemented appear reusable and sufficient to support features proposed in a meaningful set of literature applications - see Table \ref{tab:features} in Section \ref{s:metrics-features}. \section{StreaMon Processing Engine} \label{s:at-a-glance} \subsection{System components} The previously introduced abstraction can be concretely implemented by a stream processing engine whose architecture is depicted in Figure \ref{fig:arch_overview}. It consists of four modular layers descriptively organized into two subsystems, namely {\em Measurement subsystem} and {\em Logic subsystem}, detailed in sections \ref{s:metrics-features} and \ref{s:tracking}, respectively. \textbf{Event layer} - Such layer is in charge of parsing each raw captured packet, and match an {\em event} among those user-programmed via the StreaMon API. The matched event identifies a user-programmed \textit{primary key} which permits to retrieve an {\em eventually} stored state. The event layer is further in charge of supplementary technical tasks (see section \ref{s:tracking}), such as handling special timeout events, deriving further secondary keys, etc. \textbf{Metric layer} - StreaMon operates on a per-packet basis and does {\em not} store any (raw) traffic in a local database. The application programmer can instantiate a number of {\em metrics} derived by a basic common structure, implemented as computation/memory efficient multi-hash data structures (i.e., Bloom-type sketches), updated at every packet arrival. \textbf{Feature layer} - this layer permits to compute user-defined arithmetic functions over (one or more) metric outputs. Whereas metrics carry out the bulky task of accounting {\em basic} statistics in a scalable and computation/memory efficient manner, the features compute {\em derived} statistics tailored to the specific application needs, at no (noticeable) extra computational/memory cost. \textbf{Decision layer} - this final processing stage implements the actual application logic. This layer keeps a list of \textit{conditions} expressed as mathematical/logical functions of the feature vector provided by the previous layer and any other possible secondary status. Each condition will trigger a set of specified and pre-implemented \textit{actions} and a state \textit{transition}. \begin{figure}[!tb] \centering \includegraphics[width=.5\textwidth]{figs/Streamon_arch.pdf} \caption{StreaMon processing engine architecture} \label{fig:arch_overview} \vspace{-.5cm} \end{figure} \begin{figure}[t] \centering \includegraphics[width=.5\textwidth]{figs/xml.png} \caption{Excerpt of XML based StreaMon code showing: (i) metric element allocation, (ii) feature compositions, (iii) an event logic description. In particular this picture shows a timeout event handler, described in terms of metric operations, feature extractions, conditions, actions and state transition} \label{fig:xml} \end{figure} \subsubsection*{StreaMon's programming language} Application programmers describe their desired monitoring operations through an high-level XML-like language, which permits to specify custom (dynamic) states, configure measurement metrics, formalize when (e.g.in which state and for which event) and how (i.e. by performing which operations over available metrics and state information) to extract features, and under which conditions trigger relevant actions (e.g. send an alert or data to a central controller). We remark that a monitoring application formally specified using our XML description does not require to be {\em compiled} by application developers, but is run-time installed, thus significantly simplifying on-field deployment. Figure \ref{fig:xml} shows an excerpt of a StreaMon application code. \subsubsection*{HW acceleration} StreaMon allows the seamless integration of HW accelerated metrics, e.g. mandated by stringent performance requirements. This remains transparent to the application programmer, which can thus port the same application from a SW based probe to a dedicated HW platform {\em with no changes in the application program}. Seamless HW integration is technically accomplished by performing {\em all} metrics, parsing, and event matching HW-accelerated computations in a front-end (in our specific case, an FPGA), and by bringing the relevant results up to the user plane through a HW/SW interface, by appending the meta-data generated by the HW tasks to the packet. \subsection{Measurement Subsystem} \label{s:metrics-features} The StreaMon Measurement subsystem provides a fast, computation/memory efficient, set of n highly configurable built-in modules that allows a programmer to deploy and compute a wide range of traffic features in stream mode. \textbf{Multi-Hash Metric module (MH)} - From a high level point of view, StreaMon metric modules are functional blocks exporting a simple interface to update and retrieve metric values associated to a given key. Even though in principle such modules can be implemented with any compact data structure compatible to our proposed stream mode approach (and indeed the current architecture support pluggable user defined metric modules), several metrics of practical interest can be derived from a basic structure depicted in Figure \ref{fig:bf-metric} and extending the construction proposed in \cite{info10Bianchi}. It permits to count (or time-average, see below) {\em distinct} values (called variations in \cite{info10Bianchi}) associated to a same key; for example: the number of {\em distinct} IP destinations contacted by a same IP source, or the number of distinct DNS names associated to a same IP address. The MH module is implemented using Bloom filter extensions \cite{broder02, info10Bianchi, ccr11Bianchi}. Processing occurs in three stages: (i) \emph{extract} from the packet\footnote{ Even if, for simplicity of explanation, we account such an extraction to the MH module, for obvious performance reasons we perform packet parsing and the consequent extraction of the flowkeys DFK and MFK once for all, for all metrics, in the Event Layer. Indeed, different metrics may make usage of a same packet field.} two binary strings, \emph{Detector FlowKey} (DFK) and \emph{Monitor FlowKey} (MFK), that will be used in to query the subsequent filters; (ii) \emph{detect} whether the DFK has already appeared in a past time window (\emph{Variation Detector filter}, VD), and if this is the case, iii) \emph{account} to the MFK key some packet-related quantity (e.g. number of bytes, or simply add 1 if packet count is targeted) to the third stage's \emph{Variation Monitor} (VM) filter. The reader interested in the rationale behind such construction and in further design details may refer to \cite{info10Bianchi}. Finally note that, as a special case, the MH module can be configured to either i) provide an ordinary count, by disabling the VD stage, or ii) perform a match for checking whether some information associated to the incoming packet is first seen (in a given time window), by disabling the VM stage. \begin{figure}[!t] \centering \includegraphics[width=.45\textwidth]{figs/bf-metric.pdf} \vspace*{-.5cm} \caption{Metric module structure} \vspace*{-.5cm} \label{fig:bf-metric} \end{figure} \begin{table*}[t] \hfill{} \begin{footnotesize} \begin{tabular}{| l | p{3.7cm}| p{11.8cm}|} \hline \textbf{Paper}&\textbf{Description}&\textbf{Brief description of reference features supported by StreaMon}\\ \hline \cite{exposure}&Passive DNS anal. x malicious domain detection&(i) Short life and Access Ratio per domain; (ii) Multi-homing/multi-address per domain; (iii) AVG, STD, total number, CDF of the TTL for a given domain\\ \hline \cite{complex}&Traffic charact. via Complex Network modeling&CDF, STD, Max/Min of: (i) total number of endpoints and correspondent host per endpoint; (ii) bytes exchanged between two endpoints\\ \hline \cite{fastflux}&DNS analysis for fastflux domain detection&\textit{Arithmetic functions of:} (i) number of unique A records returned in all DNS lookups, (ii) number of A records in a DNS response, (iii) number of name server (NS) records in one single lookup, (iv) number of unique ASNs for each A record\\ \hline \cite{scidive}&Stateful VOIP attack detection&\textit{Stateful rules:} (i) RTP traffic after SIP BYE? (ii) source IP address changed in time window? (iii) RTP sequence numbers evolve correctly?\\ \hline \cite{ddos02}&DDOS detection through ICMP and L4 analysis&Sent/received TCP packet ratio; Sent/received for different ICMP messages; UDP bitrate; No. different UDP connections per destination; Connection packet count\\ \hline \cite{p2pSVM}&SVM-based method for P2P traffic classification&\textit{Per distinct host and port:} (i) ratio between TCP and UDP traffic; (ii) AVG traffic speed; (iii) ratio between TCP and UDP AVG packet length; \textit{Per distinct IP:} (i) ARP request count, (ii) ratio between sent/received {TCP, UDP} packet, (iii) ratio between \{TCP, UDP\} and total traffic; \textit{Per distinct port:} traffic duration\\ \hline \cite{snare}&Mail spammer detection via Network level analysis& (i) AVG/STD msg len in past 24h; (ii) AVG distance to 20 nearest IP neighbors of the sender; (iii) AVG/STD of geodesic distance between sender and recipient; (iv) Message body len; (v) AVG, STD and total no. different recipients\\ \hline \end{tabular} \end{footnotesize} \caption{} \label{tab:features} \end{table*} \textbf{Programming the Measurement subsystem} - StreaMon programmers can deploy two types of ``counting'' data structures: (i) an ordinary \emph{Count Sketch} that sums a value associated to the MFK (CBF) or (ii) a \emph{Time decaying Bloom Filter} (TBF) \cite{ccr11Bianchi} that performs a smoothed (exponentially decaying) average of a value associated to the MFK. At {\em deployment time}, an MH module is configured by: (i) enabling/disabling the VD or the VM; (ii) specifying the counting data structure in the VM; (iii) setting the MH parameters, i.e., number of hash functions, total memory allocated, swapping threshold \cite{info10Bianchi} or memory time window in the VD, TBF's smoothing parameter \cite{ccr11Bianchi}; (iv) chain multiple MH module (so as to update a metric only if a former one was updated). At {\em run time}, the programmer may dynamically change flowkeys and updating quantities. In particular, for each possible event and (if needed) flow status (i.e., XFSM entry), a list of Metric Operations (MOP) is defined, where a MOP is a set/get primitive that defines the flowkeys (as a packet fields combinations) and quantities to be monitored/accounted. Finally, a set of StreaMon metrics are combined into Features by simply defining mathematical functions $F_i=f_i(\overline{M})$, where $\overline{M}$ is a Metric vector. \textbf{What practical features can be supported?} - We believe that the above metrics fulfill the needs of a non negligible fraction of real world monitoring applications. Indeed, \cite{ccr11Bianchi} shows a concrete example of a real world application reimplemented by just using modules derived from our general MH module described above. Indeed, limiting to the StreaMon's measurement subsystem (we will significantly extend the framework in the next section), table \ref{tab:features} shows features considered in a number of works taken from a somewhat heterogeneous set (different targets, operating at different network layers, different classification approaches) which, according to our analysis, are readily deployed through suitable configuration of the above metrics/features. Indeed, most applications either require to track/match (1) features that are directly retrievable from a single packet and do not have memory of past values (e.g.: short life of a DNS domain, message body length, all the features in \cite{fastflux} except the first one); and/or (2) require counting or averaging over a various set of parameters (eventually uniquely accounted - e.g. variations), which are readily instantiated with an MH module\footnote{ For instance, the {\em number of bytes exchanged between two hosts} can be obtained by setting as VMK is the concatenation of source and destination IP address and as updating quantity the length field of the packet; similarly, unique counts such as the \emph{total number of distinct TTL for a given DNS domain} is obtained by setting as VDK the concatenation of Domain Name and TTL in the DNS response and as VMK is the Domain Name.}; and/or (3) require logical/mathematical combinations of different statistics, which is the goal of our Feature Layer. Traffic features not covered by the families listed above are those which require a stateful correlation of different flows status. Such metrics (like the ones exploited in \cite{scidive}) are supported by StreaMon but require the stateful framework described in Section \ref{s:tracking}. \subsection{Logic subsystem} \label{s:tracking} StreaMon's \emph{Logic Subsystem} is the result of the interwork between the {\em Event Layer} and the {\em Decision Layer}. It provides the application developer with the ability to \emph{customize the tracking logic} associated to a monitored entity {\em subject to specific user-defined conditions}, so as to provide a verdict and/or perform analyses beyond those provided by the Monitoring Subsystem. \textbf{Event Layer} - The \emph{Event Layer} generates the events triggering the StreaMon processing chain, identifies the monitored entity (primary key), and retrieves the specific event context (in particular the flow status). This layer is further composed of three functional component: (i) the \textit{capture engine} responsible for ``capturing'' the triggering event, either a ``packet arrival'' or a ``timeout expiration'';(ii) the \textit{timeout manager} in charge of keeping track of all registered timeouts and manage their expiration; (iii) the \textit{status table}, a (dleft hash) table storing the status associated to the processed flows - the primary key status - and the so-called \textit{secondary support table}, a table storing states not associated to the primary key but required by the application state machine to take some decision. The triggering event is associated to a \textit{primary key}, i.e. the monitored entity (flow, host, etc) under investigation for which the monitoring application will provide a final verdict, like ``infected'' or ``legitimate''. For intercepted packets, the primary key is a combination of packet fields parsed by the (extensible) protocol dissector implemented in this layer. For locally generated timeout expiration events, the primary key is retrieved from the timeout context. The primary key is used to retrieve the current status of the flow: no match in the state table is considered to be a \emph{default} state. If the primary key is not directly retrievable from the packet, and instead the flow status is related to some other flow previously processed, a secondary support table storing the references to the entries in the status table is used. Such secondary support table is called \emph{related status table}. For example, this can be the case of an application keeping track of SIP URIs, (i.e.: the primary key is the SIP user ID in SIP INVITE messages) and consider as suspicious all UDP packets related to a data connection with a suspicious SIP user. In case of UDP packets, neither of the packet fields can be used to retrieve the flow status. Instead, a secondary support table is used to keep a reference between the socket 5-tuple and the status entry of the associated SIP initiator user. If the application does not take into account the flow status, the primary extraction is skipped. \textbf{Decision Layer} - The Decision Layer is the last processing unit of the StreaMon architecture and receives the event context carrying an indication of the current flow status and a set of traffic features container. This layer keeps a list of \emph{Decision Entries} (DEs) defined as the 3-tuple \emph{(enabling Condition (C), Output actions (O), State Transition (ST))}. For each triggering event, and according to the current flow status, the decision layer verifies the enabling conditions and executes the actions and the state transition associated to the matched condition. Since secondary support tables can be updated, StreaMon support \emph{variable conditions}, i.e. in which the comparison operands may change during time. The first matched condition will trigger the execution of a action set (like DROP, ALERT, SET\_TIMEOUT etc..). \textbf{Programming the Logic Subsystem} - Programmers describe an XFSM specifying (i) states and triggering events; (ii) for each state, which metrics and features are updated and which auxiliary information is collected/processed (secondary table); (iii) which conditions trigger a state transition and the associated actions. \begin{table} \begin{center} \begin{footnotesize} \begin{tabular}{ | l | l | } \hline \multicolumn{2}{ |c| }{\textbf{Logic Subsystem built-in primitives}}\\ \hline \multirow{3}{*}{\textbf{Operators}} & \texttt{SUM, DIFF, DIV, MULT, MOD} \\ & \texttt{EQ, NEQ, LT, GT}, \texttt{SQRT, LOG, POW} \\ & \texttt{AND, OR, NOT, XOR} \\ \hline \multirow{3}{*}{\textbf{Actions}} & \texttt{SET\_TIMEOUT}, \texttt{UPDATE\_TIMEOUT}\\ & \texttt{SAVE\_TIMEOUT\_CTX}, \texttt{DROP}, \texttt{ALLOW}, \texttt{MARK} \\ & \texttt{NEXT\_STATUS}, \texttt{UPDATE\_TABLE}, \texttt{PRINT}, \texttt{EXPORT}\\ \hline \end{tabular} \end{footnotesize} \end{center} \vspace*{-.5cm} \caption{} \vspace*{-.5cm} \label{tab:prim} \end{table} To identify the triggering event, StreaMon keeps a list of user-defined ``event descriptors'', expressed as a 3-tuple \textit{(type, descriptor, primary key)}. For example, in case of intercepted packets, an event can be defined as: $(packet; (udp.dport==53); ip.src)$. Moreover, programmers are given primitives to define for each event and state: (i) a set of metric operations (MOP) as described in section \ref{s:metrics-features}; (ii) a set of conditions expressed as an arithmetic function of features and secondary support table values. For each condition, programmers define a set of built-in actions and a state transition. Table \ref{tab:prim} summarizes the logic subsystem supported condition operators and actions. \section{Simple Use case examples} \label{ss:usecases} \begin{figure*}[t] \centering \includegraphics[width=1.02\textwidth]{figs/usecases-xfsm.pdf} \caption{Application XFSMs: (a) Conficker use case; (b) DDOS use case} \label{fig:xfsms} \end{figure*} We use the following simple examples to highlight the flexibility of StreaMon in supporting heterogeneous features commonly found in real-world monitoring applications. The input data traces are obtained by properly merging a packet trace gathered from a regional Internet provider with either (i) real malicious traffic extracted from traces captured in our campus network (use case \ref{ss:use-p2p} and \ref{ss:use-conficker}) or (ii) \emph{synthetic} traces properly generated in our laboratories (use case \ref{ss:use-ddos}). \subsection{P2P traffic classification} \label{ss:use-p2p} This example shows how straightforward is the implementation of three transport layer traffic features described in \cite{kar04}, for detecting peer to peer protocols that use UDP and TCP connections. The (stateless) application considers the following packet events (which will ignore well known UDP and TCP ports.):\\ \begin{footnotesize} $E_1: if(ip.proto==UDP)\&\&(udp.port \neq 25, 53, 110, ...)$ \\ $E_2: if(ip.proto==TCP)\&\&(tcp.port \neq 25, 53, 110, ...)$ \\ \end{footnotesize} This application extracts the following traffic features $F_1=M_1 \& M_2$; $F_2=| M_3 - M_4|$ where:\\ $\{M_1, M_2\}$: VD enabled - return 1 for each IP src/dst pair which previously opened a \{UDP, TCP\} socket, 0 otherwise;\\ $\{M_3, M_4\}$: VD enabled and VM type CBF - count the number of different \{TCP source ports, hosts\} connected to the same destination IP address. Metrics are read/updated on the basis of the matched event, as follows: \begin{center} \begin{footnotesize} \begin{tabular}{ | l | l | } \hline $E_1$ & $E_2$ \\ \hline $set(M_1, ip.src|ip.dst)$ & $get(M_1, ip.src|ip.dst)$; \\ $get(M_2, ip.src|ip.dst)$ & $set(M_2, ip.src|ip.dst)$ \\ $get(M_3, ip.dst)$ & $set(M_3, tcp.sport|ip.dst, ip.dst)$ \\ $get(M_4, ip.dst)$ & $set(M_4, ip.src|ip.dst, ip.dst)$ \\ & $get(M_i, ip.src), i=3,4$ \\ \hline \end{tabular} \end{footnotesize} \end{center} The application detects a p2p client if the following condition holds after a transitory period: $(F_1==1\&\&( F_2 < 10))$. \subsection{Conficker botnet detection} \label{ss:use-conficker} Conficker is one of the largest Botnets found in recent years \cite{torpig}. A multi-step detection algorithm can attempt to track the two following (separated) phase: (1) a bot tries to contact the C\&C Server, and (2) a single bot tries to contact and infect other hosts. To contact the C\&C Server, infected hosts perform a huge number of DNS queries (with a high NXDomain error probability) to resolve randomly generated domains. In the \emph{infection} phase, every host tries to open a TCP connection to the ports 445 of random IPs. Our Conficker detector will use the following metrics (VD disabled and VM of type TBF ): number of total DNS queries per host ($M_1$), number of DNS NXDomain per host ($M_2$) and number of TCP SYN and SYNACK to/from port 445 (respectively $M_3$ and $M_4$). These metrics are combined into the following features: $F_1 = M_2/M_1$, $F_2=M_4/M_3$. For a DNS NXDomain response, the condition $F_1 > 0.25$ is checked. If the condition is true, the state of the actual flow changes to \emph{alert} and an \emph{event timeout} is set. In the alert state the application updates $M_3$ and $M_4$ and, when this timeout expires, these metrics are used to compute $F_2$ and to verify the related condition: if the condition is true, then the host is considered \emph{infected} and goes to the next state, otherwise it returns to the default state. The application XFSM is graphically described (with simplified syntax) in Figure \ref{fig:xfsms}.a. Figure \ref{fig:conficker} shows the trend of features used in this configuration, for an host infected by Conficker (A) and for a \emph{clean} host (B). In the first case, the value of $F_1$ is relatively high since the beginning of the monitoring (due to the presence of reverse DNS queries, easily filtered by the application); the value increases when the host starts to perform \emph{Conficker queries}. The value of $F_2$ instead is very low, clearly denoting a port scan. Also for the host B the presence of rDNS queries increases the value of $F_1$ and this involve a change of state, and the application starts to analyze TCP feature. However the $F_2$ value (nearly 100\% of SYNACKs are received) reveals that this is clearly not a network scan. Testing this configuration in a 90-hours trace with 53 different hosts in idle state, we obtained 100\% of detection (8 infected hosts detected) without false positive or false negative. \begin{figure}[t] \centering \includegraphics[width=.48\textwidth]{measurements/conficker.pdf} \caption{Conficker features temporal evolution} \vspace*{-.5cm} \label{fig:conficker} \end{figure} \subsection{DDOS detection and mitigation} \label{ss:use-ddos} In this section we sketch a simple algorithm which can be used as an initial base for detecting and mitigating DDOS attacks. The algorithm is driven by the number of SYN packets received by possible DDOS targets. The XFSM of this configuration is depicted in \ref{fig:xfsms}.b, and is governed by the following two events: \\ \begin{footnotesize} $E_1: if(ip.proto==TCP)\&\&(tcp.flags==SYN)$ \\ $E_2: timeout expired$ \\ \end{footnotesize} Metric $M_1$ (VD=off, VM=TBF) tracks the number of TCP SYN addressed to a same target in 60 seconds (with 240s TBF's memory). All external servers for which $F_1=M_1$ is under a given threshold (100 in this example) are in default state (because they are obviously not under attack and thus do not need an explicit status). When $F_1$ exceeds this threshold, the target goes in \emph{monitored state}, a timeout is set and the current value of $F_1$ value is stored into a secondary support table with key ip.dst. As soon as this timeout expires, the difference between the current feature value and the one stored is computed. If the condition $if(F_1(t) > 1.2 * F_1(t-1))$, holds for two consecutive times, i.e. the rate of TCP SYN has increased twice for more than 20\%, the flow goes into an \emph{attack} state. Our use case example mimics this mitigation strategy by considering legitimate all the source hosts that have already shown meaningful activity and contacted the server {\em before the actual emergence of the attack}, and thus likely dropping most of the TCP SYN having a spoofed source IP address. In our use case example, a second metric $M_2$ tracks the TCP SYN rate, smoothed over a chosen time window, at which which each host contacts each destination IP addresses ($M_2$ is configured with VD disabled and VM of type TBF with smoothing window 240s, and it is updated with MFK $(IP_{src}, IP_{dst})$. \begin{figure}[t] \centering \includegraphics[width=.48\textwidth]{measurements/ddos.pdf} \caption{DDOS temporal representation: (i) DDOS target host status (red curve), (ii) $F_1$ value for the DDOS target (blu curve) and (iii) $F_2$ for two different legitimate traffic sources connecting to the DDOS target server} \vspace*{-.5cm} \label{fig:DDOS} \end{figure} Figure \ref{fig:DDOS} shows the temporal evolution of the state of a server, in an experiment where we performed a DDoS attack, with spoofed TCP SYN packets, after about 140s. The figure also reports the measured TCP SYN rate, as well as the traffic generated by two hosts performing regular queries towards the server: one starting at time 0, and the other starting right in the middle of the attack. When the actual attack is detected, the mitigation strategy starts filtering traffic. Thanks to the second tracked metric, $M_2$, the user starting activities before the DDoS attack is not filtered; on the contrary, connections from sources not previously detected via the $M_2$ metric are blocked, as shown by the green curve. New connections can be accepted as soon the server leaves the \emph{attack} state (see the $F_2$ growth of the second host). \subsection{HW accelerated detection of non standard encrypted traffic} \label{s:entropy} This example shows how HW metrics can be integrated in {\em StreaMon}. Since one of the task performed during deep packet inspection is the collection of statistics on byte frequencies, offsets for common byte-values, packet information entropy, we illustrate an use case that is a simplified version of the approach described in \cite{entropy}, in which encrypted flows are detected by combining two traffic features: (i) the bit information entropy of a packet; (ii) the percentage of printable characters, $i.e.$ ASCII characters in the range $[32 \ldots 127]$. Both features are computationally demanding for a software implementation, whereas they are best delegated to an FPGA implementation (just 250 logic cells for our implementation). This use case implements a simple stateless StreaMon application with the following characteristics: (i) it considers as event the reception of a UDP or TCP packet with length greater than 100 bytes; (ii) the primary key is the socket 5-tuple of the packet; (iii) it makes use of two FPGA precomputed metrics: popcount (number of bit set to 1) $M_1$ and printable chars percentage $M_2$ (which is directly mapped into a StreaMon feature $F_1=M_1$); (v) the packet entropy Feature $F_2$ is obtained as $- \frac{1}{N} \sum_{i=0}^1 n_i \cdot (log(n_i)-log(N)$, where $N=len, n_0= N - M_1, n_1 = M_1$. An encrypted flow is detected if: $(F_2 < 0.75) \& (-3\sigma < F1- H_U(l)< 3\sigma)$ where $H_U(l)$ is the entropy of a packet with uniform distribution of 1 in the payload and length $l$ and $\sigma$ is the standard deviation. Figure \ref{fig:entropy} shows $M_1$ and $M_2$ to the simple case of HTTP (unencrypted flow) and SSL (encrypted flow) traffic, and as in \cite{entropy} justify the condition expressed above. \begin{figure}[t] \centering \includegraphics[width=.45\textwidth]{figs/entropy-use-case.png} \caption{(a) percentage of printable character as function of the percentage of packets (b) percentage of bit set to 1 as function of the percentage of packets} \vspace*{-.5cm} \label{fig:entropy} \end{figure} \subsection{Port knocking} \label{s:pknock} This use case shows how a stateful firewall application can be easily implemented with StreaMon. In particular, this use case implements a port knocking mechanism to enable SSH access. The application's XFSM is depicted in figure \ref{fig:portkn}. SSH access will be granted only to those clients guessing the correct port sequence 5000, 4000, 7000 (a short sequence only for presentation convenience). The application considers four events (TCP SYN received respectively on port 5000, 4000, 7000, other ports) and four states (default, 5000contacted, 4000contacted, allowed). A state transition is triggered only if the client (identified by the IP source address) contacts the expected port in the sequence. If after a state transition the next expected port is not contacted in five seconds, the flow status is rolled back to default. In addition, to avoid random port scanning for guessing the correct sequence, a metric $M_1$ (VD=ON, VM=TBF with smoothing window 5s) counting the SYN rate is used. If a port scan is detected ($if (F_1 = M_1) > 40$) the flow status is updated to "attack", the host is blocked for 2 minutes and an alert is generated. M1 is updated with $DFK = ip.src|ip.dst|tcp.dport$ and $MFK = ip.src$ when a SYN to an unexpected port is contacted. \section{Performance evaluation} \begin{figure}[t] \centering \includegraphics[width=.5\textwidth]{figs/portk.pdf} \vspace*{-.5cm} \caption{Port Knocking use case XFSM} \vspace*{-.5cm} \label{fig:portkn} \end{figure} \subsection{Implementation Overview} We experimented StreaMon on two platforms. A SW-only deployment leverages off-the-shelf NIC drivers, integrating the efficient PFQ packet capturing kernel module \cite{pfq}. To test HW acceleration, we also deployed StreaMon using a 10 gbps Combo FPGA card as packet source. In what follows, unless otherwise specified (specifically, section \ref{s:fpga} which deals with HW accelerated metrics implemented over the Combo FPGA card) we focus on the SW-only configuration. StreaMon start-up sequence is summarized as follows. The application program is given as input to a pre-processor script as a XML formatted textual file. The program is parsed, StreaMon process is configured and executed while in parallel the interpreted feature and condition expressions are transformed them into C++ code and build a "on-demand DLL"\footnote{ Note that this is an optimization step which is devised to on-the-fly compile (transparent to the programmer), the application code, so as to avoid a "feature/condition interpreter", which would have lowered the overall performance and would not have permitted line rate feature computation and condition verification. Since no recompilation of the StreaMon code is ever needed, but user defined features are integrated as a DLL, dynamic deployment of user programs is made possible. }; StreaMon implementation takes advantages from multicore platforms by implementing parallel processing chains as shown in the implementation architecture depicted in figure \ref{fig:multicore} (for the SW-only setup). The PFQ driver is used as packet filter and configured with a number of software queues equal to the number of cores. Traffic flows are dispatched from the physical queues of the NIC by the PFQ steering function. For each PFQ, a separate StreaMon chain is executed by a single thread with CPU affinity fixed to one of the available core and share the metric banks, status tables and timeout tables (the concurrent access to the shared data is protected by spinlocks). The throughput measurement has been performed on a Intel Xeon X5650 (2.67 GHz, 6 cores) Linux server with, 16 GB ram and Intel 82599eb 10Gbit optical interfaces. \begin{figure}[!t] \centering \includegraphics[width=.45\textwidth]{figs/multicore.pdf} \caption{StreaMon prototype implementation architecture} \label{fig:multicore} \end{figure} \subsection{Performance Analysis} \label{s:perf} \subsubsection{Measurement subsystem performance} Figure \ref{fig:metric-measurement} shows the processing throughput expressed as percentage of the maximum throughput (6,468 Gbps) obtained with one PFQ source block without the overhead of StreaMon, expressed as function of the number of metrics (1, 4, 8, 16, 24, 32) in case of (1, 2, 3, 6) CPU core parallel processing. The input test data is a small portion of a long trace captured within a local internet provider infrastructure and its parameters are summarized in the following table: \begin{center} \begin{footnotesize} \begin{tabular}{ | c | c | c | c |} \hline \textbf{pkt no.} & \textbf{AVG pkt len} & \textbf{Host no.} & \textbf{TX rate}\\ \hline 6320928 & 632.82 bytes & 31827 & 6.47 Gbps \\ \hline \end{tabular} \end{footnotesize} \end{center} where \textit{pkt len} and \textit{TX} are the average packet length and replayed transmission bitrate respectively. As expected, the throughput decreases with the number of metrics and grows with the number of cores (even though the growth is lower than expected due to the simple thread concurrency management adopted in this prototype). It is important to underline that such graph shows the worst metric computational scenario in which: (i) all metrics are sequentially updated and retrieved for each packet (single event and stateless logic) and are configured with both the VD (a BF pair) and the MV (DLEFT) enabled; (ii) all flow keys are different (n metrics, n * 2 flow keys). Nevertheless, the results are promising, as for example in case of 16 metrics, for which we observe the following average bitrate (Gbps): \begin{center} \begin{footnotesize} \begin{tabular}{ | c | c | c | c |} \hline \textbf{1 core} & \textbf{2 cores} & \textbf{3 cores} & \textbf{6 cores}\\ \hline 0.566 & 0.983 & 1.367 & 2.315 \\ \hline \end{tabular} \end{footnotesize} \end{center} \begin{figure}[!t] \centering \includegraphics[width=.45\textwidth]{measurements/metric.pdf} \caption{System throughput evaluation} \label{fig:metric-measurement} \end{figure} Without discussing here the advantages of a stateful pre-filtering (which will be shown in section \ref{ss:use-perf}), we want to underline that, in the typical case in which not all the metrics are updated for each packet and require distinct flow keys, we can experience a better relation between the processing throughput and the number of metrics. For this reason, we compared the average throughput obtained by executing the traffic classification algorithm described in \cite{complex} (the most computational demanding algorithm cited in \ref{tab:features}, in terms of number of distinct metrics and flow keys, and portion of processed traffic). The following table shows the average bitrate (in Gbps) obtained by \cite{complex} in case of TCP traffic classification (\cite{complex}: 6 metrics, 3 flowkeys) and the average bitrate of the worst case configuration (6 metrics and 12 flow keys). \begin{center} \begin{footnotesize} \begin{tabular}{ | c | c | c | c| } \hline & \textbf{1 core} & \textbf{2 cores} & \textbf{3 cores} \\ \hline \textbf{\cite{complex}} & 2.792 & 4.316 & 5.199 \\ \hline \textbf{worst case} & 1.243 & 2.812 & 2.008 \\ \hline \end{tabular} \end{footnotesize} \end{center} \begin{table*} \hfill{} \begin{center} \begin{footnotesize} \begin{tabular}{ | c | c | c | c | c | l |} \hline \textbf{case} & \textbf{pkts}. & \textbf{AVG plen [bytes]} & \textbf{Ms, FKs.}& \textbf{TX [Gbps]} & \textbf{RX [\# core: Gbps]}\\ \hline \ref{ss:use-p2p} & 7676958 & 404.97 & 8, 4 & 3.658 & 1: 1.956, 2: 2.954, 3: 3.654\\ \hline \ref{ss:use-conficker} & 11852112 & 329.77 & 4, 2 & 2.462 & 1: 2.462 \\ \hline \ref{ss:use-ddos} & 9387240 & 215.38 & 2, 2 & 2.067 & 1: 2.003, 2: 2.067\\ \hline \end{tabular} \caption{Use case trace parameters and throughput} \vspace*{-.5cm} \label{tab:usecases} \end{footnotesize} \end{center} \end{table*} \subsubsection{Use case performance evaluation} \label{ss:use-perf} Table \ref{tab:usecases} shows the details of the test traces used for three among the use cases presented in Section \ref{ss:usecases} and the their performance evaluation, where \emph{pkt. no} is the total number of packets, \emph{AVG plen} is the average packet length, \emph{M} is the total number of metrics, \emph{FK} is that the transmission bitrate is fixed by the injector probe and depends on the average packet length. As already underlined, the use case applications experience a better throughput with respect to the bitrate obtained with the same number of metrics in the worst case measurement showed in \ref{fig:metric-measurement}. Note that due to the different performances of the COMBO card processor and the limitation to single core processing\footnote{This is due the fact that the COMBO card capture driver does not allow to open multiple instances of the same communication channel toward the FPGA NIC}, the encrypted flow detection use case performance is omitted to avoid confusion, as it would present results not directly comparable with the remaining ones. \subsubsection{FPGA accelerated primitives} \label{s:fpga} As already discussed, our current StreaMon prototype supports seamless integration of HW precomputed primitives, and in particular we have implemented packet entropy computation (see section \ref{s:entropy}) and event parsing on a INVEA-TECH Combo FPGA card \cite{invea}. The interface between the Combo card and StreaMon core software is realized via a footer appended to the ethernet frame. The Combo card capture the packet traveling in the network under inspection, performs the required operation and transfer the resulting metrics and signals as a list of Type Len Value (TLV) to the PC hosting the capture card. This list of TLV is parsed by the Event Layer and will be available to the remaining layers. Our next step is to implement a full deep packet inspection module. To confirm its necessity, we have so far implemented a simplified content matching packet payload inspection primitive in both SW and HW. Figure \ref{fig:string_matching} compares the two implementations for a toy example of just 32 content strings, showing the severe performance degradation of the SW implementation even in such small scale example. At the same time, the figure shows that, by offloading content matching primitives to the HW front-end, our prototype achieves almost optimal throughput performance. \begin{figure}[!t] \centering \includegraphics[width=.4\textwidth]{measurements/hw-string.png} \caption{FPGA accelerated and SW string matching comparison} \vspace*{-.5cm} \label{fig:string_matching} \end{figure} \section{Related Work} In the literature, several monitoring platforms have targeted monitoring applications' programmability. A Monitoring API for programmable HW network adapters is proposed in \cite{trimintzios06dimapi}. On top of such probe, network administrators may implement custom C++ monitoring applications. One of the developed applications is Appmon \cite{antoniades06appmon}. It uses deep packet inspection to classify observed flows and attribute these flows to an application. Flow \emph{states} are stored in an hash table and retrieved when an \emph{old} flow is observed again. This way to handle states bears some resemblance with that proposed in this work, which however makes usage of (much) more descriptive eXtended Finite State Machines. CoMo \cite{ian04} is another well known network monitoring platform. We share with CoMo the (for us, side) idea of extensible plug-in metric modules, but besides this we are quite orthogonal to such work, as we rather focus on how to combine metrics with features and states using higher level programming techniques (versus CoMo's low level queries). Bro \cite{bro} provides a monitoring framework relying on event-based programming language for real time statistics and notification. Despite the attempt to define a versatile high level language, Bro is not designed to expose a clear and simple abstraction for monitoring application development and leave full programmability to its users (which we believe results in a more descriptive and yet more complex programming language). The Real-Time Communications Monitoring (RTCMon) framework \cite{rtcmon} permits development of monitoring applications, but again the development language is a low level one (C++), and (unlike us) any feature extraction and state handling must be dealt with inside the custom application logic developed by the programmer. CoralReef \cite{keys01coralarchitecture}, FLAME \cite{flame02} and Blockmon \cite{Block12} are other frameworks which grant full programmability by permitting the monitoring application developers to ``hook'' their custom C/C++/Perl traffic analysis function to the platform. On a different line, a number of monitoring frameworks are based on suitable extensions of Data Stream Management Systems (DSMS). PaQueT \cite{lig08paquet}, and more recently BackStreamDB \cite{Back12}, are programmable monitoring frameworks developed as an extension of the Borealis DSMS \cite{borealis05}. Ease of programming and high flexibility is provided by permitting users to define new \emph{traffic} metrics by simply performing queries to the DSMS. The DSMS is configured through an XML file that is processed to obtain a C++ application code. Gigascope \cite{cranor03gigascope} is another stream database for network monitoring that provide an architecture programmable via SQL-like queries. Opensketch \cite{opensk} is a recent work proposing an efficient generic data plane based on programmable metric sketches. If on the one hand we share with Opensketch the same measurement approach, on the other hand its data plane abstraction delegates any decision stage and logic adaptation the control plane and, with reference to our proposed abstraction, does not go beyond the functionalities of our proposed Measurement Subsystem. On the same line, ProgME \cite{progme} is a programmable measurement framework which revolves around the extended and more scalable notion of dynamic flowset composition, for which it provides a novel functional language. Even though equivalent dynamic tracking strategies might be deployed over Openflow based monitoring tools, by exploiting multiple tables, metadata and by delegating "monitoring intelligence" to external controllers, this approach would require to fully develop the specific application logic and to forward all packets to an external controller, (like in Openflow based monitoring tool Fresco \cite{fresco}), which will increase complexity and affect performance. Finally, while our work is, to the best of our knowledge, the first which exploits eXtended Finite State Machines (XFSM) for programming custom monitoring logic, we acknowledge that the idea of using XFSM as programming language for networking purposes was proposed in a completely different field (wireless MAC protocols programmability) by \cite{bia12}. \section{Conclusion} \label{s:conclusion} The StreaMon programmable monitoring framework described in this paper aims at making the deployment of monitoring applications as fast an easy as configuring a set of pre-established metrics and devising a state machine which orchestrates their operation while following the evolution of attacks and anomalies. Indeed, the major contribution of the paper is the design of a pragmatic application programming interface for developing stream-based monitoring tasks, which does not require programmers to access the monitoring device internals. Despite their simplicity, we believe that the wide range of features accounted in the proposed use cases suggest that StreaMon's flexibility can be exploited to develop and deploy several real world applications. \footnotesize \bibliographystyle{abbrv}
\section*{Acknowledgments}\end{small}} \newcommand\altaffilmark[1]{$^{#1}$} \newcommand\altaffiltext[1]{$^{#1}$} \voffset=-0.6in \title[Galaxies on FIRE: Feedback \&\ Inefficient Star Formation]{Galaxies on FIRE (Feedback In Realistic Environments): Stellar Feedback Explains Cosmologically Inefficient Star Formation \vspace{-0.5cm}} \vspace{-0.2cm} \author[Hopkins et al.]{ \parbox[t]{\textwidth}{ Philip F.~Hopkins\thanks{E-mail:<EMAIL>}\altaffilmark{1,2}, Du\v{s}an Kere\v{s}\altaffilmark{3}, Jos{\'e} O{\~{n}}orbe\altaffilmark{4}, Claude-Andr{\'e} Faucher-Gigu{\`e}re\altaffilmark{2,5}, Eliot Quataert\altaffilmark{2}, Norman Murray\altaffilmark{6,7}, \&\ James S.~Bullock\altaffilmark{4} } \vspace*{6pt} \\ \altaffiltext{1}{TAPIR, Mailcode 350-17, California Institute of Technology, Pasadena, CA 91125, USA} \\ \altaffiltext{2}{Department of Astronomy and Theoretical Astrophysics Center, University of California Berkeley, Berkeley, CA 94720} \\ \altaffiltext{3}{Department of Physics, Center for Astrophysics and Space Science, University of California at San Diego, 9500 Gilman Drive, La Jolla, CA 92093} \\ \altaffiltext{4}{Department of Physics and Astronomy, University of California at Irvine, Irvine, CA 92697, USA} \\ \altaffiltext{5}{Department of Physics and Astronomy and CIERA, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208, USA} \\ \altaffiltext{6}{Canadian Institute for Theoretical Astrophysics, 60 St.\ George Street, University of Toronto, ON M5S 3H8, Canada} \\ \altaffiltext{7}{Canada Research Chair in Astrophysics\vspace{-0.5cm}} \\ \vspace{-0.5cm} } \date{Submitted to MNRAS, November, 2013\vspace{-0.6cm}} \begin{document} \maketitle \label{firstpage} \vspace{-0.2cm} \begin{abstract} \vspace{-0.2cm} We present a series of high-resolution cosmological simulations$^{\ref{foot:movie}}$ of galaxy formation to $z=0$, spanning halo masses $\sim10^{8}-10^{13}\,M_{\sun}$, and stellar masses $\sim10^{4}-10^{11}\,M_{\sun}$. Our simulations include fully explicit treatment of the multi-phase ISM \&\ stellar feedback. The stellar feedback inputs (energy, momentum, mass, and metal fluxes) are taken directly from stellar population models. These sources of feedback, with {\em zero} adjusted parameters, reproduce the observed relation between stellar and halo mass up to $M_{\rm halo}\sim10^{12}\,M_{\sun}$. We predict weak redshift evolution in the $M_{\ast}-M_{\rm halo}$ relation, consistent with current constraints to $z>6$. We find that the $M_{\ast}-M_{\rm halo}$ relation is insensitive to numerical details, but is sensitive to feedback physics. Simulations with only supernova feedback fail to reproduce observed stellar masses, particularly in dwarf and high-redshift galaxies: radiative feedback (photo-heating and radiation pressure) is necessary to destroy GMCs and enable efficient coupling of later supernovae to the gas. Star formation rates agree well with the observed Kennicutt relation at all redshifts. The galaxy-averaged Kennicutt relation is very different from the numerically imposed law for converting gas into stars, and is determined by self-regulation via stellar feedback. Feedback reduces star formation rates and produces reservoirs of gas that lead to rising late-time star formation histories, significantly different from halo accretion histories. Feedback also produces large short-timescale variability in galactic SFRs, especially in dwarfs. These properties are not captured by common ``sub-grid'' wind models. \end{abstract} \begin{keywords} galaxies: formation --- galaxies: evolution --- galaxies: active --- stars: formation --- cosmology: theory\vspace{-0.5cm} \end{keywords} \vspace{-1.1cm} \section{Introduction} \label{sec:intro} \begin{figure*} \centering \begin{tabular}{cc} \includegraphics[width={0.487\textwidth}]{m12v_mr_Dec5_2013_3_s0120_t090_gas_N3c.png} & \hspace{-0.4cm} \includegraphics[width={0.49\textwidth}]{m12v_mr_Dec5_2013_3_s0440_t000_gas_N3c.png} \\ \end{tabular} \vspace{-0.1cm} \caption{Gas in a representative simulation of a Milky Way-mass halo ({\bf m12i} in Table~\ref{tbl:sims}). Image shows the projected gas density, log-weighted ($\sim4\,$dex stretch). Magenta shows cold molecular/atomic gas ($T<1000\,$K). Green shows warm ionized gas ($10^{4}\lesssim T \lesssim 10^{5}\,$K). Red shows hot gas ($T\gtrsim 10^{6}\,$K).$^{\ref{foot:gas.coloring}}$ Each image shows a box centered on the main galaxy. {\em Left:} Box $200\,$kpc (physical) on a side at high redshift. The galaxy has undergone a violent starburst, leading to strong outflows of hot and warm gas that have blown away much of the surrounding IGM (even outside the galaxy). Note that the ``filamentary'' structure of cool gas in the IGM is clearly affected by the outflows. {\em Right:} Near present-day, with a $\sim50\,$kpc box. A more relaxed, well-ordered disk has formed, with molecular gas tracing spiral structure, and a halo enriched by diffuse hot outflows. \label{fig:demo.image.1}} \end{figure*} \begin{figure} \centering \plotonesize{m12v_mr_Dec5_2013_3_s0440_t000_star_N3c.png}{1.01} \caption{Stars in the {\bf m12i} simulation at $z\sim0$, in a box $50\,$kpc on a side near present-time. Image is a mock $u/g/r$ composite. The disk is approximately face-on, and the spiral structure is visible. (The image uses {\small STARBURST99} to determine the SED of each star particle given its known age and metallicity, then ray-traces the line-of-sight flux following \citet{hopkins:lifetimes.letter}, attenuating with a MW-like reddening curve with constant dust-to-metals ratio for the abundances at each point.) \label{fig:demo.image.2}} \end{figure} \begin{figure} \centering \begin{tabular}{c} \includegraphics[width={0.9\columnwidth}]{m10_hr_Dec9_2013_s0021_t090_gas_N3c_r20.png} \\ \includegraphics[width={0.9\columnwidth}]{m10_hr_Dec9_2013_s0121_t090_gas_N3c.png} \\ \end{tabular} \caption{Gas, as Fig.~\ref{fig:demo.image.1}, for a dwarf galaxy ({\bf m10} in Table~\ref{tbl:sims}). {\em Top:} $40\,$kpc (physical) box, at high redshift. {\em Bottom:} $20\,$kpc box at intermediate redshift. Strong outflows are still present, though they are more spherical, because the galaxy halo is itself small and embedded {\em within} a much larger filament. \label{fig:demo.image.3}} \end{figure} It is well-known that feedback from stars is a critical, yet poorly-understood, component of galaxy formation. Within galaxies, star formation is observed to be inefficient in both an instantaneous and an integral sense. Instantaneously, the Kennicutt-Schmidt (KS) relation implies gas consumption timescales of $\sim50$ dynamical times \citep{kennicutt98}, while the total fraction of GMC mass converted into stars is only a few percent \citep{zuckerman:1974.gmc.constraints,williams:1997.gmc.prop,evans:1999.sf.gmc.review,evans:2009.sf.efficiencies.lifetimes}. Without strong stellar feedback, however, gas inside galaxies cools efficiently and collapses on a dynamical time, predicting order-unity star formation efficiencies on all scales \citep{hopkins:rad.pressure.sf.fb,tasker:2011.photoion.heating.gmc.evol,bournaud:2010.grav.turbulence.lmc,dobbs:2011.why.gmcs.unbound,krumholz:2011.rhd.starcluster.sim,harper-clark:2011.gmc.sims}. In an integral sense, without strong stellar feedback, gas in cosmological models cools rapidly and inevitably turns into stars, predicting galaxies with far larger masses than are observed \citep[e.g.][and references therein]{katz:treesph,somerville99:sam,cole:durham.sam.initial,springel:lcdm.sfh,keres:fb.constraints.from.cosmo.sims}. Decreasing the instantaneous star formation efficiency does not eliminate this integral problem: the amount of baryons in real galactic disks is much lower than that predicted in models absent strong feedback \citep[essentially, the Universal baryon budget; see][]{white:1991.galform,keres:fb.constraints.from.cosmo.sims}. Constraints from intergalactic medium (IGM) enrichment require that many of those baryons must have entered galaxy halos and disks at some point to be enriched, before being expelled \citep{aguirre:2001.igm.metal.evol.sims,pettini:2003.igm.metal.evol,songaila:2005.igm.metal.evol,martin:2010.metal.enriched.regions}. Galactic super-winds with mass-loading $\dot{M}_{\rm wind}$ of many times the star formation rate (SFR) are therefore generally required to reproduce observed galaxy properties \citep[e.g.][]{oppenheimer:outflow.enrichment}. Such winds have been observed ubiquitously in local and high-redshift star-forming galaxies \citep{martin99:outflow.vs.m,martin06:outflow.extend.origin,heckman:superwind.abs.kinematics,newman:z2.clump.winds,sato:2009.ulirg.outflows,chen:2010.local.outflow.properties,steidel:2010.outflow.kinematics,coil:2011.postsb.winds}. However, until recently, numerical simulations have been unable to produce winds with large-mass loading factors from an a priori model (let alone the correct scalings of wind mass-loading with galaxy mass or other properties), nor to simultaneously predict the instantaneous inefficiency of star formation within galaxies. This is particularly true of models which invoke only energetic feedback via supernovae (SNe), which is efficiently radiated in the dense gas where star formation actually occurs \citep[see e.g.][and references therein]{guo:2010.hod.constraints,powell:2010.sne.fb.weak.winds,brook:2010.low.ang.mom.outflows,nagamine:2010.dwarf.gal.cosmo.review,bournaud10}. More recent simulations, using higher resolution and invoking stronger feedback prescriptions, have seen strong winds, but have generally found it necessary to include simplified prescriptions for ``turning off cooling'' in the SNe-heated gas and/or include some adjustable parameters representing ``pre-SNe'' feedback \citep[see][]{governato:2010.dwarf.gal.form,maccio:2012.cuspcore.outflows,teyssier:2013.cuspcore.outflow,stinson:2013.new.early.stellar.fb.models,agertz:2013.new.stellar.fb.model}. This is physically motivated since feedback processes other than SNe -- protostellar jets, HII photoionization, stellar winds, and radiation pressure -- both occur and are critical in suppressing star formation in dense gas, as well as ``pre-processing'' gas prior to SNe explosions so that SNe occur at densities where thermal heating can have much larger effects \citep{evans:2009.sf.efficiencies.lifetimes,hopkins:rad.pressure.sf.fb,tasker:2011.photoion.heating.gmc.evol,lopez:2010.stellar.fb.30.dor,stinson:2013.new.early.stellar.fb.models,kannan:2013.early.fb.gives.good.highz.mgal.mhalo}. And in fact, there have been many studies with enormously higher resolution (enough to evolve each star explicitly) and a full treatment of the radiation-magnetohydrodynamics and time dependence of these multiple feedback mechanisms. Because of computational limitations, however, these have necessarily been restricted to very small systems, either single molecular clouds/star clusters \citep[e.g.][]{krumholz:2007.rhd.protostar.modes,krumholz:2011.rhd.starcluster.sim,offner:2009.rhd.lowmass.stars,offner:2011.rad.protostellar.outflows,harper-clark:2011.gmc.sims,bate:2012.rmhd.sims}, or the ``first stars'' \citep[e.g.][]{wise:2012.rad.pressure.effects,pawlik:2013.rad.feedback.first.stars,muratov:2013.popIII.star.feedback}. But these studies, {\em without exception}, have found that the non-linear interaction of the feedback mechanisms above -- especially the dual roles of HII photoionization and radiation pressure in concert with SNe -- is absolutely critical to explain the generation of large local outflows, the self-regulation of star formation, and the shape of the stellar initial mass function. Despite these breakthroughs, given limited resolution and the complexity of the baryonic physics, many cosmological models have treated galactic wind generation and the inefficiency of star formation in a tuneable, ``sub-grid'' fashion. This is not to say that the models have not tremendously improved our understanding of galaxy formation! They have demonstrated that stellar feedback can {\em plausibly} lead to (globally) inefficient star formation, constrained the parameter space of allowed feedback models, made predictions for the critical role of outflows and recycling in enriching the IGM, provided possible baryonic solutions to apparent dark matter ``problems'' \citep[e.g.][]{pontzen:2011.cusp.flattening.by.sne}, demonstrated the need for ``early'' feedback from radiative mechanisms beyond SNe alone, and generally created the framework for our interpretation of observations. However, with wind models often relying on adjustable parameters, the integrated efficiency of star formation in galaxies is to some extent tuned ``by hand'' and the predictive power is inherently limited. This is particularly true for studies of gas in the circum-galactic medium (CGM), a current area of much observational progress -- measurements of the CGM are sensitive to the phase structure of the gas, which is not faithfully represented in models which simply ``turn off'' hydrodynamics or cooling, or mimic strong feedback via pure thermal energy injection or ``particle kicks'' \citep[see e.g.][for an explicit demonstration of this]{hummels:2013.cgm.vs.obs}. Accurate treatment of star formation and galactic winds ultimately requires realistic treatment of the stellar feedback processes that maintain the multi-phase ISM. Motivated by this philosophy (and building on the studies with single-star resolution), in \citet{hopkins:rad.pressure.sf.fb} (Paper {\small I}) and \citet{hopkins:fb.ism.prop} (Paper {\small II}), we developed a new set of numerical models to follow stellar feedback on scales from sub-GMC star-forming regions through galaxies. These simulations include the energy, momentum, mass, and metal fluxes from stellar radiation pressure, HII photo-ionization and photo-electric heating, SNe Types I \&\ II, and stellar winds (O-star and AGB). Critically, the feedback is directly tied to the young stars, with the energetics and time-dependence taken from stellar evolution models. In our previous work, we showed, in isolated galaxy simulations, that these mechanisms produce a quasi-steady ISM in which GMCs form and disperse rapidly, with phase structure, turbulence, and disk and GMC properties in good agreement with observations \citep[for various comparisons, see][]{narayanan:2012.mw.x.factor,hopkins:dense.gas.tracers,hopkins:clumpy.disk.evol,hopkins:2013.accretion.doesnt.drive.turbulence}. In Paper {\small I}, \citet{hopkins:virial.sf}, and \citet{hopkins:stellar.fb.mergers} we showed that this leads naturally to ``instantaneously'' inefficient SF (predicting the KS-law), regulated self-consistently by feedback and {\em independent} of the numerical prescription for star formation in very dense gas. In \citet{hopkins:stellar.fb.winds} (Paper {\small III}) and \citet{hopkins:2013.merger.sb.fb.winds} we showed that the same feedback models reproduce the galactic winds invoked in previous semi-analytic and cosmological simulations, and that the {\em combination} of multiple feedback mechanisms is critical to produce massive, multi-phase galactic winds. However, our simulations have thus far been limited to idealized studies of isolated galaxies and galaxy mergers. These previous calculations thus cannot follow accretion from or interaction of outflows with the IGM, realistic galaxy merger histories, and many other important processes. In this paper, the first of a series, we present the FIRE (Feedback In Realistic Environments) simulations:\footnote{\label{foot:movie}Movies and summaries of key simulation properties are available at\\ {\scriptsize\url{http://www.tapir.caltech.edu/~phopkins/Site/Movies_cosmo.html}}\\ and the {\small FIRE} project website:\\ \url{http://fire.northwestern.edu}} a suite of fully cosmological ``zoom-in'' simulations developed to study the role of feedback in galaxy formation. To test the models and understand feedback in a wide range of environments, we study a wide range in galaxy halo and stellar mass (as opposed to focusing just on MW-like systems), and follow evolution fully to $z=0$. Our suite of calculations includes several of the highest-resolution galaxy formation simulations that have been run to $z=0$. Our simulations utilize a significantly improved numerical implementation of SPH (which has resolved historical discrepancies with grid codes), as well as the full physical models for feedback and ISM physics introduced and tested in Paper {\small I}-Paper {\small III}. Here, we explore the consequences of stellar feedback for the inefficiency of star formation, perhaps the most basic consequence of stellar feedback for galaxy formation. In companion papers, we will investigate the properties of outflows and their interactions with the IGM, the effect of those outflows on dark matter structure, the differences between numerical methods in treating feedback, the role of feedback in determining galaxy structure, and many other open questions. In \S~\ref{sec:ICs}-\ref{sec:sims}, we describe our methodology. \S~\ref{sec:ICs} describes the initial conditions for the simulations; \S~\ref{sec:sims:physics} outlines the implementation of the key baryonic physics of cooling, star formation, and feedback (a much more detailed description is given in Appendix~\ref{sec:appendix:algorithms}); \S~\ref{sec:sims} briefly describes the improvements in the numerical method compared to past work (again, more details are in Appendix~\ref{sec:appendix:sims}). And in Appendix~\ref{sec:appendix:test.isolated} we test and compare these algorithms with higher-resolution simulations of isolated (non-cosmological) galaxies. We describe our results in \S~\ref{sec:results}. We examine the predicted galaxy stellar masses (\S~\ref{sec:results:masses}), and how this depends on both numerical algorithms (\S~\ref{sec:results:numerics}) and feedback physics (\S~\ref{sec:results:feedback}), as well as how it compares to previous theoretical work (\S~\ref{sec:results:previous}). We show that the treatment of feedback physics overwhelmingly dominates these results, and discuss the distinct roles of multiple independent feedback mechanisms. We also explore the predictions for the Kennicutt-Schmidt relation (\S~\ref{sec:results:ks}), the shape of galaxy star formation histories (\S~\ref{sec:results:sfh}), the star formation ``main sequence'' (\S~\ref{sec:results:mainsequence}), and the ``burstiness'' of star formation (\S~\ref{sec:results:burstiness}). We summarize our important conclusions and discuss future work in \S~\ref{sec:discussion}. \vspace{-0.5cm} \section{Initial Conditions \&\ Galaxy Properties} \label{sec:ICs} The simulations presented here are a series of fully cosmological ``zoom-in'' simulations of galaxy formation; some images of the gas and stars in representative stages are shown in Figs.~\ref{fig:demo.image.1}-\ref{fig:demo.image.3}.\footnote{\label{foot:gas.coloring}Both gas and stellar images are true three-color volume renderings generated by ray-tracing lines of sight through the simulation (with every gas or star particle a source, respectively). For the stars, the physical luminosities and dust opacities in each band are used to generate the observed intensity map. For the gas, we construct synthetic ``bands'' where the particle emissivity is uniform if it falls within the temperature range specified, and zero otherwise, and the particle opacity is uniform across bands.} The technique is well-studied; briefly, a large cosmological box is simulated at low resolution to $z=0$, and then the mass within and around halos of interest at that time is identified, traced back to the starting redshift, and the Lagrangian region containing this mass is re-initialized at much higher resolution (with gas added) for the ultimate simulation \citep{porter:1985.cosmo.sim.zoom.outline,katz:1993.zoomin.technique}. We consider a series of systems with different masses. Table~\ref{tbl:sims} describes the initial conditions. All simulations begin at redshifts $\sim100-125$, with fluctuations evolved using perturbation theory up to that point.\footnote{Initial conditions were generated with the {\small MUSIC} code \citep{hahn:2011.music.code.paper}, using second-order Lagrangian perturbation theory.} \begin{footnotesize} \ctable[ caption={{\normalsize Simulation Initial Conditions}\label{tbl:sims}},center,star ]{lccccccl}{ \tnote[ ]{Parameters describing the initial conditions for our simulations (units are physical): \\ {\bf (1)} Name: Simulation designation. \\ {\bf (2)} $M_{\rm halo}^{0}$: Approximate mass of the $z=0$ ``main'' halo (most massive halo in the high-resolution region). \\ {\bf (3)} $m_{b}$: Initial baryonic (gas and star) particle mass in the high-resolution region, in our highest-resolution simulations. \\ {\bf (4)} $\epsilon_{b}$: Minimum baryonic gravity/force softening (minimum SPH smoothing lengths are comparable or smaller). Recall, force softenings are adaptive (mass resolution is fixed); for more details see Appendix~\ref{sec:appendix:sims}.\\ {\bf (3)} $m_{dm}$: Dark matter particle mass in the high-resolution region, in our highest-resolution simulations. \\ {\bf (4)} $\epsilon_{dm}$: Minimum dark matter force softening (fixed in physical units at all redshifts). } }{ \hline\hline \multicolumn{1}{c}{Name} & \multicolumn{1}{c}{$M_{\rm halo}^{0}$} & \multicolumn{1}{c}{$m_{b}$} & \multicolumn{1}{c}{$\epsilon_{b}$} & \multicolumn{1}{c}{$m_{dm}$} & \multicolumn{1}{c}{$\epsilon_{dm}$} & \multicolumn{1}{c}{Merger} & \multicolumn{1}{c}{Notes} \\ \multicolumn{1}{c}{\ } & \multicolumn{1}{c}{[$h^{-1}\,M_{\sun}$]} & \multicolumn{1}{c}{[$h^{-1}\,M_{\sun}$]} & \multicolumn{1}{c}{[$h^{-1}$pc]} & \multicolumn{1}{c}{[$h^{-1}\,M_{\sun}$]} & \multicolumn{1}{c}{[$h^{-1}$pc]} & \multicolumn{1}{c}{History} & \multicolumn{1}{c}{\ } \\ \hline {\bf m09} & 1.9e9 & 1.8e2 & 1.0 & 8.93e2 & 20 & normal & isolated dwarf \\ {\bf m10} & 0.8e10 & 1.8e2 & 2.0 & 8.93e2 & 20 & normal & isolated dwarf \\ {\bf m11} & 1e11 & 5.0e3 & 5.0 & 2.46e4 & 50 & quiescent & -- \\ {\bf m12v} & 5e11 & 2.7e4 & 7.0 & 1.38e5 & 100 & violent & several $z<2$ mergers \\ {\bf m12q} & 1e12 & 5.0e3 & 7.0 & 1.97e5 & 100 & late merger & -- \\ {\bf m12i} & 1e12 & 3.5e4 & 10 & 1.97e5 & 100 & normal & large ($\sim10\,R_{\rm vir}$) box \\ {\bf m13} & 1e13 & 2.6e5 & 15 & 1.58e6 & 150 & normal & ``small group'' mass \\ \hline\hline\\ } \end{footnotesize} \begin{figure} \centering \plotonesize{mg_mh_z0.pdf}{0.99} \vspace{-0.2cm} \caption{Galaxy stellar mass-halo mass relation at $z=0$. {\em Top}: $M_{\ast}(M_{\rm halo})$. {\em Bottom:} $M_{\ast}$ relative to the Universal baryon budget of the halo ($f_{b}\,M_{\rm halo}$). Each simulation (points) from Table~\ref{tbl:sims} is shown; large point denotes the most massive halo in each box. We compare the relation if all baryons became stars ($M_{\ast}=f_{b}\,M_{\rm halo}$; dotted) and the observationally inferred relationship as determined in \citet[][magenta]{moster:2013.abundance.matching.sfhs} and \citet[][cyan]{behroozi:2012.abundance.matching.sfhs} (dashed lines denote extrapolation beyond the observed range).$^{\ref{foot:moster.behroozi.halo.defns}}$ The agreement with observations is excellent at $M_{\rm halo}\lesssim10^{13}\,M_{\sun}$, including dwarf though MW-mass galaxies. We stress that there are {\em zero} adjusted parameters here: stellar feedback, with known mechanisms taken from stellar population models, is sufficient to explain galaxy stellar masses at/below $\sim L_{\ast}$. \label{fig:mg.mh.z0}} \end{figure} \begin{figure*} \centering \plotsidesize{mg_mh_allz.pdf}{0.99} \vspace{-0.2cm} \caption{$M_{\ast}-M_{\rm halo}$ relation as Fig.~\ref{fig:mg.mh.z0} (points follow the legend therein), at different redshifts. Observational constraints are also shown at each redshift (each pair of lines shows the $\pm1\,\sigma$ fit to the observations at that redshift). With no tuned parameters, the simulations predict $M_{\ast}-M_{\rm halo}$ and, by extension, the stellar MF and galaxy clustering, at all $z$. Redshift evolution in $M_{\ast}-M_{\rm halo}$ is weak, with the sense that low-mass dwarfs become higher-$M_{\ast}$, leading to a steeper faint-end galaxy MF, in agreement with constraints from reionization \citep[see][and references therein]{kuhlen:2012.reionization.escape.fractions}. \label{fig:mg.mh.z}} \end{figure*} The specific halos we re-simulate are chosen to represent a broad mass range and be ``typical'' in most properties (e.g.\ sizes, formation times, and merger histories) relative to other halos of the same $z=0$ mass. The simulations {\bf m09} and {\bf m10} are constructed using the methods from \citet{onorbe:2013.zoom.methods}; they are isolated dwarfs. Simulations {\bf m11}, {\bf m12q}, {\bf m12i}, and {\bf m13} are chosen to match a subset of initial conditions from the AGORA project \citep{kim:2013.AGORA}, which will enable future comparisons with a wide range of different codes. These are chosen to be somewhat quiescent merger histories, but lie well within the typical scatter in such histories at each mass (and each has several major mergers). Simulation {\bf m12v}, for contrast, is chosen to have a relatively violent merger history (several major mergers since $z\sim2$), and is based on the initial conditions studied in \citet{keres:cooling.clumps.from.broken.filaments} and \citet{faucher-giguere:2011.stream.covering.factor}. In each case, the resolution is scaled with the simulated mass, so as to achieve the optimal possible force and mass resolution. It is correspondingly possible to resolve much smaller structures in the low-mass galaxies. The critical point is that in all our simulations with mass $<10^{13}\,M_{\sun}$, we resolve the Jeans mass/length of gas in the galaxies, corresponding to the size/mass of massive molecular cloud complexes. This is necessary to resolve a genuine multi-phase ISM and for our ISM feedback physics to be meaningful. Fortunately, because most of the mass and star formation in GMCs in both observations \citep{evans:1999.sf.gmc.review,blitz:2004.gmc.mf.universal} and simulated systems (Paper {\small II}) is concentrated in the most massive GMCs, the resolution studies in Paper {\small I}-Paper {\small II}\ confirm that resolving small molecular clouds makes little difference. We refer interested readers to Paper {\small II}\ for a detailed discussion of the scales that must be resolved for feedback to operate appropriately, but note here that all our simulations are designed to be approximately comparable to the ``high-resolution'' simulations of isolated galaxies and the ISM in Paper {\small I}-Paper {\small II}, within the range of resolution where the results in those studies (star formation rates, wind outflow rates, GMC lifetimes, etc.) were numerically converged (unfortunately, it is not possible to evolve cosmological simulations to $z=0$ with the ``ultra-high'' sub-pc resolution therein). In terms of the Jeans mass/length of the galaxies, our resolution is broadly comparable between different simulations. Our worst resolution in units of the Jeans length/mass occurs in the more massive galaxies at late times, when they are relatively gas poor, and so (despite the large total galaxy mass) the Jeans length can become relatively small.\footnote{The approximate Jeans (GMC) mass/length for the $z=0$ disks, assuming Toomre $Q\sim1$, increases from $\sim10^{4}\,M_{\sun}$ ($\sim10-30\,$pc) in the $\lesssim10^{10}\,M_{\sun}$ halos to $\sim10^{7}\,M_{\sun}$ ($\sim100-200\,$pc) in the $\gtrsim10^{12}\,M_{\sun}$ halos. If $Q>1$, or if the gas fractions are higher (at higher redshifts), the Jeans masses/lengths are larger as well.} Every galaxy identified in this paper contains at least $\gg 10^{5}$ bound particles. We adopt a ``standard'' flat $\Lambda$CDM cosmology with $h\approx0.7$, $\Omega_{{\rm M}}=1-\Omega_{\Lambda}\approx0.27$, and $\Omega_{b}\approx0.046$ for all runs.\footnote{Because of our choice to match some of our ICs to widely-used examples for numerical comparisons, they feature very small cosmological parameter differences. These are percent-level, smaller than the observational uncertainties in the relevant quantities \citep{planck:2013.cosmological.params} and produce negligible effects compared to differences between randomly chosen halos.} \vspace{-0.5cm} \section{Baryonic Physics} \label{sec:sims:physics} The simulations here use the physical models for star formation and stellar feedback developed and presented in a series of papers studying isolated galaxies \citep{hopkins:stellar.fb.winds,hopkins:stellar.fb.mergers,hopkins:dense.gas.tracers,hopkins:clumpy.disk.evol}, adapted for fully cosmological simulations. We summarize their properties below, but refer to Appendix~\ref{sec:appendix:algorithms} for a more detailed explanation and list of improvements. Readers interested in further details (including resolution studies and a range of tests of the specific numerical methodology) should see Paper {\small I}\ \&\ Paper {\small II}. \vspace{-0.5cm} \subsection{Cooling} \label{sec:sims:cooling} Gas follows an ionized+atomic+molecular cooling curve from $10-10^{10}\,$K, including metallicity-dependent fine-structure and molecular cooling at low temperatures, and high-temperature ($\gtrsim10^{4}\,$K) metal-line cooling followed species-by-species for 11 separately tracked species. At all times, we tabulate the appropriate ionization states and cooling rates from a compilation of {\small CLOUDY} runs, including the effect of the photo-ionizing background, accounting for gas self-shielding. Photo-ionization and photo-electric heating from local sources are accounted for as described below. \vspace{-0.5cm} \subsection{Star Formation} \label{sec:sims:starformation} Star formation is allowed only in dense, molecular, self-gravitating regions above $n>n_{\rm crit}$ ($n_{\rm crit} = 100\,{\rm cm^{-3}}$ for our primary runs, but we also tested from $\sim10-1000\,{\rm cm^{-3}}$). This threshold is much higher than that adopted in most ``zoom-in'' simulations of galaxy formation (the high value allows us to capture highly clustered star formation). We follow \citet{krumholz:2011.molecular.prescription} to calculate the molecular fraction $f_{\rm H_{2}}$ in dense gas as a function of local column density and metallicity, and allow SF only from molecular gas. We also follow \citet{hopkins:virial.sf} and restrict star formation to gas which is locally self-gravitating, i.e.\ has $\alpha\equiv \delta v^{2}\,\delta r/G\,m_{\rm gas}(<\delta r) < 1$ on the smallest available scale ($\delta r$ being our force softening or smoothing length). This forms stars at a rate $\dot{\rho}_{\ast}=\rho_{\rm mol}/t_{\rm ff}$ (i.e.\ $100\%$ efficiency per free-fall time); so that the galaxy and even kpc-scale star formation efficiency is {\em not} set by hand, but regulated by feedback (typically at much lower values). Because of this, in Paper {\small I}, Paper {\small II}, and \citet{hopkins:virial.sf} we show that the galaxy structure and SFR are basically independent of the small-scale SF law, density threshold (provided it is high), and treatment of molecular chemistry. \vspace{-0.5cm} \subsection{Stellar Feedback} \label{sec:sims:feedback} Once stars form, their feedback effects are included from several sources. Every star particle is treated as a single stellar population, with a known age, metallicity, and mass. Then all feedback quantities (the stellar luminosity, spectral shape, SNe rates, stellar wind mechanical luminosities, metal yields, etc.) are tabulated as a function of time directly from the stellar population models in STARBURST99, assuming a \citet{kroupa:imf} IMF. (1) {\bf Radiation Pressure:} Gas illuminated by stars feels a momentum flux $\dot{P}_{\rm rad} \approx (1-\exp{(-\tau_{\rm UV/optical})})\,(1+\tau_{\rm IR})\,L_{\rm incident}/c$ along the optical depth gradient, where $1+\tau_{\rm IR} = 1+\Sigma_{\rm gas}\,\kappa_{\rm IR}$ accounts for the absorption of the initial UV/optical flux and multiple scatterings of the re-emitted IR flux if the region between star and gas particle is optically thick in the IR (see Appendix~\ref{sec:appendix:algorithms}). We assume that the opacities scale linearly with gas metallicity.\footnote{There has been some debate in the literature regarding whether or not the full $\tau_{\rm IR}$ ``boost'' applies to the infrared radiation pressure when $\tau_{\rm IR}\gg1$ (see e.g.\ \citealt{krumholz:2012.rad.pressure.rt.instab}, but also \citealt{kuiper:2012.rad.pressure.outflow.vs.rt.method} and \citet{davis:2014.rad.pressure.outflows}, who find much stronger effects in the infrared). We have considered alternatives, discussed in Paper {\small I}. However, in the simulations here we never resolve the extremely high densities where $\tau_{\rm IR}\gtrsim1$ (where this distinction is important), and so if anything are {\em under}-estimating the IR radiation pressure, even compared to the most conservative studies.} (2) {\bf Supernovae:} We tabulate the SNe Type-I and Type-II rates from \citet{mannucci:2006.snIa.rates} and STARBURST99, respectively, as a function of age and metallicity for all star particles and stochastically determine at each timestep if an individual SNe occurs. If so, the appropriate mechanical luminosity and ejecta momentum is injected as thermal energy and radial momentum in the gas within a smoothing length of the star particle, along with the relevant mass and metal yield (for all followed species). When the Sedov-Taylor phase is not fully resolved, we account for the work done by hot gas inside the unresolved cooling radius (converting the appropriate fraction of the SNe energy into momentum). We discuss this in detail in Appendix~\ref{sec:appendix:algorithms}, but emphasize that it is particularly important that SNe momentum not be neglected in massive halos whose mass resolution $\sim10^{4}\,M_{\sun}$ is much larger than the ejecta mass of a single SNe. (3) {\bf Stellar Winds:} Similarly, stellar winds are assumed to shock locally and so we inject the appropriate tabulated mechanical power $L(t,\,Z)$, wind momentum, mass, and metal yields, as a continuous function of age and metallicity into the gas within a smoothing length of the star particles. The integrated mass fraction recycled in winds (including both fast winds from young stars and slow AGB winds) and SNe is $\sim0.3$. (4) {\bf Photo-Ionization and Photo-Electric Heating:} Knowing the ionizing photon flux from each star particle, we ionize each neighboring neutral gas particle (provided there are sufficient photons, given the gas density, metallicity, and prior ionization state), moving outwards until the photon budget is exhausted; this alters the heating and cooling rates appropriately. The UV fluxes are also used to determine photo-electric heating rates following \citet{wolfire:1995.neutral.ism.phases}. Extensive numerical tests of the feedback models are presented in Paper {\small II}. \vspace{-0.5cm} \section{Simulation Numerical Details} \label{sec:sims} All simulations are run using a newly developed version of TreeSPH which we refer to as ``{\small P-SPH}'' \citep{hopkins:lagrangian.pressure.sph}, in the code {\small GIZMO}.\footnote{Details of the {\small GIZMO} code, together with a limited public version, user's guide, movies and test problem examples, are available at\\ \url{www.tapir.caltech.edu/~phopkins/Site/GIZMO.html}} This adopts the Lagrangian ``pressure-entropy'' formulation of the SPH equations developed in \citet{hopkins:lagrangian.pressure.sph}; this eliminates the major differences between SPH, moving mesh, and grid (adaptive mesh) codes, and resolves the well-known issues with fluid mixing instabilities in previously-used forms of SPH \citep[e.g.][]{agertz:2007.sph.grid.mixing,sijacki:2011.gadget.arepo.hydro.tests}. The gravity solver is a heavily modified version of the {\small GADGET-3} code \citep{springel:gadget}; but {\small GIZMO} also includes substantial improvements in the artificial viscosity, entropy diffusion, adaptive timestepping, smoothing kernel, and gravitational softening algorithm, as compared to the ``previous generation.'' These are all described in detail in Appendix~\ref{sec:appendix:sims}. We emphasize that our version of SPH has been tested extensively and found to give good agreement with analytic solutions as well as well-tested grid codes on a broad suite of test problems. Many of these are presented in \citet{hopkins:lagrangian.pressure.sph}. This includes Sod shock tubes; Sedov blastwaves; wind tunnel tests (radiative and adiabatic, up to Mach $\sim10^{4}$); linear sound wave propagation; oscillating polytropes; hydrostatic equilibrium ``deformation''/surface tension tests \citep{saitoh:2012.dens.indep.sph}; Kelvin-Helmholtz and Rayleigh-Taylor instabilities; the ``blob test'' \citep{agertz:2007.sph.grid.mixing}; super-sonic and sub-sonic turbulence tests (from Mach $\sim0.1-10^{3}$); Keplerian gas ring, disk shear, and shearing shock tests \citep{cullen:2010.inviscid.sph}; the Evrard test; the Gresho-Chan vortex; spherical collapse tests; and non-linear galaxy formation tests. For additional tests showing the improvements relative to previous-generation SPH, see \citet{hu:2014.psph.galaxy.tests}. Since it is critical for the problems addressed here that a code be able to handle high dynamic range situations, the numerical method and parameters such as SPH ``neighbor number'' were not modified for these tests individually, but are similar to what we use in our production runs in this paper. In Appendix~\ref{sec:appendix:sims}, we note that we have explicitly tested many of the purely numerical elements of the gravity and hydrodynamic solvers in the simulations shown here: for example, whether to use adaptive or fixed gravitational softenings, the choice of SPH smoothing kernel, and the timestepping algorithm. However we do not discuss these in the main text because they produce extremely small ($\lesssim 10\%$-level) differences in the quantities plotted in this paper. \vspace{-0.5cm} \section{Results} \label{sec:results} \subsection{Galaxy Masses as a Function of Redshift} \label{sec:results:masses} Fig.~\ref{fig:mg.mh.z0} plots the $z=0$ stellar mass-halo mass relation for our main set of simulations from Table~\ref{tbl:sims} (highest-resolution, with all physics enabled). Note that although each high-resolution region at $z=0$ contains one ``primary'' halo (the focus of that region), there are several smaller-mass, independent halos also in that region. We therefore identify and plot all such halos.\footnote{We use the {\small HOP} halo finder \citep{eisenstein:1998.hop.halo.finder} to automatically identify halos (which combines an iterative overdensity identification with a saddle density threshold criterion to merge subhalos and overlapping halos). Halo masses are defined as the mass within a spherical aperture about the density maximum with mean density $>200$ times the critical density at each redshift (this is chosen to be similar to the choice used in abundance-matching models, which define the observations to which we compare). Stellar mass plotted is the total stellar mass within $20\,$kpc of the center of the central galaxy in the halo (we do not include satellite galaxy masses). However, we have compared with the results of a basic friends-of-finds routine or simple by-eye identification, and find that for the results here (focused on simple, integral halo quantities, and ignoring subhalos), this makes no significant difference. Likewise defining the mass within $\sim0.1\,R_{\rm vir}$ instead of $20\,$kpc makes no significant difference.} We exclude those halos that are outside the high-resolution region (more than $1\%$ mass-contaminated by low-resolution particles; although varying this between $0.5-10\%$ makes little difference to our comparisons here) or insufficiently resolved ($<0.01$ times the primary halo mass, or with $<10^{5}$ dark matter particles). We also exclude subhalos/satellite galaxies. The known sources of stellar feedback we include, with {\em no} adjustment, automatically reproduce a relation between galaxy stellar and halo mass consistent with the observations\footnote{\label{foot:moster.behroozi.halo.defns}Note that \citet{behroozi:2012.abundance.matching.sfhs} and \citet{moster:2013.abundance.matching.sfhs} use definitions of halo mass which differ slightly (by $\approx10\%$). For our purposes, this produces negligible differences in our comparison.} from $M_{\rm halo}\sim 10^{7}-10^{13}\,M_{\sun}$. Specifically, the distribution of points for all $M_{\rm halo}\le 10^{12}\,M_{\sun}$ is statistically consistent (in a $\chi^{2}$ sense) with having been drawn from the \citet{moster:2013.abundance.matching.sfhs} curve; if we allow for the observed scatter ($\sim0.15\,$dex at $\sim 10^{12}\,M_{\sun}$, the width between the plotted lines, increasing to $\sim0.3\,$dex at the lowest observed masses) then all our primary galaxies lie within the $2\,\sigma$ scatter.\footnote{We can also fit the points here to the same power-law functional form used empirically: if we do so, the best-fit slope and normalization are both within $0.5\,\sigma$ of the fit to observations in \citet{moster:2013.abundance.matching.sfhs} (the error bar is dominated by the small-number statistics in our halo sampling). The simulations with $M_{\rm halo}\ll 10^{10}\,M_{\sun}$ are statistically inconsistent with the extrapolation of the flatter slope from \citet{behroozi:2012.abundance.matching.sfhs}, but this is entirely below the region actually observed, where \citet{behroozi:2012.abundance.matching.sfhs} and \citet{moster:2013.abundance.matching.sfhs} agree well, and there the simulations do not significantly ``prefer'' either fit.} Despite the fact that this relation implies a non-uniform (and even non-monotonic) efficiency of star formation as a function of galaxy mass, we do {\em not} need to invoke different physics or distinct parameters at different masses. This is particularly impressive at low masses, where the integrated stellar mass must be suppressed by factors of $\sim1000$ relative to the Universal baryon fraction. Unfortunately, at high masses ($>10^{13}\,M_{\sun}$), the large Lagrangian regions (hence large number of required particles) limit the resolution we can achieve; we have experimented with some low-resolution test runs which appear to produce overly massive galaxies, but higher-resolution studies are required to determine if that owes to a need for additional physics or simply poor numerical resolution. Interestingly, the scatter in $M_{\ast}$ at fixed $M_{\rm halo}$ may decrease weakly with mass, from $\sim0.5$\,dex in dwarf galaxies ($M_{\rm halo}\lesssim 10^{10}\,M_{\sun}$) to $\sim0.1-0.2$ dex in massive ($\sim L_{\ast})$ galaxies. But given the limited number of halos we study here, further investigation allowing more diverse merger/growth histories is needed. Fig.~\ref{fig:mg.mh.z} shows the $M_{\ast}-M_{\rm halo}$ relation at various redshifts. At each $z$, we compare with observationally constrained estimates of the $M_{\ast}-M_{\rm halo}$ relation. Implicitly, if they agree in $M_{\ast}(M_{\rm halo})$, our models are consistent with the observed stellar MF (given, of course, the limited statistics by which we are ``sampling'' the MF). At high redshifts, the halos we simulate are of course lower-mass, so eventually we have no high-mass galaxies; this limits the extent to which our results can be compared to observations above $z\sim2$. \vspace{-0.5cm} \subsection{Other (Basic) Galaxy Properties} We wish to focus here on galaxy masses and star formation histories. Companion papers (in preparation) will examine the galaxy morphologies and other observables in more detail. It is important, when studying those properties, to construct a meaningful comparison (e.g.\ using the same methods and wavelengths observed), and this is non-trivial. Moreover, it is by no means obvious that these properties are as robust to numerical details as the galaxy stellar masses (discussed further below), and it is completely outside the scope of this paper to fairly explore those dependencies. That said, we can briefly note the basic properties of the specific simulations in Table~\ref{tbl:sims} at $z=0$, with the caveat that these {\em may not be robust} to changes in either the initial conditions (the particular halo simulated) or our numerical methods. Morphologically, at $z=0$, run {\bf m09} resembles an ultrafaint dwarf; {\bf m10} a thick, but rotating dwarf irregular; and {\bf m11} a more ``fluffy'' dwarf spheroidal. Runs {\bf m12v}, {\bf m12q}, {\bf m12i} produce bulge+disk systems, with {\bf m12v} showing a prominent bulge at all times $z\lesssim2$; {\bf m12q} is more disk-dominated until a late major merger at $z< 0.5$ destroys the disk; and run {\bf m12i} produces a stellar disk with little bulge. Run {\bf m13} is totally bulge-dominated. Each galaxy has an approximately flat rotation curve outside of the central couple kpc; those with $M_{\ast}<10^{10}$ slowly rise with radius to $V_{\rm max}$, and the more massive systems are flat to within the central $\sim$\,kpc, except for {\bf m12v} (where the compact bulge leads to a central-kpc spike at $\sim 250\,{\rm km\,s^{-1}}$). The galaxy sizes, measured as the half-stellar mass effective radii, are $(0.3,\,0.52,\,3.5,\,2.8,\,3.2,\,4.2,\,4.8)\,{\rm kpc}$ for ({\bf m09, m10, m11, m12v, m12q, m12i, m13}), consistent with the observed stellar size-mass relation \citep{shen:size.mass,wolf:2010.disperson.gal.masses}. \vspace{-0.5cm} \subsection{(Lack of) Dependence on Numerical Methods} \label{sec:results:numerics} \begin{figure} \centering \plotonesize{mg_mh_fb.pdf}{1.01} \vspace{-0.5cm} \caption{$M_{\ast}-M_{\rm halo}$ relation at $z=0$, as Fig.~\ref{fig:mg.mh.z0}. {\em Top:} Simulations with different numerical parameters: we show the effects of varied resolution, artificial viscosity, and the algorithmic implementation of feedback. We also compare a completely different version of SPH (with a different set of hydrodynamic equations), which is known to differ significantly in certain idealized hydrodynamic test problems. These have little effect on our predictions. {\em Bottom:} Effect of physical variation in stellar feedback properties. We compare runs with no stellar feedback, with no supernovae (but stellar winds, radiation pressure, and photo-ionization heating included), or with no radiative feedback (radiation pressure and local HII-heating). ``No feedback'' runs generally predict $M_{\ast}\sim f_{b}\,M_{\rm halo}$, in severe conflict with the observations.$^{\ref{foot:nofb.lowmhalo}}$ Removing radiative {\em or} SNe feedback also produce order-of-magnitude too-large stellar masses. The non-linear {\em combination} of feedback mechanisms (not any one in isolation) is critical to drive winds and regulate galaxy masses. \label{fig:mg.mh.fb}} \end{figure} In Fig.~\ref{fig:mg.mh.fb} we investigate how the $M_{\ast}-M_{\rm halo}$ relation depends on numerical parameters and feedback. First we repeat Fig.~\ref{fig:mg.mh.z0} for simulations with different {\em purely numerical} parameters. These can and do, indeed, have significant quantitative effects -- they can easily shift the predicted stellar masses by factors $\sim2-3$. However, we stress that they do not {\em qualitatively} change our conclusions. Modest changes in resolution (our ``low-resolution'' runs correspond to one power of two step in spatial resolution, and a corresponding factor of $2^{3}=8$ change in mass resolution) lead to significant, but not order-of-magnitude, changes in $M_{\ast}$: generally we obtain larger $M_{\ast}$ by factors of $\sim1.5$ at high masses ($M_{\rm halo}\gtrsim 10^{11}\,M_{\sun}$) and $\sim2-3$ at the lowest masses ($M_{\rm halo}<10^{10}\,M_{\sun}$) at lower resolution, owing to a combination of (a) artificially enhanced mixing and thus cooling of diffuse gas, since ISM phases are less well-resolved, and (b) the fact that the coupling of feedback energy and momentum is necessarily spread over larger mass elements. If we downgrade our resolution more substantially -- by a factor of $\sim100$ in mass, or $>10$ in spatial scale (i.e.\ using the $>100\,$pc spatial resolution which is typical of most previous cosmological simulations), the results diverge more substantially: galaxy masses at $z\sim0$ are a factor of $\sim3-5$ higher at high masses and $\sim10$ higher at low masses. This makes sense, because at that resolution, we simply cannot meaningfully resolve even the most massive structures in the ISM.\footnote{We have run a couple tests with $30$ times higher particle numbers than our production-quality runs (for {\bf m12i} and {\bf m10}), to $z=2$, and found that the stellar masses at this time and earlier vary by $\sim 10-50\%$ from those quoted here. However, this appears to be primarily stochastic, rather than systematic, so we suspect the masses will not change much further at still higher resolution.} Some of our numerical tests are not plotted here because their effects are not significant. We have, for example, re-run several simulations with twice and five times larger dark matter softening lengths (same baryonic softening); using or de-activating adaptive gravitational softenings (which ensure there are always $\sim100$ neighbor particles in the softening kernel); varying the number of SPH ``neighbors'' in the hydrodynamic kernel and number of SPH particles to which energy and momentum are coupled; using a single timestep or Strang-split integration scheme in the code; varying the Courant factor of the hydrodynamic solver; changing the order of operator-splitting for the cooling and feedback steps; or forcing equal vs.\ allowing separate gravitational softenings for baryons and dark matter. These produce very small ($<10\%$) differences. We also varied the sizes of the high-resolution ``zoom-in'' Lagrangian regions of the halos; the results here are insensitive to the region size if we choose sizes $\gtrsim2\,R_{\rm vir}$ (at the redshift of interest), but the cooling of halo gas and star formation are artificially suppressed if the high-resolution region is much smaller. Changing the small-scale star formation prescription in the simulations has very little effect on our predictions. This is expected based on all of our previous studies using isolated (non-cosmological) simulations (for explicit examples where we vary the density threshold and instantaneous ``efficiency'' of star formation in dense gas by factors of $>1000$, as well as the density, temperature, chemical, and virial-state dependence of star formation, see Paper {\small I}, Paper {\small II}, and \citealt{hopkins:virial.sf}). Globally, star formation is {\em feedback-regulated}: a certain number of young stars are required to balance gravitational collapse/dissipation, independent of how those stars form. So long as cooling can proceed, they will form (what will change, via this self-regulation, if we change e.g.\ the density threshold above which stars form, is the amount of gas which ``builds up'' above that threshold; see \citealt{hopkins:dense.gas.tracers}). In Fig.~\ref{fig:mg.mh.fb} we show examples where we vary the density threshold for star formation by factors of $\sim100$ (producing $<0.2$\,dex random/non-systematic changes in stellar mass), or impose a much stricter local virial criterion for star formation (local virial parameter $<0.5$ instead of $1$; producing a $\sim20\%$ difference in stellar mass); we have also experimented with removing the virial parameter entirely or dividing the instantaneous efficiency of star formation in dense gas by a factor of $100$ (both produce $<10\%$ changes). We have also investigated different purely algorithmic methods for coupling the same feedback physics. Subtle differences in the algorithmic implementation of feedback have little {\em systematic} effect on the stellar mass, provided the same mechanisms are included; however they can only be compared statistically, since the stochastic nature of feedback means that even very subtle changes can produce significant differences in the exact time history of bursts, for example. Of what we have considered, the most important parameter is how we implement the momentum gained during the Sedov-Taylor phase of SNe remnant expansion when the cooling radius is unresolved (see Appendix~\ref{sec:appendix:algorithms}). For example, one of the ``mod.\ SNe coupling algorithm'' examples changes the particle weights (using a standard SPH kernel weight -- effectively mass-weighted in the smoothing kernel -- instead of a volumetric weighting) used to determine the coupling of SNe energy and momentum in the kernel. This can have dramatic effects on test problems: for a SNe in an infinitely thin, adiabatic disk with a low-density exterior, a mass-weighting couples all the momentum in the disk plane, instead of the vertical direction (the correct solution). Nevertheless we see this has relatively weak ($\sim20\%$) effects on the stellar mass and star formation history (in part because, in the average over many SNe over large volumes, all that matters is the total feedback input); however, it can significantly effect the morphological structure of e.g.\ the dense gas in a thin disk. We have also experimented with different functional forms for the ratio of the SNe energy and momentum coupling (producing small effects). The ``mod.\ RP algorithm'' choice discretizes our radiation pressure term (which is usually a continuous force) into intentionally very large ($>500\,{\rm km\,s^{-1}}$) ``kicks'' (this keeps the same total momentum flux, but makes each such particle ``kicked'' unbound) -- unsurprisingly this suppresses star formation further, but only by a factor of $\sim 3$. In the ``mod.\ RP+SNe'' choice we discretize the radiation pressure into smaller kicks ($=5\,{\rm km\,s^{-1}}$) and see this has little effect (as expected). In all cases the results lie within the (rather large) range allowed by observations. In a companion paper, \citet{keres:2013.fire.cosmo.vs.numerics} consider the detailed effects of substantial changes to each aspect of our numerical method described in Appendix~\ref{sec:appendix:sims}. Here, we simply show a few basic comparisons. Considerable attention has recently been paid to differences between the results of grid codes and older SPH methods (such as that in \citealt{springel:entropy}) for certain problems (especially sub-sonic fluid mixing instabilities; see \citealt{agertz:2007.sph.grid.mixing,kitsionas:2009.grid.sph.compare.turbulence,bauer:2011.sph.vs.arepo.shocks,vogelsberger:2011.arepo.vs.gadget.cosmo,sijacki:2011.gadget.arepo.hydro.tests,keres:2011.arepo.gadget.disk.angmom}). The numerical method used for our standard simulations has been specifically shown to resolve most of these discrepancies (giving results quite similar to grid codes in test problems); this is verified in \citet{hopkins:lagrangian.pressure.sph} for standard test problems and \citet{keres:2013.fire.cosmo.vs.numerics} for cosmological simulations. However we have re-run some of our simulations using the \citet{springel:entropy} formulation of SPH (described in Appendix~\ref{sec:appendix:sims}), which shows the most pronounced forms of these discrepancies. Despite the known differences between such methods for certain test problems, we find in Fig.~\ref{fig:mg.mh.fb} ({\em top} panel) that it makes little difference for the predicted galaxy masses. The older SPH method gives slightly lower $M_{\ast}(M_{\rm halo})$ (by about $\approx0.15\,$dex), primarily because cooling of diffuse ``hot halo'' gas is suppressed by less-efficient mixing. But for this specific question, the effect is quite small compared to the effects of including the appropriate stellar feedback physics. We also show an experiment where adopt an entirely distinct artificial viscosity prescription (see Appendix~\ref{sec:appendix:sims} for details), which produces negligible differences. It is important to stress that our conclusion here -- that our results depend only weakly on the numerical details -- applies to the galaxy stellar masses and other lowest-order, integrated quantities. In future work, we will study other properties of the simulations, such as the galaxy morphologies, which can (and do) depend on some parameters much more sensitively. For example, the modifications to the SNe coupling algorithm described above, which produce very little systematic change in our predicted stellar masses and star formation histories, produce surprisingly large changes to the angular momentum content and thickness of disks in the more massive galaxies. Finally, we show these results in part to stress, emphatically, that while there are {\em always} numerical choices in any code, there has been no ``tuning'' of these parameters for our study here. Certainly none of these has been ``fit'' or ``adjusted'' to match any observations, and all the choices above are held constant across our standard set of simulations, using values calibrated from simple test problems \citep[e.g.][]{hopkins:lagrangian.pressure.sph}. \begin{figure} \centering \plotonesize{mg_mh_altsims.pdf}{1.01} \vspace{-0.5cm} \caption{Comparison of the $M_{\ast}(M_{\rm halo})$ relation (as Fig.~\ref{fig:mg.mh.z0}) predicted by other published simulations in the literature using sub-grid stellar feedback models. We compile these results (colored points), where available, at low-$z$ ($z=0-0.5$; {\em top}) and high-$z$ ($z=2-3$; {\em bottom}). We compare against the simulations presented here (gray points) with explicit feedback, and observational constraints (lines). Even sub-grid models which are ``successful'' near $\sim L_{\ast}$ at $z\sim0$ over-predict $M_{\ast}(M_{\rm halo})$ by an order-of-magnitude relative to our explicit feedback simulations and observations at both low masses ($M_{\rm halo}\lesssim10^{10}\,M_{\sun}$) and/or high redshifts ($z\gtrsim2$). The exceptions appear to be the newest generation of sub-grid models which have been explicitly adjusted to mimic the effects of radiative feedback as well as SNe, seen in our explicit feedback models: this includes \citet{stinson:2013.new.early.stellar.fb.models,aumer:2013.new.cosmo.zooms.rad.pressure.fb,ceverino:2013.rad.fb,trujillo-gomez:2013.rad.fb.dwarfs}. \label{fig:mg.mh.altsims}} \end{figure} \vspace{-0.5cm} \subsection{(Strong) Dependence on Feedback} \label{sec:results:feedback} The lower panel in Fig.~\ref{fig:mg.mh.fb} shows the effect of varying the {\em physics} of feedback: now, we see dramatic differences in $M_{\ast}(M_{\rm halo})$. Removing all feedback (every mechanism listed in \S~\ref{sec:sims:feedback}), gas cools and collapses on a dynamical time $t_{\rm dyn}$ within the disk, forming stars at a rate $\dot{M}_{\ast} \sim M_{\rm gas}/t_{\rm dyn} \sim \dot{M}_{\rm gas}$ where $\dot{M}_{\rm gas}$ is the inflow rate from the halo. Most of the baryons are turned into stars.\footnote{\label{foot:nofb.lowmhalo}Even with no feedback, at very low masses $M_{\rm halo}\lesssim 10^{9}\,M_{\sun}$, some suppression of SF occurs after reionization because we still include a photo-ionizing background. However the predicted stellar mass is still larger than observed by at least an order of magnitude.} If we turn off SNe feedback, but retain all other forms of feedback, the results are nearly as bad: again, $M_{\ast}$ is severely overpredicted in both dwarfs and MW-mass systems. In the $\lesssim 10^{10}$ halos, with no SNe, other forms of feedback may still suppress SF significantly (so $M_{\ast}\ll f_{b}\,M_{\rm halo}$), but the masses are still much too large relative to those observed by factors of $\sim 100$. We also note that, as many previous studies have pointed out \citep{murray:momentum.winds,mckee:2007.sf.theory.review,shetty:2008.sf.feedback.model,cafg:sf.fb.reg.kslaw}, it is ultimately the {\em momentum} injected by SNe, not just the thermal energy, which regulates star formation. So as expected, if we artificially turn off the SNe momentum (coupling only thermal energy, as is common in many cosmological simulations), then in our simulations of massive ($>10^{12}\,M_{\sun}$) halos, this is nearly as bad as removing SNe entirely. In the lowest-mass dwarfs, the discrepancy is not so severe (factor $\lesssim2$ changes in the SFH), because the mass resolution ($\sim 100\,M_{\sun}$) is such that the early expansion phases of SNe remnants (in which the thermal energy begins to be converted into momentum) are well-resolved. If we remove radiative feedback entirely (both radiation pressure and local photo-ionization and photo-electric heating, as described in \S~\ref{sec:sims:feedback}), but retain SNe (and stellar winds), we see a nearly identical failure (to the no-SNe case) in both dwarfs and massive galaxies: while $M_{\ast} < f_{b}\,M_{\rm halo}$, far too many stars form. As we showed in Paper {\small II}, these mechanisms are critical to disrupt the dense regions of GMCs in which young stars are born, {\em before} SNe explode, and thus allowing the SNe to heat larger, lower-density volumes of gas (which can both avoid over-cooling and feel the collective effects of many SNe rather than just one), and therefore actually generate significant galactic outflows. The same result is found (on smaller scales) in much higher-resolution simulations of either single star clusters or the first stars, which directly treat the radiation-hydrodynamics with each single star as a source \citep[e.g.][]{offner:2009.rhd.lowmass.stars,krumholz:2011.rhd.starcluster.sim,tasker:2011.photoion.heating.gmc.evol,wise:2012.rad.pressure.effects}. Interestingly, in the dwarfs, if we turn off {\em only} radiation pressure, or {\em only} photo-ionization heating, the effect is much less severe: the predicted stellar mass is still significantly larger, but it is $>100$ times larger when both are removed. Radiation pressure can, to some extent, ``make up for'' the loss of photo-heating, and vice versa. This should actually not be surprising: the most massive GMCs in dwarf galaxies have local characteristic velocities $< 10\,{\rm km\,s^{-1}}$, thus either HII heating or UV radiation pressure alone can disrupt them (though we expect, under these conditions, HII heating should dominate), and this is completely consistent with both observations of star-forming regions \citep[e.g.][]{lopez:2010.stellar.fb.30.dor} and numerical radiation-hydrodynamic simulations of low-density, low-velocity clouds \citep{harper-clark:2011.gmc.sims,sales:2013.phototion.fb.strong}. And indeed this tradeoff between photo-heating and radiation pressure in small clouds is exactly what we saw in our ultra-high resolution simulations of isolated dwarfs of the same mass (discussed extensively in Paper {\small II}; see Figs.~7, 9, and 14-19 therein). In the massive systems, on the other hand, the radiation pressure term becomes more important than the HII heating. We see this in tests with both {\bf m12q} and {\bf m12v}. Even when the difference in stellar mass is not large (e.g.\ the {\bf m12v} case), the lack of radiation pressure feedback is particularly evident in the dense, early-forming center of the galaxy, where in the runs without radiation pressure feedback an enormous central density ``spike'' appears, leading to a very large circular velocity of $\sim 400\,{\rm km\,s^{-1}}$ in the central regions of these systems. At these densities, HII photo-heating is dynamically insignificant. If we disable stellar wind feedback (specifically, retaining stellar winds as a source of mass and metals, but associating no energy or momentum with that mass), and retain all other feedback, we see relatively weak effects. This is not surprising: their momentum flux is comparable to but not larger than other sources, and their energetics are much less than SNe. But they are obviously an extremely important source of mass and metals in the ISM. \vspace{-0.5cm} \subsection{Comparison to Previous Work} \label{sec:results:previous} In Fig.~\ref{fig:mg.mh.altsims}, we compare our results (grey points) at low and high redshifts, to those from previous simulations spanning a wide range of galaxy properties and numerical methods \citep{pelupessy:2004.dwarf.gal.sims.burst.sfhs, stinson:2007.dwarf.gal.sims.burst.sfh,stinson:2013.new.early.stellar.fb.models, mashchenko:2008.dwarf.sne.fb.cusps, valcke:2008.dwarf.gal.cosmo.sims, governato:2010.dwarf.gal.form,governato:2012.dwarf.form.sims, oser:2010.twophase.galform.sims, feldmann:bgg.size.evol.in.hydro.sims, brooks:2011.disk.scaling.law.sims, guedes:2011.cosmo.disk.sim.merger.survival, sawala:2011.dwarf.gal.sims.cores, scannapieco:2011.mw.mass.gal.form, de-rossi:2013.dwarf.cosmo.sims, okamoto:2013.pseudobulge.cosmo.sim, kannan:2013.early.fb.gives.good.highz.mgal.mhalo}. {\em All} of these simulations include some form of sub-grid model designed to mimic the ultimate effects of stellar feedback, although the prescriptions adopted differ substantially between each. Most of these models are specifically tuned to reproduce reasonable scaling for MW-mass systems at $z\sim0$. However, two discrepancies are immediately evident. First, nearly all the previous models predict much larger stellar masses in dwarf galaxies with $M_{\rm halo}\lesssim10^{11.5}\,M_{\sun}$, compared to either our simulations or the observational constraints. Second, even simulations which produce excellent agreement with the observations at $z=0$ tend to predict far too much star formation at high redshift (take e.g.\ the simulation in \citealt{guedes:2011.cosmo.disk.sim.merger.survival}, which produces a MW-like system with many properties consistent with observations at $z=0$, but has turned nearly all its baryons into stars at $z\gtrsim2$). These are similar to the discrepancies that appear when we re-run our simulations excluding radiative feedback. And indeed, nearly all of the models from the literature in Fig.~\ref{fig:mg.mh.altsims}, even given various freely adjustable parameters, are designed and motivated only to reproduce the effects of supernova feedback, which we have shown is insufficient to explain the observations. In fact, the only sub-grid models, to our knowledge, which currently do not produce such discrepancies (and agree broadly with our simulations both at low masses and high redshifts) are the recent generation of models in \citet{stinson:2013.new.early.stellar.fb.models,aumer:2013.new.cosmo.zooms.rad.pressure.fb,ceverino:2013.rad.fb,trujillo-gomez:2013.rad.fb.dwarfs} (for some additional results from these see \citealt{kannan:2013.early.fb.gives.good.highz.mgal.mhalo}). These new models (all of which have been developed recently) are {\em specifically} designed/tuned to mimic the effects of radiative feedback (albeit indirectly), and to reproduce via simple sub-resolution prescriptions (including turning off cooling) some of the most important effects of radiation pressure and photo-heating which were studied in our previous work \citep{hopkins:fb.ism.prop}.\footnote{There have also been interesting results from the re-tuned wind model of \citet{oppenheimer:outflow.enrichment} used more recently in slightly different forms in \citet{torrey:2013.arepo.cosmo.sim.vs.obs}, \citet{marinacci:2014.disk.form.arepo.subgrid}, and \citet{hirschmann:2013.mgal.mhalo.subgrid}. However, in this model, the wind outflow rates are set {\em explicitly} by-hand (and in fact the most recent scalings used were adjusted based on comparison to the sub-grid models including radiative feedback), and then tuned to reproduce the observed mass function. So this is essentially what we attempt to {\em predict} here.} Whether this is unique or not remains to be tested; the phase structure and other properties of outflows and the CGM in such models can be very different from those predicted here, even for the same mass-loading efficiencies (discussed further below). It will be particularly interesting to see whether other recently-developed sub-grid models such as that in \citet{agertz:2013.new.stellar.fb.model}, also incorporating the effects of radiative feedback but via very different prescriptions, will also agree well with observations at both low and high redshifts. In any case, these comparisons -- and the results from this new generation of sub-grid models -- highlight that some accounting for non-SNe feedback is critical. \begin{figure} \centering \plotonesize{zoom_kslaw}{1.01} \vspace{-0.5cm} \caption{Kennicutt-Schmidt law, observed \citep[][yellow shaded range]{kennicutt98,bigiel:2008.mol.kennicutt.on.sub.kpc.scales,genzel:2010.ks.law,daddi:2010.ks.law.highz} and simulated (points as Fig.~\ref{fig:mg.mh.z0}). We emphasize that this is a prediction: the instantaneous SF efficiency per dynamical time in dense gas is $100\%$ in the simulations, but the emergent KS-law, as a consequence of feedback, has an efficiency a factor $\sim50$ lower. As shown in Paper {\small I}\ and \citet{hopkins:virial.sf}, this is insensitive, with resolved feedback models, to the small-scale star formation law, and entirely determined by stellar feedback. ``No feedback'' models lie a factor $\sim50$ above the observations; ``no radiation'' and ``no SNe'' models (Fig.~\ref{fig:mg.mh.fb}) lie a factor $\sim10$ above observations. \label{fig:ks}} \end{figure} \begin{figure} \centering \plotonesize{zoom_sfh_z0_subgrid.pdf}{1.01} \caption{Example star formation history (SFH) for the {\bf m12v} simulation, in our standard (explicit feedback) model compared to different sub-grid feedback treatments. We show the formation history of all stars in the simulated box at $z=0$ (smoothed in $10^{7}$\,yr bins); this is not qualitatively different from the SFR versus time of the largest ``main'' galaxy in the box at each time. ``No feedback'' models force the galaxy to lie on the KS-law (SF is ``slow'') but do not expel gas; gas piles up until the SFR balances the halo accretion rate, with a broad peak from $z\sim2-6$. ``Sub-grid wind'' models ``kick'' gas at a rate proportional to the SFR; we show examples with different efficiencies and implementations. By design, model ``2'' produces a nearly identical $z=0$ stellar mass $M_{\ast}$ to our explicit feedback model. These sub-grid models (by construction) lower the SFR, but in both cases leave the qualitative behavior of the SFH identical. Explicit feedback models not only suppress the total $M_{\ast}$ formed, but change the shape of the SFH. SF is more ``bursty'' on small timescales, and the SFR is flatter in time (more biased to late times, without the broad high-$z$ peak). \label{fig:sfh.subgrid}} \end{figure} \begin{figure*} \centering \plotsidesize{zoom_sfh_z0_vsobs.pdf}{1.01} \caption{SFH for each ``main'' (largest) $z=0$ galaxy in our standard (explicit-feedback) simulations. Lines show the mean SFR averaged on timescales of $10^{8}$\,yr (black solid) and $10^{9}$\,yr (red dashed). With explicit feedback, SFRs are highly variable below the galaxy dynamical time. Moreover, the (averaged) SFRs tend to be flat and/or rising with time. In contrast, with no feedback, the SFH has a sharp rise and fall peaking at $z\sim2-6$. In the least massive dwarfs ({\bf m09}; $M_{\ast}<10^{6}\,M_{\sun}$ and $V_{\rm vir}(z=0)<20\,{\rm km\,s^{-1}}$), the SFR is strongly suppressed after reionization once a combination of the ionizing background and some small amount of feedback from the early star formation is able to expel most of the halo gas and prevent new cooling. We compare our ${\bf m11}$, ${\bf m12}$, and ${\bf m13}$ runs to the observationally inferred ``mean'' tracks (colored lines) for the main galaxies in halos of the same $z=0$ mass, from \citet[][magenta]{moster:2013.abundance.matching.sfhs} and \citet[][cyan]{behroozi:2012.abundance.matching.sfhs}. In each case the lines bracket the $1\,\sigma$ range/scatter in the observed galaxy population. Our {\bf m11} and {\bf m12} runs agree very well with these constraints; however, in the most massive systems ({\bf m13}), the galaxy never ``quenches,'' and the SFR remains high in conflict with observations below $z\sim1$. \label{fig:sfh.z0}} \end{figure*} \begin{figure*} \centering \hspace{-1.1cm} \plotsidesize{mg_mh_sfr.pdf}{1.04} \caption{{\em Top:} SFR versus galaxy stellar mass at different redshifts. We compare the observed (best-fit) relations from the compilation in \citet[][black dashed]{behroozi:2012.abundance.matching.sfhs} and \citet[][blue dashed]{zahid:2012.mass.metallicity.sfr.mass.compilations} (the systematic offset is typical of different calibrations). Allowing for the typical factor $\sim2$ systematic observational calibration uncertainty, the agreement is good at all $z$. However, magenta $+$'s compare low-resolution ($100$\,pc softening) runs of some massive halos which produce too-massive galaxies at $z=0$: there is little offset between these simulations and our fiducial models. The observed relation is simply a consequences of galaxies having relatively ``flat'' star formation histories. {\em Bottom:} Specific SFR of galaxies with different $M_{\ast}$, versus redshift. Observations are compiled in \citet[][Table~5]{behroozi:2012.abundance.matching.sfhs} and \citet{torrey:2013.arepo.cosmo.sim.vs.obs}. The dynamic range here is smaller so the plot appears noisier, but the information is identical to that at {\em top}. SSFRs at $z\gtrsim2$ are relatively flat, indicating rising SF histories at high-$z$. \label{fig:sfrs}} \end{figure*} \vspace{-0.5cm} \subsection{Instantaneous Suppression of Star Formation (at Fixed Gas Densities)} \label{sec:results:ks} We now examine galaxy star formation rates. In the previous section, we showed that the {\em integrated} SF is suppressed with feedback. But equally important is that feedback suppresses {\em instantaneous} SFRs in galaxies. This is manifest in the Kennicutt-Schmidt (KS) relation, shown in Fig.~\ref{fig:ks}.\footnote{We define $\Sigma_{\rm SFR} = \dot{M}_{\ast}/\pi\,R_{\rm SFR}^{2}$ (where $\dot{M}_{\ast}$ is the total SFR and $R_{\rm SFR}$ is the half-SFR radius) and $\Sigma_{\rm gas}=M_{\rm gas}/\pi\,R_{\rm SFR}^{2}$ (where $M_{\rm gas}$ is the gas mass within the $90\%$ SFR radius). Defining both $\dot{M}_{\ast}$ and $M_{\rm gas}$ within $R_{\rm SFR}$ or the stellar effective radius shifts the points along the relation.} We plot the simulations at all redshifts (the redshift evolution is insignificant), and compare to observations at a range of redshifts (which also find little or no evolution).\footnote{We compile the observed local galaxies in \citet{kennicutt98} and \citet{bigiel:2008.mol.kennicutt.on.sub.kpc.scales}, and high-redshift galaxies in \citet{genzel:2010.ks.law} and \citet{daddi:2010.ks.law.highz}; shaded region shows the $90\%$ inclusion range at each $\Sigma_{\rm gas}$ from the compilation. As discussed in those papers, there is no significant offset between the high and low-redshift systems at fixed $\Sigma_{\rm gas}$.} The predicted KS law agrees well with observations at all redshifts. As shown in Paper {\small I}-Paper {\small III}, this emerges naturally as a consequence of feedback, and is {\em not} put in by hand. Recall that the instantaneous SF efficiency (SF per dynamical time) in dense gas in the simulations is $100\%$; however the global SF efficiency is $\sim2\%$. This difference arises because at $\sim2\%$ efficiency, feedback injects sufficient momentum to offset dissipation (indeed, given the same feedback, we obtain the identical KS law independent of the details of our small-scale SF law; see \citealt{hopkins:rad.pressure.sf.fb,hopkins:virial.sf}). If we instead consider simulations with weak/no feedback, the global KS relation is severely over-predicted (efficient cooling leads to global efficiencies $\sim100\%$). In most cosmological simulations, this is offset ``by hand'' by simply enforcing a large-scale SF law that is sufficiently ``slow'' that it agrees with the observations; however we see that this is already (implicitly) a sub-grid feedback model. Including explicit feedback obviates the need for these prescriptions, meaning that the instantaneous SF properties are truly predictive, and not simply a consequence of our chosen small-scale SF law. \vspace{-0.5cm} \subsection{Global Star Formation Histories} \label{sec:results:sfh} In Fig.~\ref{fig:sfh.subgrid}, we examine the SFH of one MW-mass galaxy. We compare this to common sub-grid models. First, a ``no-feedback'' model following \citet{springel:multiphase}; this includes only a sub-grid model for the effects of stellar feedback on the ISM structure (an ``effective equation of state'') which ensures, by design (via tuned parameters) that the galaxy lies on the Kennicutt-Schmidt relation and has reasonable gas densities. However, without galactic winds, gas from inflows quickly builds up and the SFR rises until $\dot{M}_{\ast}\approx \dot{M}_{\rm inflow}$, and nearly all the baryons are turned into stars. The galaxy at $z=0$ is far too massive, and most of its stars are old (formed at $z\gtrsim2$, with the SFR peaking at $z\sim 5$).\footnote{As noted in the many previous simulations with this sub-grid model \citep[e.g.][]{springel:lcdm.sfh}, the ``effective equation of state'' approach does not allow cooling below $\sim10^{4}\,$K, so if those simulations properly included molecular or fine-structure cooling (with the appropriate high resolution), the SFH might peak even earlier.} We then add a sub-grid wind model, in which gas is ``kicked'' out of the galaxy (forced to free-stream to ensure it escapes the disk) at a rate proportional to the SFR: here the mass-loading is equal to the SFR (``sub-grid wind 1''). This suppresses the SFR (as it is intended to do), by about a uniform factor $\sim2$, as expected. However this still leaves a too-massive galaxy, with most of its stars formed very early. Next, we consider a stronger wind model (``sub-grid wind 2''): the mass-loading is doubled, with the free-steaming length fixed. This further suppresses the SFR -- in this model the final stellar mass agrees reasonably well with our explicit-feedback simulation. However, the sub-grid model still produces a SFR which peaks at very high redshifts $z\sim2-6$. The problem is that in all the sub-grid models -- regardless of the absolute suppression of the integrated SFR or position on the Kennicutt-Schmidt relation -- the {\em shape} of the galaxy SFH still closely resembles the shape of the halo inflow rate vs.\ time \citep[for examples of this with other sub-grid models, see][]{oppenheimer:metal.enrichment.momentum.winds,scannapieco:2011.mw.mass.gal.form,stinson:2013.new.early.stellar.fb.models,puchwein:2013.by.hand.wind.model}. These broadly peaked SF histories are disfavored by a variety of observations. They produce too-massive galaxies at high redshift, as discussed above. But they also produce galaxies with SF histories at high-$z$ that disagree with direct observational constraints \citep[see][]{papovich:highz.sb.gal.timescales,reddy:z2.lbg.spitzer,stark:2009.lbg.sfhs}. With our full, explicit feedback model included, we see that the {\em shape} of the SF history is qualitatively changed, and is more consistent with observations. At all times, SFRs are much more time-variable (this is discussed below). At the highest $z\gtrsim6$, halo and stellar masses both grow efficiently (albeit with some offset).\footnote{We caution that for our massive galaxies ($M_{\rm halo}\gtrsim10^{12}\,M_{\sun}$; with particle masses $\sim10^{4}\,M_{\sun}$), at high redshifts ($z\gtrsim4$), the progenitor galaxies have small baryonic masses and so are not as well resolved. As a result, the SF histories at these masses and redshifts depend more sensitively on the details of how feedback is coupled, even though the later-time SFRs and final stellar masses are robust to these variations. See Appendix~\ref{sec:appendix:algorithms} for details.} This is the ``rapid assembly'' phase, before/during reionization, in which feedback -- while able to eject some gas from the galaxy and provide some overall suppression and variability of $\dot{M}_{\ast}$ -- does not appear to dominate the gas dynamics (the central potential and mass of the halo grow on timescales comparable to the galaxy dynamical time; so $\dot{M}_{\ast}\propto \dot{M}_{\rm halo}$). But from $z\sim2-6$, feedback acts strongly, and there appears to be a maximum, steady-state SFR which is constant or slowly increasing with time at which the galaxy is able to cycle new material into a fountain and so maintain equilibrium. This ``quasi-equilibrium'' SFR scales with the {\em central} potential of the galaxy (see Paper {\small III}), as traced by quantities such as the central halo density or $V_{\rm max}$ (the maximum circular velocity), {not} the halo mass or virial velocity. The central potential depth increases only weakly over this time as halos accrete material on their outskirts. Below $z\sim2$, a competition ensues between slowing halo accretion rates and more highly-enriched halo gas raising cooling rates. Individual mergers also have a more dramatic effect on SF histories. In Fig.~\ref{fig:sfh.z0}, we show the SFH for each main $z=0$ galaxy in our simulations,\footnote{We define this as the formation rate vs.\ time (essentially a histogram of formation times) of all stars which end up in a $10\,$kpc aperture centered on the final ($z=0$) main galaxy in the simulation, averaged in $10^{8}\,$yr bins. Since most of these stars form ``in situ,'' the results are similar if we instead identify the most massive progenitor galaxy at all times and plot its galaxy-integrated SFR at each time.} and see that all cases with $10^{9}\lesssim M_{\rm halo} \lesssim 10^{13}\,M_{\sun}$ exhibit similar (relatively flat or slowly rising) SFHs.\footnote{Interestingly, the {\bf m12q} simulation shows a much higher high-$z$ SFR than {\bf m12v} or {\bf m12i}. This is in part because the particular choice of a ``quiescent'' halo led, in this case, to a halo with relatively little growth at late times ($z\lesssim3$), hence a particularly early ``formation time.''} In the most massive halos, some decline occurs when $M_{\rm halo}\gtrsim10^{12}\,M_{\sun}$, as the cooling time of virialized gas becomes longer relative to the dynamical time (the system transitions to ``hot mode'' accretion and filamentary infall is suppressed). However, we stress that the galaxies are clearly not ``quenched'' -- every system we simulate is still very much a star-forming, blue galaxy at $z\sim0$ (our {\bf m13} simulation would need a SFR $\ll 1\,M_{\sun}\,{\rm yr^{-1}}$ at $z=0$ to be ``red and dead'' by most definitions, but its SFR is $\sim5\,M_{\sun}\,{\rm yr^{-1}}$). In very low mass halos (e.g.\ our {\bf m09}, with $V_{\rm vir}(z=0)<20\,{\rm km\,s^{-1}}$), cooling is strongly suppressed after reionization. \vspace{-0.5cm} \subsection{Specific SFRs and the SF ``Main Sequence''} \label{sec:results:mainsequence} Fig.~\ref{fig:sfrs} compares the galaxy-integrated SFRs in all our simulated systems (including non-main halos) with observations of the SFR or specific SFR (SFR$/M_{\ast}$) as a function of galaxy stellar mass, at various redshifts. The simulations agree well with the SFR ``main sequence'' (SFR$-M_{\ast}$ relation) observed at all $z$ (observations plotted include compilations from \citealt{erb:lbg.gasmasses,noeske:2007.sfh.part1,daddi:2007.sfr.z2.exhaustion.time,elbaz:2007.sfr.mass.relation,stark:2009.lbg.sfhs} and others in \citealt{behroozi:2012.abundance.matching.sfhs} (see Table~5 therein) and \citealt{zahid:2012.mass.metallicity.sfr.mass.compilations}. The scatter is also similar to that observed. There may be some slight tension (the predictions being slightly high at $z=0$ and low at $z=2$), but these are well within the range of systematic uncertainties owing to different SFR calibrations (we show a couple such examples). By extension, the simulations similarly agree with the evolution in specific SFRs of galaxies as a function of mass. Evolution in specific SFRs and SFR$-M_{\ast}$ towards higher SSFR at high-$z$ simply reflects rising gas fractions (as it must, since the simulations lie on the same KS-law in Fig.~\ref{fig:ks}). The ``flattening'' of SSFR at high-$z$ implies SF histories of individual galaxies are rising with time (as we see directly); physically it follows from the saturation of gas fractions at large values, and rapid growth of halo mass at these times. The SFR$-M_{\ast}$ relation is, to lowest order, just $\dot{M}_{\ast} \sim M_{\ast}/t_{\rm Hubble}(z)$ -- this must be trivially true in any scenario where SFRs are relatively flat and/or rising with time (typical of star-forming galaxies). For this reason we see the same relation even in our simulations without feedback, as have other simulations with different feedback prescriptions \citep[see][]{keres:fb.constraints.from.cosmo.sims,dave:2011.mf.vs.z.winds}. And we see that even the very massive halos (which produce ``too large'' an $M_{\ast}$ at low redshifts) lie on the extension of the observed relation (the problem is that they continue on the relation, rather than ``quenching'' and moving below it, as observed at high masses). \begin{figure} \centering \hspace{-0.3cm} \plotonesize{zoom_sfh_bursty.pdf}{1.02} \vspace{-0.2cm} \caption{Variability of the SFHs shown in Fig.~\ref{fig:sfh.z0}, quantified versus timescales $\Delta t_{\rm avg}$. For each ``main'' galaxy in each simulation, we show the logarithmic dispersion in the SFR $\sigma_{\rm SFR}$ about its mean on longer timescales, when the SFR is time-averaged over the timescale $\Delta t_{\rm avg}$. The variability rises substantially on timescales $\sim10^{7}-10^{8}$\,yr (galaxy dynamical times), owing to a combination of fountain dynamics, local structure in the galaxies, and stochastic effects from individual star forming regions. The short-timescale variability is a factor of $\sim2-3$ in $\sim L_{\ast}$ galaxies, but rises to order-of-magnitude level in dwarfs (where individual star clusters and bursts have a more dramatic effect). \label{fig:sfh.burstiness}} \end{figure} \vspace{-0.5cm} \subsection{Quantifying Burstiness/Variability in SFRs} \label{sec:results:burstiness} In Fig.~\ref{fig:sfh.subgrid}, we showed that the SFRs are significantly more time-variable in models with explicit/resolved feedback as compared to sub-grid feedback models. We quantify this in Fig.~\ref{fig:sfh.burstiness}. We measure the dispersion in the SFR smoothed over various time intervals. Unsurprisingly, the scatter is larger on small timescales. On $\gg10^{8}\,$yr timescales, the variability is always small (SFHs are ``smooth'') -- this is more a function of the evolution of the halo over a Hubble time. Some such long-timescale variability is driven by mergers and global gravitational instabilities, but much of the short-timescale variability is not connected to these phenomena. Rather, on smaller timescales (comparable to the galaxy dynamical time) the dynamics of fountains, feedback, and individual giant molecular clouds and star clusters becomes important, so the scatter increases down to timescales $\sim10^{6}\,$yr (comparable to the massive stellar evolution timescale).\footnote{Note that even in the Milky Way, a large fraction of the observed star formation is associated with just the few most massive GMCs, so cloud-to-cloud variations can have significant effects on the global SFR \citep{murray:2010.sfe.mw.gmc}.} The short-timescale scatter is modest ($\sim0.3\,$dex) for massive systems ($M_{\rm halo}\gtrsim10^{12}\,M_{\sun}$), but rises in smaller halos (where even single star clusters can have large effects) to $\sim1\,$dex at $M_{\rm halo}\lesssim 10^{10}\,M_{\sun}$.\footnote{We have studied this in our resolution tests and found it is relative robust to spatial resolution, though the variability {\em increases} artificially on small timescales if the mass resolution is poor (factor $\sim10-100$ larger particle masses than we use), since single star particles then represent very massive star clusters.} \vspace{-0.5cm} \section{Discussion \&\ Conclusions} \label{sec:discussion} \subsection{Key Results and Predictions} We present a series of cosmological zoom-in simulations$^{\ref{foot:movie}}$ of galaxies with $M_{\rm halo}\sim10^{9}-10^{13}\,M_{\sun}$ and $M_{\ast}\sim10^{4}-10^{11}\,M_{\sun}$. At this time, several of these runs represent the highest-resolution in both mass and force resolution of any fully cosmological runs to $z=0$. But the most important improvement, compared to previous simulations, is that we for the first time include a fully explicit treatment of both the multi-phase (cold molecular through atomic, ionized, and hot diffuse) ISM and stellar feedback. Our treatment of the ISM is enabled both by our resolution and improved treatment of cooling and heating physics (e.g.\ molecular and metal line cooling, photo-ionization and photo-electric heating with self-shielding). Our stellar feedback model utilizes explicit time-dependent energy, momentum, mass and metal fluxes taken {\em directly} from stellar population models, without free/adjustable parameters. As such, the SFRs in our simulations, the resulting outflows, and galaxy stellar masses are {\em not} the result of tuning or ``by hand'' adjusting feedback efficiencies. Critically, we include not just thermal energy from SNe, but the momentum and energy associated with SNe Types Ia \&\ II, stellar winds (young star \&\ AGB), local photo-ionization and photo-electric heating, and radiation pressure from UV and IR photons. In addition, our formulation of SPH resolves the historical numerical problems with this method, especially important for cooling in hot halo atmospheres (see Appendix~\ref{sec:appendix:sims}; \citealt{hopkins:lagrangian.pressure.sph}; Keres et al.\ in prep). \, \ \ \\ Our key conclusions include: \begin{itemize} \item{Stellar feedback -- from known sources including SNe (energy {\em and} momentum), stellar winds, radiation pressure (primarily optical/UV), and photo-heating -- is both necessary and sufficient to explain the observed relation between galaxy stellar mass and halo mass, and by extension the shape of the galaxy mass function and clustering, at stellar masses $M_{\ast}\lesssim 10^{11}\,M_{\sun}$. This appears to be true at all redshifts.} \item{No {\em one} feedback mechanism alone is sufficient: the effects add non-linearly, and the common approximation in simulations of including only SNe feedback severely over-predicts galaxy masses (especially at low masses and/or high redshifts). The effects are even worse if the feedback momentum is ignored (if only thermal energy is considered).} \item{The $M_{\ast}-M_{\rm halo}$ relation evolves very weakly with redshift (because outflow efficiencies depend mostly on the {central} binding energy within the galaxy). At $z\gtrsim2$, weak evolution towards higher $M_{\ast}(M_{\rm halo})$ at low masses is equivalent to a steepening faint-end slope of the galaxy luminosity function, similar to what is inferred observationally \citep{bouwens:highz.sfh,stark:2009.lbg.sfhs}.} \item{Stellar feedback and standard cooling physics explain low galaxy stellar masses, but do not appear sufficient to explain ``quenching'' (late time suppression of star formation in massive halos) -- none of our massive systems are ``red and dead.''} \item{Our simulations reproduce the observed Kennicutt-Schmidt relation. This is despite the fact that we assume a small-scale SF efficiency of $100\%$ in self-gravitating dense gas. As such, the KS law and instantaneous SFRs are truly predicted, not simply a consequence of our sub-grid SF law. The low star formation efficiency we find is a consequence of stellar feedback, {\em not} the microphysics of how stars form in dense gas. Absent feedback, efficient cooling leads to a global SF efficiency of $\sim100\%$ per dynamical time. With feedback -- from the same mechanisms that produce large-scale outflows and regulate galaxy formation -- the SF efficiency self-regulates at $\sim2\%$, the level where feedback injects sufficient momentum to offset dissipation.} \item{Realistic feedback changes the {\em shape} of galaxy star formation histories. In particular, feedback from stellar radiation (both photo-heating and radiation pressure) is critical for disrupting dense, cold gas, and so is especially important for suppressing star formation in high redshift galaxies. This leads to much flatter, or gently rising, star formation histories in sub-$L_{\ast}$ galaxies. Most previous sub-grid models give qualitatively different results, in conflict with observations.} \item{The observed star formation ``main sequence'' and specific SFRs emerge naturally from the shape of the galaxies' star formation histories (from $M_{\ast}\sim10^{8}-10^{11}$ and $z\sim0-6$). This includes ``flat'' SSFR evolution at $z\sim2-6$. However these are relatively insensitive to feedback, since any broadly flat or rising SF history predicts $M_{\ast}(z)\sim t_{\rm Hubble}(z)\,\langle \dot{M}_{\ast}(z)\rangle$, consistent with the observations.} \item{Dwarf galaxies exhibit much more ``bursty'' SF histories, with large variability in their SFRs on short timescales ($\sim1\,$dex scatter on $\lesssim10^{7}\,$yr timescales). This is because star formation and star cluster formation, and their associated feedback, are stochastic. The variability is not driven by mergers or global gravitational instabilities. Massive ($\sim L_{\ast}$) galaxies are much less variable ($\sim0.3\,$dex scatter in SFRs). This may translate into significantly larger scatter in $M_{\ast}(M_{\rm halo})$ at dwarf masses compared to $\sim L_{\ast}$ galaxies.} \end{itemize} \vspace{-0.5cm} \subsection{Numerical Methods} We see relatively weak dependence on simulation resolution, which is perhaps surprising given the small-scale structure present in the ISM. However, in Papers {\small I-III} \&\ Appendix~\ref{sec:appendix:test.isolated}, we presented extensive resolution studies of isolated disk galaxies simulated using the same prescriptions but numerical resolution varied from values comparable to those here, to order-of-magnitude superior mass and spatial resolution. We showed that the galaxy-averaged SFR is one of the very first quantities to converge, and is consistent to within factor $\sim2$ even for relatively poor resolution: this is because it traces the {\em integral} effect of feedback balancing turbulent dissipation. That said, we {\em do} see qualitative changes in behavior if the resolution falls below that needed to meaningfully resolve ISM phase structure, at which point ``self-regulation'' by feedback loses meaning. As a rule of thumb, the simulations must at least resolve the Toomre/Jeans length and mass (the size of the largest GMCs) in each star-forming disk, and the results are especially numerically stable if the mass resolution can be ideally pushed to $< 10^{4}\,M_{\sun}$. Even in this regime, quantities such as the phase structure of dense gas and outflows are much more sensitive to resolution, and will be discussed in more detail in future work. Given the same feedback model, we also see little difference in the stellar mass buildup between our standard simulations, run with a numerical algorithm designed and shown to eliminate essentially all major differences between grid (Eulerian) and smoothed-particle (Lagrangian) hydrodynamics methods, and an older version of SPH that exhibits large differences in test problems. Thus we expect little or no difference between the results here and those from adaptive-grid or moving-mesh codes, if the same feedback and ISM physics could be included. This owes to two key points: first, the differences between numerical methods in this respect, even where significant, are generally much smaller than the orders-of-magnitude differences owing to the inclusion or exclusion of the relevant physics. The stellar mass content of galaxies is set by the total amount of feedback injected, and so it is unsurprising that the time-averaged star formation rate is insensitive to changes in the detailed phase structure of the gas around galaxies. Second, the numerical differences primarily affect mixing instabilities in multi-phase, sub-sonic, pressure-dominated gas. As such, many comparison studies have shown that while the numerical differences can be important for details of the structure of hot halos in massive galaxies, they are generally unimportant inside cold star-forming gas, or in sub-$L_{\ast}$ halos, where the flows of interest tend to be highly super-sonic and gravity-dominated \citep[see e.g.][]{kitsionas:2009.grid.sph.compare.turbulence,price:2010.grid.sph.compare.turbulence,bauer:2011.sph.vs.arepo.shocks,sijacki:2011.gadget.arepo.hydro.tests}. As one considers more detailed galactic properties, we expect the differences between numerical methods to manifest as discrepancies in the cooling properties, phase structure, or distribution of heavy elements in the CGM, and to impact the way in which both inflowing cool gas and feedback-driven outflows interact with gas in galactic halos. For these reasons, an accurate numerical scheme is critical if one hopes to study the detailed structure of both gas in and around galaxies with realistic feedback. A much more extensive comparison of numerical methods is presented in a companion paper \citep{keres:2013.fire.cosmo.vs.numerics}. \vspace{-0.5cm} \subsection{Future Work} This is a first exploration of cosmological simulations with explicit stellar feedback models, and many open questions remain. We have studied the effects of realistic stellar feedback on galaxy star formation histories and stellar masses; however, a complete understanding of this self-regulation requires a much more detailed examination of the dynamics of galactic outflows. In companion papers, we will study how outflows are generated, and how these interact with the circum-galactic and inter-galactic medium. It will be particularly important to build new observational diagnostics and explore whether or not different feedback mechanisms lead to different observable properties in the ISM, CGM, and IGM \citep{faucher-giguere:2014.fire.neutral.hydrogen.absorption}. Complementary questions regarding the morphology of galaxies -- how the sizes, bulge-to-disk ratios, kinematics, and other properties of the simulated systems here depend on different feedback mechanisms -- will be developed as well. The resolution and explicit treatment of the ISM in these simulations make possible many additional studies. Going forward, it will also be important to examine the role of additional physics. Some other physics is probably needed to explain the ``quenching'' of star formation in massive systems ($M_{\rm halo}\gg10^{12}\,M_{\sun}$). AGN feedback is a plausible candidate, which we have studied in previous work using idealized sub-grid models for the ISM. But the consequences could easily be completely different in a resolved multi-phase medium. Other physics such as magnetic fields, anisotropic conduction, and cosmic rays may be important as well, and their consequences are just beginning to be explored \citep[e.g.][]{jubelgas:2008.cosmic.ray.outflows,hanasz:2013.cosmic.ray.winds,salem:2013.cosmic.ray.outflows}. \vspace{-0.7cm} \begin{small}\section*{Acknowledgments}\end{small} We thank the many friends and peers who discussed this work in progress and sent helpful suggestions after the first draft was posted to the arXiv. The simulations here used computational resources granted by the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number OCI-1053575; specifically allocations TG-AST120025 (PI Keres), TG-AST130039 (PI Hopkins), and TG-TG-AST090039 (PI Quataert). Collaboration between institutions for this work was partially supported by a workshop grant from UC-HiPACC. Partial support for PFH was provided by NASA through Einstein Postdoctoral Fellowship Award Number PF1-120083 issued by the Chandra X-ray Observatory Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of the NASA under contract NAS8-03060, and by the Gordon and Betty Moore Foundation through Grant \#776 to the Caltech Moore Center for Theoretical Cosmology and Physics. JO also thanks the financial support of the Fulbright/MICINN Program and NASA Grant NNX09AG01G. DK acknowledges support from the Hellman Fellowship fund at the UC San Diego and NASA ATP grant NNX11AI97G. CAFG is supported by a fellowship from the Miller Institute for Basic Research in Science and by NASA through Einstein Postdoctoral Fellowship Award number PF3-140106 and grant number 10-ATP10-0187. EQ is supported in part by NASA ATP Grant 12-ATP12-0183, a Simons Investigator award from the Simons Foundation, the David and Lucile Packard Foundation, and the Thomas Alison Schneider Chair in Physics. \\ \vspace{-0.2cm}
\section{Be stars and the Be phenomenon} Be stars are Main Sequence or slightly evolved non-supergiant, late-O, B, or early-A stars that show or have shown emission in at least one Balmer line \citep{collins1987}. Emission can also appear in other lines of the spectrum as well as in the continuum, particularly as an excess of light in the infrared domain. The emission is due to the presence of a circumstellar disk, fed by episodic ejections of material from the surface of the star to a Keplerian orbit. This is called the ``Be phenomenon''. However, how these outbursts producing the disk can occur was not understood until now. There are $\sim$2000 Be stars known as of today and listed in the BeSS (Be Star Spectra) database of Be stars \citep{neiner_BeSS}. Be stars represent about 20\% of all B-type stars in our galaxy. Only B stars of sufficiently high rotational velocity at the Zero Age Main Sequence (ZAMS) can become Be stars \citep{martayan2007}, and this velocity depends on the metallicity of the protostellar environment. In addition, there seems to be a strong dependence of the proportion of Be stars compared to B stars as a function of spectral type. Indeed, the Be phenomenon is mostly observed around the sub-types B1-B2 \citep[see, e.g., Fig.~1 of][]{balona2000}. B2 is also the spectral type at which pulsations of both $\beta$\,Cep and Slowly Pulsating B (SPB) types can occur at the same time. Light and line-profile short-term variability (of the order of one day) due to non-radial pulsations is indeed detected from the ground or with Hipparcos in almost all (86\%) early-Be stars, in 40\% of mid-types (B4-5e), and in only 18\% of late-Be stars according to \cite{hubert1998}. Thanks to high-precision photometric space-based missions, we actually find that all Be stars pulsate, whatever their spectral type, but with a lower amplitude for cooler stars. Short-term variability also occurs owing to rotational modulation caused for example by surface spots. Variability on other time scales is also common. Cyclic variations of the order of months or years are associated with the wind, which is variable and stronger than for B stars. In addition, variations in the disk emission line profiles with time-scales of a few years to decades are observed: first, the intensity of emission lines slowly decreases as the disk dissipates in the interstellar medium; second, a denser zone can exist in the disk, which slowly precesses and produces global one-armed disk oscillation. Finally, abrupt emission increases are due to short-lived (days, tens of days), sometimes recurrent, or/and long-lived (months) outbursts. Depending on the frequency of outbursts and on the speed of disk dissipation, the disk can sometimes completely disappear and reappear following a new outburst. For a complete review of Be stars, we refer the reader to \cite{porter2003}. Be stars are known to be very fast rotators. This leads to many rotational effects, in particular Be stars are flattened by the centrifugal force. However, their velocities are not high enough to reach the critical limit at which the centrifugal force compensates gravity at the equator. Indeed, galactic Be stars rotate on average at 88\% of the critical angular velocity \citep{fremat2005}. Thus, at least in most cases, while rotation is certainly an important ingredient in igniting outbursts, it cannot by itself explain the ejection of matter from the star that leads to the formation of the decretion disk. Another mechanism is required to provide the additional angular momentum at the surface needed to eject matter. Several explanations have been put forward to provide this additional angular momentum. \cite{oudmaijer2010} showed that about 30\% of Be stars are in binary systems with a close companion, which is similar to the binary rate of B stars. The binary companion certainly plays a very important role for those 30\% of Be stars (e.g., in Be/X-ray binaries). For single stars, results from the MiMeS project \citep[e.g.][]{neiner2011} showed that the magnetic field of Be stars, if it is present, is weak, and that it only influences possible co-rotating clouds close to the stellar surface and not the Keplerian Be disk \citep{neiner_omeori2012}. Therefore pulsations appear to be the most likely explanation, in addition to the rapid rotation, of the Be phenomenon. \section{Correlation between pulsations and outbursts in HD\,49330} \label{hd49330} CoRoT allowed us to observe the hot (B0IVe) Be star HD\,49330 during an outburst \citep{huat2009}. By completing a photometric analysis of the precise CoRoT data of that star acquired over $\sim$136 days, we were able to detect over 300 different frequencies attributable to variations, including at least thirty independent ones. These include high frequencies as well as several groups of low frequencies, which are typical of the pressure (p) and gravity (g) pulsation modes observed in $\beta$ Cep and SPB stars, respectively. Some of the frequencies have also been detected in photospheric line profile variations with the help of simultaneous spectroscopic observations \citep{floquet2009}. Thanks to these CoRoT data, we have discovered a correlation between both amplitude changes and the presence/absence of certain frequencies of pulsation, with the different phases of an outburst: (i) the amplitude of the main p-mode frequencies decreases before and during the outburst and increases again after it; (ii) several groups of g-mode frequencies appear just before the outburst, their amplitude reaches a maximum during the outburst, and they disappear as soon as it has finished. These frequencies appear to have complex structures, which could represent pulsation modes with a short lifetime. The detailed characterization of the pulsation modes and fundamental stellar parameters of HD\,49330, and the presence of both p-modes and g-modes, provide strong constraints on its seismic modeling. We tried to model HD\,49330 with $\kappa$-driven pulsations using the Tohoku oscillation code \citep{saiocox1980,lee1995} that accounts for the combined action of Coriolis and centrifugal accelerations on stellar pulsations as needed for Be star modeling. However, whatever models of pulsations and stellar structure we used, we find that p-modes and g-modes cannot be excited at the same time by the $\kappa$ mechanism in HD\,49330 using stellar parameters determined spectroscopically. \section{Discovery of stochastically excited gravito-inertial modes in HD\,51452} HD\,51452 is a hot Be (B0IVe) star observed with CoRoT, which has a rotation frequency $f_{\rm rot} \sim 1.22$ c~d$^{-1}$. For such a hot star, the $\kappa$ mechanism can create p modes of pulsations or possibly low-order g modes with frequencies above $\sim$1.5 c~d$^{-1}$. Nevertheless g-modes are detected in the CoRoT light curve of HD\,51452 with frequencies below $\sim$1.5 c~d$^{-1}$. In addition, a multiplet of frequencies is detected, with its main peak in the domain below $\sim$1.5 c~d$^{-1}$, with a frequency spacing of about 0.6 c~d$^{-1}$ \citep[see][]{neiner_51452}. These frequencies below $\sim$1.5 c~d$^{-1}$ cannot be explained with the $\kappa$ mechanism, even when taking stellar flattening into account. They could, however, be stochastic gravito-inertial (gi) modes. In particular sub-inertial gi-modes would be below $\sim$2.44 c~d$^{-1}$. There are two types of gi-mode: (1) those usually called g-modes, which are gravity modes modified by the Coriolis acceleration, and which show a regular pattern in period when they are asymptotic \citep{lee1997,ballot2010}, and (2) rotational (r) modes, which are mainly driven by the Coriolis acceleration, are sub-inertial and show regular patterns in frequency \citep{provost1981,saio1982,lee2006}. Since we observed at least one multiplet in frequency in the sub-inertial domain for HD\,51452, those peaks may be interpreted as r-modes. Since no specific frequency or period spacing is found for the other frequency peaks, they could be any type of gi-mode. Convective regions, such as the convective core and the sub-surface convection zone in massive stars, are indeed able to stochastically excite oscillation modes \citep{cantiello2009,belkacem2010} and particularly g-modes \citep{samadi2010,shiode2013}. The latter become gi-modes in fast rotators such as HD\,51452 because of the action of the Coriolis acceleration \citep[see e.g.][]{lee1997,DR2000,Mathis2009,ballot2010}. This excitation is also observed in realistic numerical simulations of convective cores surrounded by a stably stratified radiative envelope \citep[see][]{browning2004}. Gravito-inertial waves are excited through their couplings with volumetric turbulence in convective regions (where pure g-modes in a slowly rotating star are evanescent, while gi-modes in a rapidly rotating star become inertial) and by the impact of structured turbulent plumes at the interfaces between convective and radiative regions. \cite{samadi2010} examined the stochastic excitation of gravity modes by turbulent convection in massive stars. They found that the excitation of low $n$-order g-modes occurs in the core while the asymptotic g-modes are mostly excited in the outer convective zone. The mode amplitudes that they deduced, however, are well below the detection threshold of the CoRoT satellite for a massive star. On the contrary, recent work by \cite{shiode2013} showed analytically that taking stochastic excitation by penetration into account produces g-modes of detectable amplitude. In both works however, no rotation is considered. In addition, three-dimensional (3D) simulations of a convective core in a A star \citep{browning2004} or B star (Augustson et al., in preparation) showed efficient stochastic excitation of g-waves. HD\,51452 is a Be star, rotating close to its breakup velocity. Therefore, the calculation of the excitation of gi-modes (including r-modes) in this star requires the study of the influence of very fast rotation. This application can be derived from the work by \cite{belkacem2009} and is the study of a forthcoming paper (Mathis et al. 2013). The detection of gi-modes in HD\,51452 presented in \cite{neiner_51452}, however, already suggests that fast rotation enhances the amplitude of gi-modes and thus r-modes. The fact that HD\,51452 is a very hot Be star excludes the possibility that the observed gi-modes are excited by the $\kappa$-mechanism. In view of these results, however, it might be necessary to reconsider our interpretation of several other rapidly rotating B or Be stars, for which the $\kappa$-mechanism seemed like an obvious excitation mechanism but for which stochastic excitation might also be at work. In particular the g-modes observed in HD\,49330 (see Sect.~\ref{hd49330}) could be stochastically excited, which would explain why the seismic models with $\kappa$-driven modes did not work. \section{Transport of angular momentum by waves in Be stars} Considering the arguments presented above for HD\,51452, the results of our modeling of HD\,49330 with the $\kappa$-mechanism, as well as the fact that the power spectrum of that star shows broad frequency groups rather than sharp frequency peaks around 1-2 c~d$^{-1}$ \citep{huat2009}, we propose that HD\,49330 hosts stochastic g-waves. During the quiet phase, stochastic gi-waves can be excited in the convective core. If they are sub-inertial, as observed in the Be star HD\,51452, these waves transport more angular momentum to the subsurface layers than $\kappa$-driven modes because their frequency is lower. The net deposit of angular momentum indeed depends on the thermal dissipation, i.e., it is proportional to $1/f^4$ for a given rotation \citep{zahn1997, Mathis2009}. When enough angular momentum has accumulated in the outer layers of the star, these layers get unstable and could emit transient g-waves, which we detect. Possibly, the g-waves excited in the core may then also become visible. The surface layers reach the critical velocity, in particular at the equator where the rotation was already the closest to critical. The destabilisation of the surface layers thus ignites the outburst and breaks the cavity in which the p-modes were propagating. This explains both the disappearance of the p-modes during the precursor and outburst phases of HD\,49330 and the ejection of material from the surface into the disk, i.e., the occurrence of the outburst, as observed by \cite{huat2009} with CoRoT. Relaxation then occurs, recreating the cavity and letting the p-modes reappear while the transient g-waves disappear. Each time stochastic g-waves from the core accumulate enough angular momentum in the outer layers of the star, this outburst phenomenon will occur again. The idea that non-radial oscillations excited in massive stars are able to efficiently transport angular momentum and to allow the surface of a Be star to reach its critical velocity has already been proposed by \cite{ando1986}, \cite{lee_saio}, and \cite{lee2006}. However, in these previous works, gi-waves were excited by the $\kappa$-mechanism. In the case of HD\,49330, we propose that gi-waves are stochastically excited in turbulent convective regions. \cite{pantillon2007} and \cite{rogers2012} demonstrated that stochastic gi-waves are able to transport angular momentum in the same way as $\kappa$-driven waves. The type of excitation does not influence the transport. It depends mostly on the amplitude and frequency of the mode \citep{zahn1997, Mathis2009}. The higher the amplitude and the lower the frequency, the more transport there is. The Tohoku models of HD\,49330 indeed show that the rate of local angular momentum change due to pulsational transportation integrated over the mean sphere increases drastically in the few percents of the stellar radius just below the surface, i.e., there is a strong net deposit of angular momentum in the surface layers, because this is where the pulsations are damped. This confirms that pulsations increase angular momentum in the outermost layers of the star. The Tohoku pulsation code is currently being modified to include stochastically excited pulsations. This new version will be used to test our scenario (Neiner, Saio et al., in preparation). \section{Conclusion} All Be stars observed with CoRoT pulsate, whatever their spectral type. A scenario similar to the one observed in HD\,49330 could thus occur in all other Be stars, thus providing an explanation of the long-lasting mystery of the origin of Be outbursts and disks, for single Be stars. Stochastic waves are certainly excited in the convective core of any massive star, but with amplitudes that may be undetectable (even with space-based facilities) if the star rotates slowly. From the observations of stochastically excited gi-modes in HD\,51452 and the results of the seismic modeling of HD\,49330, we however propose that rotation enhances the amplitude of stochastic gi-waves/modes. Since these waves are of low frequencies, they transport more angular momentum than other waves/modes of higher frequency for a given amplitude. Therefore we conclude that a hot slowly rotating pulsating B star will only excite detectable $\kappa$-driven p modes, i.e., it will be a $\beta$\,Cep star, but if it rotates fast, it will also excite stochastic gi-waves/modes with a larger amplitude than the $\beta$\,Cep star and it will become a hot Be star. A cool slowly rotating pulsating B star will excite $\kappa$-driven g-modes, i.e., it will be a SPB star, but if it rotates fast, it will also excite stochastic gi-waves/modes with a larger amplitude than the SPB star and it will become a cool Be star. We conclude that single Be stars bring their surface layer above the critical velocity thanks to three ingredients: (1) rapid rotation itself, (2) the transport of angular momentum by pulsations (of all types, but mostly prograde g-modes), and (3) the enhancement of the amplitude of stochastic gi-waves thanks to the rapid rotation, which allows more transport of angular momentum. The fact that a B star becomes a Be star therefore depends strongly on its rotation rate, pulsations, and convection (in competition with its stellar wind). \acknowledgements The CoRoT space mission, launched on December 27th 2006, has been developed and is operated by Centre National d'Etudes Spatiales (CNES), with the contribution of Austria, Belgium, Brazil, the European Space Agency (RSSD and Science Program), Germany and Spain. CN acknowledges fundings from SIROCO (Seismology, Rotation and Convection with the CoRoT Satellite) ANR (Agence Nationale de la Recherche) project, CNES and PNPS (Programme National de Physique Stellaire).
\section{Introduction} The idea of intertwining operator $V$ such that $VP = QV$ for $P$ and $Q$ ordinary differential operators goes back to Gelfand, Levitan, Marchenko, Naimark, Delsarte and Lions (see \cite{levitan}, \cite{Lions}). It was picked up again by C. F. Dunkl, R\"{o}sler and K. Trim\`eche (see \cite{Dunkl}, \cite{Rosler} \cite{Trimeche}) who established some fundamental ideas related to the class of differential difference operators. In this work we investigate in the rank one case the particular cases of complex reflection Dunkl operator $T(k)$, associated with complex reflection group $G(m,1,1)$, on the set of radial rays $U=\cup_{j=1}^m\varepsilon^j\mathbb{R}$, which is given by \cite{Dunklref} \begin{equation} T(k)f(x):=\frac{df(x)}{dx}+\sum_{i=1}^{m-1}\frac{k_i}{x}\sum_{j=0}^{m-1}\varepsilon^{-ij}f(\varepsilon^jx),\,\, \varepsilon=e^{\frac{2i\pi}{m}}\,\,\,\text{and}\,\,\,\, k_i\in\mathbb{C}. \label{diffh} \end{equation} In particular, when $m=2$, $T(k)$ coincides with the following Dunkl operator on the real line \begin{equation} Tf(x):=\frac{df(x)}{dx}+\frac{\nu+1/2}{x}(f(x)-f(-x)).\label{Dunklop} \end{equation} First, we indicate briefly some results involving intertwining operators. In \cite{Dunkl}, C. F. Dunkl has proved that there exists a linear isomorphism $V$, called the Dunkl intertwining operator, from the space of polynomials on $\mathbb{R}$ of degree $n$ onto itself, satisfying the transmutation relation \begin{equation} T\circ V=V\circ \frac{d}{dx}\label{trans1},\,\,\,V(1)=1. \end{equation} In \cite{Rosler}, R\"{o}sler has obtained an integral representation of $V$ and K. Trim\`eche \cite{Trimeche} extended it to a topological isomorphism from $\mathcal{E}(\mathbb{R}),$ the space of even $C^{\infty}$-functions on $\mathbb{R}$, onto itself satisfying the relation \eqref{trans1} and obtained the following form \begin{equation} V(f):=\mathcal{R}_{\nu}(f_e)+\frac{d}{dx}\circ\mathcal{R}_{\nu}\circ I(f_o), \end{equation} where $f_e$ and $f_o$ are respectively the even and odd parts of the function $f$, \begin{equation} I(f)(x):=\int_0^{x}f(t)dt \label{rightinv} \end{equation} and $\mathcal{R}_{\nu}$ is the Riemann-Liouville operator given by \begin{equation} \mathcal{R}_{\nu}(f)(x):=\frac{\Gamma(\nu+1)}{\Gamma(\frac{1}{2})\Gamma(\nu+\frac{1}{2})}\int_0^1 (1-t)^{\nu-\frac{1}{2}}t^{-\frac{1}{2}}f(xt^{\frac{1}{2}})dt.\label{Riemann1} \end{equation}\\ The goal of this paper is to provide a similar construction for an intertwining operator $V_m$ between the complex Dunkl operator $T(k)$ and the derivative operator $\frac{d}{dx}$. Our construction is based on some hyper-Bessel operator and Riemann-Liouville type transform.\\ The remaining sections of this paper are organized as follows. In Section 2, we first recall notations and some results for Dunkl operator, we establish a new representation for the complex Dunkl operator by using circular matrices. In section 3, we discuss some results satisfied by the hyper-Bessel functions which can be found in the literature. In section 4, we give a new intertwining operator between $T(k)$ and $\frac{d}{dx}$. \section{Complex Dunkl operators of type $G(m,1,N)$} Let $m \in \mathbb{N}$ $(m\geq 2)$. We denote by $G$ the cyclic group generated by $\varepsilon=e^{\frac{2i\pi}{m}}$ and by \begin{equation}U=\cup _{j=1}^m \varepsilon^j\mathbb{R}\end{equation}a set of radial rays in complex plane. For $i=1,\,...,\,m,$ we define the operators \begin{equation} p_{i}(f)(x)=\frac{1}{m}\sum_{j=0}^{m-1}\varepsilon^{-ij}f(\varepsilon^jx). \end{equation} These obey \begin{equation} id=\sum_{i=1}^{m}p_{i},\,\,\,\,\,p_{i}p_{j}=\delta_{ij}p_{i}. \end{equation} Then, the elements $p_{i}$ are idempotents which are generalizations of the primitive idempotents (1 - s)/2 and (1 + s)/2 for a real reflection $s$. \begin{definition}A function $f:\,U \rightarrow \mathbb{C}$ is called of type $j$ with respect to $G$, if $$f(\varepsilon x)=\varepsilon^jf(x),$$ hold for every $x\in U.$ \end{definition} \begin{lemma}Let $f$ be a function $f:U\rightarrow \mathbb{C}$. Then, $f$ can be decomposed uniquely in the form \begin{equation*} f=\sum_{j=0}^{m-1}f_{j}\,, \end{equation*} where the component function $f_{j}$ is of type $j$, given by \begin{align} f_{j}=p_{j}(f).\label{idem} \end{align} \end{lemma} \begin{example} Let $\kappa=e^{\frac{i\pi}{m}}.$ By using the previous Lemma we obtain easily the following decomposition of the exponential function $e^{\kappa x}$ $$e^{\kappa x}= \cos_m(x)+\sum_{l=1}^{m-1}\kappa^l \sin_{m,l}(x),$$ where the hyper-trigonometric functions $\cos_m(x)$ and $\sin_m(x)$ are given by \cite{Erdely} \begin{align} \cos_{m}(x):=\sum_{n=0}^{\infty}(-1)^n\frac{x^{nm}}{(nm)!}\,\,\,\text{and}\,\,\,\,\sin_{m,l}(x):=\sum_{n=0}^{\infty}(-1)^n\frac{x^{nm+l}}{(nm+l)!}.\label{cosine} \end{align} The function $y(x)=\cos_m(\lambda x)$ is the unique $C^{\infty}$-solution of the system \begin{equation*}\left\{ \begin{array}{c} y^{(m)}(x)=-\lambda^m y(x),\\ y(0)=1,\,y^{(1)}(0)=\,...\,=y^{(m-1)}(0)=0. \end{array} \right. \end{equation*} \end{example} We denote by $\mathcal{E}(U)$ the space of $C^{\infty}$-complex valued functions on $U$ equipped with the topology of uniform convergence on compacts of the functions and all their derivatives, is a Frechet space and we denote by $\mathcal{E}_j(U)$ the subspace of $\mathcal{E}(U)$ of functions of type $j$ with respect to the group $G$. Of course we have \begin{align*} \mathcal{E}(U)=\bigoplus_{j=0}^{m-1}\mathcal{E}_{j}(U).\end{align*} Let $\nu = (\nu_{1},\,...,\,\nu_{m-1},0)\in \mathbb{C}^{m}$ and $k=(k_1,\,...\,k_{m-1},\,0),$ with $k_j=m\nu_j+m-j$.\\ The complex reflection Dunkl operator associated to cyclic $G$ generated by $\varepsilon=e^{{2i\pi}{m}}$ is defined by (\cite{Dunklref}, \cite{Bouzaffour}) \begin{equation} T(k)f(x):=\frac{df(x)}{dx}+\sum_{i=1}^{m-1}\frac{k_i}{x}\sum_{j=0}^{m-1}\varepsilon^{-ij}f(\varepsilon^jx). \end{equation} \begin{proposition}The operator $T(k)$ can be written in the the form \begin{align*} T(k)=\frac{d}{dx}+\frac{\omega_k}{x}, \end{align*} where $$\omega_k(f)=<\Omega\Lambda (f),\,k>,$$ $\Omega$ is the Fourier $m\times m$ matrix, which is given by $\Omega=(\varepsilon^{-(i-1)(j-1)})_{i,j}$ and $\Lambda(f)(z)$ is the vector valued function form $U$ into $\mathbb{C}^{m}$, given by $\Lambda f(x)=^t(f(x),\,f(\varepsilon x),\,...,\,f(\varepsilon^{m-1}x)).$ \end{proposition} \begin{proof}Put \begin{equation*} \omega_k:=\sum_{i=1}^{m-1}k_{i}p_{i}. \end{equation*} A simple calculation shows that (see \cite{Martin}) \begin{equation*} \omega_k(f)=<\Omega \Lambda(f),\,k> \end{equation*} and $$T(k)f=\frac{df}{dx}+\frac{\omega_k(f)}{x}.$$ \end{proof} \begin{lemma} 1) If $f\in \mathcal{E}(U),$ then $T(k) (f)\in \mathcal{E}(U).$\\ 2) For $j=1,\,...,\,m-1,$ we have $$p_j\circ \frac{d}{dx}=\frac{d}{dx} \circ p_{j+1}.$$ Furthermore, if $f\in \mathcal{E}_j(U),$ then $T(k)( f)\in \mathcal{E}_j(U).$ \end{lemma} \begin{proof} This follows immediately from the fact that:\\ For $i=1,\,...,\,m-1,$ \begin{align*} p_{i}(f)(x)=\frac{1}{m}\sum_{j=0}^{m-1}\varepsilon^{-ij}f(\varepsilon^jx) =x\int_0^1p_{i-1}(f^{(1)})(xt)dt. \end{align*} \end{proof} \section{The hyper-Bessel functions} Let $\nu=(\nu_1,\,...,\,\nu_{m-1})\in \mathbb{R}^{m-1},$ satisfying $\nu_k\geq -1+\frac{k}{m},$ we denote by \begin{align*}&|\nu|:=\nu_1+...+\nu_{m-1},\\& \nu+\mathbf{n}:=(\nu_1+n,\,...,\,\nu_{m-1}+n)\,(n\in \mathbb{N}),\\& \Gamma(\nu):=\Gamma(\nu_1)...\,\Gamma(\nu_{m-1}).\end{align*} The normalized hyper-Bessel function with vector index $\nu$ is defined by (see, \cite{Klyu}, \cite{Dularue}, \cite{Dimov2}) \begin{align} \mathcal{J}_{\nu,m}(x):&=(\frac{x}{m})^{-|\nu|}\Gamma(\nu+\mathbf{1})J_{\nu,m}(x)\label{normalized} =\sum_{n=0}^{\infty}\frac{(-1)^n\Gamma(\nu+\mathbf{1})} {n!\Gamma(\nu+\mathbf{n}+\mathbf{1})}(\frac{ x}{m})^{nm }\nonumber.\end{align} Here $J_{\nu,m}(x)$ is the hyper-Bessel function \cite{Dularue}. The function $\mathcal{J}_{\nu,m}(\lambda x)$ is a unique $C^{\infty}$-solution of the following problem \cite{Klyu} \begin{equation} \left\{ \begin{array}{l l} B_m(f)(x)=-\lambda^m f(x), \\ f(0)=1,\,f^{(1)}(0)=\,...\, =f^{(m-1)}(0)=0. \end{array} \right. \end{equation} where the hyper-Bessel is given by \begin{equation} B_m=\prod_{j=1}^{m-1}(\frac{d}{dx}+\frac{m\nu_j+m-j}{x})\frac{d}{dx}. \end{equation} The simplest higher order hyper-Bessel operator is the operator of $m$-fold differentiation \begin{align*} \frac{d^m}{dx^m} =x^{-m}(x\frac{d}{dx})(x\frac{d}{dx}-1)...(x\frac{d}{dx}-m+1). \end{align*} For $m=2$ and $a_1=2\nu+1,$ $(\nu>-1/2)$ the hyper-Bessel operator generalizes the well known second order differential operator of Bessel $B_2$ given by where \begin{equation} B_2:=\frac{d^2}{dx^2}+\frac{2\nu+1}{x}\frac{d}{dx},\label{operaBes} \end{equation} and the corresponding normalized Bessel function is given by \begin{equation*} \mathcal{J}_{\nu,2}(x):=\frac{2^{\nu}\Gamma(\nu+1)}{x^{\nu}}J_{\nu}(x), \end{equation*} where $J_{\nu}(x)$ is the classical Bessel function (see, \cite{Wat}). From Corollary 2 in \cite{Klyu} we obtain the following differential recurrence relations for the normalized hyper-Bessel functions $\mathcal{J}_{\nu,m}(x)$ \begin{align} &\frac{d}{dx}\mathcal{J}_{\nu,m}(x)=-\frac{(\frac{x}{m})^{m-1}}{(\nu_1+1)\,...\,(\nu_{m-1}+1)}\mathcal{J}_{\nu+\mathbf{1},m}(x),\label{recurrence1}\\& (\frac{d}{dx}+\frac{m\nu_k}{x})\mathcal{J}_{\nu,m}(x)=\frac{m\nu_k}{x}\mathcal{J}_{\nu-e_k,m}(x),\label{recurrence2} \end{align} where $e_k, \, (1\leq k\leq m-1)$ are the standard basis of $\mathbb{R}^{m-1}$. \section{Intertwining operator} Let $\nu=(\nu_1,\,...,\,\nu_{m-1})\in \mathbb{C}^{m-1}$ such that $\Re(\nu_j)>0$. We define the fractional integrals $\mathcal{R}_{\nu,m}$ of Riemann-Liouville type for $f\in \mathcal{E}_m(U)$ ($\mathcal{E}_m(U)$ the subspace of $\mathcal{E}(U)$ of functions of type $m$) by \begin{align} \mathcal{R}_{\nu,m}f(x)&:=\frac{m^{3/2}\Gamma(\nu+\mathbf{1})}{(2\pi)^{(m-1)/2}}\int_0^1G_{m-1,m-1}^{m-1,0}\left(\left. \begin{matrix} \nu_{1}, \nu_{2}, ..., \nu_{m-1}\\ -\frac{1}{m},\,...,\, -\frac{m-1}{m}\end{matrix}\right \vert t\right)f(xt^{\frac{1}{m}})dt\label{Dimovstrans}, \end{align} where $G_{p,q}^{m,n}\left(\left. \begin{matrix} \ a_{1}, a_{2}, ..., a_{p}\\ \ b_1,b_{2},...,b_q\end{matrix}\right \vert z\right) $ is the Meijer's function (see \cite{Erdely}). This operator intertwines the hyper-Bessel operator $B_m$ and the $m$-th differential operator $\frac{d^m}{dz^m}$ \begin{equation} B_m\circ \mathcal{R}_{\nu,m}=\mathcal{R}_{\nu,m}\circ \frac{d^m}{dz^m}, \label{intertB} \end{equation} and maps the hyper-cosine function $\cos_m(\lambda x)$ \eqref{cosine} of order $m\geq 2$ into a normalized hyper-Bessel function $\mathcal{J}_{\nu,m}$ $$\mathcal{J}_{\nu,m}(\lambda x)=\mathcal{R}_{\nu,m}(\cos_m(\lambda\,.))(x).$$For $m=2,$ $\mathcal{R}_{\nu,m}$ is reduced to the so called Riemann-Liouville transform $\mathcal{R}_{k}$ defined in \eqref{Riemann1}. The operator $\mathcal{R}_{\nu}^m$ can be written also as a product of the Erd\'elyi-Kober integrals \begin{equation} \mathcal{R}_{\nu,m}f(x)=\frac{m^{3/2}\Gamma(\nu+\mathbf{1})}{(2\pi)^{(m-1)/2}}\prod_{k=1}^mI_{m-1}^{(\frac{k}{m},\,\nu_k+1-\frac{k}{m})}f(x), \end{equation} where the Erd\'elyi-Kober fractional integrals is defined by \begin{equation} I^{\alpha,\,\beta}_{\gamma}f(x):=\int_{0}^1\frac{(1-t)^{\alpha-1}t^{\beta}}{\Gamma(\alpha)} f(xt^{\frac{1}{\gamma}})dt,\,\,\, Re(\alpha)>0,\,\,Re(\beta)>0,\,\,Re(\gamma)>0. \end{equation} By Theorem 3.5.7 in \cite{kiryakova2} and by similar argument as \cite{Trimeche}, we can show that the operator $\mathcal{R}_{\nu,m}$ is a topological isomorphism from $\mathcal{E}_m(U)$ onto itself and its inverse is given by \begin{equation} \mathcal{R}_{\nu,m}^{-1}f(x)=\frac{(2\pi)^{(m-1)/2}}{m^{3/2}\Gamma(\nu+\mathbf{1})}\prod_{k=1}^m \prod_{j=1}^{n_k}(-1+j+\frac{k}{m}+\frac{1}{m}x\frac{d}{dx})I_{m-1}^{(\nu_k,\,n_k-\nu_k+\frac{k}{m}+1)}f(x), \end{equation}where \begin{equation} n_k=\left\{ \begin{array}{l l} [\nu_k-\frac{k}{m}+1]+1,\,\,\,\text{if}\,\,\,\, \nu_k-\frac{k}{m}\ \text{is non integer}, \\ \nu_k-\frac{k}{m}+1,\,\,\,\text{if}\,\,\,\,\, \nu_k-\frac{k}{m}\ \text{is integer}. \end{array} \right. \label{SSys1} \end{equation}Let consider the operator $V_m$ defined for $f\in \mathcal{E}(U)$ by \begin{align}V_m(f)&=\sum_{j=1}^mA_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j(f), \end{align} where the operator $I$ is defined in\eqref{rightinv} and \begin{align} & A_{m}=1, \,\,\,\,A_{m-1}=\frac{d}{dx}, \,\,\,\, A_j=\prod_{k=j+1}^{m-1}(\frac{d}{dx}+\frac{m\nu_{k}+m-k}{x})\frac{d}{dx},\,\,\, 1\leq j\leq m-2.\end{align} The operator $V_m$ is well defined on the space $\mathcal{E}(U),$ since for $f\in\mathcal{E}(U),$ we have $$I^{m-j}\circ p_j(f) \in \mathcal{E}_m(U).$$ \begin{theorem} The operator $V_m$ satisfy the following intertwining relation on the space $\mathcal{E}(U)$ $$T(k)\circ V_m=V_m \circ \frac{d}{dx}.$$ \end{theorem} \begin{proof} Let $f \in \mathcal{E}(U).$ It is clearly that for $j=1,\,...,\,m$, the function $$A_j\circ I^{m-j}\circ p_j(f)\in \mathcal{E}_j(U).$$ Then, \begin{align*}T(k)\circ V_m(f)&=\frac{d}{dz}\circ \mathcal{R}_{\nu,m}\circ p_m(f) + \sum_{j=1}^{m-1}(\frac{d}{dx}+\frac{k_j}{x})\circ A_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j(f)\\&=\frac{d}{dx}\circ \mathcal{R}_{\nu,m}\circ p_m(f) + B_{m}\circ \mathcal{R}_{\nu,m}\circ I^{m-1}\circ p_1(f)+ \sum_{j=2}^{m-1} A_{j-1}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j(f).\end{align*} On the other hand from \eqref{intertB}, we can write \begin{align*} B_{m}\circ \mathcal{R}_{\nu,m}\circ I^{m-1}\circ p_1=\mathcal{R}_{\nu,m}\circ \frac{d^m}{dz^m}\circ I^{m-1}p_1= \mathcal{R}_{\nu,m}\circ \frac{d}{dx}\circ p_1=A_m\circ\mathcal{R}_{\nu,m}\circ p_m \circ\frac{d}{dx}. \end{align*} Similarly \begin{align*} \frac{d}{dx}\circ \mathcal{R}_{\nu,m}\circ p_m=A_{m-1}\circ\mathcal{R}_{\nu,m}\circ I\circ p_{m-1}\circ \frac{d}{dx}.\end{align*} So that \begin{align*} \sum_{j=2}^{m-1} A_{j-1}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j&=\sum_{j=1}^{m} A_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j-1}\circ p_{j+1},\\&= \sum_{j=1}^{m-2} A_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ\frac{d}{dx}\circ p_{j+1}\\&= \sum_{j=1}^{m-2} A_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j\circ \frac{d}{dx}. \end{align*} Thus, \begin{align*}T(k)\circ V_m(f)=\sum_{j=1}^{m} A_{j}\circ \mathcal{R}_{\nu,m}\circ I^{m-j}\circ p_j\circ \frac{d}{dx}(f)=V_m\circ \frac{d}{dx}(f). \end{align*} \end{proof} \begin{theorem} Under the condition \begin{equation}k_j=m\nu_j+m-j\geq 0,\,\,j=1,\,...,\,m-1.\label{const}\end{equation}The following system \begin{equation} \left\{ \begin{array}{l l} T(k)f(x)=\kappa\lambda f(x), \\ f(0)=1 . \end{array} \right. \label{SSys1} \end{equation} has the following solution \begin{align} \mathcal{D}_{\nu}(\lambda,\, x)=\mathcal{J}_{\nu}( \lambda x)+\sum_{j=1}^{m-1}\frac{(\kappa\lambda)^{j}}{m^j(\nu_1+1)\,...\,(\nu_{m-j}+1)} \mathcal{J}_{(\nu_1+1,\dots,\nu_{j}+1,\nu_{j+1},...,\nu_{m-1})}(\lambda x).\label{Dunklkernel} \end{align} \end{theorem} \begin{proof}According to Theorem 4.1, $V_m$ intertwines $B_m$ and $\frac{d}{dx}$ in $\mathcal{E}(\mathbb{R})$. We apply the intertwines operator $V_m$ to the initial value problem \begin{equation} \left\{ \begin{array}{l l} f^{'}(x)= \kappa\lambda f(x), \\ f(0)=1. \end{array} \right. \label{SSys2} \end{equation} Then if $f$ is a solution of \eqref{SSys2} then $V_m(f)$ is a solution \eqref{SSys1}. Therefore $\mathcal{D}(\lambda,x)=V_m(e^{\kappa\lambda\,.})(x)$ is a solution of the system \eqref{SSys1}. Using \eqref{recurrence1} and \eqref{recurrence2} we can write $\mathcal{D}(\lambda,x)$ in form \eqref{Dunklkernel}.\end{proof}
\section{Introduction}\label{s:intro} Linear Quadratic optimal control problems (LQ in the following) are a standard topic in control theory and dynamical systems, and are very popular in applications. They consist in a linear control system with quadratic Lagrangian. We briefly recall the general features of a LQ problem, and we refer to \cite[Chapter 16]{AAAbook} and \cite[Chapter 7]{MR1425878} for further details. We are interested in \emph{admissible trajectories}, namely curves $x:[0,t_1]\to \mathbb{R}^n$ such that there exists a control $u \in L^2([0,t_1],\mathbb{R}^k)$ such that \begin{equation} \dot{x} = Ax +Bu, \qquad x(0) = x_0, \qquad x(t_1) = x_1,\qquad x_0,x_1,t_1 \text{ fixed}, \end{equation} that minimize a quadratic functional $\phi_{t_1}: L^2([0,t_1],\mathbb{R}^k) \to \mathbb{R}$ of the form \begin{equation} \phi_{t_1}(u) = \frac{1}{2}\int_0^{t_1} \left(u^*Ru + x^*P u + x^* Q x \right) dt. \end{equation} The condition $R\geq 0$ is necessary for existence of optimal control. We also assume $R>0$ (for the singular case we refer to \cite[Chapter 9]{MR1425878}). Without loss of generality we may reduce to the case \begin{equation} \phi_{t_1}(u) = \frac{1}{2}\int_{0}^{t_1} \left(u^* u - x^*Qx \right)dt. \end{equation} Here $A,B,Q$ are constant matrices of the appropriate dimension. The vector $Ax$ represents the \emph{drift} field, while the columns of $B$ represent the controllable directions. The meaning of the \emph{potential} term $Q$ will be clear later, when we will introduce the Hamiltonian associated with the LQ problem. We assume that the system is \emph{controllable}, namely there exists $m >0$ such that \begin{equation} \rank(B,AB,\ldots,A^{m-1}B) = n. \end{equation} This hypothesis implies that, for any choice of $t_1,x_0,x_1$, the set of controls $u$ such that the associated trajectory $x_u :[0,t_1] \to \mathbb{R}^n$ connects $x_0$ with $x_1$ in time $t_1$ is non-empty. It is well known that the optimal trajectories of the LQ system are projections $(p,x) \mapsto x$ of the solutions of the Hamiltonian system \begin{equation} \dot{p} = -\partial_x H (p,x), \qquad \dot{x} = \partial_p H (p,x), \qquad (p,x) \in T^*\mathbb{R}^n = \mathbb{R}^{2n}, \end{equation} where the Hamiltonian function $H: \mathbb{R}^{2n} \to \mathbb{R}$ is defined by \begin{equation}\label{eq:Hamiltonian} H(p,x) = \frac{1}{2}(p,x)^* \H \begin{pmatrix} p \\ x \end{pmatrix}, \qquad \H = \begin{pmatrix} BB^* & A \\ A^* & Q \end{pmatrix}. \end{equation} We denote by $P_t : \mathbb{R}^{2n} \to \mathbb{R}^{2n}$ the flow of the Hamiltonian system, which is defined for all $t \in \mathbb{R}$. To exploit the natural symplectic setting on $T^*\mathbb{R}^n = \mathbb{R}^{2n}$, we employ canonical coordinates $(p,x)$ such that the symplectic form $\omega = \sum_{i=1}^n dp_i \wedge dx_i$ is represented by the matrix $\Omega = \left(\begin{smallmatrix} 0 & \mathbb{I}_n \\ -\mathbb{I}_n & 0 \end{smallmatrix}\right)$. The flow lines of $P_t$ are precisely the integral lines of the \emph{Hamiltonian vector field} $\vec{H} \in \text{Vec}(\mathbb{R}^{2n})$, defined by $dH(\cdot) = \omega(\,\cdot\,,\vec{H})$. More explicitly \begin{equation} \vec{H}_{(p,x)}=\begin{pmatrix} -A^* & -Q \\ BB^* & A \end{pmatrix}\begin{pmatrix} p \\ x \end{pmatrix} = -\Omega\H \begin{pmatrix} p \\ x \end{pmatrix}. \end{equation} By the term Hamiltonian vector field, we denote both the linear field $\vec{H}$ and the associated matrix $-\Omega\H$. The Hamiltonian flow can be explicitly written in terms of the latter as \begin{equation} P_t = e^{-t\Omega\H}, \end{equation} where the r.h.s. is the standard matrix exponential. \subsection*{Conjugate times} We stress that not all the integral lines of the Hamiltonian flow lead to minimizing solutions of the LQ problem, since they only satisfy first order conditions for optimality. For this reason, they are usually called \emph{extremals}. Sufficiently short segments, however, are optimal, but they lose optimality at some time $t>0$, called the \emph{first conjugate time}. In the following, we give a geometrical definition of conjugate time, in terms of curves in the Grassmannian of Lagrangian subspaces of $\mathbb{R}^{2n}$. We say that a subspace $\Lambda \subset \mathbb{R}^{2n}$ is \emph{Lagrangian} if $\omega|_\Lambda \equiv 0$, and $\dim \Lambda = n$. A notable example of Lagrangian subspace is the \emph{vertical} subspace, that is $\mathcal{V} := \{(p,0)|\,p \in \mathbb{R}^n\}$. \begin{definition} The Jacobi curve $J(\cdot)$ is the following family of Lagrangian subspaces of $\mathbb{R}^{2n}$ \begin{equation} J(t):= e^{t\Omega\H}\mathcal{V}, \qquad \mathcal{V} :=\{(p,0)|\,p \in \mathbb{R}^n\}. \end{equation} From the geometrical viewpoint, $J(\cdot)$ is a smooth curve in the submanifold of the Grassmannian of the $n$-dimensional subspaces of $\mathbb{R}^{2n}$ defined by the Lagrangian subspaces. \end{definition} \begin{definition} We say that $t$ is a conjugate time if $J (t) \cap \mathcal{V} \neq 0$. The \emph{multiplicity} of the conjugate time $t$ is the dimension of the intersection. \end{definition} In the language introduced by V. Arnold, these are times of \emph{verticality} of the Jacobi curve. It is not hard to show that $t$ is a conjugate time if and only if there exist solutions of the Hamilton equations such that $x(0) = x(t) = 0$. We briefly recall the connection between conjugate times and second order conditions for optimality. The solutions of the LQ problems can be seen as constrained minima of the quadratic functional $\phi_{t_1}$ on $\mathcal{U}(x_0,x_1) \subset L^2([0,t_1],\mathbb{R}^k)$ given by all the controls $u$ such that $x_u(0) = x_0$ and $x_u(t_1) = x_1$. It is easy to check that $\mathcal{U}(x_0,x_1) = u^* + \mathcal{U}(0,0)$ for any $u^* \in \mathcal{U}(x_0,x_1)$, that is $\mathcal{U}(x_0,x_1)$ is an affine space over the vector space $\mathcal{U}(0,0)$. For this reason, the behaviour of $\phi_{t_1}$, restricted to $\mathcal{U}(0,0)$ provides all the informations about optimality. It is a well known fact that the number of conjugate times in the interval $(0,t_1)$, counted with their multiplicity, is equal to the negative inertia index of the quadratic form $\phi_{t_1} : \mathcal{U}(0,0) \to \mathbb{R}$ (this can be proved directly with the techniques in~\cite[Propositions 16.2, 16.3]{AAAbook}, see also~\cite[Theorem I.2]{NoteCime} for a more general setting). The occurrence of conjugate times implies that an extremal cannot be a minimizer, since one can find a small variation of $u^*$ that decreases the value of $\phi_{t_1}$. The first conjugate time determines existence and uniqueness of minimizing solutions of the LQ problem, as specified by the following proposition. \begin{prop} Let $\bar{t}$ be the first conjugate time, namely $\bar{t} := \inf\{t>0|\,J(t) \cap \mathcal{V} \neq 0\}$. \begin{itemize} \item For $t_1<\bar{t}$, for any $x_0,x_1$ there exists a unique minimizer connecting $x_0$ with $x_1$ in time $t_1$. \item For $t_1>\bar{t}$, for any $x_0,x_1$ there exists no minimizer connecting $x_0$ with $x_1$ in time $t_1$. \item For $t_1= \bar{t}$, existence of minimizers depends on the initial data. \end{itemize} \end{prop} In this paper we completely characterise the occurrence of conjugate times for a controllable LQ problem. In particular, we prove the following result. \begin{mainthm}\label{t:main} The conjugate times of a controllable linear quadratic optimal control problem obey the following dichotomy: \begin{itemize} \item If the Hamiltonian field $\vec{H}$ has at least one odd-dimensional Jordan block corresponding to a pure imaginary eigenvalue, the number of conjugate times in the interval $[0,T]$ grows to infinity for $T\to \pm\infty$. \item If the Hamiltonian field $\vec{H}$ has no odd-dimensional Jordan blocks corresponding to a pure imaginary eigenvalue, there are no conjugate times. \end{itemize} \end{mainthm} In Sec.~\ref{s:main}, we also provide estimates for the first conjugate time, in terms of the (signed) eigenvalues of $\vec{H}$ (see Corollaries~\ref{c:estimate1} and~\ref{c:estimate2}). Before passing to a more detailed description of curves of Lagrangian subspaces, we stress that the concept of Jacobi curves is not limited to LQ optimal control problems and can be defined for way more general geometrical structures, such as control systems with Tonelli Lagrangian including, among the others Riemannian, sub-Riemannian, Finsler and sub-Finsler manifolds. In these more general settings, however, we cannot exploit the natural linear structure of $\mathbb{R}^n$, and the Jacobi curve is a curve of subspaces of the tangent space to the cotangent bundle, associated with a fixed ``geodesic'' (i.e. locally minimizing curve) of the underlying structure. We refer the interested reader to \cite{curvature,geometryjacobi1,agrafeedback}. The plan of the paper is as follows. In Sec.~\ref{s:preliminaries} we recall some basic facts about geometry of curves in Lagrange Grassmannian, and the main technical tool: the Maslov index. Then, in Sec.~\ref{s:main} we prove the main result. \section{Main results}\label{s:main} We start this section by defining, more precisely, the class of dynamical systems under investigation. Let $(\Sigma,\sigma)$ be a symplectic vector space. \begin{definition}\label{d:LQ} A LQ optimal control problem is a pair $(H,\mathcal{V})$, where $H : \Sigma \to \mathbb{R}$ is a quadratic form (the Hamiltonian) and $\mathcal{V} \subset \Sigma$ is a Lagrangian subspace, such that $H|_\mathcal{V} \geq 0$. \end{definition} By choosing appropriate Darboux coordinates, $\Sigma = \mathbb{R}^{2n}$, $\omega = \Omega$, $\mathcal{V} = \{(p,0)|\,p \in \mathbb{R}^{n}\}$ and the Hamiltonian is \begin{equation}\label{eq:Hamiltonian2} H(p,x) = \frac{1}{2}(p,x)^* \H \begin{pmatrix} p \\ x \end{pmatrix}, \qquad \H = \begin{pmatrix} BB^* & A \\ A^* & Q \end{pmatrix}. \end{equation} Thus, Definition~\ref{d:LQ} is a coordinate-free characterization of the systems introduced in Sec.~\ref{s:intro}. With the pair $(H,\mathcal{V})$ we associate the Jacobi curve $J(t) = e^{t\Omega\H}\mathcal{V}$, which is a smooth curve in the Lagrange Grassmannian $\L(\Sigma)$. The assumption $H|_\mathcal{V} \geq 0$ is equivalent to the monotonicity of $J(\cdot)$. \begin{lem} The Jacobi curve of the system $(H,\mathcal{V})$ is monotone and equiregular. \end{lem} \begin{proof} Let $z \in J(t)$, then there exists $z_0 \in \mathcal{V}$ such that $z = e^{t\Omega\H}z_0$. The last formula also provides a smooth extension of $z$ belonging to the Jacobi curve for times close to $t$. Then, by definition of the quadratic form $\dot{J}(t)$, we obtain \begin{equation} \dot{J}(t)[z] = \omega(z,\dot{z}) = \omega\left(e^{t\Omega\H}z_0,e^{t\Omega\H}\Omega\H z_0\right) = \omega(z_0,\Omega\H z_0) = -z_0^* BB^* z_0 \leq 0, \end{equation} where we have used the fact that the Hamiltonian flow is a one-parameter group of symplectomorphisms. This proves that $\dot{J}(t) \leq 0$ as a quadratic form and the curve is monotone. Now observe that $J(t+\varepsilon) = e^{t\Omega\H} J(\varepsilon)$. This imples, by definition of $i$-th extension, that \begin{equation} J^{(i)}(t) = e^{t\Omega\H} J^{(i)}(0), \qquad i \geq 0, \end{equation} hence the $i$-th extensions have the same dimension for all $t$, and the curve is equiregular. \end{proof} \begin{lem} The system $(H,\mathcal{V})$ is controllable if and only if the Jacobi curve $J(\cdot)$ is ample. \end{lem} \begin{proof} By definition, the system $(H,\mathcal{V})$, which can be written as in Eq.~\eqref{eq:Hamiltonian2}, is controllable if \begin{equation} \rank(B,AB,\ldots,A^{m-1}B) = n. \end{equation} It is sufficient to prove that this is equivalent to ampleness at $t=0$, since ampleness at all $t$ follows from the equiregularity of the curve. Indeed, for small $t$, $J(t) = \{(p,S(t)p)|\,p \in \mathbb{R}^n\}$. We explicitly compute $S(t)$ as follows. Observe that \begin{equation} J(t) = e^{t\Omega\H}\begin{pmatrix} p \\ 0 \end{pmatrix} = \begin{pmatrix} \phi_{11}(t) & \phi_{12}(t) \\ \phi_{21}(t) & \phi_{22}(t) \end{pmatrix} \begin{pmatrix} p \\ 0 \end{pmatrix}, \qquad p \in \mathbb{R}^n. \end{equation} It is clear that $S(t) = \phi_{21}(t) \phi_{11}(t)^{-1}$. Then we can compute iteratively the derivatives of $S(t)$ at $t=0$, and we obtain, for any $m >0$ \begin{equation} \rank\{\dot{S}(0),\ddot{S}(0),\ldots,S^{m-1}(0)\} = \rank\{B,AB,\ldots,A^{m-1}B\}. \end{equation} Therefore controllability is equivalent to ampleness of the curve at $t=0$. \end{proof} We employ the symbol $\mathcal{H}$ to denote the set of controllable dynamical systems $(H,\mathcal{V})$ or, with no risk of confusion, the associated Hamiltonian vector fields $\vec{H}$. Since the associated Jacobi curve is monotone, ample and equiregular, Lemma~\ref{l:ampletrasv} and Corollary~\ref{c:ampletrasv} apply. This has important consequences on conjugate times. \begin{definition} We say that $\Gamma\subset \Sigma$ is an $\vec{H}$-invariant subspace if $P_t( \Gamma )= \Gamma$ for all $t \in \mathbb{R}$. \end{definition} \begin{prop} \label{lem_lagrangian_invariant_no_conj_points} Let $\vec{H} \in \mathcal{H}$. Suppose there exists an $\vec{H}$-invariant Lagrangian subspace $\Gamma \subset \Sigma$, then the Jacobi curve $J(\cdot)$ has no conjugate times. \end{prop} \begin{proof} Indeed, by Lemma~\ref{l:ampletrasv}, the Jacobi curve remains transversal to $\Gamma$ for all times. Then, by Corollary~\ref{c:ampletrasv}, the only intersection with $\mathcal{V} = J(0)$ can occur at $t=0$. \end{proof} Notice that the Lagrangian hypothesis is crucial. Indeed, Proposition~\ref{lem_lagrangian_invariant_no_conj_points} is false if the $\vec{H}$-invariant subspace is simply isotropic. \subsection{Proof of the main result} Now we are ready to prove Theorem~\ref{t:main}. By ``eigenvalues of the Hamiltonian'' we will mean the eigenvalues of $\Omega\H$, that is the matrix representing the Hamiltonian vector field $\vec{H}$. The proof is based on the following steps: \begin{itemize} \item[(i)] Assuming $\vec{H}$ diagonalizable, with pure imaginary spectrum, there are infinitely many conjugate times (Proposition~\ref{prop_pure_imaginary_spectrum}); \item[(ii)] Assuming $\vec{H}$ diagonalizable, with at least one pure imaginary eigenvalue, there are infinitely many conjugate times (Proposition~\ref{prop_at_least_one_pure_imag}). \item[(iii)] For a general $\vec{H}$, with at least one Jordan block of odd order corresponding to a pure imaginary eigenvalue, there are infinitely many conjugate times (Proposition~\ref{prop_oddjordan}). \end{itemize} \begin{itemize} \item[(iv)] For a general $\vec{H}$, if all Jordan blocks corresponding to pure imaginary eigenvalues are of even order, there are no conjugate times (Proposition~\ref{prop_noconjtimes}). \end{itemize} We directly prove (i). Then, with the techniques of Sec.~\ref{s:red}, we reduce (ii) and (iii) to the ``extremal'' case (i). We start by recalling an important property of the spectrum of Hamiltonian matrices as $\vec{H}$. If $\lambda$ is an eigenvalue, then also $\pm\lambda, \pm \bar{\lambda}$ are eigenvalues with the same multiplicity, where the bar denotes complex conjugation. Then, eigenvalues always appear in pairs (if $\lambda = \beta$ or $\lambda = i\beta$ for $\beta \in \mathbb{R}$) or in quadruples otherwise. We denote by $E_{\lambda} \subseteq \mathbb{R}^{2n}$ the real invariant subspace corresponding to the eigenvalues $\lambda,\bar{\lambda}$ of $\vec{H}$. This is the real vector space generated by the generalized eigenvectors $\xi$, $\bar{\xi}$ corresponding to the eigenvalues $\lambda$ and $\bar\lambda$, respectively. More precisely \begin{equation} E_{\lambda}:= \spn\{u,v \in \mathbb{R}^{2n} |\, u + iv \in \ker(\vec{H}-\lambda\mathbb{I})^k,\,k \geq 0\}. \end{equation} It is clear that $E_{\lambda} = E_{\bar{\lambda}}$. \begin{lem} \label{lem_H_orthog} Let $\lambda$ and $\lambda^\prime$ be eigenvalues of $\vec{H}$ (not necessarily distinct). If $\lambda + \lambda^\prime \neq 0$ and $\bar\lambda + \lambda^\prime \neq 0$ then $E_{\lambda}\Omega E_{\lambda^\prime}=0$. \end{lem} \begin{proof} For simplicity, we prove the theorem assuming $\vec{H}$ to be diagonalizable. Recall that $\vec{H} = -\Omega\H$ and $\Omega^2 = -\mathbb{I}$. Let $\xi$ and $\xi^\prime$ be eigenvectors corresponding to $\lambda$ and $\lambda^\prime$ respectively. Since $\Omega^2 = -\mathbb{I}$, we have $\xi^\prime\mathbf{H}\xi=\lambda\xi^\prime\Omega\xi$ and $\xi\mathbf{H}\xi^\prime=\lambda^\prime\xi\Omega\xi^\prime$ so $(\lambda+\lambda^\prime)\xi\Omega\xi^\prime=0$. Analogously, we obtain $\xi'H \bar{\xi} = \bar{\lambda} \xi' \Omega \bar\xi$ and $\bar{\xi}H\xi' = \lambda'\bar\xi\Omega\xi'$. Then $(\bar\lambda+\lambda^\prime)\bar\xi\Omega\xi^\prime=0$. Since $\lambda + \lambda^\prime \ne 0$ and $\bar\lambda + \lambda^\prime \ne 0$ it follows that $E_{\lambda}\Omega E_{\lambda^\prime}=0$. The above result still holds if $\vec{H}$ is not diagonalizable (see \cite[Lemma D.1, Chapter II]{MeyerHall}). \end{proof} \begin{rem} In particular if $\lambda=\alpha+i\beta$, with $\alpha\ne 0$ then $\Omega\rvert_{E_\lambda}\equiv 0$, i.e. $E_{\alpha+i\beta}$ is isotropic if $\alpha \neq 0$. \end{rem} It follows that the invariant subspaces associated with purely imaginary eigenvalues, non-purely imaginary eigenvalues, and $E_0$ are pairwise $\Omega$-orthogonal. This, together with the non-degeneracy of $\Omega$, implies the following decomposition in $\Omega$-orthogonal symplectic subspaces \begin{equation} \mathbb{R}^{2n} = \underbrace{E_0 \oplus \left(\bigoplus_{\alpha\neq 0} E_{\alpha+i\beta}\right)}_{\text{non pure imaginary}} \oplus\underbrace{\left( \bigoplus_{\beta \neq 0} E_{i\beta} \right)}_{\text{pure imaginary}}. \end{equation} In the following, with the term ``pure imaginary eigenvalue'' we understand all the eigenvalues $\lambda = i\beta$, with $\beta \neq 0$. \begin{lem}\label{lem_lagrangian_invariant} There exists an $\vec{H}$-invariant, Lagrangian subspace $\Gamma_+$ of the symplectic space $\displaystyle{\bigoplus_{\substack{\lambda \text{ non pure} \\ \text{imaginary}}}E_\lambda}$. \end{lem} \begin{proof} If zero is not an eigenvalue of $\vec{H}$, we take $\Gamma_+=\displaystyle{\bigoplus_{\alpha>0}}E_{\alpha+i\beta}$, which is $\vec{H}$-invariant by definition. If zero is an eigenvalue of $\vec{H}$, let us consider the corresponding invariant subspace $E_0$, with $\dim E_0=2m$. Choose an isotropic $m$-dimensional subspace $\Gamma_0\subset E_0$ (which is indeed $\vec{H}$-invariant). Hence $\Gamma_+=\Gamma_0\oplus\displaystyle{\bigoplus_{\substack{\alpha>0}}}E_{\alpha+i\beta}$ satisfies the required properties. \end{proof} \subsubsection{Diagonalizable case} In this section, we assume $\vec{H}$ to be diagonalizable. \begin{prop} \label{prop_pure_imaginary_spectrum} Let $\vec{H} \in \mathcal{H}$. Suppose that $\vec{H}$ is diagonalizable and has a pure imaginary spectrum. Then the Jacobi curve $J(\cdot)$ has infinitely many conjugate times. \end{prop} \begin{proof} If $\vec{H}$ has only pure imaginary eigenvalues, it is well known (see e.g. \cite[Appendix 6]{MathMethCM}) that there exists a symplectic change of coordinates such that the Hamiltonian can be written as \begin{equation}\label{eq:normalpureim} H(p,x)=\frac{1}{2}\sum_{j=1}^n \omega_j(p_j^2+x_j^2),\qquad \omega_1\ge \omega_2\ge \dots \ge \omega_n. \end{equation} Notice that the eigenvalues of $\vec{H}$ are $\pm i\omega_j$, $j=1,\dots,n$. The signs of the $\omega_j$ are precisely the signs of $H$ on the real eigenspaces $E_{i\omega_j}$. The following two lemmas are crucial. \begin{lem}[Givental' \cite{Givental}]\label{lem_cond_omega_j+omega_{n-j+1}} There exists a Lagrangian subspace $\Lambda\subset \mathbb{R}^{2n}$ such that $H\rvert_\Lambda >0$ if and only if $\omega_j+\omega_{n-j+1}>0$, $j=1,\dots,n$. \end{lem} \begin{lem}[Fa{\u\i}busovich \cite{ExUnRiccEq}]\label{lem_controllable_admits_positive_lag} Under the controllability assumption (or, equivalently, the ampleness of the Jacobi curve), there exists a Lagrangian subspace $\Lambda \subset \mathbb{R}^{2n}$ such that $H\rvert_\Lambda>0$. \end{lem} Lemmas~\ref{lem_controllable_admits_positive_lag} and~\ref{lem_cond_omega_j+omega_{n-j+1}} imply the following inequality: \begin{equation}\label{eq:inequality} \sum_{j=1}^n \omega_j>0. \end{equation} Now, let us define a new curve $L(t):=P_t (L_0)$ in $\L(\mathbb{R}^{2n})$, where $L_0:=\{(p,0):p\in \mathbb{R}^n\}\subset \mathbb{R}^{2n}$, $L_0\in \L(\mathbb{R}^{2n})$. \begin{rem} Notice that, in order to bring the Hamiltonian to the normal form of Eq.~\eqref{eq:normalpureim}, we have done a symplectic change of basis. Thus, in general, $L_0 \neq \mathcal{V}$. \end{rem} If we reorder coordinates in such a way that $(p,x) \mapsto (p_1,x_1,\ldots,p_n,x_n)$, we can write \begin{equation} L(t)=\begin{pmatrix}r(t\omega_1) & & \\ & \ddots & \\ & & r(t\omega_n)\end{pmatrix}L_0, \end{equation} where $r(t\omega_j)$ is a rotation of angle $t\omega_j$ in the $2$-dimensional subspace $(p_j,x_j)$. Observe that, given $t>0$ we can choose $\varepsilon>0$ sufficiently small such that $L(\varepsilon)\cap L_0=L(t+\varepsilon)\cap L_0= 0$. Therefore the Maslov index $L_{[\varepsilon,t+\varepsilon]}\cdot \mathcal{M}_{L_0}$ is well defined, since the endpoints of the curve are transversal to the train. We employ the shorthand $L_{(0,t)}\cdot\mathcal{M}_{L_0} = L_{[\varepsilon,t+\varepsilon]}\cdot \mathcal{M}_{L_0}$, for any $\varepsilon$ sufficiently small, and similar notation is understood every time a small variation of the end-times is required. We now prove that the index $L_{(0,+\infty)}\cdot\mathcal{M}_{L_0}$ is infinite. Intersections with the train occur at each half-rotation in each $2$-dimensional subspace $(p_j,x_j)$, with a sign given by the sign of $\omega_j$. Therefore, by a direct computation, we have \begin{equation} L_{(0,T)}\cdot\mathcal{M}_{L_0} = \sum_{j=1}^{n}\lfloor \frac{T\omega_j}{\pi}\rfloor > \sum_{j=1}^{n}\frac{T\omega_j}{\pi}-n. \end{equation} Inequality~\eqref{eq:inequality} implies that there are no compensations in the sum of the signs in the computation of the Maslov index. Indeed, let $N>0$ fixed. Since $\sum_{j=1}^n \omega_j>0$ we can take $T \ge \frac{(N+n)\pi}{\sum_{j=1}^{n}\omega_j}$ so that \begin{equation}\label{ineq_conj_time} L_{(0,T)}\cdot\mathcal{M}_{L_0} > N. \end{equation} This implies that the Maslov index of the curve $L(t) = P_t (L_0)$ with the train $\mathcal{M}_{L_0}$ grows to infinity for $T \to \infty$. On the other hand, the number of conjugate times (counted with multiplicity) is the Maslov index of the Jacobi curve $J(t) = P_t (\mathcal{V})$ with the train $\mathcal{M}_{\mathcal{V}}$. Thus, by combining Proposition~\ref{p:change1} and~\ref{p:change2}, we obtain \begin{equation} \lvert J_{(0,T)}\cdot \mathcal{M}_{\mathcal{V}} - L_{(0,T)}\cdot \mathcal{M}_{L_0} \rvert \le 2n. \label{ineq_prop5&6} \end{equation} Therefore \begin{equation} J_{(0,T)}\cdot\mathcal{M}_\mathcal{V} > \frac{\sum_{j=1}^n\omega_j}{\pi}T-3n. \end{equation} Thus $J(\cdot)$ has infinitely many conjugate times. \end{proof} As a corollary of the proof of Proposition~\ref{prop_pure_imaginary_spectrum}, we can give an estimate for the first conjugate time of a LQ optimal control problem. \begin{cor}\label{c:estimate1} Suppose the Hamiltonian can be written as $H(p,x)=\frac{1}{2}\sum_{j=1}^n \omega_j(p_j^2+x_j^2)$. Then, if $T \geq \frac{(N+3n-1)\pi}{\sum_{j=1}^n\omega_j}$ there are at least $N$ conjugate times (counted with multiplicity) in the interval $(0,T]$. In particular, the first conjugate time $\bar{t}$ satisfies $\bar{t}\le \frac{3n\pi}{\sum_{j=1}^{n}\omega_j}$. \end{cor} Now we are ready to discuss the case in which both pure and non pure imaginary eigenvalues occur in the spectrum of $\vec{H}$. \begin{prop}\label{prop_at_least_one_pure_imag} Let $\vec{H} \in \mathcal{H}$. Assume that $\vec{H}$ is diagonalizable and has at least one pure imaginary eigenvalue. Then the associated Jacobi curve $J(\cdot)$ has infinitely many conjugate times. \end{prop} \begin{proof} We reduce the problem to the extremal case of Proposition~\ref{prop_pure_imaginary_spectrum}. Consider $\Gamma_+$ as in Lemma~\ref{lem_lagrangian_invariant}, and let $\dim\Gamma=k$ (we drop the index $+$ from now on). Recall that $\Gamma$ is an $\vec{H}$-invariant isotropic subspace of $\Sigma=\mathbb{R}^{2n}$. We will consider the Lagrange Grassmannian of the reduced space $\Sigma^\Gamma=\Gamma^\angle/\Gamma$. Notice that, by Lemma~\ref{l:ampletrasv}, the Jacobi curve remains transversal to $\Gamma$ for all times. Thus, by Lemma~\ref{l:ampleproj}, the reduced Jacobi curve $J^\Gamma(\cdot)$ is a smooth, ample, monotone curve in $\L(\Sigma^\Gamma)$. By construction, we have \begin{equation} \Gamma^\angle=\Gamma \oplus \displaystyle{\bigoplus_{\substack{\lambda \text{ pure} \\ \text{imaginary}}}E_\lambda}, \qquad \Sigma^\Gamma=\displaystyle{\bigoplus_{\substack{\lambda \text{ pure} \\ \text{imaginary}}}E_\lambda}. \end{equation} Therefore we reduced the problem to the case of purely imaginary spectrum, and we can apply Proposition~\ref{prop_pure_imaginary_spectrum} to conclude that $J^\Gamma(\cdot)$ has infinitely many conjugate times. Notice that conjugate times for $J^\Gamma(\cdot)$ are intersections with $\mathcal{V}^\Gamma:=\pi(\mathcal{V}) = (\Gamma^\angle \cap\mathcal{V}) / \Gamma$. This means that the original curve has infinitely many intersections with $\mathcal{V}^\Gamma\oplus \Gamma$. More precisely, as we obtained in the proof of Proposition~\ref{prop_pure_imaginary_spectrum}, and recalling that $\dim \Sigma^\Gamma = 2(n-k)$ we have \begin{equation} J_{(0,T)} \cdot\mathcal{M}_{\mathcal{V}^\Gamma\oplus\Gamma} >\frac{\sum_{j=1}^{n-k}\omega_j}{\pi}T - 3(n-k). \end{equation} By applying again Proposition~\ref{p:change1}, we obtain \begin{equation} \lvert J_{(0,T)}\cdot \mathcal{M}_{\mathcal{V}} - J_{(0,T)}\cdot \mathcal{M}_{\mathcal{V}^\Gamma\oplus \Gamma} \rvert \le n. \end{equation} Therefore \begin{equation} J_{(0,T)} \cdot\mathcal{M}_{\mathcal{V}} >\frac{\sum_{j=1}^{n-k}\omega_j}{\pi}T - 4n +3k. \end{equation} Then $J(\cdot)$ has infinitely many conjugate times as well. \end{proof} Again, we give an estimate for the number of conjugate times as a separate corollary. \begin{cor}\label{c:estimate2} Suppose the Hamiltonian, restricted to $\displaystyle{\bigoplus_{\substack{\lambda \text{ pure} \\ \text{imaginary}}}E_\lambda}$, can be written as $H(p,x)=\frac{1}{2}\sum_{j=1}^{n-k} \omega_j(p_j^2+x_j^2)$. Then if $T \geq \frac{(N+4n-3k-1)\pi}{\sum_{j=1}^{n-k}\omega_j}$, there are at least $N$ conjugate times (counted with multiplicity) in the interval $(0,T]$. In particular, the first conjugate time $\bar{t}$ satisfies $\bar{t}\le \frac{(4n-3k)\pi}{\sum_{j=1}^{n-k}\omega_j}$. \end{cor} \subsubsection{General case} Now, let us consider an arbitrary $\vec{H}$. We approach the problem with the same basic techniques devised for the diagonalizable case. Let $\lambda=i\beta$, $\beta \neq 0$ a pure imaginary eigenvalue of $\vec{H}$. Recall that, by Lemma~\ref{lem_H_orthog}, $E_{i\beta}$ is $\Omega$-orthogonal to all the others $E_{\lambda'}$, with $\lambda' \neq \pm i \beta$. Therefore $E_{i\beta}$ is symplectic. It is well known that there exists a symplectic change of coordinates on $E_{i\beta}$ such that the Hamiltonian $H\rvert_{E_{i\beta}}$ has one of the following normal forms (see \cite{Ciampi,Williamson} and \cite[Appendix 6]{MathMethCM}). \begin{enumerate} \item[(a)] If $\pm i\beta$ correspond to a pair of Jordan blocks of even order $2k$: \begin{multline}\label{ham_normal_form_2k} H(p,x) = \pm\frac{1}{2}\left[\sum_{j=1}^k\left(\frac{1}{\beta^2} x_{2j-1}x_{2k-2j+1}+x_{2j}x_{2k-2j+2}\right) - \beta^2\sum_{j=1}^{k}p_{2j-1}x_{2j} + \sum_{j=1}^k p_{2j}x_{2j-1} - \right. \\ - \left.\sum_{j=1}^{k-1}\left(\beta^2 p_{2j+1}p_{2k-2j+1}+p_{2j+2}p_{2k-2j+2}\right)\right]. \end{multline} \item[(b)] If $\pm\lambda$ correspond to a pair of Jordan blocks of odd order $2k+1$: \begin{multline}\label{ham_normal_form_2k+1} H(p,x) = \pm\frac{1}{2}\left[\sum_{j=1}^k\left(\beta^2 p_{2j}p_{2k-2j+2}+x_{2j}x_{2k-2j+2}\right)\right. -\sum_{j=1}^{2k}p_jx_{j+1} - \\ - \left.\sum_{j=1}^{k+1}\left(\beta^2 p_{2j-1}p_{2k-2j+3}+x_{2j-1}x_{2k-2j+3}\right)\right]. \end{multline} \end{enumerate} Notice that the dimension of $E_\lambda$ is $4k$ or $4k+2$, respectively. \begin{lem}\label{lem_Jordan_blocks_isotropic_subspace} Let $\lambda=i\beta$ a pure imaginary eigenvalue of $\vec{H}$. Thus \begin{enumerate} \item[(a)] If the Jordan block corresponding to $\lambda$ has even order $2k$ then there exists a Lagrangian $\vec{H}$-invariant subspace $\Gamma\subset E_{\lambda}$ (of dimension $2k$). \item[(b)] If the Jordan block corresponding to $\lambda$ has odd order $2k+1$ then there exists an isotropic $\vec{H}$-invariant subspace $\Gamma\subset E_{\lambda}$ of dimension $2k$. \end{enumerate} \end{lem} \begin{proof} Let us consider the first case. As we recall above, $H\rvert_{E_\lambda}$ can be written as in Eq.~\eqref{ham_normal_form_2k}. Then, a careful inspection shows that $\vec{H}|_{E_\lambda} = -\Omega\H|_{E_\lambda}$ has the structure, in coordinates $(p,x) \in \mathbb{R}^{4k}$, displayed in Fig.~\ref{f:a}. \begin{figure} \centering \subfigure[\label{f:a}]{\input{fig1.tex}}\qquad\qquad\qquad\qquad \subfigure[\label{f:b}]{\input{fig2.tex}} \caption{Block structure of the normal form of $\vec{H}|_{E_\lambda}$ for a pair of Jordan blocks of even order (case a) and odd order (case b). In case (a), $\dim E_{\lambda} = 4k$, and each box denotes the presence of a non-vanishing $2\times 2$ block. In case (b), $\dim E_{\lambda} = 4k+2$, and each box denotes the presence of a non-vanishing $1\times 1$ block. All other entries are zero.} \end{figure} Notice that, for what follows, we do not need to know the explicit form of each box. If $k$ is even, we choose $\Gamma=\{(p,x)\in \mathbb{R}^{4k}|\, p_{k+1}=\ldots=p_{2k}=x_1=\ldots=x_{k}=0\}$ and if $k$ is odd we set $\Gamma=\{(p,x)\in \mathbb{R}^{4k}|\, p_{k+2}=\ldots=p_{2k}=x_1=\ldots=x_{k+1}=0\}$. It is a simple check that, in both cases, $\Gamma$ is a $2k$-dimensional $\vec{H}$-invariant space, which is also isotropic by construction, and thus Lagrangian (since $\dim E_\lambda = 4k$). Now, suppose that the Jordan block corresponding to $\lambda$ has odd order $2k+1$. Thus $H|_{E_\lambda}$ can be written as in Eq.~\eqref{ham_normal_form_2k+1} and $\vec{H}\rvert_{E_\lambda} = -\Omega\H|_{E_\lambda}$ has the structure, in coordinates $(p,x) \in \mathbb{R}^{4k+2}$, displayed in Fig.~\ref{f:b}. Once again, we stress that we do not need the explicit form of each box. By choosing $\Gamma=\{(p,x)\in \mathbb{R}^{2n}|\, p_1=\ldots=p_{k+1}=x_{k+1}=\ldots=x_{2k+1}=0\}$, we get the required subspace. \end{proof} \begin{prop}\label{prop_oddjordan} Let $\vec{H} \in \mathcal{H}$. Suppose there exists at least one Jordan block of odd order corresponding to a pure imaginary eigenvalue of $\vec{H}$. Thus the Jacobi curve has infinitely many conjugate times. \end{prop} \begin{proof} We will reduce the problem to the diagonalizable case by studying the curve in a reduced space $\Sigma^\Gamma=\Gamma^\angle/\Gamma$. Let $\pm\lambda_1,\dots, \pm\lambda_m$ be the pure imaginary eigenvalues of $\vec{H}$ and let us consider, for each $i$, the quotient spaces $E_{\lambda_i}^{\Gamma_i}:=E_{\lambda_i}\cap\Gamma_i^\angle / \Gamma_i$, where the subspaces $\Gamma_i\subset E_{\lambda_i}$ are as in Lemma \ref{lem_Jordan_blocks_isotropic_subspace}. Notice that $\dim E_{\lambda_i}^{\Gamma_i}=0$ or $2$ depending on whether the Jordan block corresponding to $\lambda_i$ is even or odd, respectively. Now set $\Gamma=\Gamma_1\oplus\dots\oplus\Gamma_m\oplus\Gamma_+ $, where $\Gamma_+$ as in Lemma~\ref{lem_lagrangian_invariant}. Hence $\Sigma^\Gamma=E_{\lambda_1}^{\Gamma_1}\oplus\dots \oplus E_{\lambda_m}^{\Gamma_m}$, so if there is at least one $\lambda_i$ for which the corresponding Jordan block has odd order then $\vec{H}\rvert_{\Sigma^\Gamma}$ has nonempty pure imaginary spectrum and it is diagonalizable. Moreover, since the original Jacobi curve is ample and monotone, the reduced Jacobi curve $J^\Gamma(\cdot)$ is ample and monotone too by Lemma~\ref{l:ampleproj}. Thus the result follows from Proposition \ref{prop_pure_imaginary_spectrum}. \end{proof} \begin{prop}\label{prop_noconjtimes} Let $\vec{H} \in \mathcal{H}$. If all Jordan blocks of $\vec{H}$ corresponding to pure imaginary eigenvalues are of even order, the Jacobi curve has no conjugate times. \end{prop} \begin{rem} This proposition applies, in particular, when there are no pure imaginary eigenvalues. \end{rem} \begin{proof} By Lemma~\ref{lem_lagrangian_invariant_no_conj_points} it is enough to find an $\vec{H}$-invariant Lagrangian subspace $\Gamma\subset\Sigma$. Let $\pm\lambda_1,\dots, \pm\lambda_m$ be the pure imaginary eigenvalues of $\vec{H}$. By Lemma~\ref{lem_Jordan_blocks_isotropic_subspace} there exists a Lagrangian $\vec{H}$-invariant subspace $\Gamma_i\subset E_{\lambda_i}$ for each $i$. Set $\Gamma=\Gamma_1\oplus\dots\oplus\Gamma_m\oplus\Gamma_+ $, where $\Gamma_+$ is as in Lemma~\ref{lem_lagrangian_invariant}. \end{proof} \section{Curves in the Lagrange Grassmannian}\label{s:preliminaries} Let $(\Sigma, \omega)$ be a $2n$-dimensional symplectic vector space. Recall that subspace $\Lambda \subset \Sigma$ is called \emph{Lagrangian} if it has dimension $n$ and $\omega|_{\Lambda}\equiv 0.$ The \emph{Lagrange Grassmannian} $\L(\Sigma)$ is the set of all $n$-dimensional Lagrangian subspaces of $\Sigma$. \begin{prop}\label{p:lagrass} $\L(\Sigma)$ is a compact $n(n+1)/2$-dimensional submanifold of the Grassmannian of $n$-planes in $\Sigma$. \end{prop} \begin{proof} Let $\Delta \in \L(\Sigma)$, and consider the set $\Delta^{\pitchfork}:= \{\Lambda \in \L(\Sigma)\,|\, \Lambda \cap \Delta =0 \}$ of all Lagrangian subspaces transversal to $\Delta$. Clearly, the collection of these sets for all $\Delta \in \L(\Sigma)$ is an open cover of $\L(\Sigma)$. Then it is sufficient to find submanifold coordinates on each $\Delta^{\pitchfork}$. Let us fix any Lagrangian complement $\Pi$ of $\Delta$ (which always exists, though it is not unique). Every $n$-dimensional subspace $\Lambda \subset \Sigma$ that is transversal to $\Delta$ is the graph of a linear map from $\Pi$ to $\Delta$. Choose an adapted Darboux basis on $\Sigma$, namely a basis $\{e_i,f_i\}_{i=1}^n$ such that \begin{gather} \Delta = \spn\{f_1,\ldots,f_n\}, \qquad \Pi = \spn\{e_1,\ldots,e_n\}, \\ \omega(e_i,f_j) - \delta_{ ij} = \omega(f_i,f_j) = \omega(e_i,e_j) = 0, \qquad i,j=1,\ldots,n. \end{gather} In these coordinates, the linear map is represented by a matrix $S_{\Lambda}$ such that \begin{equation} \Lambda \cap \Delta=0 \Leftrightarrow \Lambda=\{(p,S_{\Lambda} p)| \,p \in \Pi\simeq \mathbb{R}^{n}\}. \end{equation} Moreover it is easy to see that $\Lambda \in \L(\Sigma) $ if and only if $ S_{\Lambda}=S_{\Lambda}^{*}$. Hence, the open set $\Delta^{\pitchfork}$ of all Lagrangian subspaces transversal to $\Delta$ is parametrized by the set of symmetric matrices, and this gives smooth submanifold coordinates on $\Delta^\pitchfork$. This also proves that the dimension of $\L(\Sigma)$ is $n(n+1)/2$. Finally, as a closed subset of a compact manifold, $\L(\Sigma)$ is compact. \end{proof} Fix now $\Lambda \in \L(\Sigma)$. The tangent space $T_{\Lambda }\L(\Sigma)$ to the Lagrange Grassmannian at the point $\Lambda$ can be canonically identified with the set of quadratic forms on the space $\Lambda$ itself, namely \begin{equation} T_{\Lambda}\L(\Sigma)\simeq Q(\Lambda). \end{equation} Indeed, consider a smooth curve $\Lambda(\cdot)$ in $\L(\Sigma)$ such that $\Lambda(0)=\Lambda$, and denote by $\dot{\Lambda}\in T_{\Lambda}\L(\Sigma)$ its tangent vector. For any point $z\in \Lambda$ and any smooth extension $z(t)\in \Lambda(t)$, we define the quadratic form \begin{equation} \dot{\Lambda}:= z \mapsto \omega(z,\dot{z}), \end{equation} where $\dot{z}= \dot{z}(0)$. A simple check shows that the definition does not depend on the extension $z(t)$. Finally, if in local coordinates $\Lambda(t)=\{(p,S(t)p)|\,p\in \mathbb{R}^{n}\}$, the quadratic form $\dot{\Lambda}$ is represented by the matrix $\dot S(0)$. In other words, if $z \in \Lambda$ has coordinates $p \in \mathbb{R}^n$, then $\dot{\Lambda}[z] = p^*\dot{S}(0)p$. \subsection{Transversality properties} In this section we introduce some important properties of curves in the Lagrange Grassmannian. Then we discuss the specific case of a Jacobi curve. Let $J(\cdot)\in \L(\Sigma)$ be a smooth curve in the Lagrange Grassmannian. For $i \in \mathbb{N}$, consider \begin{equation} J^{(i)}(t)=\tx{span}\left\{\frac{d^{j}}{dt^{j}}\ell(t)\bigg| \ \ell(t)\in J(t),\, \ell(t) \text{ smooth},\, 0\leq j \leq i\right\}\subset \Sigma, \qquad i\geq 0. \end{equation} \begin{def} \label{d:amplestar} The subspace $J^{(i)}(t)$ is the \emph{i-th extension} of the curve $J(\cdot)$ at $t$. The flag \begin{equation} J(t) = J^{(0)}(t)\subset J^{(1)}(t)\subset J^{(2)}(t)\subset \ldots\subset \Sigma, \end{equation} is the \emph{associated flag of the curve} at the point $t$. The curve $J(\cdot)$ is called: \begin{itemize} \item[(i)] \emph{equiregular} at $t$ if $\text{dim }J^{(i)}(\cdot)$ is locally constant at $t$, for all $i \in \mathbb{N}$, \item[(ii)] \emph{ample} at $t$ if there exists $N\in \mathbb{N}$ such that $J^{(N)}(t)=\Sigma$, \item[(iii)] \emph{monotone increasing} (resp. \emph{decreasing}) at $t$ if $\dot{J}(t)$ is non-negative definite (resp. non-positive definite) as a quadratic form. \end{itemize} \end{def} In coordinates, $J(t)=\{(p,S(t)p)|\ p\in \mathbb{R}^{n}\}$ for some smooth family of symmetric matrices $S(t)$. The curve is ample at $t$ if and only if there exists $N\in \mathbb{N}$ such that \begin{equation} \text{rank} \{\dot S(t), \ddot S(t),\ldots, S^{(N)}(t)\}=n. \end{equation} We say that the curve is equiregular, ample or monotone (increasing or decreasing) if it is equiregular, ample or monotone for all $t$ in the domain of the curve. A crucial property of ample, monotone curves is described in the following lemma. \begin{lem}\label{l:ampletrasv} Let $J(\cdot) \in \L(\Sigma)$ a monotone, ample curve at $t_0$. Then, for any fixed Lagrangian subspace $\Lambda$, there exists $\varepsilon > 0$ such that $J(t) \cap \Lambda = 0$ for $0<|t - t_0| <\varepsilon$. \end{lem} In other words, ample, monotone curves can intersect any fixed Lagrangian subspace $\Lambda$ only at a discrete set of times. \begin{proof} Without loss of generality, assume $t_0 =0$. Choose a Lagrangian splitting $\Sigma = \Lambda \oplus \Pi$, such that, for $|t|<\varepsilon$, the curve is contained in the chart defined by such a splitting. In coordinates, $J(t)=\{(p,S(t)p)|\ p\in \mathbb{R}^{n}\}$, with $S(t)$ symmetric. The curve is monotone, then $\dot{S}(t)$ is a semidefinite symmetric matrix. Without loss of generality, we assume $\dot{S}(t) \geq 0$. Assume that $J(0) \cap \Lambda \neq 0$. In coordinate, this means that $S(0)$ has some vanishing eigenvalues. We now show that the whole spectrum of $S(t)$ is strictly increasing in $t$, hence it moves away from zero for $t$ sufficiently small. Notice that $S(t) - S(0) = \int_{0}^t \dot{S}(\tau)d\tau \geq 0$, by the monotonicity assumption. Then, for any $z \in \mathbb{R}^n$, consider the smooth function $t \mapsto f_z(t):=z^*[S(t)-S(0)]z$, which is non-decreasing and vanishes at $t=0$. Moreover, $f_z(t)$ cannot be constantly zero on any interval of the form $[0,\delta)$, otherwise $z$ would be in the kernel of $S(t)-S(0)$ for all $t \in [0,\delta)$ and, a fortiori, in the kernel of all the derivatives $S^{(N)}(0)$, which is absurd by the ampleness hypothesis. Therefore, $f_z(t) >0$ for $0<t<\varepsilon$. Since $z$ is arbitrary \begin{equation}\label{eq:monotonicity} S(t) > S(0), \qquad 0<t < \varepsilon. \end{equation} Now, denote by $\lambda_1(t)\geq \ldots\geq\lambda_n(t)$ the eigenvalues of $S(t)$ at each fixed $t$. Then, by the Courant min-max principle, we have the following variational characterisation \begin{equation} \lambda_k(t) = \max\{\min\{x^*S(t)x|\, x \in U \subset \mathbb{R}^n,\, |x| =1 \}|\, \dim U = k\}, \qquad k=1,\ldots,n. \end{equation} Thus, by Eq.~\eqref{eq:monotonicity}, each eigenvalue is strictly increasing for $0<t<\varepsilon$. So, even if $S(0)$ has a non-trivial kernel, it becomes non-degenerate for sufficiently small small $t>0$. The same argument shows that this is true also for $t<0$. \end{proof} Observe that, if $\Lambda = J(0)$, then $S(0) = 0$ in any chart given by the splitting $\Sigma = J(0) \oplus \Pi$. Therefore, the proof of Lemma~\ref{l:ampletrasv} implies that all the eigenvalues of $S(t)$ are strictly non-zero for all $|t|<\varepsilon$, $t\neq 0$. If the curve is also monotone and ample, the only restriction on $\varepsilon$ comes from the fact that $J(t)$ must belong to the given coordinate chart. In particular, the eigenvalues of $S(t)$ are strictly positive for all $t>0$ (and strictly negative for $t<0$) at least until the first intersection of $J(\cdot)$ with $\Pi$ occurs. This means that $J(\cdot)$ cannot have further intersections with $J(0)$ until it crosses $\Pi$. Thus, we obtain the following. \begin{cor}\label{c:ampletrasv} Let $J(\cdot) \in \L(\Sigma)$ a monotone, ample curve, such that $J(\cdot) \cap \Pi = 0$, for some Lagrangian subspace $\Pi$. Then $J(\cdot)$ has no self-intersections, namely $J(t_1) \cap J(t_2) = 0$ for all $t_1 \neq t_2$. \end{cor} \subsection{Reduction}\label{s:red} Let $(\Sigma,\omega)$ be a symplectic vector space, and let $\Gamma \subset \Sigma$ be an isotropic subspace, namely $\omega|_{\Gamma}\equiv 0$. For any subspace $V \subset \Sigma$, we denote by the symbol $V^\angle$ the corresponding $\omega$-orthogonal subspace. \begin{definition} The reduction of $(\Sigma,\omega)$ with respect to an isotropic subspace $\Gamma$ is the symplectic space $(\Sigma^\Gamma,\omega)$, where \begin{equation} \Sigma^\Gamma:=\Gamma^\angle / \Gamma. \end{equation} \end{definition} The definition is well posed, since $\omega$ descends to a well-defined symplectic form on the quotient. Moreover, if $\dim\Sigma = 2n$ and $\dim\Gamma = k$, then $\Sigma^\Gamma$ is a $2(n-k)$-dimensional symplectic space. The projection $\pi^\Gamma : \L(\Sigma) \to \L(\Sigma^\Gamma)$, defined by $\Lambda \mapsto \Lambda \cap \Gamma^\angle /\Gamma$, is not even continuous in general. Nevertheless, the following lemma holds true. \begin{lem}\label{l:smoothproj} The restriction of $\pi^\Gamma$ to $\Gamma^\pitchfork := \{\Lambda \in \L(\Sigma)|\, \Lambda\cap\Gamma = 0\}$ is smooth. \end{lem} \begin{proof} Let $\Lambda \in \Gamma^\pitchfork$. We can always find a Lagrangian space $\Pi$ which contains $\Gamma$ and such that $\Pi\cap \Lambda = 0$. The proof is now trivial in charts given by a Darboux basis on the splitting $\Pi \oplus \Lambda$. Indeed, in these charts, the projection corresponds to take a $n-k \times n-k$ block of the representative matrix. \end{proof} The next lemma provides condition under which a monotone, ample Jacobi curve remains monotone and ample upon projection. \begin{lem}\label{l:ampleproj} Let $J(\cdot) \in \L(\Sigma)$ a monotone, ample (at $t_0$) curve such that $J(\cdot) \in \Gamma^\pitchfork$. Then the projection $J^\Gamma(\cdot):=\pi^\Gamma(J(\cdot))$ is a monotone, ample (at $t_0$) curve in $\L(\Sigma^\Gamma)$. \end{lem} \begin{proof} By Lemma~\ref{l:smoothproj}, the projection $J^\Gamma(\cdot)$ is still smooth. We prove the Lemma by analysing the coordinate presentation of the curve. Without loss of generality, we choose $t_0 = 0$. We find $\Pi \in \L(\Sigma)$ such that $\Gamma \subset \Pi$, and $\Sigma = J(0)\oplus \Pi$. Therefore, we introduce Darboux coordinates $(p,x) \in \mathbb{R}^{2n}$ such that, for small $t$ \begin{equation} \Pi = \{(0,x)|\,x\in\mathbb{R}^n\}, \qquad J(t) = \{(p,S(t)p)|\,p\in \mathbb{R}^n\}, \qquad S(0) =0. \end{equation} Moreover, if $\dim \Gamma = k$, we split $\mathbb{R}^n = \mathbb{R}^k \oplus \mathbb{R}^{n-k}$, and we write $x = (x_1,x_2)$ and $p=(p_1,p_2)$. Thus \begin{equation} \Gamma = \{((0,0),(x_1,0))|\,x_1 \in \mathbb{R}^k\}, \qquad \Gamma^\angle = \{((0,p_2),(x_1,x_2))|\, p_2,x_2 \in \mathbb{R}^{n-k}, \, x_1 \in \mathbb{R}^k\}. \end{equation} Accordingly, the matrix $S(t)$ splits as $\left(\begin{smallmatrix} S_{11}(t) & S_{12}(t) \\ S_{12}^*(t) & S_{22}(t) \end{smallmatrix}\right)$. In terms of these coordinates, and analogous coordinates on $\Sigma^\Gamma = \Gamma^\angle / \Gamma$, we obtain that the matrix representing the reduced curve is $S^\Gamma(t) := S_{22}(t)$ (which is a $n-k\times n-k$ symmetric matrix). More precisely \begin{equation} J^\Gamma(t) = \{(p_2,S_{22}(t) p_2)|\, p_2 \in \mathbb{R}^{n-k}\}. \end{equation} The original curve is monotone (say non-decreasing), then $\dot{S}(t) \geq 0$. Therefore, also $\dot{S}_{22}(t) \geq 0$, and $J^\Gamma(\cdot)$ is monotone too. We now prove that the reduced curve is still ample at $0$ if the original curve was. We assume $S(t)$ to be real-analytic, otherwise, it is sufficient to replace $S(t)$ with its Taylor polynomial of sufficiently high order. From the proof of Lemma~\ref{l:ampletrasv}, $S(t) > S(0) =0$ for $t>0$ sufficiently small. Thus, for all $y \in \mathbb{R}^{n-k}$, the function $t \mapsto y^*S_{22}(t)y$ is zero at $t=0$, and strictly positive for $t>0$. But an analytic function with these properties has at least a non-vanishing (strictly positive) derivative. Hence, for some $i >0$, $y^*S^{(i)}_{22}(0)y > 0$. Since this construction holds for any $y \in \mathbb{R}^{n-k}$, this implies \begin{equation} \rank\{\dot{S}_{22}(0),\ldots,S_{22}^{(N)}(0)\} = n-k, \end{equation} for some sufficiently large $N >0$. \end{proof} \subsection{Maslov index and conjugate times} \input{maslov.tex}
\section{Introduction} \label{sec:Introduction} The spin states of quantum dots (QDs) are promising platforms for quantum computation.\cite{loss:pra98, kloeffel:annurev13} In particular, remarkable progress has been made with $S$-$T_0$ qubits in lateral GaAs double quantum dots (DQDs),\cite{petta:sci05, foletti:nphys09, shulman:sci12, levy:prl02, klinovaja:prb12} where a qubit is based on the spin singlet ($S$) and triplet ($T_0$) state of two electrons in the DQD. In this encoding scheme, rotations around the $z$ axis of the Bloch sphere can be performed on a subnanosecond timescale \cite{petta:sci05} through the exchange interaction, and rotations around the $x$ axis are enabled by magnetic field gradients across the QDs.\cite{foletti:nphys09} The lifetimes of $S$-$T_0$ qubits have been studied with great efforts. When the qubit state precesses around the $x$ axis, dephasing mainly results from Overhauser field fluctuations, leading to short dephasing times $T_2^{*} \sim 10\mbox{ ns}$.\cite{khaetskii:prl02, merkulov:prb02, coish:prb04, petta:sci05, johnson:nat05, bluhm:prl10} This low-frequency noise can be dynamically decoupled with echo pulses,\cite{petta:sci05, barthel:prl10, bluhm:nphys11, medford:prl12} and long decoherence times $T_2 > 200\mbox{ $\mu$s}$ have already been measured.\cite{bluhm:nphys11} In contrast to $x$-rotations, precessions around the $z$ axis dephase predominantly due to charge noise.\cite{dial:prl13, higginbotham:prl14} Rather surprisingly, however, recent Hahn echo experiments by Dial \textit{et al.} \cite{dial:prl13} revealed a relatively short $T_2 \simeq \mbox{0.1--1 $\mu$s}$ and a power-law dependence of $T_2$ on the temperature $T$. The origin of the observed decoherence is so far unknown, although the dependence on $T$ suggests that lattice vibrations (phonons) may play an important role. In this work, we calculate the phonon-induced lifetimes of a $S$-$T_0$ qubit in a lateral GaAs DQD. Taking into account the spin-orbit interaction (SOI) and the hyperfine coupling, we show that one- and two-phonon processes can become the dominant decay channels in these systems and may lead to qubit lifetimes on the order of microseconds only. While the decoherence and relaxation rates due to one-phonon processes scale with $T$ for the parameter range considered here, the rates due to two-phonon processes scale with $T^2$ at rather high temperatures and obey power laws with higher powers of $T$ as the temperature decreases. Among other things, the qubit lifetimes depend strongly on the applied magnetic field, the interdot distance, and the detuning between the QDs. Based on the developed theory, we discuss how the lifetimes can be significantly prolonged. The paper is organized as follows. In Sec.~\ref{sec:SystemAndBasisStates} we present the Hamiltonian and the basis states of our model. In the main part, Sec.~\ref{sec:LargeDetuning}, we discuss the calculation of the lifetimes in a biased DQD and investigate the results in detail. In particular, we show that two-phonon processes lead to short dephasing times and identify the magnetic field direction at which the lifetimes peak. The results for unbiased DQDs are discussed in Sec.~\ref{sec:SmallDetuning}, followed by our conclusions in Sec.~\ref{sec:Conclusions}. Details and further information are appended. \section{System, Hamiltonian, and Basis States} \label{sec:SystemAndBasisStates} We consider a lateral GaAs DQD within the two-dimensional electron gas (2DEG) of an AlGaAs/GaAs heterostructure that is grown along the [001] direction, referred to as the $z$ axis. Confinement in the $x$-$y$-plane is generated by electric gates on the sample surface, and the magnetic field $\bm{B}$ is applied in-plane to avoid orbital effects. When the DQD is occupied by two electrons, the Hamiltonian of the system reads \begin{eqnarray} H &=& \sum_{j=1,2} \Bigl( H_0^{(j)} + H_{Z}^{(j)} + H_{\rm SOI}^{(j)} + H_{\rm hyp}^{(j)} + H_{\rm el-ph}^{(j)} \Bigr) \nonumber \\ & & + H_{C} + H_{\rm ph} , \end{eqnarray} where the index $j$ labels the electrons, $H_0$ comprises the kinetic and potential energy of an electron in the DQD potential, $H_Z$ is the Zeeman coupling, $H_{\rm SOI}$ is the SOI, $H_{\rm hyp}$ is the hyperfine coupling to the nuclear spins, $H_{\rm el-ph}$ is the electron-phonon coupling, $H_C$ is the Coulomb repulsion, and $H_{\rm ph}$ describes the phonon bath. \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{spectrum.pdf} \caption{The energy spectrum of the DQD calculated for the parameters described in the text. The $S$-$T_0$ qubit is formed by the eigenstates of type $|(1,1)S\rangle$ and $|(1,1)T_0\rangle$. } \label{fig:spectrum} \end{center} \end{figure} The electron-phonon interaction has the form \begin{equation} H_{\rm el-ph} = \sum_{\vec{q},s} W_s(\vec{q}) a_{\vec{q}s} e^{i\vec{q}\cdot\vec{r}} + \mbox{h.c.} , \label{eq:HelphBasicFormMainText} \end{equation} where $\vec{r}$ is the position of the electron, $\vec{q}$ is a phonon wave vector within the first Brillouin zone, $s \in \{l, t_1, t_2\}$ stands for the longitudinal ($l$) and the two transverse ($t_1, t_2$) phonon modes, and ``h.c.'' is the hermitian conjugate. The coefficient $W_s(\vec{q})$ depends strongly on $\bm{q}$ and $s$, and is determined by material properties such as the relative permittivity $\epsilon_r$, the density $\rho$, the speed $v_l$ ($v_t$) of a longitudinal (transverse) sound wave, and the constants $\Xi$ and $h_{14}$ for the deformation potential and piezoelectric coupling, respectively. The annihilation operator for a phonon of wave vector $\bm{q}$ and mode $s$ is denoted by $a_{\vec{q}s}$. The Hamiltonian \begin{equation} H_{\rm SOI} = \alpha \left( p_{x'} \sigma_{y'} - p_{y'} \sigma_{x'} \right) + \beta \left( p_{y'} \sigma_{y'} - p_{x'} \sigma_{x'} \right) \end{equation} contains both Rashba and Dresselhaus SOI. Here $p_{x'}$ and $p_{y'}$ are the momentum operators for the $x'$ and $y'$ axes, respectively. The latter coincide with the crystallographic axes [100] and [010], respectively, and $\sigma_{x'}$ and $\sigma_{y'}$ are the corresponding Pauli operators for the electron spin. We take into account the coupling to states of higher energy by performing a Schrieffer-Wolff transformation that removes $H_{\rm SOI}$ in lowest order.\cite{khaetskii:prb00, aleiner:prl01, golovach:prl04, golovach:prb08, stano:prb05, stano:prl06, raith:prl12} The resulting Hamiltonian $\widetilde{H}$ is equivalent to $H$, except that $H_{\rm SOI}$ is replaced by \begin{equation} \widetilde{H}_{\rm SOI} \simeq g \mu_B (\vec{r}_{\rm SOI} \times \vec{B}) \cdot \vec{\sigma}, \end{equation} where $g$ is the in-plane $g$ factor, $\bm{\sigma}$ is the vector of Pauli matrices, and \begin{equation} \vec{r}_{\rm SOI} = \left( \frac{y'}{l_R} + \frac{x'}{l_D} \right) \vec{e}_{[100]} - \left( \frac{x'}{l_R} + \frac{y'}{l_D} \right) \vec{e}_{[010]}. \end{equation} Here $x'$ and $y'$ are the coordinates of the electron along the main crystallographic axes, whose orientation is provided by the unit vectors $\vec{e}_{[100]}$ and $\vec{e}_{[010]}$, respectively. The spin-orbit lengths are defined as $l_R = \hbar/(m_{\rm eff} \alpha)$ and $l_D = \hbar/(m_{\rm eff} \beta)$, where $m_{\rm eff}$ is the effective electron mass in GaAs and $\alpha$ ($\beta$) is the Rashba (Dresselhaus) coefficient. For our analysis, the most relevant effect of the nuclear spins is the generation of an effective magnetic field gradient between the QDs, which is accounted for by $H_{\rm hyp}$. We note that this magnetic field gradient may also result from a nearby positioned micromagnet.\cite{laird:prl07, pioroladriere:nphys08, brunner:prl11} For details of $H$ and $\widetilde{H}$, see Appendix~\ref{sec:Hamiltonian}. The $S$-$T_0$ qubit in this work is formed by the basis states $\ket{(1,1)S}$ and $\ket{(1,1)T_0}$, where the notation $(m,n)$ means that $m$ ($n$) electrons occupy the left (right) QD. In first approximation, these states read \begin{eqnarray} \ket{(1,1)S} &=& \ket{\Psi_{+}} \ket{S} , \\ \ket{(1,1)T_0} &=& \ket{\Psi_{-}} \ket{T_0} , \end{eqnarray} with \begin{equation} \ket{\Psi_{\pm}} = \frac{\ket{\Phi_L^{(1)} \Phi_R^{(2)}} \pm \ket{\Phi_R^{(1)} \Phi_L^{(2)}}}{\sqrt{2}} , \end{equation} where the $\Phi_{L,R}(\bm{r})$ are orthonormalized single-electron wave functions for the left and right QD, respectively (see also Appendix~\ref{sec:BasisStates}).\cite{burkard:prb99, stepanenko:prb12} The spin singlet is \begin{equation} \ket{S} = \frac{\ket{\uparrow\downarrow} -\ket{\downarrow\uparrow}}{\sqrt{2}} , \end{equation} whereas \begin{equation} \ket{T_0} = \frac{\ket{\uparrow\downarrow} + \ket{\downarrow\uparrow}}{\sqrt{2}} , \end{equation} with the quantization axis of the spins along $\bm{B}$. Analogously, one can define the states $\ket{(1,1)T_+} = \ket{\Psi_{-}} \ket{\uparrow \uparrow}$ and $\ket{(1,1)T_-} = \ket{\Psi_{-}}\ket{\downarrow \downarrow}$, which are energetically split from the qubit by $\pm g \mu_B |\bm{B}|$. For our analysis of the phonon-induced lifetimes, a simple projection of $\widetilde{H}$ onto this 4D subspace of lowest energy is not sufficient, because \begin{equation} \sum_j \Bigl( \bra{\Psi_+}H_{\rm el-ph}^{(j)} \ket{\Psi_+} - \bra{\Psi_-}H_{\rm el-ph}^{(j)} \ket{\Psi_-} \Bigr) = 0 . \end{equation} That is, corrections from higher states must be taken into account in order to obtain finite lifetimes.\cite{golovach:prb08, meunier:prl07} The spectrum that results from the states considered in our model is plotted in Fig.\ \ref{fig:spectrum}. Depending on the detuning $\epsilon$ between the QDs, the lifetimes of the qubit are determined by admixtures from $\ket{(2,0)S}$, $\ket{(0,2)S}$, or states with excited orbital parts. \section{Regime of Large Detuning} \label{sec:LargeDetuning} \subsection{Effective Hamiltonian and Bloch-Redfield theory} \label{secsub:EffHamAndBlochRedfield} We first consider the case of a large, positive detuning $\epsilon$ at which the energy gap between $\ket{(0,2)S}$ and the qubit states is smaller than the orbital level spacing $\hbar \omega_0$. In this regime, contributions from states with excited orbital parts are negligible, and projection of $\widetilde{H}$ onto the basis states $|(1,1)T_0\rangle$, $|(1,1)S\rangle$, $|(1,1)T_+\rangle$, $|(1,1)T_-\rangle$, $|(0,2)S\rangle$, and $|(2,0)S\rangle$ yields \begin{widetext} \begin{equation} \widetilde{H} = \begin{pmatrix} P_T & \frac{\delta b_B}{2} & 0 & 0 & 0 & 0 \\ \frac{\delta b_B}{2} & V_+-V_-+P_T & \frac{\Omega}{\sqrt{2}} & -\frac{\Omega}{\sqrt{2}} & -\sqrt{2}t+P_{S}^\dagger & -\sqrt{2}t+P_{S} \\ 0 & \frac{\Omega}{\sqrt{2}} & E_Z + P_T & 0 & 0 & 0 \\ 0 & -\frac{\Omega}{\sqrt{2}} & 0 & -E_Z+P_T & 0 & 0 \\ 0 & -\sqrt{2}t+P_{S} & 0 & 0 & -\epsilon+U-V_-+P_{SR}& 0\\ 0 & -\sqrt{2}t+P_{S}^\dagger &0 & 0 & 0 & \epsilon+U-V_-+P_{SL} \\ \end{pmatrix} + H_{\rm ph} . \label{eq:matrix} \end{equation} \end{widetext} Here $P_T$, $P_S$, $P_S^\dagger$, $P_{SL}$, and $P_{SR}$ are the matrix elements of the electron-phonon interaction, $t$ is the tunnel coupling, $U$ is the on-site repulsion, $V_{\pm} = \langle \Psi_{\pm} | H_C | \Psi_{\pm}\rangle$, $E_Z = g \mu_B |\bm{B}|$, \begin{eqnarray} \Omega = g \mu_B &\bigl(& \langle\Phi_L|(\vec{r}_{\rm SOI}\times\vec{B})_z|\Phi_L\rangle \nonumber \\ & & - \langle\Phi_R|(\vec{r}_{\rm SOI}\times\vec{B})_z|\Phi_R\rangle \bigr) , \end{eqnarray} and $\delta b_B=2\langle(1,1)S|H_{\rm hyp}|(1,1)T_0\rangle$ (see also Appendix \ref{secsub:HyperfineInteraction}). We note that the energy in Eq.\ (\ref{eq:matrix}) was globally shifted by $\bra{(1,1)T_0} \bigl( H_0^{(1)}+ H_0^{(2)} + H_C \bigr) \ket{(1,1)T_0}$. Furthermore, we mention that the state $\ket{(2,0)S}$ is very well decoupled when $\epsilon$ is large and positive. In Eq.~(\ref{eq:matrix}), $\ket{(2,0)S}$ is mainly included for illustration purposes, allowing also for large and negative $\epsilon$ and for an estimate of the exchange energy at $\epsilon \simeq 0$. In order to decouple the qubit subspace $\{|(1,1)S\rangle, |(1,1)T_0\rangle\}$, we first apply a unitary transformation to $\widetilde{H}$ that diagonalizes $\widetilde{H} - \sum_{j} H_{\rm el-ph}^{(j)}$ exactly. Then we perform a third-order Schrieffer-Wolff transformation that provides corrections up to the third power in the electron-phonon coupling, which is sufficient for the analysis of one- and two-phonon processes. The resulting effective Hamiltonian can be written as $H_{\rm q} + H_{\rm q-ph}(\tau) + H_{\rm ph}$ in the interaction representation, where the time is denoted by $\tau$ to avoid confusion with the tunnel coupling. Introducing the effective magnetic fields $\vec{B_{\rm eff}}$ and $\vec{\delta B}(\tau)$ and defining $\vec{\sigma^\prime}$ as the vector of Pauli matrices for the $S$-$T_0$ qubit, \begin{equation} H_{\rm q}= \frac{1}{2} g \mu_B \vec{B_{\rm eff}} \cdot \vec{\sigma^\prime} \end{equation} describes the qubit and \begin{equation} H_{\rm q-ph}(\tau) = \frac{1}{2} g \mu_B \vec{\delta B}(\tau)\cdot \vec{\sigma^\prime} \end{equation} describes the interaction between the qubit and the phonons. The time dependence results from \begin{equation} H_{\rm q-ph}(\tau) = e^{i H_{\rm ph} \tau/\hbar} H_{\rm q-ph} e^{-i H_{\rm ph} \tau/\hbar} . \end{equation} For convenience, we define the basis of $\vec{\sigma^\prime}$ such that $B_{\textrm{eff},x} = 0 = B_{\textrm{eff},y}$. Following Refs.\ \onlinecite{golovach:prl04, borhani:prb06}, the decoherence time ($T_2$), the relaxation time ($T_1$), and the dephasing contribution ($T_{\varphi}$) to $T_2$ of the qubit can then be calculated via the Bloch-Redfield theory (see also Appendix~\ref{sec:BlochRedfieldTheory}), which yields \begin{gather} \frac{1}{T_2} = \frac{1}{2 T_1} + \frac{1}{T_\varphi}, \\ \frac{1}{T_1} = J_{xx}^+(\omega_Z)+J_{yy}^+(\omega_Z) , \\ \frac{1}{T_{\varphi}} = J_{zz}^+(0) , \end{gather} where $\hbar \omega_Z = J_{\rm tot} = |g\mu_B \vec{B_{\rm eff}}|$ and \begin{equation} J_{ii}^+(\omega)=\frac{g^2\mu_B^2}{2 \hbar^2}\int_{-\infty}^{\infty}\cos(\omega \tau)\langle\delta B_i(0)\delta B_i (\tau)\rangle d\tau . \end{equation} The correlator $\langle \delta B_i(0)\delta B_i (\tau) \rangle$ is evaluated for a phonon bath in thermal equilibrium and depends strongly on the temperature $T$. \subsection{Input parameters} \label{secsub:InputParameters} The material properties of GaAs are $g = -0.4$, $m_{\rm eff} = 6.1 \times 10^{-32} \mbox{ kg}$, $\epsilon_r \simeq 13$, $\rho = 5.32 \mbox{ g/cm$^3$}$, $v_l \simeq 5.1\times 10^3 \mbox{ m/s}$ and $v_t \simeq 3.0\times 10^3 \mbox{ m/s}$ (see also Appendix \ref{secsubsub:DisplacementOperator}),\cite{cleland:book, adachi:properties, ioffe:data} $h_{14} \simeq -0.16 \mbox{ As/m$^2$}$,\cite{adachi:properties, huebner:pss73, ioffe:data} and $\Xi\approx -8 \mbox{ eV}$.\cite{adachi:gaas, vandewalle:prb89} In agreement with $\omega_0 /(2 \pi) = 30\mbox{ GHz}$,\cite{dial:prl13} we set $l_c = \sqrt{\hbar/(m_{\rm eff} \omega_0)} \simeq 96 \mbox{ nm}$, which is the confinement length of the QDs due to harmonic confining potential in the $x$-$y$ plane. For all basis states, the orbital part along the $z$ axis is described by a Fang-Howard wave function \cite{fang:prl66} of width $3 a_z = \mbox{6 nm}$ (see Appendix \ref{sec:BasisStates}). Unless stated otherwise, we set $l_R=2\ \mu\rm m$ and $l_D=1\ \mu\rm m$,\cite{hanson:rmp07, khaetskii:prb01, winkler:book} where $l_D$ is consistent with the assumed $a_z$ (see also Appendix \ref{sec:InputParameters}).\cite{winkler:book} We note, however, that adapting $a_z$ to $l_D$ is not required, because changing the width of the 2DEG by several nanometers turns out not to affect our results. All calculations are done for $|\vec{B}| = 0.7 \mbox{ T}$,\cite{bluhm:prl10, shulman:sci12} $\delta b_B = -0.14 \mbox{ $\mu$eV}$, in good agreement with, e.g., Refs.\ \onlinecite{bluhm:prl10, dial:prl13}, and an interdot distance of $2a = 400 \mbox{ nm}$. For Figs.\ \ref{fig:spectrum}--\ref{fig:angular_dependence} (large $\epsilon$), we use $U = 1 \mbox{ meV}$, $t = 7.25\mbox{ $\mu$eV}$, and $V_+ = 40\mbox{ $\mu$eV}$.\cite{stepanenko:prb12} We choose here $V_- = 39.78\mbox{ $\mu$eV}$ such that the resulting energy splitting $J_{\rm tot}(\epsilon)$ between the qubit states is mostly determined by the hyperfine coupling at $\epsilon \to 0$, as commonly realized experimentally.\cite{petta:sci05, dial:prl13} The detuning $\epsilon$ is then set such that $0 < U- V_\pm -\epsilon < \hbar \omega_0$ and $J_{\rm tot} = 1.43 \mbox{ $\mu$eV}$, and we note that this splitting is within the range studied in Ref.\ \onlinecite{dial:prl13}. \subsection{Temperature dependence} \label{secsub:TemperatureDependence} \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{tem_dep.pdf} \caption{(a) Temperature dependence of the decoherence time ($T_2$, blue) and relaxation time ($T_1$, red) for the parameters in the text. The solid line corresponds to a power-law fit to $T_2$ for $\mbox{0.1 K} \leq T \leq \mbox{0.2 K}$, which yields $T_2 \propto T^{-3}$ and good agreement with recent experiments.\cite{dial:prl13} We note that $T_2 \ll T_1$. (b) The decoherence time due to one-phonon ($1/\Gamma_2^{\rm 1p}$) and two-phonon processes ($1/\Gamma_2^{\rm 2p}$) and the full decoherence time $T_2 = 1/\Gamma_2= 1/(\Gamma_2^{\rm 1p} + \Gamma_2^{\rm 2p})$ as a function of temperature. We note that $1/\Gamma_2^{\rm 2p}$ changes its behaviour from $\propto C_1+C_2 T^{-5}$ to $\propto T^{-2}$, where $C_1$ and $C_2$ are constants, whereas $1/\Gamma_2^{\rm 1p} \propto T^{-1}$ for the range of $T$ considered here. } \label{fig:tem_dep} \end{center} \end{figure} Figures \ref{fig:spectrum}--\ref{fig:different_spin_orbit} consider $\vec{B}$ applied along the $x$ axis that connects the two QDs, assuming that the $x$ axis coincides with the crystallographic $[110]$ direction. The geometry $x \parallel [110]$ is realized in most experiments,\cite{barthel:prl10, medford:prl12, higginbotham:prl14} particularly because GaAs cleaves nicely along [110]. In stark contrast to previous theoretical studies of phonon-limited lifetimes, where $T_2 = 2 T_1$,\cite{golovach:prl04, bulaev:prl05, trif:prl09, maier:prb13, hachiya:arX} Fig.\ \ref{fig:tem_dep}(a) reveals $T_2 \ll T_1$ at $30\mbox{ mK} \leq T \leq 1\mbox{ K}$ considered here, which implies $T_\varphi \ll T_1 $. In the discussion below we therefore focus on the details of the temperature dependence of $\Gamma_2 = 1/T_2$. We note, however, that the contributions to $\Gamma_2$ and $\Gamma_1 = 1/T_1$ from one-phonon processes scale similarly with $T$, and analogously for two-phonon processes. Defining $\Gamma_2^{\rm 1p}$ ($\Gamma_2^{\rm 2p}$) as the decoherence rate due to one-phonon (two-phonon) processes, Fig.\ \ref{fig:tem_dep}(b) illustrates $\Gamma_2^{\rm 2p} \gg \Gamma_2^{\rm 1p}$, and so $\Gamma_2 = \Gamma_2^{\rm 1p} + \Gamma_2^{\rm 2p} \simeq \Gamma_2^{\rm 2p}$. In the considered range of temperatures, we find $\Gamma_2^{\rm 1p} \propto T$. This behavior results from the fact that $\hbar \omega_{Z}/(k_B T)<1$ for our parameters, where $k_B$ is the Boltzmann constant. Therefore, the dominant terms in the formula for $\Gamma_2^{\rm 1p}$ are proportional to Bose-Einstein distributions defined as \begin{equation} n_B(\omega) = \frac{1}{e^{\hbar \omega/(k_B T)} - 1 } \label{eq:BoseEinsteinDistribution} \end{equation} and may all be expanded according to $n_B(\omega) \simeq k_B T/(\hbar \omega)$, keeping in mind that the $n_B(\omega)$ contributing to $\Gamma_2^{\rm 1p}$ are evaluated at $\omega = \omega_Z$ because of energy conservation. The time $1/\Gamma_2^{\rm 2p}$ due to two-phonon processes smoothly changes its behaviour from $C_1+C_2 T^{-5}$ at $T \sim 40\mbox{ mK}$ to $T^{-2}$ with increasing temperature, where $C_n$ are constants. This transition is explained by the fact that, in the continuum limit, the rate corresponds to an integral over the phonon wave vector $\vec{q}$, where the convergence of this integral is guaranteed by the combination of the Bose-Einstein distribution and the Gaussian suppression that results from averaging over the electron wave functions. More precisely, the decay rate is obtained by integrating over the wave vectors of the two involved phonons. Due to conservation of the total energy, however, considering only one wave vector $\vec{q}$ is sufficient for this qualitative discussion. For $\Gamma_2^{\rm 2p}$, we find that the dominating terms decay with $\bm{q}$ due to factors of type \begin{equation} f_s(\bm{q}) = e^{-(q_x^2+q_y^2)l_c^2} \hspace{0.02cm} n_B(\omega_{\bm{q}s}) \left[n_B(\omega_{\bm{q}s}) + 1 \right] , \end{equation} where $q_x$ and $q_y$ are the projections of $\bm{q}$ onto the $x$ and $y$ axis, respectively, and $\hbar \omega_{\bm{q}s} = \hbar v_s |\bm{q}|$ is the phonon energy. Whether the Bose-Einstein part or the Gaussian part from $f_s(\bm{q})$ provides the convergence of the integral depends on $l_c$, $v_s \in \{ v_l, v_t \}$, and mainly $T$, as the latter can be changed significantly. When the Gaussian part $\exp[-(q_x^2+q_y^2)l_c^2]$ cuts the integral, $\Gamma_2^{\rm 2p} \propto T^2$ due to the expansion $n_B(n_B+1)\simeq (k_B T)^2/(\hbar \omega_{\bm{q}s})^2$ that applies in this case. When $n_B(n_B+1)$ affects the convergence of the integral, terms with higher powers of $T$ occur. The resulting temperature dependence is rather complex, but is usually well described by $1/\Gamma_2^{\rm 2p} = C_m + C_n T^{-\nu}$ with $\nu \geq 2$ for different ranges of $T$ [see Fig.\ \ref{fig:tem_dep}(b)]. The temperature ranges for the different regimes are determined by the details of the setup and the sample. For the parameters considered here, a power-law approximation $T_2 \propto T^{\eta}$ for $T = \mbox{100--200 mK}$ yields $\eta \simeq -3$ mainly because of the dephasing due to two-phonon processes (see Figs.\ \ref{fig:tem_dep} and \ref{fig:different_spin_orbit}), which agrees well with the experimental data of Ref.~\onlinecite{dial:prl13}. \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{different_spin_orbit_temp_dep.pdf} \caption{Dependence of the decoherence time $T_2$ on the temperature for the parameters in the text and different spin-orbit lengths. Keeping the splitting $J_{\rm tot}$ between the qubit states constant, the values chosen for the detuning $\epsilon$ are \mbox{0.896 meV} (black), \mbox{0.912 meV} (blue), \mbox{0.918 meV} (green), and \mbox{0.933 meV} (red), increasing with increasing SOI. Within the range $T = \mbox{100--200 mK}$, $T_2 \propto T^{-3}$ in all cases. We note that the best quantitative agreement with the experiment \cite{dial:prl13} is obtained for the strongest SOI (red), where $l_R=1\ \mu\rm m$ and $l_D=0.5\ \mu\rm m$.} \label{fig:different_spin_orbit} \end{center} \end{figure} Figure \ref{fig:different_spin_orbit} shows the resulting temperature dependence of $T_2$ for different spin-orbit lengths. Remarkably, the calculation yields short $T_2$ even when SOI is completely absent. Keeping $J_{\rm tot} = 1.43\mbox{ $\mu$eV}$ fixed by adapting the value of $\epsilon$, one finds that $T_2$ decreases further with increasing SOI. As seen in Eq.~(\ref{eq:matrix}), $\widetilde{H}_{\rm SOI}$ couples $\ket{(1,1)S}$ to the triplet states $\ket{(1,1)T_+}$ and $\ket{(1,1)T_-}$. An important consequence of the resulting admixtures is that greater detunings are required in order to realize a desired $J_{\rm tot}$. In Fig.~\ref{fig:different_spin_orbit}, for instance, $\epsilon$ increases from $0.896\mbox{ meV}$ (no SOI) to $0.933\mbox{ meV}$ ($l_R = 1\mbox{ $\mu$m}$, $l_D = 0.5\mbox{ $\mu$m}$). As explained below, increasing $\epsilon$ decreases the lifetimes because it enhances the effects of $\ket{(0,2)S}$ through reduction of the energy gap (see also Fig.\ \ref{fig:spectrum}). \subsection{Origin of strong dephasing} \label{secsub:OriginStrongDephasing} The results discussed thus far have revealed two special features of the phonon-mediated lifetimes of $S$-$T_0$ qubits in biased DQDs. First, $T_2 \ll T_1$, as seen in Fig.~\ref{fig:tem_dep}(a). Second, the strong decay does not require SOI, as seen in Fig.~\ref{fig:different_spin_orbit}. These features have not been observed in previous calculations for, e.g., spin qubits formed by single-electron\cite{khaetskii:prb01, golovach:prl04} or single-hole\cite{bulaev:prl05, trif:prl09} or two-electron\cite{golovach:prb08} states in GaAs QDs, hole-spin qubits in Ge/Si nanowire QDs,\cite{maier:prb13} or electron-spin qubits in graphene QDs.\cite{hachiya:arX} Therefore, we discuss the dominant decay mechanism for $S$-$T_0$ qubits in DQDs in further detail and provide an intuitive explanation for our results. Assuming again a large, positive detuning $\epsilon$, with $0 < U - V_{\pm} - \epsilon < \hbar \omega_0$, and setting $\Omega = 0$ (no SOI), the states $|(1,1)T_+\rangle$, $|(1,1)T_-\rangle$, and $|(2,0)S\rangle$ of Eq.\ (\ref{eq:matrix}) are practically decoupled from the qubit. The relevant dynamics are then very well described by \begin{equation} \widetilde{H} = \begin{pmatrix} 0 & \frac{\delta b_B}{2} & 0 \\ \frac{\delta b_B}{2} & V_+-V_- & -\sqrt{2}t+P_{S}^\dagger \\ 0 & -\sqrt{2}t+P_{S} & -\epsilon+U-V_- + \widetilde{P}\\ \end{pmatrix} + H_{\rm ph} , \label{eq:Simple3x3MatrixMainText} \end{equation} with $\ket{(1,1)T_0}$, $\ket{(1,1)S}$, and $\ket{(0,2)S}$ as the basis states and \begin{equation} \widetilde{P} = P_{SR} - P_T . \end{equation} In the absence of SOI, the hyperfine interaction ($\delta b_B$) is the only mechanism that couples the spin states and enables relaxation of the $S$-$T_0$ qubit. We note that even when $\Omega$ is nonzero the relaxation times $T_1$ are largely determined by the hyperfine coupling instead of the SOI for the parameters considered in this work. At sufficiently large temperatures, where $T_2 \ll T_1$, $\delta b_B$ is negligible in the calculation of $T_2$, leading to pure dephasing, $T_2 = T_\varphi$. In addition, the matrix element $P_{S}$ turns out to be negligible for our parameters. Following Appendix~\ref{sec:SimpleModelDephasing}, we finally obtain \begin{equation} \frac{1}{T_2} = \frac{1}{T_\varphi} = \frac{2 t^4}{\hbar^2 (\Delta_S^\prime)^6} \int_{-\infty}^{\infty} \langle\widetilde{P}^2(0)\widetilde{P}^2(\tau)\rangle d\tau \label{eq:Simple3x3DecRateFinalMainText} \end{equation} from this simple model, where \begin{equation} \Delta_S^\prime = \sqrt{(U - V_{+} - \epsilon)^2 + 8 t^2} \end{equation} corresponds to the energy difference between the eigenstates of type $\ket{(1,1)S}$ and $\ket{(0,2)S}$ (using $\delta_B = 0$). We note that terms of type $a_{\bm{q}s}^\dagger a_{\bm{q}s}$ and $a_{\bm{q}s} a_{\bm{q}s}^\dagger$ must be removed from $\widetilde{P}^2$ in Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}), as the Bloch-Redfield theory requires $\langle \bm{\delta B}(\tau) \rangle$ to vanish (see also Appendix~\ref{sec:SimpleModelDephasing}).\cite{slichter:book} In Fig.~\ref{fig:ComparisonSimpleModel}, we compare $T_2$ from Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}) with $T_2$ derived from Eq.~(\ref{eq:matrix}) for $\Omega = 0$ (see also Fig.~\ref{fig:different_spin_orbit}), and find excellent agreement at $T \gtrsim 50\mbox{ mK}$ where relaxation is negligible. \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{simplified_3_x_3.pdf} \caption{Decoherence time $T_2$ as a function of temperature from two different models. The dotted line is also shown in Fig.~\ref{fig:different_spin_orbit} and was calculated via Eq.~(\ref{eq:matrix}), using the parameters in the text with $\Omega = 0$ (no SOI) and $\epsilon = 0.896\mbox{ meV}$. The crosses result from Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}), using exactly the same parameters. We note that the associated $J_{\rm tot}$ differ only slightly. The remarkable agreement demonstrates that the simple model of Sec.~\ref{secsub:OriginStrongDephasing} accounts for the dominant decay mechanism. At $T \lesssim 50\mbox{ mK}$, the curves start to deviate because relaxation is no longer negligible. When the hyperfine coupling in Eq.~(\ref{eq:Simple3x3MatrixMainText}) is not omitted, excellent agreement is obtained also at low temperatures. } \label{fig:ComparisonSimpleModel} \end{center} \end{figure} The above analysis provides further insight and gives explanations for the results observed in this work. First, Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}) illustrates that dephasing requires two-phonon processes and cannot be achieved with a single phonon only. As dephasing leaves the energy of the electrons and the phonon bath unchanged, the single phonon would have to fulfill $\omega_{\bm{q}s} = 0 = |\bm{q}|$. However, phonons with infinite wavelengths do not affect the lifetimes, which can be explained both via $e^{i\bm{q}\cdot\bm{r}} \to 1$ [see Eq.~(\ref{eq:HelphBasicFormMainText})] and via the vanishing density of states at $\omega_{\bm{q}s} \to 0$ for acoustic phonons in bulk. Thus, $\Gamma_2^{\rm 1p} = \Gamma_1^{\rm 1p}/2$ in all our calculations, where $\Gamma_1^{\rm 1p}$ is the relaxation rate due to one-phonon processes. Second, as discussed above, we find that the hyperfine interaction in combination with electron-phonon coupling presents an important source of relaxation in this system.\cite{raith:prl12} Third, the strong dephasing at large detuning $\epsilon$ results from two-phonon processes between states of type $\ket{(1,1)S}$ and $\ket{(0,2)S}$. This mechanism is very effective because the spin state remains unchanged. Therefore, the dephasing requires neither SOI nor hyperfine coupling, and we note that Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}) reveals a strong dependence of $T_\varphi$ on the tunnel coupling $t$ and the splitting $\Delta_S^\prime$. Hence, the short $T_\varphi$ in the biased DQD can be interpreted as a consequence of the Pauli exclusion principle. When the energy of the right QD is lowered ($\epsilon > 0$), the singlet state of lowest energy changes from $\ket{(1,1)S}$ toward $\ket{(0,2)S}$, since the symmetric orbital part of the wave function allows double-occupancy of the orbital ground state in the right QD. The triplet states, however, remain in the (1,1) charge configuration. While this feature allows tuning of the exchange energy and readout via spin-to-charge conversion on the one hand,\cite{petta:sci05} it enables strong dephasing via electron-phonon coupling on the other hand: effectively, phonons lead to small fluctuations in $\epsilon$; due to Pauli exclusion, these result in fluctuations of the exchange energy and, thus, in dephasing. This mechanism is highly efficient in biased DQDs, but strongly suppressed in unbiased ones, as we show in Sec.~\ref{sec:SmallDetuning} and Appendix~\ref{sec:SimpleModelDephasingZeroDetunSinglets}. \subsection{Angular dependence} \label{secsub:AngularDependence} \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{angular_dependence.pdf} \caption{Dependence of the relaxation $(T_1)$ and decoherence time $(T_2)$ on the angle $\theta_B$ between the in-plane magnetic field $\bm{B}$ and the $x$ axis that connects the QDs. When $\bm{B} \perp x$ ($\theta_B=\pi/2$), both $T_1$ and $T_2$ exhibit a maximum. Red (black) corresponds to the spin-orbit lengths $l_R=2\ \mu\rm m$ and $l_D=1\ \mu\rm m$ ($l_R=1\ \mu\rm m$ and $l_D=0.5\ \mu\rm m$). For the stronger SOI, the lifetimes increase by almost two orders of magnitude. For details, see text.} \label{fig:angular_dependence} \end{center} \end{figure} We also calculate the dependence of $T_1$ and $T_2$ on the angle between $\bm{B}$ and the $x$ axis, assuming that $x \parallel [110]$. The results for $T = \mbox{100 mK}$ and $J_{\rm tot} = 1.43 \mbox{ $\mu$eV}$ are plotted in Fig.~\ref{fig:angular_dependence}. Remarkably, the phonon-induced lifetimes of the qubit are maximal when $\bm{B} \perp x$ and minimal when $\bm{B} \parallel x$. The difference between minimum and maximum increases strongly with the SOI, and for $l_R=1\mbox{ $\mu$m}$ and $l_D=0.5\mbox{ $\mu$m}$ we already expect improvements by almost two orders of magnitude. These features can be understood via the matrix elements of the effective SOI,\cite{stano:prl06, golovach:prb08, raith:prl12} \begin{equation} \Omega = F_{\rm SOI}(a,l_c) E_Z \frac{l_D\cos{(\theta_B - \theta)}+l_R\cos{(\theta_B + \theta)}}{l_D l_R}, \end{equation} where $\theta_B$ ($\theta$) is the angle between $\bm{B}$ (the $x$ axis) and the crystallographic axis [110], and $F_{\rm SOI}(a,l_c)$ is a function of $a$ and $l_c$. From this result, we conclude that there always exists an optimal orientation for the in-plane magnetic field for which the effective SOI is suppressed and, thus, for which the phonon-mediated decay of the qubit state is minimal (comparing the lifetimes at fixed $J_{\rm tot}$). Remarkably, one finds for $x \parallel [110]$ ($\theta = 0$) that this suppression always occurs when $\bm{B} \perp x$ ($\theta_B = \pi/2$), independent of $l_R$ and $l_D$. In the case where $\Omega = 0$, the finite $T_2$ in our model results from admixtures with $|(0,2)S\rangle$, as explained in Sec.~\ref{secsub:OriginStrongDephasing}. Due to the hyperfine interaction, these admixtures also lead to finite $T_1$. We wish to emphasize, however, that suppression of the effective SOI only results in a substantial prolongation of the lifetimes when the spin-orbit lengths are rather short, as the dominant decay mechanism in biased DQDs is very effective even at $\Omega = 0$. \section{Regime of Small Detuning} \label{sec:SmallDetuning} All previous results were calculated for a large detuning $\epsilon \sim U - V_\pm$. Now we consider an unbiased DQD, i.e., the region of very small $\epsilon$. The dominant decay mechanism in the biased DQD is strongly suppressed at $\epsilon \simeq 0$, where the basis states $\ket{(2,0)S}$ and $\ket{(0,2)S}$ are both split from $\ket{(1,1)S}$ by a large energy $U - V_{+}$. Adapting the simple model behind Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}) to an unbiased DQD yields \begin{equation} \frac{8 t^4}{\hbar^2 (U - V_{+})^6} \int_{-\infty}^{\infty} \langle\widetilde{P}^2(0)\widetilde{P}^2(\tau)\rangle d\tau \end{equation} as the associated dephasing time (see Appendix~\ref{sec:SimpleModelDephasingZeroDetunSinglets} for details). Comparing the prefactor with that of Eq.~(\ref{eq:Simple3x3DecRateFinalMainText}) results in a remarkable suppression factor below $10^{-4}$ for the parameters in this work. As explained in Appendix~\ref{sec:SimpleModelDephasingZeroDetunSinglets}, this suppression factor may also be estimated via $(\Delta_S^\prime)^4/(U-V_{+})^4$ for fixed $J_{\rm tot}$, where $\Delta_S^\prime$ is the splitting between the eigenstates of type $\ket{(1,1)S}$ and $\ket{(0,2)S}$ at large $\epsilon$ and $U-V_{+}$ is the above-mentioned splitting at $\epsilon\simeq 0$. \begin{figure}[tb] \begin{center} \includegraphics[width=0.85\linewidth]{excited_states.pdf} \caption{Temperature dependence of the decoherence time $(T_2)$ and its one-phonon ($1/\Gamma_2^{\rm 1p}$) and two-phonon ($1/\Gamma_2^{\rm 2p}$) parts for the detuning $\epsilon \simeq 0$, where excited states are taken into account. For this plot $U = 1\mbox{ meV}$, $V_+ = 50\mbox{ $\mu$eV}$, $V_- = 49.5\mbox{ $\mu$eV}$, $t = 24\mbox{ $\mu$eV}$, $J_{\rm tot} = 1.41 \mbox{ $\mu$eV}$, and the other paramters as described in the text. We note that $T_2 \simeq 2 T_1$. } \label{fig:excited_states} \end{center} \end{figure} Consequently, the lifetimes $T_1$ and $T_2$ in the unbiased DQD are no longer limited by $|(2,0)S\rangle$ or $|(0,2)S\rangle$, but by states with an excited orbital part (see Fig.~\ref{fig:spectrum}). We therefore extend the subspace by the basis states $|(1^*,1)S\rangle$, $|(1^*,1)T_0\rangle$, $|(1^*,1)T_+\rangle$, and $|(1^*,1)T_-\rangle$, and proceed analogously to the case of large detuning (see Appendixes \ref{sec:BasisStates} and \ref{sec:HamiltonianSmallDetuning} for details). The asterisk denotes that the electron is in the first excited state, leading to an energy gap of $\hbar \omega_0$ compared to the states without asterisk. Setting $\bm{B} \parallel x \parallel [110]$, the orbital excitation is taken along the $x$ axis, because states with the excitation along $y$ turn out to have negligible effects on the qubit lifetimes. From symmetry considerations, states with the excited electron in the right QD should only provide quantitative corrections of the lifetimes by factors on the order of 2 and are therefore neglected in this analysis. The resulting temperature dependence of $T_2$, $1/\Gamma_2^{\rm 1p},$ and $1/\Gamma_2^{\rm 2p}$ is shown in Fig.\ \ref{fig:excited_states}. The plotted example illustrates that two-phonon processes affect $T_2$ only at rather high temperatures when $\epsilon$ is small, leading to $T_2 \propto T^{-1}$ for a wide range of $T$ due to single-phonon processes. In stark contrast to the biased DQD, we find $T_2 \simeq 2 T_1$. Remarkably, the absolute value of $T_2$ is of the order of milliseconds, which exceeds the $T_2$ at large $\epsilon$ by 2--3 orders of magnitude. For $\bm{B} \perp x$, $x \parallel [110]$, and typical sample temperatures $T \sim 0.1\mbox{ K}$, we find that the lifetimes can be enhanced even further. \section{Conclusions and outlook} \label{sec:Conclusions} In conclusion, we showed that one- and two-phonon processes can be major sources of relaxation and decoherence for $S$-$T_0$ qubits in DQDs. Our theory provides a possible explanation for the experimental data of Ref.~\onlinecite{dial:prl13}, and we predict that the phonon-induced lifetimes are prolonged by orders of magnitude at small detunings and, when the SOI is strong, at certain orientations of the magnetic field. Our results may also allow substantial prolongation of the relaxation time recently measured in resonant exchange qubits.\cite{medford:prl13} While the model developed in this work applies to a wide range of host materials, the resulting lifetimes depend on the input parameters and, thus, on the setup and the heterostructure. By separately neglecting the deformation potential coupling ($\Xi = 0$) and the piezoelectric coupling ($h_{14 }= 0$), we find that the qubit lifetimes of Figs.~\ref{fig:tem_dep}--\ref{fig:excited_states} for GaAs DQDs are limited by the piezoelectric electron-phonon interaction, the latter providing much greater decay rates than the deformation potential coupling. Consequently, the phonon-limited lifetimes of singlet-triplet qubits may be long in group-IV materials such as Ge or Si,\cite{maune:nat12, prance:prl12, zwanenburg:rmp13} where the piezoelectric effect is absent due to bulk inversion symmetry. Essentially, there are two different schemes for manipulating singlet-triplet qubits in DQDs electrically. The first and commonly realized approach is based on biased DQDs and uses the detuning to control the exchange energy.\cite{petta:sci05} Alternatively, the exchange energy can be controlled by tuning the tunnel barrier\cite{loss:pra98} rather than the detuning. Our results suggest that the second approach is advantageous, as it applies to unbiased DQDs for which the phonon-mediated decay of the qubit state is strongly suppressed. In addition, one finds $d J_{\rm tot}/d\epsilon \propto \epsilon$ at very small detunings $\epsilon$,\cite{burkard:prb99} which implies that not only $d J_{\rm tot}/d\epsilon \simeq 0$ but also $\langle d J_{\rm tot}/d\epsilon\rangle \simeq 0$ at $\epsilon \simeq 0$, where $\langle \cdots \rangle$ now stands for the average over some random fluctuations of $\epsilon$. Therefore, singlet-triplet qubits in unbiased DQDs are also protected against electrical noise. The latter, for instance, turned out to be a major obstacle for the implementation of high-fidelity controlled-phase gates between $S$-$T_0$ qubits.\cite{shulman:sci12} Keeping in mind that two-qubit gates for singlet-triplet qubits may also be realized with unbiased DQDs,\cite{klinovaja:prb12} we conclude that operation at $\epsilon \simeq 0$ with a tunable tunnel barrier is a promising alternative to the commonly realized schemes that require nonzero detuning. As single-qubit gates for $S$-$T_0$ qubits correspond to two-qubit gates for single-electron spin qubits, the regime $\epsilon \simeq 0$ is also beneficial for many other encoding schemes. \begin{acknowledgments} We thank Peter Stano, Fabio L. Pedrocchi, Mircea Trif, James R. Wootton, Robert Zielke, Hendrik Bluhm, and Amir Yacoby for helpful discussions and acknowledge support from the Swiss NF, NCCR QSIT, S$^3$NANO, and IARPA. \end{acknowledgments}
\section{Introduction} The detection of water vapor (H$_2$O) and carbon dioxide (CO$_2$) in the stratospheres of the Giant Planets and Titan by \citet{Feuchtgruber1997}, \citet{Coustenis1998}, \citet{Samuelson1983} and \citet{Burgdorf2006} has raised several questions: what are the sources of oxygen to their upper atmospheres? And do the sources vary from planet to planet? Oxygen-rich deep interiors of the Giant Planets cannot explain the observations because these species are trapped by condensation below their tropopause (except CO$_2$ in Jupiter and Saturn). Therefore, several sources in their direct or far environment have been proposed: icy rings and/or satellites \citep{Strobel1979}, interplanetary dust particles \citep{Prather1978} and large comet impacts \citep{Lellouch1995}. While the relative similarity of the infall fluxes inferred for H$_2$O by \citet{Feuchtgruber1997} may indicate that interplanetary dust particles (IDP) could be the source for all Giant Planets \citep{Landgraf2002}, infrared and far-infrared observations have unveiled a quite different picture. ISO, Cassini, Odin and Herschel observations have proven that the Jovian stratospheric H$_2$O and CO$_2$ originate from the Shoemaker-Levy~9 (SL9) comet impacts \citep{Lellouch2002,Lellouch2006,Cavalie2008b,Cavalie2012,Cavalie2013}, while Herschel has recently shown that the external flux of water at Saturn and Titan is likely due to the Enceladus geysers and the water torus they feed \citep{Hartogh2011, Moreno2012}. The situation is even more complex for carbon monoxide (CO). Because CO does not condense at the tropopauses of Giant Planets, oxygen-rich interiors are also a potential source. An internal component has indeed been observed in the vertical profile of CO in Jupiter by \citet{Bezard2002} and in Neptune, originally by \citet{Marten1993} and \citet{Guilloteau1993}, while an upper limit has been set on its magnitude by \citet{Cavalie2009} and \citet{Fletcher2012} for Saturn. The measurement of the tropospheric mole fraction of CO can be used to constrain the deep O/H ratio \citep{Lodders1994}, which is believed to be representative of condensation processes of the planetesimals that formed the Giant Planets \citep{Owen1999, Gautier2005}. On the other hand, large comets seem to be the dominant external source, as shown by various studies: \citet{Bezard2002} and \citet{Moreno2003} for Jupiter, \citet{Cavalie2010} for Saturn and \citet{Lellouch2005,Lellouch2010}, \citet{Hesman2007}, \citet{Fletcher2010} and \citet{Luszcz-Cook2013} for Neptune. The first detection of CO in Uranus was obtained by \citet{Encrenaz2004} from fluorescent emission at 4.7\,$\mu$m. Assuming a uniform distribution throughout the atmosphere, a mixing ratio of 2\dix{-8} was derived. The authors tentatively proposed that CO was depleted below the tropopause, suggesting that CO would have an external origin. Despite this first detection almost a decade ago, the situation has remained unclear. Ground-based heterodyne spectroscopy has been used unsuccessfully in the past to to try and detect CO in Uranus. \citet{Rosenqvist1992} have first set an upper limit of 4\dix{-8} and subsequent attempts to detect CO have failed so far \citep{Marten1993,Cavalie2008a}. In this paper, we present observations of CO in Uranus carried out with the Herschel Space Observatory \citep{Pilbratt2010} in 2011-2012, that led to the first detection of CO in Uranus in the submillimeter range. In the following sections, we will describe the observations, their modeling and the new constraints on the origin of CO in Uranus and its deep O/H ratio that we have derived. \section{Observations \label{Observations}} \begin{table*} \caption{Summary of the Herschel-HIFI observations of CO in Uranus. } \label{Obs_list} \begin{center} \begin{tabular}{ccccccc} \hline Date & OD & Obs. ID & $\nu$ [GHz] & $\Delta t$ [h] & $\theta_{\mathrm{HIFI}}$ [\arcsec] & $\theta_{\mathrm{Uranus}}$ [\arcsec] \\ \hline 2011-07-01 & 779 & 1342223423 & 921.800\,GHz & 1.82 & 23.0 & 3.53 \\ 2012-06-15 & 1128 & 1342247027 & 921.800\,GHz & 2.54 & 23.0 & 3.47 \\ 2012-06-15 & 1128 & 1342247028 & 921.800\,GHz & 2.54 &23.0 & 3.47 \\ 2012-06-15 & 1128 & 1342247029 & 921.800\,GHz & 2.54 &23.0 & 3.47 \\ \hline \end{tabular} \end{center} \small{\underline{Note:} OD means operational day, $\nu$ is the CO line center frequency, $\Delta t$ is the total integration time, $\theta_{\mathrm{HIFI}}$ is the Herschel-HIFI telescope beamwidth, and $\theta_{\mathrm{Uranus}}$ is the equatorial diameter of Uranus.} \end{table*} We observed the CO(8-7) line at 921.800\,GHz with the HIFI instrument \citep{Degraauw2010} aboard Herschel \citep{Pilbratt2010} on July 1, 2011, as part of the Guaranteed Time Key Program ``Water and related chemistry in the solar system'', also known as ``Herschel solar system Observations'' (HssO; \citealt{Hartogh2009}). The CO(8-7) line was targeted in Uranus for $\sim$2 hours. The resulting spectrum led us to a tentative detection of CO in Uranus at the level of $\sim$2.5$\sigma$ (on the line peak) at native resolution and encouraged us to perform a deeper integration of this line. We obtained a $\sim$8-hour integration (split into three observations of equal length) of the same line on June 15, 2012, as part of the Herschel Open Time 2 program OT2\_tcavalie\_6. The observations have been performed in double-beam switch mode with the Wide Band Spectrometer (WBS) at a nominal spectral resolution of 1.1\,MHz (more details given in \tab{Obs_list}). We have processed the data with HIPE 9 \citep{Ott2010} up to Level 2 for the H/V polarizations and stitched the WBS subbands together. The baseline ripple frequencies caused by the strong continuum emission of Uranus have been determined by a normalized periodogram \citep{Lomb1976} and the 3-4 strongest sine waves have been removed. Those sine waves are caused by the hot and cold black bodies and by the local oscillator chain of the instrument and have periods of 90-100\,MHz \citep{Roelfsema2012}. The double-sideband response of HIFI has been corrected by assuming a sideband ratio of 1, i.e., a single sideband gain of 0.5 \citep{Roelfsema2012}, and identical continuum levels in both sidebands. The uncertainty on the sideband ratio is 12\% (3\% on the single sideband gain), and the continuum levels in the two sidebands should be different by less than 1\%, according to our model. Finally, we have coadded the H/V polarizations after weighting them according to their respective noise levels (the V spectra were always noisier than the H spectra). We have obtained a clear detection at the level of 7$\sigma$ on the line peak, at a smoothed resolution of 8\,MHz using a gaussian lineshape, on the combined 8-hour observation shown in \fig{empirical_models}. Because we have not performed any absolute calibration, we will analyze this line in terms of line-to-continuum ratio ($l/c$). The observed continuum levels differ by 6\% in the H and V polarizations, so we have to account for an additional uncertainty of 3\% on the continuum level of our averaged spectrum. We note that we also targeted the CO (3-2) and (6-5) lines (at 345.796\,GHz and 691.473\,GHz, respectively) in Uranus using the HARP receiver array and the D-band receiver (respectively) of the James Clerk Maxwell Telescope (JCMT) on October 15-16 and November 2, 2009, as part of the M09BI02 project. These observations resulted in the determination of an upper limit of 6\dix{-8} uniform with altitude up to the CO homopause for the CO mole fraction and will not be discussed further. \section{Models and results \label{Modeling}} \subsection{Radiative transfer model} We have performed all spectral line computations with the forward radiative transfer model detailed in \citet{Cavalie2008a,Cavalie2013}, adapted to Uranus. This line-by-line model accounts for the elliptical geometry of the planet and its rapid rotation. Opacity due to H$_2$-He-CH$_4$ collision-induced absorption \citep{Borysow1985,Borysow1988,Borysow1986} was included. \citet{Orton2007} published H$_2$-H$_2$ collision-induced coefficient tables which better reproduce the continuum of Uranus between 700 and 1100\,\unitcm{-1} as observed by Spitzer, but these coefficients are not significantly different in the wavelength range of our observations. We have used the JPL Molecular Spectroscopy catalog \citet{Pickett1998} and the H$_2$/He pressure-broadening parameters for CO lines from \citet{Sung2004} and \citet{Mantz2005}, i.e. a collisional linewidth at 300\,K of 0.0661\,\unitgamma~for the CO(8-7) line and a temperature dependence exponent of 0.638. We have used the thermal profiles of \citet{Feuchtgruber2013} and \citet{Orton2013a}. They have the same tropopause temperature (53\,K) but the profile of \citet{Feuchtgruber2013} is continuously warmer than the profile of \citet{Orton2013a} in the stratosphere (by 2\,K at 10\,mbar, 5\,K at 1\,mbar, and 11\,K at 0.1\,mbar). We will present results for both thermal profiles hereafter. All synthetic lines have been smoothed to the working resolution of 8\,MHz using a gaussian lineshape. The CO line is optically thin with $\tau$$=$0.04-0.25 (depending on models) and probes the stratosphere of Uranus between the 0.1 and 5\,mbar levels, allowing us to derive information on the CO abundance. The S/N of the observations results in error bars of 14\%. By adding this uncertainty quadratically with other uncertainty sources (sideband ratio, continuum levels), we end up with an uncertainty of 19\% on the results presented hereafter. \subsection{Empirical models \label{empirical}} We have tested two classes of empirical models: (i) uniform profiles (referred to as ``Uniform'' hereafter), and (ii) uniform profile in the stratosphere down to a cutoff pressure level (refered to as ``Stratospheric'' hereafter). These profiles are not physically plausible (mainly due to the low homopause in Uranus, see next subsection), but have been considered for comparison with \citet{Encrenaz2004} and \citet{Teanby2013}. Our results are described hereafter and are summarized in \tab{model_results}. The uniform distribution of \citet{Encrenaz2004} with a CO mole fraction of 2\dix{-8} overestimates the observed line core by a factor of 2.5-3. The observed line can be fitted with an ``Uniform'' profile in which the CO mole fraction is 7.2\dix{-9} with \citet{Feuchtgruber2013}'s profile, or 9.3\dix{-9} with \citet{Orton2013a}'s profile. The line can be fitted equally well with a ``Stratospheric'' profile in which the CO is constant above the 100\,mbar level with a CO mole fraction of 7.1\dix{-9} with the thermal profile of \citet{Feuchtgruber2013}, or 9.0\dix{-9} with \citet{Orton2013a}'s profile. For comparison with other papers (e.g. \citealt{Encrenaz2004,Cavalie2008a,Teanby2013}), we have set this transition level to 100\,mbar, but our computations show this level could be located anywhere between $\sim$3 and 1000\,mbar, our results in terms of mole fraction would be affected by less than 10\%. If set above the 3\,mbar level, then more CO would be needed. From these empirical models, it is not possible to favor an internal or an external origin for CO in the atmosphere of Uranus, because the models cannot be distinguished (see \fig{empirical_models}). Fortunately, \citet{Teanby2013} recently published Herschel-SPIRE observations at CO line wavelengths. These observations are sensitive to the 10-2000\,mbar range, with a contribution function peak at 200\,mbar (see their Fig.~2b), and they did not result in any detection. Those authors have set a stringent upper limit of 2.1\dix{-9} on the CO mole fraction in their internal source model. This is $\sim$3-5 times lower than required by our observation. It is thus a clear indication that the HIFI line is caused by external CO. \begin{figure}[!h] \begin{center} \includegraphics[width=9cm,keepaspectratio]{empirical_models.ps} \end{center} \caption{Herschel-HIFI observation of the CO(8-7) line in Uranus on June 15, 2012, expressed in terms of line-to-continuum ratio ($l/c$, black line). This line can be modeled successfully with either empirical models: (i) a ``Uniform'' profile with a constant mole fraction of 7.2\dix{-9} throughout the atmosphere (red line), and (ii) a ``Stratospheric'' profile with a constant mole fraction of 7.1\dix{-9} above the 100\,mbar level and zero below it (blue line). The spectrum resulting from the \citet{Encrenaz2004} uniform source profile is also shown for comparison (grey line, labeled ``E04''). The synthetic lines have been obtained with the thermal profile of \citet{Feuchtgruber2013}.} \label{empirical_models} \end{figure} \begin{table} \caption{Summary of the empirical and diffusion model results.} \label{model_results} \begin{center} \begin{tabular}{cccc} \hline \multicolumn{4}{c}{Empirical model}\\ \hline Thermal profile & Uniform & \multicolumn{2}{c}{Stratospheric} \\ \hline Feuchtgruber & 7.2\dix{-9} & \multicolumn{2}{c}{7.1\dix{-9}} \\ Orton & 9.3\dix{-9} & \multicolumn{2}{c}{9.0\dix{-9}} \\ \hline \hline \multicolumn{4}{c}{Diffusion model}\\ \hline Thermal profile & Internal source & \multicolumn{2}{c}{External source} \\ & $y_{\mathrm{CO}}$ & \Flux{CO} & $y_0$ \\ \hline Feuchtgruber & 1.9\dix{-8} & 2.2\dix{5} & 3.1\dix{-7} \\ Orton & 2.7\dix{-8} & 2.7\dix{5} & 3.9\dix{-7} \\ \hline \end{tabular} \end{center} \small{\underline{Note:} All results are mole fractions, except \Flux{CO} (in cm$^{-2}\cdot$s$^{-1}$). The cut-off level in the ``Stratospheric'' empirical model is at 100\,mbar. The internal source value of $y_{\mathrm{CO}}$~in the diffusion model enables fitting the CO line core amplitude, but the line is too broad and additional broad wings incompatible with the data are generated.} \end{table} \subsection{Diffusion model} Uranus has the lowest homopause amongst the Giant Planets \citep{Orton1987,Moses2005}. It is located around the 1\,mbar level, where submillimeter observations generally probe. Therefore, we have computed more realistic CO vertical profiles by accounting for diffusion processes to see how our previous results are impacted by the low homopause of Uranus. Such modeling also shows how the various external sources can be parametrized. The vertical profile of CO primarily depends on the sources of CO, but it is also influenced by other oxygen sources. Indeed, O produced by H$_2$O photolysis reacts with CH$_3$ and other hydrocarbons to produce CO \citep{Moses2000}. The magnitude of the H$_2$O flux is still quite uncertain, mostly due to limitations in the knowledge of the thermal structure and the eddy mixing at the time of \citet{Feuchtgruber1997}'s observations. For the sake of simplicity, (photo-)chemical processes have been ignored. We have adapted to Uranus the 1D time-dependent model of \citet{Dobrijevic2010,Dobrijevic2011} and removed all photochemical processes. \citet{Orton2013b} have constrained the stratospheric $K_{zz}$ within [1000:1500]\,\,cm$^{2}\cdot$s$^{-1}$~(vertically constant) with CH$_4$ and C$_2$H$_6$ Spitzer observations. We have taken their best fit value (1200\,\,cm$^{2}\cdot$s$^{-1}$) in our model. Three sources of CO, representing simple cases, have been tested: (i) an internal source, (ii) a steady flux of micrometeorites (IDP), and (iii) a single large comet impact\footnote{This does not exclude a combination of internal and external sources, or any intermediate situation between a continuum of micrometeoritic impacts and a single impact event.}. The three sources tested are controlled by a few parameters: (i) the tropospheric CO mole fraction $y_{\mathrm{CO}}$~for an internal source, (ii) the flux \Flux{CO} at the upper boundary of the model atmosphere for a steady source, and (iii) the equivalent mole fraction of CO $y_0$ deposited by a comet and averaged over the planet. We have assumed that all the CO was deposited at levels higher than 0.1\,mbar in analogy to the SL9 impacts \citep{Lellouch1995,Moreno2003} and that the impact time $\Delta t$$\sim$300\,years as it roughly corresponds to the diffusion time down to 1\,mbar in Uranus in our model, but other combinations of deposition time and level are possible. To infer the mass and diameter of the comet, we have assumed the comet density was 0.5\,g$\cdot$cm$^{-3}$ \citep{Weissman2004,Davidsson2007} and that the comet yielded 50\% CO at impact \citep{Lellouch1997}. The vertical profiles and resulting spectra corresponding to the three sources, as obtained with the thermal profile of \citet{Feuchtgruber2013}, are displayed in \fig{CO_HIFI_2012}. The best fits to the spectrum are obtained for external source models. Despite resulting in different vertical profiles, a steady flux \Flux{CO}$=$2.2\dix{5}\,\,cm$^{-2}\cdot$s$^{-1}$~and a 640\,m diameter comet depositing 3.5\dix{13}\,g of CO ($y_0$$=$3.1\dix{-7}) result in lines that cannot be distinguished from one another with our observations. Such impact at Uranus occurs every $\sim$500~years with a factor of 6 uncertainty \citep{Zahnle2003}. Such timescales are fully compatible with our assumption on $\Delta t$. With the thermal profile of \citet{Orton2013a}, we obtain slightly higher values because of lower stratospheric temperatures: \Flux{CO}$=$2.7\dix{5}\,\,cm$^{-2}\cdot$s$^{-1}$~and $y_0$$=$3.9\dix{-7} (i.e. a 700\,m diameter comet). All fit parameters are listed in \tab{model_results}. These values remain to be confirmed by more rigorous (photochemical) modeling and higher S/N data. The amplitude of the CO core emission is reproduced with an internal source model in which $y_{\mathrm{CO}}$$=$1.9\dix{-8} (see \fig{CO_HIFI_2012}). With \citet{Orton2013a}'s thermal profile, $y_{\mathrm{CO}}$$=$2.7\dix{-8}. We note that $\sim$3 times more tropospheric CO is needed in this model, compared to the ``Uniform'' empirical model value derived in Sect.~\ref{empirical}. This is due to the fact that the observed emission line probes the mbar level, i.e. where the CO vertical profile sharply decreases because of the low homopause in the atmosphere of Uranus. As a result, a stronger internal source is required to reach a sufficient level of abundance of CO around the mbar level. The main outcome of this model is that it now overestimates the line core width and results in an additional broad absorption, because CO is much more abundant in the lower stratosphere than in the external source models (by already 2 orders of magnitude at 10\,mbar). The absence of such a broad CO absorption in the data cannot be caused by our sinusoidal ripple removal procedure, because we have removed sine waves of much shorter period than the total width of such broad absorption wings. Because the width of the line core is not fitted and because there is no broad absorption in the spectrum, and because the derived $y_{\mathrm{CO}}$ values are an order of magnitude larger than the upper limit set by Herschel-SPIRE observations of this region of the atmosphere \citep{Teanby2013}, the internal source model can be ruled out. Thus, as long as there is no significant photochemical source of CO in the stratosphere, the HIFI line is caused by external CO. \begin{figure*}[!h] \begin{center} \includegraphics[width=18cm,keepaspectratio]{Uranus_OT2.ps} \end{center} \caption{Left: Herschel-HIFI observation of the CO(8-7) line in Uranus on June 15, 2012, expressed in terms of line-to-continuum ratio ($l/c$, black line). For each source, the models that best fit the emission core are displayed: an internal source yielding a mole fraction of 1.9\dix{-8} in the upper troposphere (red line), a steady external flux (due to IDP or a local source) of 2.2\dix{5}\,\,cm$^{-2}\cdot$s$^{-1}$~(blue line), and a comet with a diameter of 640\,m depositing 3.4\dix{13}\,g of CO above the 0.1\,mbar level $\sim$300\,years ago (green line). These models have been computed with the thermal profile of \citet{Feuchtgruber2013}. The internal source model overestimates the line core width and produces a broad absorption that is not observed in the data. The external source models can barely be differentiated. Right: Vertical profiles associated to the spectra.} \label{CO_HIFI_2012} \end{figure*} \subsection{An upper limit on the deep O/H ratio in Uranus} Thermochemistry in the deep interior of Uranus links the CO abundance to the one of H$_2$O and thus to the internal O/H ratio \citep{Fegley1988,Lodders1994} with the following net thermochemical equilibrium reaction \begin{equation*} \mathrm{H}_2\mathrm{O}+\mathrm{CH}_4=\mathrm{CO}+3\mathrm{H}_2. \end{equation*} The upper tropospheric mole fraction of CO is fixed at the level where the thermochemical equilibrium is quenched by vertical diffusion. The upper limit of \citet{Teanby2013} on the internal source ($y_{\mathrm{CO}}$$=$2.1\dix{-9}) can be further used to try and constrain the deep atmospheric O/H ratio in Uranus. Their observations probe between 10 and 2000\,mbar, i.e. well below the homopause level (see \fig{CO_HIFI_2012} right). As a consequence, this upper limit is valid even if the authors have not accounted for the low homopause of Uranus. We have adapted to Uranus the thermochemical model developed by \citet{Venot2012} to constrain the O/H ratio. This model accounts for C, N and O species. We have extended our thermal profile to high pressures following the dry adiabat (the \citealt{Feuchtgruber2013} and \citealt{Orton2013a} are similar in the upper troposphere and thus give similar deep tropospheric profiles) and we have constrained the O/H and C/H ratios by fitting the following upper tropospheric mole fractions with errors lower than 4\%: 0.152 for He \citep{Conrath1987}, 0.016 for CH$_4$ \citep{Baines1995,Sromovsky2008}, and the 2.1\dix{-9} upper limit for CO. The level at which CO is quenched depends not only on the temperature profile and the deep O/H ratio, but also on the deep $K_{zz}$. Assuming Uranus' interior is convective, we can estimate $K_{zz}$ from the planet's internal heat flux \citep{Stone1976}. Following \citet{Pearl1990}, $K_{zz}$$\sim$10$^8$\,cm$^{2}\cdot$s$^{-1}$, within one order of magnitude \citep{Lodders1994}. The resulting tropospheric vertical profiles for this nominal model are shown in \fig{Thermo}. The elemental ratios in this model are 501 times solar for O/H and 18 times solar for C/H (with solar abundances, $\Sun$ hereafter, taken from \citealt{Asplund2009}). The N species have no significant impact on the C and O species. We have also computed the elemental ratios in a series of additional models to evaluate the influence of parameters like $K_{zz}$ and the upper tropospheric CH$_4$ mole fraction on the O/H ratio. The results are displayed in \tab{elemental_ratios}. We find that the deep O/H is lower than $\sim$500$\Sun$ (nominal model), but could be even below 340$\Sun$ to be in agreement with the CO tropospheric upper limit in all cases. On Neptune, \citet{Luszcz-Cook2013} find that ``an upwelled CO mole fraction of 0.1 ppm implies a global O/H enrichment of at least 400, and likely more than 650, times the protosolar value''. \begin{table} \caption{Summary of the thermochemical model results. The values have been obtained so as to reach the 2.1\dix{-9} upper limit of \citet{Teanby2013} for the CO upper tropospheric mole fraction.} \label{elemental_ratios} \begin{center} \begin{tabular}{cccccc} \hline Model & $K_{zz}$ & $y_{\mathrm{CH}_4}$ & C/H & $y_{\mathrm{CO}}$ & O/H \\ & \,cm$^{2}\cdot$s$^{-1}$ & \dix{-2} & $\times\Sun^{(1)}$ & \dix{-9} & $\times\Sun^{(2)}$ \\ \hline Nominal & 10$^8$ &1.6$^{(3)}$ & 18 & 2.1 & 501 \\ CH$_4$-rich & 10$^8$ & 3.2$^{(4)}$ & 40 & 2.1 & 417 \\ low $K_{zz}$ & 10$^7$ & 1.6 & 13 & 2.1 & 631 \\ high $K_{zz}$ & 10$^9$ & 1.6 & 23 & 2.1 & 339 \\ \hline \end{tabular} \end{center} \small{$^{(1)}$ Solar C/H volume ratio: 2.69\dix{-4} \citep{Asplund2009}\\ $^{(2)}$ Solar O/H volume ratio: 4.90\dix{-4} \citep{Asplund2009}\\ $^{(3)}$ \citet{Baines1995} and \citet{Sromovsky2008}\\ $^{(4)}$ \citet{Fry2013}, \citet{Sromovsky2011}, and \citet{Karkoschka2009}} \end{table} \begin{figure*}[!h] \begin{center} \includegraphics[width=12cm,keepaspectratio]{thermochemistry.ps} \end{center} \caption{Molar fraction profiles in the troposphere of Uranus obtained with \citet{Venot2012}'s model, targeting the 2.1\dix{-9} upper limit on the upper tropospheric CO mole fraction obtained by \citet{Teanby2013}. The temperature profile in the troposphere is shown in black solid line. Thermochemical equilibrium profiles are plotted in black with the same layout as their corresponding species. CO and CO$_2$ are quenched around 2-3\dix{6}\,mbar. H$_2$O departs from thermochemical equilibrium because of condensation and causes the increase of other species mole fractions (the sum of all mole fractions is normalized to unity at all levels). The model parameters are: O/H$=$501$\Sun$, C/H$=$18$\Sun$, and $K_{zz}$$=$$10^8$\,\,cm$^{2}\cdot$s$^{-1}$.} \label{Thermo} \end{figure*} \section{Discussion and conclusion \label{Discussion}} We have detected the CO(8-7) line at 921.800\,GHz in Uranus with Herschel and we have constrained its possible sources. Herschel-HIFI (this work) and Herschel-SPIRE \citep{Teanby2013} results show that the average CO mole fraction is decreasing from the stratosphere to the troposphere. This suggests the deep interior is not the source of the observed CO. Our diffusion model calculations confirms that the internal source hypothesis is not valid and shows that Uranus has an external source of CO as long as there is not a significant photochemical source of CO in the stratosphere. The data can be successfully fitted by an empirical model in which CO has a mole fraction of 7.1-9.0\dix{-9} above the 100\,mbar level (value depending on the chosen thermal profile). There is a contradiction between this model mole fraction values and the mole fraction reported by \citet{Encrenaz2004} (3\dix{-8} in their external source model). Regarding this apparent discrepancy, we note that modelling LTE emission from CO rotational lines is much simpler than inferring an abundance from non-LTE fluorescence (e.g. \citealt{Lopez-Valverde2005}). At any rate, a reanalysis of the \citet{Encrenaz2004} data in the light of CO distributions proposed in this paper should be performed. Comet and steady source models, in which diffusion processes are accounted for, give very similar fit to the data. These results should be confirmed with more elaborate models, i.e. photochemical models, and more sensitive observations. Oxygen photochemistry computations, taking into account nearly concomitant measurements of the thermal profile \citep{Feuchtgruber2013,Orton2013a}, of the influx of H$_2$O (Jarchow et al., in prep.), and of the influx of CO$_2$ \citep{Orton2013b} would enable us to draw better constraints on the external source of oxygen. It would certainly reduce the external flux of CO or the mass of the impacting comet we have obtained from a simple diffusion model, because the chemical conversion of H$_2$O into CO would already provide a significant part of the observed stratospheric column of CO. The internal source upper limit derived by \citet{Teanby2013} ($y_{\mathrm{CO}}$$=$2.1\dix{-9}), which also goes in contradiction with the detection level of \citet{Encrenaz2004} (2\dix{-8} in their internal source model), was used to derive an upper limit on the deep O/H ratio of Uranus. Our thermochemical simulations show that the deep O/H ratio is lower than 500$\Sun$ to provide a $y_{\mathrm{CO}}$~value lower than 2.1\dix{-9}. A dedicated probe, as in the mission concepts proposed by \citet{Arridge2013} and \citet{Mousis2013} in response to the ESA 2013 Call for White Papers for the Definition of the L2 and L3 Missions in the ESA Science Programme, or radio observations might be the only way to measure the deep O/H ratio in Uranus. \begin{acknowledgements} T. Cavali\'e acknowledges funding from the Centre National d'\'Etudes Spatiales (CNES). T. Cavali\'e and F. Selsis acknowledge support from the European Research Council (Starting Grant 209622: E$_3$ARTHs). O. Venot acknowledges support from the KU Leuven IDO project IDO/10/2013 and from the FWO Postdoctoral Fellowship programme. G. Orton acknowledges funding from the National Aeronautics and Space Administration to the Jet Propulsion Laboratory, California Institute of Technology. L.~N. Fletcher was supported by a Royal Society Research Fellowship at the university of Oxford. The authors thank M. Hofstadter for his constructive review. HIFI has been designed and built by a consortium of institutes and university departments from across Europe, Canada and the United States under the leadership of SRON Netherlands Institute for Space Research, Groningen, The Netherlands and with major contributions from Germany, France and the US. Consortium members are: Canada: CSA, U.Waterloo; France: CESR, LAB, LERMA, IRAM; Germany: KOSMA, MPIfR, MPS; Ireland, NUI Maynooth; Italy: ASI, IFSI-INAF, Osservatorio Astrofisico di Arcetri-INAF; Netherlands: SRON, TUD; Poland: CAMK, CBK; Spain: Observatorio Astron\'omico Nacional (IGN), Centro de Astrobiolog\'ia (CSIC-INTA). Sweden: Chalmers University of Technology - MC2, RSS \& GARD; Onsala Space Observatory; Swedish National Space Board, Stockholm University - Stockholm Observatory; Switzerland: ETH Zurich, FHNW; USA: Caltech, JPL, NHSC. The James Clerk Maxwell Telescope is operated by the Joint Astronomy Centre on behalf of the Science and Technology Facilities Council of the United Kingdom, the National Research Council of Canada, and the Netherlands Organisation for Scientific Research. \end{acknowledgements}
\section{Introduction} Portfolio optimization is an important problem in economic analysis and risk management \cite{key-1,key-2}, and under certain nonlinear constraints maps exactly into the problem of finding the ground states of a long-range spin glass \cite{key-3,key-4,key-5}. The main assumption is that the return of any financial asset is described by a random variable, whose expected mean and variance are interpreted as the reward, and respectively the risk of the investment. The problem can be formulated as following: given a set of financial assets, characterized by their expected mean and their covariances, find the optimal weight of each asset, such that the overall portfolio provides the smallest risk for a given overall return. The standard mean-variance optimization problem has an unique solution describing the so called ``efficient frontier'' in the $(risk,return)$-plane \cite{key-6}. The expected return is a monotonically increasing function of the standard deviation (risk), and for accepting a larger risk the investor is rewarded with a higher expected return. Recently it has been shown that the portfolio optimization problem containing short sales with obligatory deposits (margin accounts) is equivalent to the problem of finding the ground states of a long-range Ising spin glass, where the coupling constants are related to the covariance matrix of the assets defining the portfolio \cite{key-3,key-4,key-5}. As a consequence of this nonlinear constraint, the solution consists of an exponentially large number of optimal portfolios, completely different from each other, and extremely sensitive to any changes in the input parameters of the problem. Therefore, under such constraints, the concept of rational decision making becomes questionable, since the investor has an exponential number of ``options'' to choose from. Here, we discuss the portfolio optimization problem using a quadratic formulation of the nonlinear obligatory deposits constraint. From the physics point of view, finding an optimal portfolio amounts to calculating the mean-field magnetizations of a random Ising model with the constraint of a constant magnetization norm. We show that the proposed model reduces to an eigenproblem, with $2N$ solutions, where $N$ is the number of assets defining the portfolio. In support to our results, we also work out a detailed numerical example of a portfolio of several risky common stocks traded on the Nasdaq Market. \section{Nonlinear optimization model} A portfolio is an investment made in $N$ assets $A_{n}$, with the expected returns $r_{n}$, and covariances $s_{nm}=s_{mn}$, $n,m=1,2,...,N$. Let $w_{n}$ denote the relative amount invested in the $n$-th asset. Negative values of $w_{n}$ can be interpreted as short selling. The variance of the portfolio captures the risk of the investment, and it is given by: \begin{equation} s^{2}=\sum_{i=1}^{N}\sum_{j=1}^{N}w_{i}w_{j}s_{ij}=\mathbf{w}^{T}\mathbf{S}\mathbf{w}, \end{equation} where $\mathbf{w}=[w_{1},w_{2},\ldots,w_{N}]^{T}$ is the vector of weights, and $\mathbf{S}=[s_{nm}]$ is the covariance matrix. Also, another characteristic of the portfolio is the expected return: \begin{equation} \rho=\sum_{n=1}^{N}w_{n}r_{n}=\mathbf{w}^{T}\mathbf{r}, \end{equation} where $\mathbf{r}=[r_{1},r_{2},\ldots,r_{N}]^{T}$ is the vector of asset returns. The standard portfolio selection problem consists in finding the solution of the following multi-objective optimization problem {[}1,2,6{]}: \begin{equation} \min_{\mathbf{w}}\left\{ s^{2}=\mathbf{w}^{T}\mathbf{S}\mathbf{w}\right\} , \end{equation} \begin{equation} \max_{\mathbf{w}}\left\{ \rho=\mathbf{w}^{T}\mathbf{r}\right\} , \end{equation} subject to the invested wealth constraint: \begin{equation} \sum_{n=1}^{N}w_{n}=1. \end{equation} As mentioned in the introduction, this problem has an unique solution, which can be obtained using the method of Lagrange multipliers \cite{key-1,key-2,key-6}. Recently it has been shown that by replacing the invested wealth constraint (5) with an obligatory deposits constraint the problem cannot be solved analytically anymore {[}3-5{]}. The constraint consists in imposing the requirement to leave a certain deposit (margin) proportional to the value of the underlying asset, and it has the form: \begin{equation} \gamma\sum_{n=1}^{N}\left|w_{n}\right|=W, \end{equation} where $\gamma>0$ is the fraction defining the margin requirement, and $W$ is the total wealth invested. As a direct consequence of the constraint's nonlinearity, the problem has an exponentially large number of solutions: \begin{equation} n(N,\rho)\sim\exp(\omega(\rho)N), \end{equation} where $\omega(\rho)$ is a positive number depending on the portfolio return \cite{key-3,key-4,key-5}. The solutions are also completely different from each other, and extremely sensitive to any changes in the input parameters of the problem. Thus, finding the global optimum becomes prohibitive (NP-problem) for a larger $N$. Let us now to reformulate this constraint using a quadratic function: \begin{equation} \gamma\sum_{n=1}^{N}w_{n}^{2}=W. \end{equation} Thus, we impose the requirement to leave a certain deposit proportional to the quadratic value of the asset. This is equivalent to a constant norm $\left\Vert \mathbf{w}\right\Vert ^{2}=k=W/\gamma$. Also, we combine the multi-objective optimization problem into a single Lagrangian objective function as following: \begin{equation} \min_{\mathbf{w},\mu}\left\{ F(\mathbf{w},\lambda,\mu)=\lambda\mathbf{w}^{T}\mathbf{S}\mathbf{w}-(1-\lambda)\mathbf{w}^{T}\mathbf{r}-\mu(\mathbf{w}^{T}\mathbf{w}-k)\right\} , \end{equation} where $\lambda\in[0,1]$ is the risk aversion parameter, and $\mu$ is the Lagrange parameter. If $\lambda=0$ then the solution corresponds to the portfolio with maximum return, without considering the risk. In this case the optimal solution will be formed only by the asset with the greatest expected return. The case with $\lambda=1$ corresponds to the portfolio with minimum risk, regardless the value of the expected return. In this case the problem becomes: \begin{equation} \min_{\mathbf{w},\mu}\left\{ F(\mathbf{w},1,\mu)=\mathbf{w}^{T}\mathbf{S}\mathbf{w}-\mu(\mathbf{w}^{T}\mathbf{w}-k)\right\} , \end{equation} with the solutions given by the equation: \begin{equation} \nabla_{\mathbf{w}}F(\mathbf{w},\lambda,\mu)=2\mathbf{S}\mathbf{w}-2\mu\mathbf{w}=0. \end{equation} This is a standard eigenproblem: \begin{equation} \mathbf{S}\mathbf{w}=\mu\mathbf{w}, \end{equation} where $\mathbf{S}$ is a symmetric matrix with $N$ real eigenvalues, and $N$ real eigenvectors. The eigenvector corresponding to the largest eigenvalue will provide the global optimum, since it will have the lowest risk. Any value $\lambda\in(0,1)$ represents a tradeoff between the risk and return. In this case the solution corresponds to the critical point of the Lagrangian, which is also the solution of the following system of equations: \begin{equation} \nabla_{\mathbf{w}}F(\mathbf{w},\lambda,\mu)=2\lambda\mathbf{S}\mathbf{w}-(1-\lambda)\mathbf{r}-2\mu\mathbf{w}=0, \end{equation} \begin{equation} \frac{\partial F(\mathbf{w},\lambda,\mu)}{\partial\mu}=\mathbf{w}^{T}\mathbf{w}-k=0. \end{equation} One can see that the Lagrangian objective function is equivalent to the free energy of an Ising model with random couplings $J_{nm}=-2\lambda s_{nm}$ and a random magnetic field $h_{n}=(1-\lambda)r_{n}$. From the physics point of view, finding an optimal portfolio amounts to calculating the mean-field magnetizations $w_{n}$ of this random Ising model with the constraint of a constant magnetization norm. In the following we show that solving this system of equations reduces to an inhomogeneous eigenproblem. From the first equation we have: \begin{equation} \mathbf{w}=\frac{1}{2}(1-\lambda)(\lambda\mathbf{S}-\mu\mathbf{I})^{-1}\mathbf{r}. \end{equation} Introducing this result into the second equation we obtain: \begin{equation} \frac{1}{4}(1-\lambda)^{2}\mathbf{r}^{T}(\lambda\mathbf{S}-\mu\mathbf{I})^{-2}\mathbf{r}-1=0. \end{equation} The left-hand side of this equation is the Schur complement of the matrix: \begin{equation} \mathbf{M}=\left[\begin{array}{cc} (\lambda\mathbf{S}-\mu\mathbf{I})^{2} & \frac{1}{2}(1-\lambda)\mathbf{r}\\ \frac{1}{2}(1-\lambda)\mathbf{r}^{T} & 1 \end{array}\right]. \end{equation} Since this matrix must be singular (the Schur complement is zero), we have: \begin{equation} \mathrm{det}\left[(\lambda\mathbf{S}-\mu\mathbf{I})^{2}-\frac{1}{4}(1-\lambda)^{2}\mathbf{r}\mathbf{r}^{T}\right]=0, \end{equation} which reduces to: \begin{equation} \mathrm{det}\left[\frac{1}{4}(1-\lambda)^{2}\mathbf{r}\mathbf{r}^{T}-\lambda^{2}\mathbf{S}^{2}+2\lambda\mu\mathbf{S}-\mu^{2}\mathbf{I}\right]=0. \end{equation} Obviously, there is a vector $\mathbf{w}$ such that: \begin{equation} \left[\frac{1}{4}(1-\lambda)^{2}\mathbf{r}\mathbf{r}^{T}-\lambda^{2}\mathbf{S}^{2}+2\lambda\mu\mathbf{S}-\mu^{2}\mathbf{I}\right]\mathbf{w}=0. \end{equation} This is an inhomogeneous $N\times N$ eigenproblem \cite{key-7}, and it can be reduced further to a $2N\times2N$ standard eigenproblem by introducing the following quantity: \begin{equation} \mathbf{u}=\mu\mathbf{w}, \end{equation} such that we have: \begin{equation} \left[\frac{1}{4}(1-\lambda)^{2}\mathbf{r}\mathbf{r}^{T}-\lambda^{2}\mathbf{S}^{2}\right]\mathbf{w}+2\lambda\mathbf{S}\mathbf{u=\mu}\mathbf{u}. \end{equation} By combining the last two equations into a matrix representation we obtain: \begin{equation} \mathbf{\left[\begin{array}{cc} \mathbf{0} & \mathbf{I}\\ \mathbf{A} & \mathbf{B} \end{array}\right]}\left[\begin{array}{c} \mathbf{w}\\ \mathbf{u} \end{array}\right]=\mu\left[\begin{array}{c} \mathbf{w}\\ \mathbf{u} \end{array}\right], \end{equation} where \begin{equation} \mathbf{A}=\frac{1}{4}(1-\lambda)^{2}\mathbf{r}\mathbf{r}^{T}-\lambda^{2}\mathbf{S}^{2}, \end{equation} and \begin{equation} \mathbf{B}=2\lambda\mathbf{S}, \end{equation} and $\mathbf{0}$, $\mathbf{I}$ are the zero, and respectively identity matrices. This eigenproblem obviously has $2N$ eigenvalues $\mu$, that may be real or complex. If $\mathbf{w}$ is a real optimal portfolio, associated to a real eigenvalue, then the corresponding risk and return are given by: \begin{equation} s=\sqrt{\mathbf{w}^{T}\mathbf{S}\mathbf{w}},\quad\rho=\left|\mathbf{w}^{T}\mathbf{r}\right|. \end{equation} In the above equation we have defined the return as the absolute value of the scalar product. This is a consequence of the fact that the eigenvectors can be determined only up to the sign value, i.e. both $[\mathbf{w},\mathbf{u}]^{T}$ and $-[\mathbf{w},\mathbf{u}]^{T}$ are eigen-vectors corresponding to the same eigenvalue $\mu$. This means that if the return is negative, we can simply change the sign $(\mathbf{w}\rightarrow-\mathbf{w})$, such that the return becomes positive. The real and imaginary parts of the complex portfolios are also valid investment portfolios, however they are sub-optimal and may be discarded. Assuming that $\mathbf{w}=\mathbf{x}+i\mathbf{y}$, where $\mathbf{x}=\mathrm{Re}(\mathbf{w})$ and $\mathbf{y}=\mathrm{Im}(\mathbf{w})$, the risk and return of the real and respectively imaginary parts can be determined as following: \begin{equation} s_{x}=\sqrt{\mathbf{x}^{T}\mathbf{S}\mathbf{x}},\quad\rho_{x}=\left|\mathbf{x}^{T}\mathbf{r}\right|, \end{equation} \begin{equation} s_{y}=\sqrt{\mathbf{y}^{T}\mathbf{S}\mathbf{y}},\quad\rho_{y}=\left|\mathbf{y}^{T}\mathbf{r}\right|. \end{equation} \begin{figure}[ht] \centering \includegraphics[scale=0.7]{fig1.eps}\caption{Asset prices (log scale).} \end{figure} \begin{figure}[ht] \centering \includegraphics[scale=0.7]{fig2.eps} \caption{Daily returns of the assets.} \end{figure} \section{Numerical example} In order to illustrate the above results, we consider the case of a portfolio consisting of $N=12$ common stocks from IT industry: \begin{equation} \mathrm{AMZN,EBAY,FB,INT,CSCO,VZ,MSFT,ORCL,GOOG,YHOO,DELL,HPQ} \end{equation} A historical record of daily prices of these stocks for the last $T=300$ trading days was used to estimate the mean return and the covariance matrix. The daily returns of the assets are calculated as: \begin{equation} R(n,t)=\frac{P(n,t+1)-P(n,t)}{P(n,t)}, \end{equation} where $t=1,2,...,T$ is the day index, and $P(n,t)$ is the price of asset $A_{n}$ at the closing day $t$. The estimate average returns and covariances are: \begin{equation} r_{n}=\frac{1}{T}\sum_{t=1}^{T}R(n,t),\quad n=1,2,...,N, \end{equation} \begin{equation} s_{ij}=\frac{1}{T}\sum_{t=1}^{T}[R(i,t)-r_{i}][R(j,t)-r_{j}],\quad i,j=1,2,...,N. \end{equation} The time series of the asset prices for the considered time period are given in Figure 1. In Figure 2 we give also the expected daily returns of the assets. The risk aversion parameter $\lambda$ was discretized as $\lambda_{t}=t/T$, $t=0,1,...,T$, where $T=1000$. Also, the fraction defining the margin requirement was set to $\gamma=1$. Thus by solving the eigenproblem for each value of $\lambda$, one obtains $2NT=24,000$ eigenvalues and eigenvectors containing the weights of the portfolios. The risk-return, $(s,\rho)$, representation of these complex solutions is shown in Figure 3, which also shows the risk-return values of the stocks included in the portfolio. The real contributions $(s_{x},\rho_{x})$ are shown in blue, while the imaginary contributions $(s_{y},\rho_{y})$ are shown in red. The pure real solutions, corresponding to the real eigenvalues are extracted in Figure 4. Here we also show the portfolio with the maximum Sharpe ratio $\xi=\rho/s$, The Sharpe ratio represents the expected return per unit of risk [1,2]. The portfolio with maximum Sharpe ratio $\xi$ gives the highest expected return per unit of risk, and therefore is the most \textquotedbl{}risk-efficient\textquotedbl{} portfolio. \section{Conclusion} We have considered the portfolio optimization problem with the obligatory deposits constraint, when both long buying and short selling of a relatively large number of assets is allowed. Recently it has been shown that as a consequence of this nonlinear constraint, the solution consists of an exponentially large number of optimal portfolios, completely different from each other, and extremely sensitive to any changes in the input parameters of the problem, making the concept of rational decision making questionable. Here we have reformulated the problem using a quadratic obligatory deposits constraint, and we have shown that from the physics point of view, finding an optimal portfolio amounts to calculating the mean-field magnetizations of a random Ising model with the constraint of a constant magnetization norm. We have shown that the model reduces to an eigenproblem, with $2N$ solutions, where $N$ is the number of assets defining the portfolio. Also, in order to illustrate our results, we have presented a detailed numerical example of a portfolio of several risky common stocks traded on the Nasdaq Market. \begin{figure}[ht] \centering \includegraphics[scale=0.7]{fig3.eps} \caption{Complex solutions of the portfolio optimization problem with quadratic obligatory deposits constraint (blue=real and red=imaginary components). SP is the portfolio with the maximum Sharpe ratio.} \end{figure} \begin{figure}[ht5] \centering \includegraphics[scale=0.7]{fig4.eps} \caption{Real solutions of the portfolio optimization problem with quadratic obligatory deposits constraint. SP is the portfolio with the maximum Sharpe ratio.} \end{figure}
\section{Introduction} Magnetic helicity is an important quantity that reflects the topology of the magnetic field (Woltjer, 1958a,b and Taylor, 1986). Pioneering studies of magnetic helicity in solar physics have been performed by several authors focussing on the accumulation of magnetic helicity in the solar atmosphere \citep[e.g.][]{Berger84,Chae01}, the force-free $\alpha$ coefficient, and the mean current helicity density in solar active regions \citep{Seehafer90}. Besides the hemispheric sign distribution of large-scale helical features in active regions \citep{Pevtsov94,Ab97}, there can be patches of right-handed and left-handed fields corresponding respectively to positive and negative helicities, intermixed in a mesh-like pattern in the sunspot umbra and a threaded pattern in the sunspot penumbra \citep{Su09}. \cite{Zhang} showed that the individual magnetic fibrils tend to be dominated by the current density component caused by magnetic inhomogeneity, while the large-scale magnetic region tends to be dominated by the component of the current density associated with the magnetic twist. \cite{Venkatakrishnan09} pointed out that the existence of global twist for a sunspot -- even in the absence of a net current -- is consistent with a fibril structure of sunspot magnetic fields. \begin{figure*} \begin{center} \includegraphics[width=\textwidth]{vector5_hel5.ps} \end{center} \caption{ Photospheric vector magnetograms (left) and plots of $J_zB_z$ (right) for the active region NOAA 11158 between 11--15 February 2011. The arrows show the transverse component of the magnetic field. Light (dark) shades indicate positive (negative) values of $B_z$ on the left and $J_zB_z$ on the right. }\label{fig:helispec1} \end{figure*} The redistribution of magnetic helicity contained within different scales was argued to be the interchange of twist and writhe due to magnetic helicity conservation \citep[cf.][]{Zeldovich83,KB99}. Furthermore, the spectral magnetic helicity distribution is important for understanding the operation of the solar dynamo \citep{BS05}. It has been argued that, if the large-scale magnetic field is generated by an $\alpha$ effect \citep{KR80}, it must produce magnetic helicity of opposite signs at large and small length scales \citep{See96,Ji99}. We call such a magnetic field bi-helical \citep{YB03}. To alleviate the possibility of catastrophic (magnetic Reynolds number-dependent) quenching of the $\alpha$ effect \citep{GD94} and slow saturation \citep{Bra01}, one must invoke magnetic helicity fluxes from small-scale magnetic fields \citep{KMRS00,BF00,BS05,BCC09,HB12}. In the present paper, we determine the spectrum of magnetic helicity and its relationship with magnetic energy from photospheric vector magnetograms of a solar active region. We use a technique that is based on the spectral representation of the magnetic two-point correlation tensor. It is related to the method of \cite{MGS82} for determining the magnetic helicity spectrum from in situ measurements of the magnetic field in the solar wind. Their key assumption allowing for the determination of magnetic helicity spectra is that of homogeneity. This technique was recently applied to data from {\em Ulysses} to show that the magnetic field at high heliographic latitudes has opposite signs of helicity in the two hemispheres and also at large and small length scales \citep{BSBG11}; see also \cite{WBM11,WBM12} for results from corresponding simulations. In the present work, a variant is proposed where we assume local statistical isotropy in the horizontal plane to compute magnetic energy and helicity spectra. \section{Data analysis} We have analyzed data from the solar active region NOAA 11158 during 11--15 February 2011, taken by the Helioseismic and Magnetic Imager (HMI) on board the {\em Solar Dynamics Observatory} (SDO). The pixel resolution of the magnetogram is about $0.5''$, and the field of view is $250''\times 150''$. Figure \ref{fig:helispec1} shows photospheric vector magnetograms (left) and the corresponding distribution of $h_C^{(z)}=J_z B_z$ (right) from the vector magnetograms of that active region on different days. Here, $J_z=\partial B_y/\partial x-\partial B_x/\partial y$, and $J_z/\mu_0$ is the vertical component of the current density in SI units with $\mu_0$ being the vacuum permeability, while in cgs units, the current density is $J_z c/4\pi$ with $c$ being the speed of light. The superscript `$(z)$' on $h_C^{(z)}$ indicates that only the vertical contribution to the current helicity density is available. It turns out that the mean value of the current helicity density, ${\cal H}_C^{(z)}=\langle{h_C^{(z)}}\rangle$, is positive and $\approx2.7$\,G$^2$\,km$^{-1}$. Furthermore, as a proxy of the force-free $\alpha$ parameter, we determine $\alpha=J_z/B_z$, which is on the average $\langle{\alpha}\rangle\approx2.8\times 10^{-5}$\,km$^{-1}$. For future reference, let us estimate the current helicity normalized to its theoretical maximum value, henceforth referred to as {\em relative} helicity. This is not to be confused with the gauge-invariant magnetic helicity relative to that of an associated potential field \citep{Berger84}. Thus, we consider the ratio \begin{equation} r_C=\langle J_zB_z\rangle\left/ \left(\langle J_z^2\rangle\langle B_z^2\rangle\right)^{1/2}\right. \end{equation} as an estimate for the relative current helicity. For the active region NOAA 11158 we find $r_C=+0.034$. This value is based on one snapshot, but similar values have been found at other times. Let us now turn to the two-point correlation tensor, $\langle B_i({\bm x},t)\,B_j({\bm x}+{\bm\xi},t)\rangle$, where ${\bm x}$ is the position vector on the two-dimensional surface, and angle brackets denote ensemble averaging or, in the present case, averaging over annuli of constant radii, i.e., $|{\bm\xi}|={\rm const}$. Its Fourier transform with respect to ${\bm\xi}$ can be written as \begin{equation} \left\langle\hat{B}_i({\bm k},t)\hat{B}_j^*\!({\bm k}',t)\right\rangle =\Gamma_{ij}({\bm k},t)\delta^2({\bm k}-{\bm k}'), \end{equation} where $\hat{B}_i({\bm k},t)=\int B_i({\bm x},t) \,e^{i{\bm k}\cdot{\bm x}}d^2x$ is the two-dimensional Fourier transform, the subscript $i$ refers to one of the three magnetic field components, the asterisk denotes complex conjugation, and ensemble averaging will be replaced by averaging over concentric annuli in wavevector space. Following \cite{MGS82}, it is possible to determine the magnetic helicity spectrum from the spectral correlation tensor $\Gamma_{ij}({\bm k},t)$ by making the assumption of local statistical isotropy. At the end of this paper we consider the applicability of this assumption in more detail. Considering that $\bm{k}$ defines the only preferred direction in $\Gamma_{ij}$, and that $k_i\hat{B}_i=0$, the only possible structure of $\Gamma_{ij}({\bm k},t)$ is \citep[cf.][]{Moffatt78} \begin{equation} \Gamma_{ij}(\bm{k},t)=\frac{2E_M(k,t)}{4\pi k}(\delta_{ij}-\hat{k}_i\hat{k}_j) +\frac{iH_M(k,t)}{4\pi k}\varepsilon_{ijk}k_k,\label{eq:helispec5} \end{equation} where $\hat{k}_i=k_i/k$ is a component of the unit vector of $\bm{k}$, $k=|\bm{k}|$ is its modulus with $k^2=k_x^2+k_y^2$, and $E_M(k,t)$ and $H_M(k,t)$ are the magnetic energy and magnetic helicity spectra\footnote{We use this opportunity to point out a sign error in the corresponding Equation~(3) of \cite{BSBG11}. Their results were however based on the equation $H_M(k)=4\mbox{Im} \langle\hat{B}_T\hat{B}_N^*\!\rangle$, which has the correct sign. Here, $\hat{B}_T$ and $\hat{B}_N$ are transverse and normal components of the Fourier-transformed magnetic field.}, normalized such that \begin{eqnarray} {\cal E}_M(t)\equiv \frac{1}{2}\langle{\bm{B}^{2}}\rangle\!&=&\!{\int}^\infty_0\!\! E_M(k,t)\,dk\nonumber ,\\ {\cal H}_M(t)\equiv \langle{\bm{A}\cdot\bm{B}}\rangle\!&=&\!\int^\infty_0\!\! H_M(k,t)\,dk.\label{eq:helispec6} \end{eqnarray} Note that the mean energy density in ${\rm erg}/{\rm cm}^{3}$ is ${\cal E}_M/4\pi$. We emphasize that the expression for $\Gamma_{ij}({\bm k},t)$ differs from that of \cite{Moffatt78} by a factor $2k$, because we are here in two dimensions, so the differential for the integration over shells in wavenumber space changes from $4\pi k^2\,dk$ to $2\pi k\,dk$. Note that the magnetic vector potential is not an observable quantity, so the magnetic helicity might not be gauge-invariant. However, if the spatial average is over all space, or if the magnetic field falls off sufficiently rapidly toward the boundaries, both ${\cal H}_M(t)$ and $H_M(k,t)$ are gauge-invariant. Indeed, with the present analysis, $H_M(k,t)$ is manifestly gauge-invariant, because it has been computed directly from the magnetic field as obtained through the photospheric vector magnetogram. The components of the correlation tensor of the turbulent magnetic field can be written in the form \begin{eqnarray} \label{eq:helitens} &&4\pi k{\bf \Gamma}(k,\phi_k)=\\ &&\left(\begin{array}{ccc} (1-\cos^2\phi_k)2E_M & -\sin2\phi_kE_M & -ik\sin\phi_k H_M\\ -\sin2\phi_kE_M & (1-\sin^2\phi_k)2E_M & \;\;ik\cos\phi_k H_M\\ ik\sin\phi_k H_M & -ik\cos\phi_k H_M & 2E_M \end{array}\right),\nonumber \end{eqnarray} where we have defined the polar angle in wavenumber space, $\phi_k = {\rm Arctan}(k_y, k_x)$, so that $k_x=k\cos\phi_k$ and $k_y=k\sin\phi_k$. For brevity, we have also skipped the arguments $k$ and $t$ on $E_M(k,t)$ and $H_M(k,t)$. In the following we present shell-integrated spectra. However, because we consider here two-dimensional spectra, they correspond to the power in annuli of radius $k$ and are obtained as \begin{eqnarray} 2E_M(k,t)&=&2\pi k\,\mbox{Re}\left\langle \Gamma_{xx}+\Gamma_{yy}+\Gamma_{zz}\right\rangle_{\phi_k},\\ kH_M(k,t)&=&4\pi k\,\mbox{Im}\left\langle\cos\phi_k\Gamma_{yz}-\sin\phi_k\Gamma_{xz}\right\rangle_{\phi_k}, \end{eqnarray} where the angle brackets with subscript $\phi_k$ denote averaging over annuli in wavenumber space. \begin{figure} \begin{center} \includegraphics[width=\columnwidth]{phelicity.ps} \end{center} \caption{(a) $2E_M(k)$ (dotted line) and $k|H_M(k)|$ (solid line) for NOAA 11158 at 23:59:54\,UT on 13 February 2011. Positive (negative) values of $H_M(k)$ are indicated by open (closed) symbols, respectively. $2E_M^{(h)}(k)$ (red, dashed) and $2E_M^{(v)}(k)$ (blue, dash-dotted) are shown for comparison. (b) Same as upper panel, but the magnetic helicity is averaged over broad logarithmically spaced wavenumber bins. \label{fig:phelicity} }\end{figure} The realizability condition \citep{Moffatt69} implies that \begin{equation} k|H_M(k,t)|\le2E_M(k,t). \label{realizability} \end{equation} It is therefore convenient to plot $k|H_M(k,t)|$ and $2E_M(k,t)$ on the same graph, which allows one to judge how helical the magnetic field is at each wavenumber. Furthermore, to assess the degree of isotropy, we also consider magnetic energy spectra $E_M^{(h)}(k)$ and $E_M^{(v)}(k)$ based respectively on the horizontal and vertical magnetic field components, defined via \begin{eqnarray} 2E_M^{(h)}(k)&=&4\pi k\,\mbox{Re}\left\langle \Gamma_{xx}+\Gamma_{yy}\right\rangle_{\phi_k},\\ 2E_M^{(v)}(k)&=&4\pi k\,\mbox{Re}\left\langle \Gamma_{zz}\right\rangle_{\phi_k}. \end{eqnarray} Under isotropic conditions, we expect $E_M(k)\approx E_M^{(h)}(k)\approx E_M^{(v)}(k)$. We now consider magnetic energy and helicity spectra for the active region NOAA 11158. The calculated region of the field of view is $256''\times 256''$, i.e.\ $512\times 512$ pixels or $L^2=(186\,{\rm Mm})^2$. We present first the results for NOAA 11158 at 23:59:54UT on 13 February 2011; see Figure~\ref{fig:phelicity}(a). It turns out that the magnetic energy spectrum has a clear $k^{-5/3}$ range for wavenumbers in the interval $0.5\,{\rm Mm}^{-1}<k<5\,{\rm Mm}^{-1}$. The magnetic helicity spectrum is predominantly positive at intermediate wavenumbers, but we also see that toward high wavenumbers the magnetic helicity is fluctuating strongly around small values. To determine the sign of magnetic helicity at these smaller scales, we average the spectrum over broad, logarithmically spaced wavenumber bins; see the lower panel of Figure~\ref{fig:phelicity}. This shows that even at smaller length scales the magnetic helicity is still positive, again consistent with the fact that this active region is at southern latitudes. \begin{figure} \begin{center} \includegraphics[width=\columnwidth]{heli_phelicity_Bran_zhq5_enea.ps} \includegraphics[width=\columnwidth]{heli_phelicity_Bran_zhq5_kHmz.ps} \end{center} \caption{ Similar to Figure~\ref{fig:phelicity}, showing $E_M(k,t)$ (upper panel) and $k|H_M(k,t)|$ (lower panel) for the other days. \label{fig:heli_phelicity_Bran_zhq5_ene} }\end{figure} To calculate the relative magnetic helicity $r_M$, we define the integral scale of the magnetic field in the usual way as \begin{eqnarray} \ell_M=\left.\int k^{-1}E_M(k)\,dk\right/\int E_M(k)\,dk. \end{eqnarray} The realizability condition of \Eq{realizability} can be rewritten in integrated form \citep[e.g.][]{KTBN13} as \begin{eqnarray} |{\cal H}_M|=\left|\int\! H_M dk\right| \le 2\!\int \!k^{-1}E_M(k) dk \equiv 2\ell_M{\cal E}_M.\quad \end{eqnarray} In particular, we have $|{\cal H}_M(t)|\le 2\ell_M{\cal E}_M(t)$. This gives \begin{eqnarray} r_M={\cal H}_M/2\ell_M{\cal E}_M, \end{eqnarray} which obeys $|r_M|\le1$. Again, this quantity is not to be confused with the gauge-invariant helicity of \cite{Berger84}. For the active region NOAA 11158 at 23:59:54\,UT on 13 February 2011 we have $\ell_M\approx5.8\,{\rm Mm}$, ${\cal H}_M\approx3.3\times10^{4}\,{\rm G}^2\,{\rm Mm}$, and ${\cal E}_M\approx6.7\times10^{4}\,{\rm G}^2$, so $r_M\approx0.042$. The relative magnetic helicity has thus the same sign as the relative current helicity. The corresponding magnetic column energy in the two-dimensional domain of size $L^2$ is $L^2{\cal E}_M/4\pi\approx1.8\times10^{24}\,{\rm erg}\,{\rm cm}^{-1}$, which is about three times larger than the values given by \cite{SZYL13}. The magnetic column helicity is $L^2{\cal H}_M\approx1.1\times10^{33}\,{\rm Mx}^2\,{\rm cm}^{-1}$. Several estimates of the gauge-invariant magnetic helicity of NOAA 11158 using time integration of photospheric magnetic helicity injection \citep{VAMC12,LS12} and nonlinear force-free coronal field extrapolation \citep{Jing12,Georg13} suggest magnetic helicities of the order of $10^{43}\,{\rm Mx}^2$. This value would be comparable to ours if the effective vertical extent were $\approx100\,{\rm Mm}$. We should remember, however, that there is no basis for such a vertical extrapolation of our two-dimensional data. Interestingly, the magnetic energy spectra $E_M^{(h)}(k)$ and $E_M^{(v)}(k)$ based respectively on the horizontal and vertical magnetic field components agree remarkably well at wavenumbers below $k=3\,{\rm Mm}^{-1}$, corresponding to length scales larger than 2\,Mm. This suggests that our assumption of isotropy might be a reasonable one. The mutual departure between $E_M^{(h)}(k)$ and $E_M^{(v)}(k)$ at larger wavenumbers could in principle be a physical effect, although there is no good reason why the magnetic field should be mostly vertical only at small scales. If it is indeed a physical effect, it should then in future be possible to verify that this wavenumber, where $E_M^{(h)}(k)$ and $E_M^{(v)}(k)$ depart from each other, is independent of the instrument. Alternatively, this departure might be connected with different accuracies of horizontal and vertical magnetic field measurements \citep{Zhang2012}. If that is the case, one should expect that with future measurements at better resolution the two spectra depart from each other at larger wavenumbers. In that case, our spectral analysis could be used to isolate potential artefacts in the determination of horizontal and vertical magnetic fields. In Figure~\ref{fig:heli_phelicity_Bran_zhq5_ene} we show $2E_M(k)$ and $k|H_M(k)|$ for different days. It turns out that on small scales the spectra are rather similar in time, and that there are differences in the amplitude mainly on large scales. Also the sign of $H_M(k)$ remains positive for the different days. We find that the mean spectral values of magnetic energy of the active region at the solar surface is consistent with a $k^{-5/3}$ power law, which is expected based on the theory of \cite{GS95} and consistent with spectra from earlier work on solar magnetic fields \citep{Ab05,St12}, ruling out the $k^{-3/2}$ spectrum suggested by \cite{Iro63} and \cite{Kra65}. \begin{figure}[t!] \begin{center} \includegraphics[width=\columnwidth]{heli_phelicity_Bran_zhq5_Hc.ps} \end{center} \caption{Unsigned current helicity spectrum, $|H_C(k)|$.} \label{fig:phc} \end{figure} Under isotropic conditions, the current helicity spectrum, $H_C(k,t)$, is related to the magnetic helicity spectrum via \citep{Moffatt78} \begin{equation} H_C(k,t)\approx k^2H_M(k,t).\label{eq:helispec9} \end{equation} It is normalized such that $\int H_C(k)\,dk=\langle\bm{J}\cdot\bm{B}\rangle$. In Figure~\ref{fig:phc} we show $|H_C(k)|$ obtained in this way. For $k\ga1{\rm Mm}^{-1}$, the current helicity spectrum shows a $k^{-5/3}$ spectrum, which is consistent with numerical simulations of helically forced hydromagnetic turbulence \citep{BS05b,Bra09}, and indicative of a forward cascade of current helicity. Similar spectra have also been obtained for the analogous case of kinetic helicity \citep{AL77,BO97}. These results imply that the relative helicity decreases toward smaller scales; see the corresponding discussion on p.~286 of \cite{Moffatt78}. \section{Conclusions} We have applied a novel technique to estimate the magnetic helicity spectrum using vector magnetogram data at the solar surface. We have made use of the assumption that the spectral two-point correlation tensor of the magnetic field can be approximated by its isotropic representation. This assumption is partially justified by the fact that the energy spectra from horizontal and vertical magnetic fields agree at wavenumbers below $2\,{\rm Mm}^{-1}$. However, it will be important to assess the assumption of isotropy in future work through comparison with simulations. An example are the simulations of \cite{Losada}, who employed however only a one-dimensional representation of the spectral two-point correlation function. Nevertheless, the present results look promising, because the sign of magnetic helicity is the same over a broad range of wavenumbers and consistent with that theoretically expected for the southern hemisphere. This is consistent with the right-handed twist inferred from all previous studies of NOAA 11158 using different methods. Except for the smallest wavenumbers, magnetic and current helicities have essentially the same sign. Therefore, a sign change is only expected at smaller wavenumbers corresponding to scales comparable to those of the Sun itself. It would be useful to extend our analysis to a larger surface area of the Sun to see whether there is evidence for a sign change toward small wavenumbers and thus large scales reflecting the global magnetic field of the solar cycle. Such a change of sign is expected from dynamo theory \citep{Bra01} and is a consequence of the inverse cascade of magnetic helicity \citep{PFL}. Figure~\ref{fig:phelicity} gives indications of an opposite sign for $k\le0.1\,{\rm Mm}^{-1}$, which corresponds to scales that are still much smaller than those of the Sun. However, measurements of spectral power on scales comparable to those of the observed magnetogram itself are not sufficiently reliable. Our results suggest that the unsigned current helicity spectrum shows a $k^{-5/3}$ power law. This is in agreement with simulations of hydromagnetic turbulence \citep{BS05b} and implies that the turbulence becomes progressively less helical toward smaller scales. Our results suggest that at a typical scale of $\ell_M\approx6\,$Mm, the relative magnetic helicity reaches values around $0.04$. This magnetic helicity must have its origin in the underlying dynamo process, and can be traced back to the interaction between rotation and stratification. \cite{Losada} parameterized these two effects in terms of a stratification parameter Gr and a Coriolis number Co and found that the relative kinetic helicity is approximately 2\,Gr\,Co. For the Sun, they estimate ${\rm Gr}=1/6.5$, so a relative helicity of 0.04 might correspond to ${\rm Co}\approx0.1$. For the solar rotation rate, this corresponds to a correlation time of about 6 hours, which translates to a depth of about 8\,Mm. Again, more precise estimates should be obtained using realistic simulations. In addition to measuring magnetic helicity over larger regions, it will be important to apply our technique to many active regions covering both hemispheres of the Sun and different times during the solar cycle. This would allow us to verify the expected hemispheric dependence of magnetic helicity. Compared with previous determinations of the hemispheric dependence of current helicity \citep{Zhang2012}, our technique might allow us to isolate instrumental artefacts resulting from different resolutions of vector magnetograms for horizontal and vertical magnetic fields. \acknowledgments We thank the referee for detailed and constructive comments that have led to significant improvements of the manuscript. This study is supported by grants from the National Natural Science Foundation (NNSF) of China under the project grants 10921303, 11221063 and 41174153 (HZ), the NNSF of China and the Russian Foundation for Basic Research under the collaborative China-Russian project 13-02-91158 (HZ+DDS), the European Research Council under the AstroDyn Research Project No.\ 227952, and the Swedish Research Council under the project grants 2012-5797 and 621-2011-5076 (AB). \newcommand{\yraa}[3]{ #1, {RAA,} {#2}, #3} \newcommand{\yapj}[3]{ #1, {ApJ,} {#2}, #3} \newcommand{\yjfm}[3]{ #1, {JFM,} {#2}, #3} \newcommand{\ymn}[3]{ #1, {MNRAS,} {#2}, #3} \newcommand{\yana}[3]{ #1, {A\&A,} {#2}, #3} \newcommand{\yprl}[3]{ #1, {PhRvL,} {#2}, #3} \newcommand{\yprd}[3]{ #1, {PhRvD,} {#2}, #3} \newcommand{\ypre}[3]{ #1, {PhRvE,} {#2}, #3} \newcommand{\yjswsc}[3]{ #1, {JSWJC,} {#2}, #3} \newcommand{\ysov}[3]{ #1, {Sov.\ Astron.,} {#2}, #3} \newcommand{\ypf}[3]{ #1, {Phys.\ Fluids,} {#2}, #3} \newcommand{\ybook}[3]{ #1, {#2} (#3)}
\section{Shear cell design} The shear cell is composed of two parallel microscope slides, which are sand-blasted (rms roughness $1\um$) to prevent slipping, except for a small window of diameter $\approx 2~\mathrm{mm}$ to probe optically the sample. For light scattering experiments, the spacing between the two slides is controlled by three stainless steel, high-precision ball bearings. The ball bearings are embedded in a custom-made rectangular frame of polydimethylsiloxane (PDMS), which contains the sample and avoids solvent evaporation. The PDMS frame is prepared by mixing two fluid components (a base and a cross-linker) with a mass-ratio 50:1, yielding a material with an elastic modulus of about 10 kPa. For microscopy observations, the two microscope slides are separated by a $250~\mu\mathrm{m}$ thick, $16 \times 16 \mathrm{mm}^2$ double-adhesive gene frame (Thermo Scientific), which acts as a spacer and avoids evaporation. A motor is used to displace one of the plates along the $x$ direction by an amount $\delta$, thereby imposing a strain $\gamma = \delta / e$, with $e$ the sample thickness. The motor speed during the displacement is $0.05~\mathrm{mm}~\mathrm{s}^{-1}$. For both light scattering and microscopy, we measure the thickness of the sample chamber by microscopy, by measuring $e\mathrm{_{obj}}$, the vertical displacement of the microscope objective when focusing the upper and lower plates, respectively, using a $20\times$ air objective. The sample thickness is obtained as $e = n e\mathrm{_{obj}}$, where $n = 1.38$ is the sample refractive index. We measure $e\mathrm{_{obj}}$ at several locations separated by 1 cm, finding no difference to within the measurement uncertainty. This implies that the maximum deviation from parallelism over 1 cm is smaller than the measurement uncertainty ($\le 8~\mu\mathrm{m}$), corresponding to less than $5\times 10^{-4}e$ (respectively, $1.3\times 10^{-3}e$) over the region sampled by light scattering (respectively, microscopy). \section{Light scattering measurements on a purely elastic sample cyclically sheared} We test our light scattering set-up by measuring the correlation function $g_2 - 1$ for a sample whose mechanical response is purely elastic, a transparent PDMS elastomer, seeded with copper particles of diameter $3~\mu\mathrm{m}$. Here, the scattering signal is dominated by the contribution of the particles, whose microscopic configuration is essentially frozen due to the stiffness of the elastomer, whose elastic modulus is $G' \approx 500~\mathrm{kPa}$. We impose a cyclic shear deformation of amplitude $\gamma = 4.6$ \%. The inset of fig.~SM1 shows $\beta^{-1}(g_2-1)$ at short time lags, $\tau$: when $\tau$ corresponds to an odd number of half-cycles, the correlation drops to zero, due to the relative motion of the scatterers associated with the affine displacement field induced by the applied strain. For a delay time equal to an integer number of cycles, a high correlation level is recovered, an echo effect similar to that reported previously in concentrated emulsions and colloidal glasses~\cite{Hebraud1997,PetekidisPRE2002}. In the inset of Fig.~SM1 the height of the echos is unity, indicating that the scatterers have recovered exactly the initial microscopic configuration, as expected for tracer particles embedded in a purely elastic matrix, whose deformation is fully reversible. In the main figure, only data for integer values of $\tau$ are plotted, but for a very extended range of delay $\tau$. No significant loss of correlation is observed over $2000$ cycles, thus demonstrating that the setup is stable enough for the dynamics to be reliably probed over thousands of cycles. \begin{figure \includegraphics[width=10cm]{figSM1} \begin{center} \noindent \textbf{Figure SM1:} (Color online) Intensity correlation functions for an elastic PDMS elastomer. Main figure: data for integer delays $\tau$. Inset: zoom on the behavior of $\beta^{-1}(g_2-1)$ at small $\tau$. Data for both half-integer and integer delays are shown. \label{fig:SM1} \end{center} \end{figure} \section{Strain-dependence of the viscoelasticity of a colloidal polycrystal} \begin{figure}[h!] \includegraphics[width=10cm]{figSM2} \begin{center} \noindent \textbf{Figure SM2:} (Color online) Strain dependence of the storage modulus, $G'$ (green triangles), and the loss modulus, $G''$ (red circles), as measured by standard oscillatory rheology, at a fixed frequency, $f=0.5$ Hz. Same sample as for the light scattering experiments (micellar polycrystal doped with silica nanoparticles of diameter 30 nm, at a volume fraction $\varphi =1\%$). The vertical lines indicate the five strain amplitudes used in the light scattering experiments discussed in the main text. \label{fig:SM2} \end{center} \end{figure} Figure SM2 shows the strain dependence of the elastic and loss moduli measured by oscillatory rheology, for the colloidal polycrystal investigated by light scattering in the main text. The vertical lines indicate the strain amplitudes in the light scattering experiments: they correspond to an intermediate regime beyond the linear regime (which ends beyond $\gamma \approx 0.3\%$), but before fluidization occurs, for $\gamma \gtrsim 6 \%$. \\ \section{Probability distribution function of the scatterers' velocity} We show here that in the asymptotic, stationary regime the grain boundaries undergo ballistic motion and that the probability distribution function (PDF) of their velocity is a Levy stable law~\cite{bouchaud90}. In our dynamic light scattering (DLS) experiments we measure the intensity correlation function $g_2(q,\tau)-1$ (see Eq.~(1) of the main text), which is related to the intermediate scattering function $f(q,\tau)$ (also known as the dynamic structure factor) by $$g_2(\mathbf{q},\tau)-1 = \beta f(\mathbf{q},\tau)^2 \,,$$ with \begin{equation} \label{eqn:fq} f(\mathbf{q},\tau) = \frac{1}{N}\left < \sum_{j,k=1}^N \exp \left \{i\mathbf{q}\cdot[\mathbf{r}_j(0)-\mathbf{r}_k(\tau)]\right\} \right > \,, \end{equation} with $N$ the number if scatterers, $\mathbf{r}_j$ the time-dependent position of the $j$-th scatterer, and where the brackets denote an ensemble average, and the factor $\beta$ an instrumental constant $\lesssim 1$~\cite{Berne1976}. As discussed in the main text, we find that in the stationary regime correlation functions measured at different $q$ all collapse on a master curve when plotted against time scaled by $q \equiv |\mathbf{q}_x|$, for $q \ge q_\mathrm{c}$. This implies that $f(q,\tau)$ does not depend on $q$ and $\tau$ separately, but rather on the product $u = q \tau$. Recalling the compressed exponential shape of $g_2-1$, one has $f(q,\tau) = f(u) = \exp[-(uV_{0,x})^p]$, where $V_{0,x}$ represents the modulus of a characteristic velocity related to the relaxation time $\tau_R$ introduced in the main text by \begin{equation} \label{eq:V0} V_{0,x} = \frac{\frac{2^{-\frac{1}{p}}}{p}\Gamma\left (\frac{1}{p} \right )}{q\tau_R}\,, \end{equation} with $\Gamma$ the gamma function. \begin{figure}[h!] \includegraphics[width=10cm]{figSM3} \begin{center} \noindent \textbf{Figure SM3:} (Color online) Probability distribution function of the $x$ component of the velocity of the grain boundaries in the asymptotic, stationary regime, for $\gamma = 4.6\%$. The PDF is obtained by Fourier transforming the compressed exponential fit to the data shown in the inset of Fig. 2b of the main text, see Eq.~(\ref{eq:pdf}) above. The distribution function is essentially flat up to the characteristic velocity $V_{0,x}$, while it decays as a power law for large $|V_x|$, with an exponent $-p-1$ directly related to the compressing exponent $p = 1.54$ of the fit to $g_2-1$. $V_{0,x}= 5.54~\mu\mathrm{m}~\mathrm{cycle}^{-1}$ is related to the $q$-dependent decay time of $g_2-1$, $\tau_R$, by Eq.~(\ref{eq:V0}). \label{fig:SM3} \end{center} \end{figure} Under these conditions, by following~\cite{Berne1976,LucaFaraday2003} one finds that the ensemble average in Eq.~(\ref{eqn:fq}) can be recast as an average over the probability distribution function of the $x$ component of the scatterers' velocity, $P_V(V_x)$, yielding: \begin{equation} \label{eq:fu} f(u) = \int \mathbf{d}V_x P_V(V_x)\exp(-iuV_x) \,. \end{equation} By taking the inverse Fourier transform of Eq.~(\ref{eq:fu}), one finds \begin{equation} \label{eq:pdf} P_V(V_x) \propto \int \mathbf{d}u f(u)\exp(iuV_x)= \int \mathbf{d}u \exp[-(uV_{0,x})^p] \exp(iuV_x) \,. \end{equation} The last term on the r.h.s. of Eq.~(\ref{eq:pdf}) is the integral representation of the Levy stable law $L_{p,0}$~\cite{LucaFaraday2003,bouchaud90}. The Levy PDF is characterized by a flat distribution for $|V_x| \lesssim V_{0,x}$, followed by a power-law tail, $P_V(|V_x|) \sim |V_x|^{-p-1}$~\cite{bouchaud90}. We show in Fig. SM3 the PDF obtained by numerical integration of Eq.~(\ref{eq:pdf}), for the dynamics measured in the asymptotic regime shown in Fig. 2b of the main text ($\gamma = 4.6\%$, $p = 1.54$, $V_{0,x} = 5.54\times 10^{-4}~\mu\mathrm{m}~\mathrm{cycle}^{-1}$). \begin{table} [h!] \begin{center} \begin{tabular}{l|l|l} \multicolumn{1}{c|}{$\gamma$ (\%)} & \multicolumn{1}{c|}{$p$} & \multicolumn{1}{c}{$V_{0,x}~\mathrm{(}\mu\mathrm{m}~\mathrm{cycle}^{-1}\mathrm{)}$} \\ \hline 1.5 & 1.66 & $3.52 \times 10^{-4}$\\ 2.5 & 1.76 & $2.82 \times 10^{-4}$\\ 3.5 & 1.65 & $2.55 \times 10^{-4}$\\ 4.6 & 1.54 & $5.64 \times 10^{-4}$\\ 5.2 & 1.53 & $6.54 \times 10^{-4}$\\ \end{tabular} \end{center} \vspace{0.1 cm} {\noindent \textbf {Table SM4:} Compressing exponent $p$ governing the power-law tail of the velocity distribution of the grain boundaries and characteristic velocity obtained from the fits to the intensity correlation functions $g_2-1$ in the asymptotic, stationary regime, for the five values of the applied strain amplitude.} \end{table} Table SM4 summarizes the parameters of the Levy distributions of $V_x$ for all our experiments. We find that the $p$ exponent governing the slope of the tail of $P_V$ varies only slightly with the applied strain, $\gamma$. The variation of the characteristic velocity is more pronounced: the general trend is for $V_{0,x}$ to increase with $\gamma$, albeit with some scatter in the data. The order of magnitude of $V_{0,x}$ is $5 \times 10^{-4}\mu\mathrm{m}~\mathrm{cycle}^{-1}$. This is close to $V_{0,x} \sim 10^{-3}\mu\mathrm{m}~\mathrm{cycle}^{-1}$, as obtained by analyzing in real space the grain boundary displacement for a sample with a slightly larger grain size, as discussed in Sec.~\ref{sec:cm_vs_dls} below. \section{Comparison between light scattering and microscopy measurements} \label{sec:cm_vs_dls} \begin{figure \includegraphics[width=12.5cm]{figSM5} \begin{center} \noindent \textbf{Figure SM5} (Color online) a) - d): zoom into representative GB trajectories, from Fig. 1c of the main text. The trajectories are obtained by measuring the GB position at $t = 1, 112, 2617, 3130$, and 3711 shear cycles. e): GB displacement with respect to the position at $t=1$, for the trajectories shown in a)-d), as indicated by the labels. The angle $\alpha$ between consecutive segments of a trajectory is also shown for two segments of the trajectory c). f): probability distribution function of $\alpha$. The PDF is strongly peaked around $\alpha = 0$, implying that the GBs trajectories are close to straight lines, a behavior suggestive of ballistic motion and incompatible with diffusion, for which $\alpha$ would be evenly distributed (dotted line). The labels indicate the average and the standard deviation of $\alpha$. \label{fig:SM5} \end{center} \end{figure} In order to provide additional support to the analysis of the grain boundary motion performed on light scattering data, we measure the GB displacement also in microscopy experiments. Figures SM5a-d zoom into some of the trajectories shown in Fig. 1c of the main text. The trajectories are obtained by measuring the position of representative GBs at times $t = 1, 112, 2617, 3130$, and 3711 cycles. The trajectories are overlaid to the images of the polycrystal taken at $t=1$ (red) and $3711$ (blue). The images at intermediate times are not shown for clarity. For the same GBs, the displacement $(\Delta x,\Delta y)$ with respect to the position at $t=1$ is shown in Fig. SM5e. Clearly, the trajectories are close to straight lines, a behavior incompatible with random motion. To quantify the tendency of the GBs to move along straight lines, we calculate the angle $\alpha$ between successive segments of the trajectories, as exemplified in Fig. SM5e. Figure SM5f shows the PDF of $\alpha$, obtained from all the trajectory segments at all the locations shown in Fig. 1c of the main text, \textit{i.e.} 52 segments from 13 different trajectories. The PDF is strongly peaked around $\alpha = 0$, confirming that the trajectories are incompatible with diffusive motion and are rather suggestive of straight-line displacements as in ballistic motion, consistent with the DLS results. \begin{figure}[h!] \includegraphics[width=9cm]{figSM6} \begin{center} \noindent \textbf{Figure SM6:} (Color online) Cumulative distribution function of the velocity components $V_\alpha$ of the grain boundaries in the asymptotic, stationary regime ($\alpha = x$ (resp. , $y$) for the component parallel (resp., perpendicular) to the direction of the applied strain). CM and symbols: data obtained by analyzing the trajectories obtained from confocal microscopy and shown in Fig. 1c of the main text ($\gamma = 3.6\%$). Lines: data obtained by light scattering for the five strain amplitudes reported in the main text. \label{fig:SM3} \end{center} \end{figure} We measure the GB velocity over two time intervals, $t\in [2617-3130]$ and $t\in [3130-3711]$. We find comparable average velocities, consistently with the notion that the sample dynamics become stationary at large $t$, as seen by DLS. We calculate the cumulative distribution of the GB velocity, using all trajectories and both time intervals and compare it to the cumulative velocity distributions obtained from the DLS data analysis. The results are shown in Fig. SM6 for the components of the velocity parallel and perpendicular to the shear direction. Several comments are in order. First, there is an overall good agreement between the microscopy and DLS data. This agreement is particularly remarkable given that the sample composition (nanoparticle kind, size and concentration) has been separately optimized according to the specific requirements of each experiment. As a consequence, the grain size --although of the same order of magnitude-- is not identical in the samples used for microscopy and light scattering (see Figs. 1b and 1d in the main text). Second, microscopy data measured for the $x$ and $y$ direction overlap, thus indicating that plasticity is essentially isotropic. Third, the order of magnitude of the GB velocity in the stationary regime is very small, as also seen in Figs. SM3 and SM5. This highlights a key requirement of our experiments, \textit{i.e.} sensitivity to small-scale motion. In this respect, light scattering is superior to microscopy: for the former, the smallest rms displacement that can be reliably measured is $\Delta r_{\mathrm{min}} \sim 0.1\um$, corresponding to a decay of 5\% of $g_2-1$ at the largest scattering vector. For microscopy, $\Delta r_{\mathrm{min}} \sim 0.34 \um$ (corresponding to 1 pixel), more than three times larger than by DLS. An additional advantage of light scattering is a better statistics: for DLS, the probed volume is $V_{\mathrm{scatt}} \sim 1 ~\mathrm{mm}^3$, about 80 times larger than $V_{\mathrm{micro}} \sim 0.013~\mathrm{mm}^3$, the volume accessible to confocal microscopy. These advantages motivate our choice of DLS as the main quantitative probe of the GB dynamics.
\section{Introduction}\label{introduction} The \emph{growth rate} (or the $d$-function) of a finite algebra~$\m A$ is $d_{\m A}(n)$ = the least size of a generating set for $\m A^n$. For a solvable group, this rate is always linear in~$n$. On the other hand, unary algebras (which are also solvable) have exponential growth rate. In this paper we investigate the relationship between the growth rate of $\m A$ and its structural properties in the case, when $\m A$ is finite and solvable. It turns out that stronger abelianness properties yield a closer relationship between various growth-restricting conditions. For example, if the variety generated by $\m A$ is abelian, or if $\m A$ is a subdirect product of simple abelian algebras, then the growth rate is non-exponential if and only if $\m A$ has a Maltsev term, in which case the growth rate is linear. This result does not hold for nilpotent (hence for solvable) algebras. To define the abelianness properties investigated in this paper the commutator from Chapter~3 of \cite{hobby-mckenzie} is used. Some fluency in tame congruence theory is required to understand the proofs. The properties are the following: \begin{enumerate} \item[(1)] $\m a$ is solvable, \item[(2)] $\m a$ is (left) nilpotent (see \cite{sym}), \item[(3)] $\m a$ is abelian, \item[(4)] $\m a$ is a subdirect product of simple abelian algebras, or \item[(5)] $\m a$ generates an abelian variety. \end{enumerate} It is not hard to show from the definitions that these properties are related by the implications (4)$\Rightarrow$(3)$\Rightarrow$(2)$\Rightarrow$(1) and (5)$\Rightarrow$(3). It is also not too hard to show that no other implications hold except those that are formal consequences of these. (To verify this latter statement, one must find examples showing that (1)~$\not\Rightarrow$~(2), (2)~$\not\Rightarrow$~(3), (3)~$\not\Rightarrow$~(4), (3)~$\not\Rightarrow$~(5), (4)~$\not\Rightarrow$~(5), and (5)~$\not\Rightarrow$~(4), and for many of these there exists an example that is a group. In the two situations where there is no group example, (3)~$\not\Rightarrow$~(5) and (4)~$\not\Rightarrow$~(5), the algebra from Example~8.7 of \cite{strnil} shows that a simple abelian algebra can generate a nonabelian variety.) \goodbreak Now consider the following six growth-restricting conditions: \begin{enumerate} \item[(i)] $\m a$ has a Maltsev polynomial. \item[(ii)] $\m a$ has a pointed cube polynomial (see Definition~\ref{pointed_cube}). \item[(iii)] $\m A$ is a spread of its type~$\atyp$ minimal sets (see Definition~\ref{spread}). \item[(iv)] $d_{\m a}(n)\in O(n)$. \item[(v)] $d_{\m a}(n)\notin 2^{\Omega(n)}$. \item[(vi)] No finite power $\m a^n$ has a nontrivial strongly abelian homomorphic image. \end{enumerate} For arbitrary finite algebras, these conditions are related by the implications (i)$\Rightarrow$(iv), (i)$\Rightarrow$(ii)$\Rightarrow$(v), and (iii)$\Rightarrow$(iv)$\Rightarrow$(v)$\Rightarrow$(vi) and no implications hold that are not consequences of these, as we show in Theorem~\ref{basic}. In this paper we show that as we assume stronger and stronger hypotheses from the list (1)--(5), the conditions (i)--(vi) gradually become equivalent to one another. The following theorem summarizes our main results. \begin{thm}\label{main} Let~$\m A$ be a finite algebra. \begin{enumerate} \item If~$\m A$ is solvable, then (i)$\Rightarrow$(iii) by Corollary~\ref{kkv_cor} and (ii)$\Rightarrow$(iv) by Theorem~\ref{1-solv}. \item If~$\m A$ is left nilpotent, then (ii)$\Rightarrow$(i) by Theorem~\ref{nilpotent} and (vi)$\Rightarrow$(iii) by Theorem~\ref{spreadnilp}. Therefore (i) and (ii) are equivalent, (iii), (iv), (v) and (vi) are equivalent, and the first group of equivalent conditions implies the second group. The same implications (and no others, cf.\ Example~\ref{abelian_spread_not_maltsev}) hold under the stronger assumption that $\m a$ is abelian. \item If $\m A$ is a subdirect product of simple abelian algebras, then (iv)$\Rightarrow$(i) by Theorem~\ref{simplefact}. Therefore all six conditions are equivalent for $\m a$. \item If $\m A$ generates an abelian variety, then (iii)$\Rightarrow$(i) by Theorem~\ref{spreadaff}. Therefore all six conditions are equivalent for $\m a$. \end{enumerate} \end{thm} These results are expressed diagrammatically in Section~\ref{summary_sec}. Our theorem completely determines the relationships between the growth-restrict\-ing conditions (i)--(vi) for nilpotent and abelian algebras, but some questions remain open for solvable algebras. We do know that (i)$\Rightarrow$(ii)$\Rightarrow$(iv)$\Rightarrow$(v)$\Rightarrow$(vi) and (i)$\Rightarrow$(iii)$\Rightarrow$(iv) for solvable algebras, and we also know that (iii)$\not\Rightarrow$(ii)$\not\Rightarrow$(i). We know nothing else about the implications between these properties for finite solvable algebras that are not nilpotent. The original purpose of our investigation was to show that the $d$-function of a finite solvable algebra grows at a linear or exponential rate. We were not able to prove or refute this statement. We considered the other conditions in order to better understand what might force the $d$-function of a solvable algebra into $O(n)$ or $2^{\Omega(n)}$. If, for example, one could now show that (vi)$\Rightarrow$(iii) for finite solvable algebras, then one would have that the $d$-functions of such algebras grow linearly or exponentially. \section{Preliminaries}\label{quoted} \subsection{Notation} We use Big Oh notation. If $f$ and $g$ are real-valued functions defined on some subset of the real numbers, then $f\in O(g)$ means that there are positive constants $M$ and $N$ such that $|f(x)|\leq M|g(x)|$ for all $x>N$. We write $f\in \Omega(g)$ to mean that there are positive constants $M$ and $N$ such that $|f(x)|\geq M|g(x)|$ for all $x>N$. Finally, $f\in \Theta(g)$ means that both $f\in O(g)$ and $f\in \Omega(g)$ hold. \subsection{Easy Estimates} \begin{thm}[Theorem 2.2.1 of \cite{kksz-A}]\label{basic_estimates} Let $\m a$ be an algebra. \begin{enumerate} \item[(1)] $d_{\m a^k}(n) = d_{\m a}(kn)$. \item[(2)] If\/ $\m b$ is a homomorphic image of $\m a$, then $d_{\m b}(n) \leq d_{\m a}(n)$. \item[(3)] If\/ $\m b$ is an expansion of $\m a$ (with operations; equivalently, if $\m a$ is a reduct of\/~$\m b$), then $d_{\m b}(n) \leq d_{\m a}(n)$. \item[(4)] {\rm(From \cite{quick-ruskuc})} If\/ $\m b$ is the expansion of $\m a$ obtained by adjoining all constants, then \[ d_{\m a}(n) - d_{\m a}(1)\leq d_{\m b}(n) \leq d_{\m a}(n). \] \end{enumerate} \kern-10pt \qed \end{thm} \begin{thm}[From Theorem 2.2.2 of \cite{kksz-A}]\label{first_bounds} If $\m a$ is a finite algebra of more than one element and $n>0$, then \[ \lceil \log_{|A|}(n) \rceil\leq d_{\m a}(n) \leq |A|^n. \] \kern-10pt \qed \end{thm} Recall that the \emph{free spectrum} of a variety $\mathcal V$ is the function $f_{\mathcal V}(n):=|F_{\mathcal V}(n)|$ whose value at $n$ is the cardinality of the $n$-generated free algebra in $\mathcal V$. \begin{thm}[From Theorem 2.2.4 of \cite{kksz-A}]\label{spec1} If $\m a$ is a nontrivial finite algebra and $f_{\mathcal V}$ is the free spectrum of the variety $\mathcal V = {\mathcal V}(\m a)$, then \begin{enumerate} \item[(1)] if $f_{\mathcal V}(n)\in O(n^k)$ for some fixed $k\in\mathbb Z^+$, then $d_{\m a}(n)\in 2^{\Theta(n)}$; \item[(2)] if $f_{\mathcal V}(n)\in 2^{O(n)}$, then $d_{\m a}(n)\in \Omega(n)$. \qed \end{enumerate} \end{thm} \begin{cor}[Corollary 2.2.5 of \cite{kksz-A}]\label{abelian_cor} Let $\m a$ be a nontrivial finite algebra and let\/~$\m b$ be a nontrivial homomorphic image of $\m a^k$ for some $k$. \begin{enumerate} \item[(1)] If\/ $\m b$ is strongly abelian (or even just strongly rectangular), then $d_{\m a}(n)\in 2^{\Theta(n)}$. \item[(2)] If\/ $\m b$ is abelian, then $d_{\m a}(n)\in \Omega(n)$. \qed \end{enumerate} \end{cor} \begin{thm}[Theorem 2.2.6 of \cite{kksz-A}]\label{affine_prep} If $\m a^2$ is a finitely generated affine algebra, then $d_{\m a}(n)\in O(n)$. If, moreover, $\m a$ is finite and has more than one element, then $d_{\m a}(n)\in \Theta(n)$. \qed \end{thm} \subsection{Basic relationships among conditions (i)--(vi)} Recall that a polynomial $F(x,y,z)$ of $\m a$ is a \emph{Maltsev polynomial} if $F(x,y,y)\approx x \approx F(y,y,x)$ holds in $\m a$. In \cite{kksz-A}, this well known concept is generalized to the following: \begin{df}\label{pointed_cube} A term $F(x_1,\ldots,x_m)$ is an \emph{$m$-ary, $p$-pointed, $k$-cube term} for $\m a$ if there is a $k\times m$ matrix $M$ consisting of variables and $p$ distinct constant symbols, with every column of $M$ containing a symbol different from $x$, such that $\m a$ satisfies \begin{equation}\label{cube} F(M)\approx \left( \begin{matrix} x\\ \vdots\\ x \end{matrix} \right). \end{equation} \end{df} For example, \[ {F}\left(\begin{matrix} x&y&y\\ y&y&x \end{matrix} \right) \approx \left(\begin{matrix} x\\ x \end{matrix} \right), \] is a way to express that $F$ is a Maltsev term in the form (\ref{cube}) with $m=3$, $p=0$ and $k=2$. The main results from \cite{kksz-A} about the restriction on growth imposed by a pointed cube term now follow. \begin{thm}[Theorem 5.2.1 of \cite{kksz-A}]\label{pointed_polynomial} Let $\m a$ be an algebra with an $m$-ary, $p$-pointed, $k$-cube term, with at least one constant symbol appearing in the cube identities (that is, $p\geq 1$). If\/ $\m a^{p+k-1}$ is finitely generated, then all finite powers of $\m a$ are finitely generated and $d_{\m a}(n)$ is bounded above by a polynomial of degree at most $\log_w(m)$, where $w = 2k/(2k-1)$. \qed \end{thm} \begin{cor}[Corollary 5.2.4 of \cite{kksz-A}]\label{pointed_polynomial_cor} If $\m a^k$ is a finitely generated algebra with a $0$-pointed or $1$-pointed $k$-cube term, then $d_{\m a}(n)\in O(n^{k-1})$. \qed \end{cor} \begin{df}\label{pointed_cube_pol} An $m$-ary, $p$-pointed, $k$-cube term for the constant expansion of $\m a$ is called an \emph{$m$-ary, $p$-pointed, $k$-cube polynomial} for $\m a$. \end{df} \begin{rem} An algebra has a Maltsev polynomial (or a pointed cube polynomial) if and only if its constant expansion has a Maltsev term (or pointed cube term). Passing to the constant expansion does not affect the growth rate significantly, as noted in Theorem~\ref{basic_estimates}~(4). Therefore, the last two results hold for polynomials as well as for terms. In fact, replacing ``term'' with ``polynomial'' in Theorem~\ref{pointed_polynomial} and Corollary~\ref{pointed_polynomial_cor} yields correct statements. In the literature, a $0$-pointed cube term is called a cube-term. Note that a $p$-pointed cube term is not necessarily a cube-term even if all constants are unary terms. \end{rem} \begin{thm}[From Theorem 5.4.1 of \cite{kksz-A}]\label{avoid} Let $\m a$ be an algebra with $|A|>1$ whose $d$-function assumes only finite values. There is an algebra $\m b$ such that $d_{\m b}(n) = d_{\m a}(n)$ for all $n$, where $\m b$ does not have a pointed cube polynomial. \qed \end{thm} For the other concepts appearing in the conditions (i)--(vi), we direct the reader to Definition~\ref{spread} for ``spread'' and to \cite{hobby-mckenzie} for ``minimal set'' and ``strongly abelian''. We can now establish which implications between conditions (i)--(vi) from the Introduction hold for all finite algebras. \begin{thm}\label{basic} For arbitrary finite algebras, the conditions \begin{enumerate} \item[(i)] $\m a$ has a Maltsev polynomial. \item[(ii)] $\m a$ has a pointed cube polynomial. \item[(iii)] $\m A$ is a spread of its type~$\atyp$ minimal sets. \item[(iv)] $d_{\m a}(n)\in O(n)$. \item[(v)] $d_{\m a}(n)\notin 2^{\Omega(n)}$. \item[(vi)] No finite power $\m a^n$ has a nontrivial strongly abelian homomorphic image. \end{enumerate} are related by the implications \[ (i)\Rightarrow(iv),\quad (i)\Rightarrow(ii)\Rightarrow(v), \quad \textrm{and}\quad (iii)\Rightarrow(iv)\Rightarrow(v)\Rightarrow(vi). \] The following non-implications can be established with finite counterexamples: \[ (vi)\not\Rightarrow (v)\not\Rightarrow (iv)\not\Rightarrow (iii)\not\Rightarrow (ii)\not\Rightarrow (iv)\not\Rightarrow (ii) \quad \textrm{and}\quad (i)\not\Rightarrow (iii). \] These facts imply that the implications that hold are only those indicated in the following diagram and their consequences: \bigskip \begin{center} \begin{picture}(200,60) \setlength{\unitlength}{1mm} \put(0,0){$(iii)$} \put(10,0){$\Longrightarrow$} \put(20,0){$(iv)$} \put(30,0){$\Longrightarrow$} \put(40.5,0){$(v)$} \put(50,0){$\Longrightarrow$} \put(60,0){$(vi)$.} \put(22,10){\rotatebox[origin=c]{270}{$\Longrightarrow$}} \put(21,20){$(i)$} \put(30,20){$\Longrightarrow$} \put(40,20){$(ii)$} \put(42,10){\rotatebox[origin=c]{270}{$\Longrightarrow$}} \end{picture} \end{center} \medskip \end{thm} We shall prove this theorem in Section~\ref{summary_sec}. \section{A new characterization of solvability}\label{solvchar} In this section we show how elementary translations coming from idempotent polynomials characterize solvability. \begin{df}\label{t-digraph} Let $\m A$ be an algebra and $p$ an idempotent polynomial of~$\m A$ (that is, $p(x,x,\ldots,x)=x$ for every $x\in A$). The \emph{translation-digraph} $\Tr(p)$ of $p$ has vertex set~$A$, and directed edges of the form $(c,c')=\big(p(c,c,\ldots,c), p(c,\ldots,c,d,c,\ldots,c)\big)$, where $c,d\in A$. (The element $d$ occurs in exactly one of the arguments of $p$, but it can be any of the arguments.) \end{df} Recall that a \emph{neighborhood} $U$ of an algebra $\m A$ is the range of an idempotent unary polynomial of $\m A$, that is, $U=e(A)$ for a unary polynomial $e$ satisfying $e\big(e(x)\big)=e(x)$ for every $x\in A$. \begin{thm}\label{solvcharthm} Let $\m A$ be a finite algebra. Then $\m A$ is solvable if and only if for every neighborhood $U$ of~$\m A$, and every idempotent polynomial $p$ of the induced algebra~$\m A|_U$, the directed graph $\Tr(p)$ is strongly connected. \end{thm} The proof of Theorem~\ref{solvcharthm} is included at the end of this section. The essential part of the proof is to establish the theorem for classes of solvable minimal congruences. \begin{lm}\label{transl} Let $\m S$ be a finite, simple, abelian algebra and $p$ an idempotent polynomial of\/~$\m S$. Then $\Tr(p)$ is strongly connected. \end{lm} \begin{proof} Since $S$ is connected by traces, it is sufficient to prove that for every trace~$N$ and every $a,b\in N$ there is a directed path in $\Tr(p)$ connecting $a$ to~$b$. Consider the multitraces with respect to~$N$, that is, all sets of the form $q(N,N,\ldots,N)$, where $q$ is any polynomial of~$\m S$. Choose a multitrace $M$ that contains~$N$ and is maximal under inclusion. Then $p(M,\ldots,M)$ contains~$M$ (and therefore $N$), since $p$ is idempotent, and is a multitrace, so $p(M,\ldots,M)=M$. Theorem~3.10 of \cite{kkv1} shows that the induced algebra $\m S|_M$ is polynomially equivalent to the full matrix power $(\m S|_N)^{[k]}$ for some~$k$ (more precisely, $\m S|_M$ is isomorphic to an algebra on $N^k$ that is polynomially equivalent to $(\m S|_N)^{[k]}$). Consider first the case, when the type of~$\m S$ is~$\utyp$. Let $a$ correspond to the vector $(a_1,\ldots,a_k)$ and $b$ correspond to $(b_1,\ldots,b_k)$. We may assume that $a$ and $b$ differ only in one coordinate, and, to simplify the notation, that this is the first coordinate. That is, $a_i=b_i$ for $i\ge 2$. Indeed, if such pairs can be connected by a directed path, then by changing only one coordinate at a time we can connect $a$ to $b$. By the definition of a matrix power, the operation of $p$ induced on~$M$ can be represented as follows: \[ p(\wec{x}^1,\ldots,\wec{x}^\ell)= \begin{pmatrix} p_1(x_1^1,x_1^2,\ldots,x_i^j,\ldots,x_k^\ell)\\ \vdots\\ p_k(x_1^1,x_1^2,\ldots,x_i^j,\ldots,x_k^\ell) \end{pmatrix}\,, \] where $\wec{x}^j$ is (the transpose of) $(x_1^j, \ldots, x_k^j)$ and each $p_i$ is a $k\ell$-ary operation of~$\m S|_N$. Hence each $p_i$ is essentially unary. We use that $p$ is idempotent. Consider $p(\wec{x},\ldots,\wec{x})$ and move the first coordinate $x_1$ of $\wec{x}$ in~$N$ while keeping all other coordinates fixed. Then the value of $p_1$ must change, so $p_1$ must depend on one of the variables $x_1^1,x_1^2,\ldots,x_1^\ell$, and therefore on no other variable. Let this variable be $x_1^j$. The fact that $p$~is idempotent shows that $p_1(x_1^1,x_1^2,\ldots,x_k^\ell)=x_1^j$. An analogous statement holds for all other rows, in particular, each $p_i$ is a projection, hence it is idempotent. Substitute $b=(b_1,\ldots,b_k)$ to the $j$-th variable of $p$ and $a=(a_1,\ldots,a_k)$ to all other variables. Then in the first row we get $b_1$, and in the $i$-th row we get $a_i=b_i$ (if $i\ge 2$). Therefore the result of the operation is $b$, and so there is an edge $(a,b)$ in $\Tr(p)$. If the type of~$\m S$ is $\atyp$, then $\m S|_N$ is polynomially equivalent to a $1$-dimensional vector space over a field~$\m F$, and $(\m S|_N)^{[k]}$ is polynomially equivalent to the module $N^k$ over $\m F^{k\times k}$, where the matrices act on $N^k$ by multiplication. Since $p$ is idempotent, \[ p(\wec{x}^1,\ldots,\wec{x}^\ell)= M_1\wec{x}^1+\ldots+M_\ell\,\wec{x}^\ell \] holds for some matrices $M_i\in \m F^{k\times k}$ satisfying that their sum is the identity matrix. Notice that for any $\wec{x},\wec{y}\in N^k$ we have \[ M_1(\wec{x}+\wec{y})+M_2\,\wec{x}+\ldots+M_\ell\,\wec{x}= \wec{x}+M_1\wec{y}\,, \] so $(\wec{x},\wec{x}+M_1\wec{y})$ is an edge of $\Tr(p)$, and similarly, $(\wec{x},\wec{x}+M_i\,\wec{y})$ is an edge, too, for every $1\le i\le \ell$. Now let $\wec{a},\wec{b}\in N^k$. Since \[ M_1(\wec{b}-\wec{a})+\ldots+M_\ell(\wec{b}-\wec{a})=\wec{b}-\wec{a}\,, \] we have \[ \wec{a}+M_1(\wec{b}-\wec{a})+\ldots+M_\ell(\wec{b}-\wec{a})=\wec{b}\,, \] and by our remark, this yields a path of $\ell$ edges in $\Tr(p)$ connecting $\wec{a}$ to~$\wec{b}$. \end{proof} \begin{proof}[Proof of Theorem~\ref{solvcharthm}] Let $A$ be a finite solvable algebra. We argue by induction on~$|A|$. Let $U$ be a neighborhood of~$\m A$. Then the induced algebra~$\m A|_U$ is solvable, too, so we can assume that $U=A$. Choose and fix an idempotent polynomial~$p$ of~$\m A$. Let $\alpha$ be a minimal congruence of~$\m A$. By the induction assumption, $\m A/\alpha$ is strongly connected with respect to $p/\alpha$. Thus if $a,b\in A$ are given, then there exist edges $(u_1,v_1), \ldots, (u_k,v_k)$ of $\Tr(p)$ such that $a\equiv_\alpha u_1$, $v_k\equiv_\alpha b$, and each $v_i\equiv_\alpha u_{i+1}$ for $1\le i\le k-1$. Therefore it is sufficient to prove that each $\alpha$-class is strongly connected. Let $S$ be such a class. Then the induced algebra $\m A|_S$ is simple, abelian, and $p|_S$ is a polynomial of this induced algebra, since $p$ is idempotent. Therefore Lemma~\ref{transl} shows that $S$ is strongly connected indeed. For the converse, suppose that $A$ is not solvable. Then there is a prime quotient $\langle\delta,\theta\rangle$ in $\mathop{}\mathopen{}\mathrm{Con}(\m A)$ of nonabelian type. Let $U$ be a $\langle\delta,\theta\rangle$-minimal set. By \cite{hobby-mckenzie}, Lemma~2.10 and Theorems~4.15, 4.17, and 4.23, $U$ is a neighborhood, and there is a pseudo-meet operation~$p$ on~$\m A|_U$. This operation is idempotent, satisfies the identity $p(x,y)=p\big(x,p(x,y)\big)$, and there is an element $1\in U$ such that $p(1,x)=p(x,1)=x$ for every $x\in U$. These properties imply that there is no nontrivial edge in $\Tr(p)$ whose endpoint is~$1$, hence this graph is not strongly connected. Indeed, if $p(c,d)=1$, then \[ 1=p(c,d)=p\big(c,p(c,d)\big)=p(c,1)=c\,, \] and then $1=p(c,d)=p(1,d)=d$, so $c=d=1$. This calculation shows that any edge terminating at $1$ must originate from $1$. Thus the proof of Theorem~\ref{solvcharthm} is complete. \end{proof} \section{Solvable algebras with a pointed cube term}\label{solv} \begin{thm}\label{1-solv} Let $\m A$ be a finite, nontrivial solvable algebra that has a pointed cube term. Then $d_{\m a}(n)\in \Theta(n)$. In fact, the algebra $\m A^n$ is generated by those elements of $A^n$ that are constant with the possible exception of one component. \end{thm} (This establishes for solvable algebras the extra implication (ii)$\Rightarrow$(iv) among the conditions of Theorem~\ref{basic}.) \begin{proof} The quotient modulo any maximal congruence of $\m A$ is abelian, so the combination of Corollary~\ref{abelian_cor}~(2) and Theorem~\ref{basic_estimates}~(2) shows that $d_{\m a}(n)\in \Omega(n)$. Therefore it is sufficient to prove that $d_{\m a}(n)\in O(n)$. For a given~$n$, let $G$ consist of those elements of $A^n$ that are constant with the possible exception of one component. The size of~$G$ is $|A|+n|A|(|A|-1)$ for $n\ge 3$, which is linear in $n$. Thus it is enough to prove that $G$ is a generating set of~$\m A^n$. Let $\m B$ be the subalgebra generated by~$G$. Note that $B$ is symmetric under any permutation of coordinates. This will allow us to simplify notation in the proof below by rearranging coordinates at certain points. We may (and shall) assume that our cube identities involve only the variable~$x$, since all other variables can be replaced by a fixed constant from~$A$. We induct on $|A|$. Let $\alpha$ be a minimal congruence of~$\m A$. By our induction hypothesis, $(A/\alpha)^n$ is generated by $G/(\alpha^n)$, since this is the set of all sequences in $(A/\alpha)^n$ that are constant with the possible exception of one component. Therefore $B$ intersects each $\alpha^n$-class. Our proof shall proceed by ``repairing'' these representatives component by component. For a given $1\le m\le n$ call a sequence $(b_1,\ldots,b_m)$ of elements of~$A$ \emph{complete}, if for every $a_{m+1},\ldots,a_n$ there exist elements $b_{m+1},\ldots,b_n$ such that $(b_1,\ldots,b_n)\in B$ and $b_i\equiv_\alpha a_i$ for all $i>m$. We shall say informally that \emph{$(b_1,\ldots,b_m)$ can be extended into~$B$ along $a_{m+1},\ldots,a_n$.} The argument in the previous paragraph shows that the empty sequence is complete (when $m=0$), and we endeavor to show that every element of $A^n$ is complete, which means that $B=A^n$. Therefore it is sufficient to prove the following. \begin{clm}\label{1-solvclm} Suppose that each element of $A^{m-1}$ is complete. Then each element of $A^m$ is complete. \end{clm} \cproof To set up notation, we permute the coordinates in the following way. We assume that $(b_2,\ldots,b_m)$ is complete and intend to show that $(c,b_2,\ldots,b_m)$ is complete for every~$c$. Fix $a_{m+1},\ldots, a_n\in A$. We have to extend $(c,b_2,\ldots,b_m)$ into $B$ along $a_{m+1},\ldots, a_n$. The completeness of $(b_2,\ldots,b_m)$ allows us to find elements $b,b_{m+1},\ldots,b_n$ such that $b\equiv_\alpha c$,\; $b_j\equiv_\alpha a_j$ for all $j>m$ and $\wec{b}=(b_1,\ldots,b_n)\in B$. In other words, $(b,b_2,\ldots,b_m)$ can be extended into $B$ along $a_{m+1},\ldots, a_n$. Let $S=b/\alpha=c/\alpha$. The induced algebra $\m S=\m C|_S$ is simple and abelian. Consider a cube identity $F(x,\ldots, x,u_1,\ldots,u_p)\approx x$ (here we rearranged the variables appropriately and $u_1,\ldots,u_p$ are constants). For each constant $u_j$ extend $(b_2,\ldots,b_m)$ into~$B$ along $u_j$ and $a_{m+1},\ldots, a_n$. We get a vector $\wec{u}^j\in B$ whose first coordinate is denoted by~$u_j'$. Consider the polynomials \begin{align}\label{pq} p(x_1,\ldots,x_\ell) &= F^{\m A}(x_1,\ldots, x_\ell,u_1,\ldots,u_p)\\ q(x_1,\ldots,x_\ell) &= F^{\m A}(x_1,\ldots, x_\ell,u_1',\ldots,u_p')\,.\notag \end{align} The first polynomial is idempotent because of the cube identity above. Therefore $p(S,\ldots,S)\subseteq S$, and $p$~restricted to~$S$ is a polynomial of~$\m S$. Since $u_j\equiv_\alpha u_j'$, the polynomial~$q$ can also be restricted to~$S$. The congruence~$\alpha$ is abelian. This implies that $\pi(x)=q(x,x,\ldots,x)$ is a permutation of~$S$. Indeed, if $q(s,\ldots s)=q(t,\ldots,t)$ for some $s,t\in S$, then using the term condition for $F$ (which can be applied, since $s\equiv_\alpha t$ and $u_j\equiv_\alpha u_j'$) we get that $p(s,\ldots, s) =p(t,\ldots,t)$. Since~$p$ is idempotent, we see that ${s=t}$. Thus $\pi$ is indeed a permutation. Let $o$ be the order of $\pi$ in the symmetric group on~$S$ and $r(x_1,\ldots,x_\ell) = \pi^{o-1}\big(q(x_1,\ldots,x_\ell)\big)$. This is an idempotent polynomial of $\m S$. Lemma~\ref{transl} can be applied to the polynomial~$r$ to connect $b$ to~$c$ by edges in $\Tr(r)$. We plan to go along this path and extend into~$B$ step by step, so it is enough to show how to do a single edge. For the simplicity of notation we may assume that $(b,c)$ itself is an edge. That is, also by appropriately rearranging the arguments of~$F$, that $r(d,b,\ldots,b)=c$ for some element~$d\in S$. This means that $q(d,b,\ldots,b)=\pi(c)$. Now choose a second cube identity, where the first variable is not~$x$. We shall apply the term~$F$ to a matrix~$M$ whose elements we now describe. The number of rows is~$n$ and the number of columns is the arity of~$F$. We shall distinguish five types of columns of~$M$, that is, arguments of~$F$. The definitions of these types and the way we fill the entries of the matrix~$M$ are shown on Figure~\ref{FigI-V} (page~\pageref{FigI-V}). \newmdenv[linewidth=1pt]{mybox} \begin{figure}[ht] \begin{mybox} \leftline{The types of columns of~$M$:} \begin{enumerate} \item[(I)] The first column/argument is the only one to have type~I. There is $x$ in the first cube identity and a constant named $s$ in the second cube identity. \item[(II)] Type II column/argument: both cube identities have~$x$ in this argument. \item[(III)] Type III: there is $x$ in the first cube identity, and a constant in the second. We exclude the first argument (which also has this property). \item[(IV)] Type IV: there is a constant in the first cube identity, and $x$ in the second. \item[(V)] Type V: there are constants in both cube identities. \end{enumerate} \leftline{The way to fill the columns of~$M$:} \begin{enumerate} \item[(V)] In a type V argument, let $u_k$ be the constant in the first cube identity, and $w$ the constant in the second cube identity. Then put $u_k'$ (defined above) into the first coordinate and $w$ to all other coordinates. (Of course for different columns these constants may differ.) This column is in $G$. \item[(IV)] Into a type IV column, put the appropriate vector $\wec{u}^j$, also defined above. Recall that this vector extends $(b_2,\ldots,b_m)$ along $u_j,a_{m+1},\ldots,a_n$ into $B$, and its first entry is $u_j'$. \item[(III)] For a type III argument, let $t$ be the constant appearing in the second cube identity. Put $b$ to the first coordinate and $t$ everywhere else. This column is in~$G$. \item[(II)] Put the vector $\wec{b}\in B$ defined above to all type II columns. \item[(I)] To the first column (the only column of type~$I$), put the element~$d$ to the first coordinate, and $s$ (the constant in the first variable of the second cube identity) everywhere else. This is also in~$G$. \end{enumerate} \[ \begin{matrix} \text{I} & \text{II} & \text{III} & \text{IV} & \text{V}\\ d & b & b & u_j' & u_k'\\ s & b_2 & t & b_2 & w\\ s & b_3 & t & b_3 & w\\ \vdots &\vdots &\vdots &\vdots &\vdots\\ s & b_m & t & b_m & w\\ s & b_{m+1} & t & b_{m+1}' & w\\ \vdots &\vdots &\vdots &\vdots &\vdots\\ s & b_n & t & b_n' & w \end{matrix} \] \end{mybox} \caption{The elements substituted into the five types of columns}\label{FigI-V} \end{figure} Applying $F$ to this matrix the first row yields $q(d,b,\ldots,b)=\pi(c)$. All other rows can be evaluated using the second cube identity. The result is $b_j$ for $2\le j\le m$ and is $\alpha$-related to $a_j$ for $j>m$, since $b_j, b_j'\equiv_\alpha a_j$. This is almost the vector we are looking for, we only have to replace $\pi(c)$ by $c$ in the first component. In other words, it is sufficient to prove the following. \begin{subclm}\label{deletepi} Suppose that $(\pi(c_1),c_2,\ldots,c_n)\in B$ holds for some vector where $c_1\in S$. Then $(c_1,c_2,\ldots,c_n)\in B$. \end{subclm} \subcproof The equations in display (\ref{pq}) and $\pi(x)=q(x,x,\ldots,x)$ show that we have $\pi^{o-1}(x)=g(x,u_1',\ldots,u_p')$ for a term $g$ obtained from $F$ by composition, and if we replace each $u_i'$ by $u_i$, then we get that $g(x,u_1,\ldots,u_p)=x$ (the $(o-1)$st power of the identity map). The columns of the matrix \[ \begin{pmatrix} \pi(c_1) & u_1' & \ldots & u_p'\\ c_2 & u_1 & \ldots & u_p\\ \vdots &\vdots &\vdots &\vdots\\ c_n & u_1 & \ldots & u_p\\ \end{pmatrix} \] are in $B$ (for the first column we assumed this, and the others are in~$G$). Applying the term~$g$ to the rows we get that $(c_1,c_2,\ldots,c_n)\in B$, proving the subclaim and Claim~\ref{1-solvclm}. This completes the proof of Theorem~\ref{1-solv}, too. \end{proof} \section{Spreads and growth rate}\label{spreadsec} \begin{df}\label{spread} Let $\m A$ be an algebra and $\mathcal U$ a collection of subsets of~$A$. We say that a subset $S\subseteq A$ is a \emph{spread} of~$\mathcal U$ if there exists a polynomial~$p$ of~$\m A$ and (not necessarily distinct) elements $U_1,\ldots,U_k\in\mathcal U$ such that $p(U_1,\ldots,U_k)=S$. \end{df} \begin{lm}\label{spreadaffrate} Suppose that $\m A$ is a finite algebra such that all constants are term operations, $p$ is a polynomial of~$\m A$ and $A=p(U_1,\ldots, U_k)$ holds for some subsets $U_i\subseteq A$. Then \[ d_{\m A}(n)\le d_{\m U_1}(n)+\ldots+d_{\m U_k}(n)\,, \] where $\m U_i=\m A|_{U_i}$ are the algebras $\m a$ induces on these sets. \end{lm} \begin{proof} For a given~$n$ let $G_i$ be a smallest size generating set of the algebra $\m U_i^n$. The set $G:=\bigcup_{i=1}^k G_i$ has size at most $d_{\m U_1}(n)+\ldots+d_{\m U_k}(n)$, and generates a subalgebra of $\m a^n$ containing all sets of the form $U_i^n$. Since $p$ is a term operation of $\m a$, and \[ p^{\m a^n}(U_1^n,\ldots,U_k^n)=\left(p^{\m a}(U_1,\ldots,U_n)\right)^n = A^n \] we get that $G$ generates $\m a^n$. \end{proof} \begin{cor}\label{spreadafflin} If a finite algebra $\m A$ is a spread of a family of subsets on which the induced algebras have Maltsev polynomials, then the growth rate of~$\m A$ is at most linear. \end{cor} \begin{proof} Use Corollary~\ref{pointed_polynomial_cor} and Lemma~\ref{spreadaffrate}. \end{proof} Several years ago, K.~A.~Kearnes, E.~W.~Kiss and M.~A.~Valeriote proved but did not publish the fact that any finite algebra with a Maltsev polynomial is \emph{covered} by its $\langle\alpha,\beta\rangle$-minimal sets. This statement is a bit stronger than the statement that any finite algebra with a Maltsev polynomial is a spread of its minimal sets, so it is relevant here. To introduce the terminology, if $\mathcal U$ and $\mathcal V$ are sets of neighborhoods of $\m a$, then $\mathcal U$ \emph{covers} $\mathcal V$ if for every $V\in {\mathcal V}$ there are neighborhoods $U_1,\ldots,U_k\in {\mathcal U}$, idempotent unary polynomials $e_1,\ldots, e_k, f$, and other polynomials $\lambda, \rho_i$ such that $f(A)=V$, $e_i(A)=U_i$ and \begin{equation}\label{decomposition} \lambda\big(e_1\rho_1(x),\ldots,e_k\rho_k(x)\big)=f(x) \end{equation} for all $x\in A$. (Here the $e_i$'s need not be distinct.) Equation (\ref{decomposition}) expresses $V$ as a retract of the product of the $U_i$'s via polynomial maps $\lambda(x_1,\ldots,x_k)$ and $\big(e_1\rho_1(x),\ldots,e_k\rho_k(x)\big)$. It is not hard to see that if $\mathcal U$ covers $\mathcal V$ and $\mathcal V$ covers $\mathcal W$, then $\mathcal U$ covers $\mathcal W$. We say that $\mathcal U$ covers the algebra $\m a$ if $\mathcal U$ covers the set $\{A\}$. Lemma~2.10 of \cite{hobby-mckenzie} guarantees that the minimal sets of a finite algebra are neighborhoods. If some set $\mathcal U$ of minimal sets covers $\m a$, and this is witnessed as in (\ref{decomposition}) by polynomials $e_i, \lambda, \rho_i$ and $f={\rm id}$, then equation (\ref{decomposition}) yields $\lambda(U_1,\ldots,U_k)=A$, so $A$ is a spread of the minimal sets in $\mathcal U$. Hence if $\m a$ is covered by its minimal sets, then it is a spread of its minimal sets. \begin{thm}\label{kkv} If $\m a$ is a finite algebra with a Maltsev polynomial, then $\m a$ is covered by its $\langle\alpha,\beta\rangle$-minimal sets, where $\langle\alpha,\beta\rangle$ runs over all prime quotients of\/~$\Conb(\m a)$. \end{thm} \begin{proof} Let $N\subseteq A$ be a neighborhood of $\m a$ that is maximal for the property that $\m a|_N$ is covered by its minimal sets. If $N=A$, then we are done, so we assume otherwise and argue to a contradiction. By Theorem~4.31 of~\cite{hobby-mckenzie}, type~$\atyp$ minimal sets in an algebra with a Maltsev polynomial are E-minimal. Hence a minimal set of a finite algebra with a Maltsev polynomial is nothing other than a set that is minimal under inclusion among nontrivial neighborhoods. Therefore the minimal sets of $\m a|_N$ are exactly the minimal sets of $\m a$ that lie in $N$. It is a basic fact of tame congruence theory (a consequence of Lemma~2.17 of~\cite{hobby-mckenzie}) that every finite algebra is the connected union of the traces of its minimal sets for congruences, hence if $N\neq A$ then there must exist $\alpha\prec\beta$ in $\Conb(\m a)$ and an $\lb\alpha,\beta\rangle$-minimal set $U$ such that $U$ has a trace $T\subseteq U$ that properly overlaps $N$. (I.e., $T\cap N\neq\emptyset$, but $T\not\subseteq N$.) Choose $0\in T\cap N$. Thus $U$ properly overlaps $N$ and $U\cap N$ contains an element $0$ from (the body of) $U$. Let $m(x,y,z)$ denote a Maltsev polynomial of $\m a$. \begin{clm} $\m a$ has idempotent unary polynomials $e$ and $f$ such that \begin{enumerate} \item $e(A) = U$ and $e(N) = \{0\}$, and \item $f(A) = N$ and $f(U) = \{0\}$. \end{enumerate} \end{clm} Since $U$ and $N$ are neighborhoods there exist idempotent unary polynomials $e_0$ and $f_0$ such that $e_0(A)=U$ and $f_0(A)=N$. We show that these can be modified to have the extra properties in the claim. Our first step is to construct $e$. As a first case, assume that $e_0f_0\in \mathop{}\mathopen{}\mathrm{Pol}_1(\m a)|_U$ is not a permutation of $U$. The polynomial $g(x)=e_0m\big(0,e_0f_0(x),e_0(x)\big)$ has the property that $g(A)\subseteq U$ and $g(N)=\{0\}$. Take $u \in T$ such that $(u,0) \in \beta - \alpha$. Then $\big(e_0f_0(u),0\big)\in \alpha$, and we get $\big(g(u),u\big)\in \alpha$. Since $(0,u)\in\beta-\alpha$, $g(0)=0$, and $g(u)\equiv u\pmod{\alpha}$ it follows that $g(\beta)\not\subseteq \alpha$ and so $g$ is not collapsing on $U$. Therefore an appropriate power $e = g^k$ is an idempotent unary polynomial with range $U$ that collapses $N$ to $\{0\}$. Now assume that $e_0f_0$ is a permutation of $U$. Then $U\simeq f_0(U)$, so $V := f_0(U)\subseteq N$ is an $\lb\alpha,\beta\rangle$-minimal set contained in $N$ and containing $0$. Let $S=f_0(T)$ be the $\lb\alpha,\beta\rangle$-trace of $V$ containing $0$. Corollary~4.8 of \cite{kkv2} guarantees that there is an idempotent unary polynomial $e_1$ of $\m a$ such that in the quotient $\m a/\alpha$ we have $\overline{e}_1(A/\alpha) = U/\alpha$ and $\overline{e}_1(S/\alpha) = \{0/\alpha\}$. Replacing $e_1$ by $e_0e_1$ if necessary we may assume that $e_0e_1 = e_1$, so $e_1(A)\subseteq U$. Since $e_1$ maps $A$ into $U$ and it is the identity modulo $\alpha$ on $U$, it follows from the $\lb\alpha,\beta\rangle$-minimality of $U$ that $e_1(A)=U$. Now $e_1f_0$ is collapsing on~$T$, so $e_1f_0$ is not a permutation of $U$. Thus we can repeat the argument of the first case using~$e_1$ in place of $e_0$ to construct an idempotent unary polynomial $e$ such that $e(A)=U$ and $e(N)=\{0\}$. Now, given $e$ and $f_0$ as above let $f(x) = f_0m\big(f_0(x),m\big(e(x),x,f_0(x)\big),0\big)$. One calculates that $f(A)\subseteq N$, $f$ is the identity on $N$ (so $f$ is idempotent with range $N$), and $f(U)=\{0\}$. This completes the proof of the claim.\hfill\rule{1.3mm}{3mm} \medskip Now we finish the proof of the theorem. The polynomial $h(x)=m\big(e(x),0,f(x)\big)$ is the identity on $N\cup U$, so some iterate $h^{\ell}(x)$ is idempotent with range $h^{\ell}(A)=M\supsetneq N$. The equation \[ h^{\ell}m\big(e\big(h^{\ell-1}(x)\big),0, f\big(h^{\ell-1}(x)\big)\big)=h^{2\ell}(x)=h^{\ell}(x) \] is an equation of type (\ref{decomposition}) for $\m a|_M$, showing that $\m a|_M$ is covered by its subneighborhoods $N$ and $U$. Because the covering relation is transitive, $\m a|_M$ is covered by~$U$ together with the minimal sets in $N$ which cover $\m a|_N$. This shows that $\m a|_M$ is covered by the minimal sets it contains, contradicting the maximality assumption on~$N$. \end{proof} \begin{cor}\label{kkv_cor} If\/ $\m a$ is a finite solvable algebra with a Maltsev polynomial, then $\m a$~is a spread of its type $\atyp$ minimal sets. \end{cor} (This establishes for solvable algebras the extra implication (i)$\Rightarrow$(iii) among the conditions of Theorem~\ref{basic}.) \begin{proof} Theorem~\ref{kkv} proves that $\m a$ is a spread of its minimal sets. They all have type~$\atyp$, since $\m a$ is solvable and has a Maltsev polynomial. \end{proof} \section{Left nilpotent algebras}\label{abel} In the papers \cite{sym}, \cite{strnil}, \cite{nilpwab} various forms of nilpotence are discussed. An algebra~$\m A$ is left nilpotent if $[1_{\m A},\dots,[1_{\m A},[1_{\m A},1_{\m A}]]\dots]=0_{\m A}$ for a sufficiently long expression. This definition implies that the class of left nilpotent algebras is closed under subalgebras and finite direct products. \begin{thm}\label{nilpotent} A finite left nilpotent algebra has a Maltsev polynomial iff it has a pointed cube polynomial. \end{thm} (This establishes for nilpotent algebras the extra implication (ii)$\Rightarrow$(i) among the conditions of Theorem~\ref{basic}.) \begin{proof} A Maltsev polynomial is a $3$-ary, $0$-pointed, $2$-cube polynomial. To prove the converse, we can assume that all elements of~$A$ are the interpretations of nullary operation symbols. Thus if $\m A$ has a pointed cube polynomial, then it is in fact a pointed cube term, and so it is interpreted as a pointed cube term in every algebra of the variety~$\vr V$ generated by~$\m A$. Thus, Theorem~\ref{pointed_polynomial} implies that no algebra in~$\vr V$ can have exponential growth rate. However, nontrivial strongly abelian algebras have exponential growth rate by Corollary~\ref{abelian_cor}. Therefore no such algebra exists in~$\vr V$. We finish the proof by recalling Theorem~6.8 of~\cite{kkv1}, which states that if a variety~$\vr V$ is generated by a left nilpotent algebra, and there is no nontrivial strongly abelian algebra in~$\vr V$, then $\m A$ has a Maltsev polynomial. \end{proof} The following corollary follows from the fact that if an abelian algebra has a Maltsev polynomial, then it has a Maltsev term. \begin{cor}\label{affine_cor} A finite abelian algebra has a pointed cube polynomial iff it is affine. \qed \end{cor} Our next goal is to prove the following theorem. \begin{thm}\label{spreadnilp} If $\m A$ is a finite, left nilpotent algebra, and $\m A^{|A|}$ does not have a nontrivial strongly abelian quotient algebra, then $\m A$ is a spread of its type~$\atyp$ minimal sets. \end{thm} (This establishes for nilpotent algebras the extra implication (vi)$\Rightarrow$(iii) among the conditions of Theorem~\ref{basic}.) \begin{lm}\label{proj} Let $\m A$ be a finite solvable algebra and $\eta_1,\ldots,\eta_n$ congruences of~$\m A$. Then for each type~$\atyp$ covering~$\delta\prec\theta$ in $\mathop{}\mathopen{}\mathrm{Con}(\m A)$ such that $\bigwedge_{i=1}^n \eta_i\le\delta$ there is an $i$ and congruences $\eta_i\le\alpha\prec\beta$ such that $\lb\delta,\theta\rangle$ and $\lb\alpha,\beta\rangle$ are projective in~$\mathop{}\mathopen{}\mathrm{Con}(\m A)$ (and hence have the same minimal sets). \end{lm} \begin{proof} Let $\gamma$ be an arbitrary congruence of~$\m A$. We prove that either $\lb\gamma\vee\delta,\gamma\vee\theta\rangle$ or $\lb\gamma\wedge\delta,\gamma\wedge\theta\rangle$ is perspective to~$\lb\delta,\theta\rangle$. Indeed, since $\delta\prec\theta$, it is sufficient to prove that $\theta\wedge(\gamma\vee\delta)\ne\theta$ or $\delta\vee(\gamma\wedge\theta)\ne\delta$. Suppose that both inequalities fail. The quotient lattice of $\Conb(\m A)$ modulo the strongly solvability congruence is modular by Theorem~7.7~(4) of~\cite{hobby-mckenzie}. Therefore $\theta=\theta\wedge(\gamma\vee\delta)$ and $\delta=\delta\vee(\gamma\wedge\theta)$ are related by the strongly solvability congruence. This contradicts the fact that $\lb\delta,\theta\rangle$ has type~$\atyp$. Now apply this observation to $\gamma=\eta_1$. Either the statement of the lemma is satisfied with $i=1$, or $\lb\eta_1\wedge\delta,\eta_1\wedge\theta\rangle$ is perspective to~$\lb\delta,\theta\rangle$. Let $\delta_1=\eta_1\wedge\delta$ and $\theta_1$ be any upper cover of $\delta_1$ that is below $\eta_1\wedge\theta$. Then $\lb\delta_1,\theta_1\rangle$ is still perspective to $\lb\delta,\theta\rangle$, since $\delta\vee\theta_1=\theta$, but it is now a prime quotient. Perspective prime quotients have the same type, so $\lb\delta_1,\theta_1\rangle$ also has type~$\atyp$. Now apply the previous observation to $\eta_2$ and this new quotient. Again, either the statement of the lemma holds for $i=2$, or we can push our cover down below $\eta_2$. Continuing this process, if the statement of the lemma fails for every~$i$, then we get a prime quotient $\lb\delta_n,\theta_n\rangle$ that is still projective to~$\lb\delta,\theta\rangle$, and $\delta_n=\big(\bigwedge_{i=1}^n\eta_i\big)\wedge\delta$, while $\theta_n\le \big(\bigwedge_{i=1}^n\eta_i\big)\wedge\theta$. But our assumption that $\bigwedge_{i=1}^n \eta_i\le\delta$ implies that $\theta_n=\delta_n=\bigwedge_{i=1}^n \eta_i$, which is a contradiction. \end{proof} \begin{cor}\label{type2} Let $\m A$ be a finite solvable algebra and $\mathcal U$ a set containing exactly one member from each polynomial isomorphism class of type $\atyp$ minimal sets of $\m A$. If\/ $\delta\prec\theta$ is a covering of type $\atyp$ in $\Conb(\m A^n)$, then there is a $\lb \delta,\theta\rangle$-minimal set of the form $U^n$ for some $U\in \mathcal U$. \end{cor} \begin{proof} Denote by $\eta_1,\ldots,\eta_n$ the projection kernels of~$\m A^n$, and apply Lemma~\ref{proj} to this set of congruences on~$\m A^n$. We get that the covering $\delta\prec\theta$ is projective to a covering $\alpha\prec\beta$ which lies above some $\eta_i$. Thus the $\lb\delta,\theta\rangle$-minimal sets are the same as the $\lb\alpha,\beta\rangle$-minimal sets. Identifying $\m a$ with $\m A^n/\eta_i$, choose some $U\in\mathcal U$ that is an $\lb \alpha/\eta_i,\beta/\eta_i\rangle$-minimal set. Then (a) $U^n$ is the image of an idempotent unary polynomial of~$\m a^n$, (b) $\m a^n|_{U^n}$ is an E-minimal algebra of type $\atyp$ (since $\m a^n|_{U^n}$ is polynomially equivalent to $(\m a|_U)^n$ and powers of solvable E-minimal algebras are E-minimal, according to Lemma 4.10 of \cite{sym}), and (c) $\alpha|_{U^n}\neq \beta|_{U^n}$. Items (a)--(c) are enough to show that $U^n$ is a minimal set for $\lb \alpha,\beta\rangle$ and hence for $\lb\delta,\theta\rangle$. \end{proof} It is proved in~\cite{sym} that left nilpotence implies the following, weaker condition: \begin{equation}\label{tracenilp} \text{$\NC(1_{\m A},N^2;\delta)$ holds whenever $\delta\prec\theta$ and $N$ is a $\lb\delta,\theta\rangle$-trace.} \tag{$\dag$} \end{equation} This condition is clearly still stronger than solvability. One of the main results of~\cite{ham} is that in an algebra satisfying~(\ref{tracenilp}), every maximal subalgebra is a block of a congruence. We shall use this result in the following proof. \begin{proof}[Proof of Theorem~\ref{spreadnilp}] Let $\mathcal U=\{U_1, \ldots, U_k\}$ be a set containing exactly one member from each polynomial isomorphism class of type $\atyp$ minimal sets of $\m A$. Let $n=|A|$ and let $\m B$ be the subalgebra of $\m A^n$ generated by all sets of the form $U^n$, where $U\in\mathcal U$. Thus $B$ is the union of the sets $t^{\m A^n}(U_1^n,U_2^n,\ldots,U_k^n)$, where $t$ is a term of~$\m A$. Clearly $t^{\m A^n}(U_1^n,U_2^n,\ldots,U_k^n) = \left(t^{\m A}(U_1,U_2,\ldots,U_k)\right)^n$. There are two cases. If $B=A^n$, then let $\wec{a}$ be a listing of all elements of~$\m A$. Then $\wec{a}\in A^n=B$, and so there exists a term $t$ such that $\wec{a}\in t^{\m A}(U_1,U_2,\ldots,U_k)^n$. Then $t^{\m A}(U_1,U_2,\ldots,U_k)=A$, and so $\m A$ is a spread of~$\mathcal U$. In the other case, $\m B$ is a proper subalgebra of~$\m A^n$. Let $\m M$ be a maximal subalgebra of~$\m A^n$ containing~$B$. Since $\m A^n$ is left nilpotent, the result quoted above implies that $M$ is a block of some congruence~$\mu$ of~$\m A^n$. We shall prove that $\m A^n/\mu$ is strongly solvable. This is sufficient, since then the quotient modulo any maximal congruence containing~$\mu$ is strongly abelian. Suppose that there is a prime quotient~$\mu\le\delta\prec\theta$ of type~$\atyp$. Corollary~\ref{type2} states that some $U_i^n$ is a minimal set for $\lb\delta,\theta\rangle$. This is a contradiction, since $U_i^n\subseteq B\subseteq M$ is contained in a single $\mu\le\delta$-class. Thus the proof of~Theorem~\ref{spreadnilp} is complete. \end{proof} \section{Abelian varieties}\label{abvar} We show that in an abelian variety, any algebra that is a spread of affine subsets is affine. Note that if $\m A$ is a finite abelian algebra, $\lb\alpha,\beta\rangle$ is a prime quotient of type~$\atyp$, and $U\in M_{\m A}(\alpha,\beta)$, then the induced algebra~$\m A|_U$ is affine. Indeed, $\m A$ is solvable, so Lemma~4.27~(4) of \cite{hobby-mckenzie} implies that the tail of~$U$ is empty. By Theorem~4.31 of~\cite{hobby-mckenzie}, the induced algebra on~$U$ is Maltsev (and E-minimal), and as $\m A$ is abelian, it is affine. Recall that an abelian group operation on an algebra $\m A$ is \textsl{compatible}, if adding $x-y+z$ as a term makes $\m A$ an affine algebra. \begin{thm}\label{spreadaff} Let $\m A$ be an algebra and $U_1,\ldots,U_k\subseteq A$ such that $A=t(U_1,\ldots,U_k)$ for a polynomial~$t$ of~$\m A$. If all induced algebras $\m A|_{U_i}$ are affine, then the following hold. \begin{enumerate} \item If\/ $\mathsf{H}(\m A^2)$ is abelian, then there is an abelian group operation~$+$ on~$A$ that is compatible with all operations of~$\m A$. Moreover, if $\alpha\in\mathop{}\mathopen{}\mathrm{Con}(\m A)$, then $\alpha$ is a congruence of the group $(A, +)$. \item If the variety ${\mathcal V}(\m A)$ generated by $\m A$ is abelian, then $\m A$ is affine. \end{enumerate} \end{thm} (This establishes for algebras generating abelian varieties the extra implication (iii)$\Rightarrow$(i) among the conditions of Theorem~\ref{basic}). \begin{proof} For each $1\le i\le n$ choose and fix an element $0_i\in U_i$, a binary polynomial $+_i$ and a unary polynomial $-_i$ of $\m A|_{U_i}$ such that $(U_i,+_i,-_i,0_i)$ is an abelian group. To define the operation $+$ on $A$ let $a,b\in A$. Then $a=t(\wec{a})$ and $b=t(\wec{b})$ for some $a_i,b_i\in U_i$. Define \[ a+b=t(a_1 +_1 b_1,\ldots, a_k+_k b_k)\,. \] This operation is well-defined, as we now show. Assuming that $a=t(\wec{a'})$, we have to prove that $t(a_1 +_1 b_1,\ldots,\allowbreak a_k+_k b_k)= t(a_1' +_1 b_1,\ldots,\allowbreak a_k'+_k b_k)$. This is an instance of the abelianness of $\m A$, because $t(\wec{a})=t(\wec{a'})$ implies $t(a_1 +_1 0_1,\ldots, a_k+_k 0_k)=t(a_1' +_1 0_1,\ldots, a_k'+_k 0_k)$. The argument is the same for the second variable of~$+$. It is clear that this operation is associative, commutative, $0=t(0_1,\ldots, 0_k)$ is a zero element, and $-t(\wec{a})=t(-_1 a_1,\ldots, -_k a_k)$ is the (well-defined) negative of $t(\wec{a})$. To prove that $+$ is compatible it is enough to show that the congruence~$\Delta$ of $\m A^2$ obtained by collapsing the diagonal satisfies $(a,b)\equiv_\Delta (c,d) \iff a+d=b+c$. Suppose first that $a+d=b+c$, we prove that $(a,b)\equiv_\Delta (c,d)$. Let $a=t(\wec{a})$, $b=t(\wec{b})$ and $c=t(\wec{c})$. Define $d_i=b_i +_i c_i -_i a_i$, clearly $d=t(\wec{d})$. We have $(a_i,a_i)\equiv_\Delta (c_i,c_i)$, and using the unary polynomial of $\m A^2$ that maps $(x,y)$ to $(x+_i 0_i,y+_i b_i -_i a_i)$ we get that $(a_i,b_i)\equiv_\Delta (c_i,d_i)$. Applying $t$ we see that $(a,b)\equiv_\Delta (c,d)$ indeed. So far, we have only used that $\m A$ is abelian, now we need that $\m A^2/\Delta$ is abelian, too. Suppose that $(a,b)\equiv_\Delta (c,d)$. Let $a=t(\wec{a})$, $b=t(\wec{b})$, $c=t(\wec{c})$ and $d=t(\wec{d})$. We have that \[ \begin{pmatrix} t(a_1+_1 0_1,\ldots, a_k+_k 0_k)\\ t(b_1+_1 0_1,\ldots, b_k+_k 0_k) \end{pmatrix} \equiv_\Delta \begin{pmatrix} t(c_1+_1 0_1,\ldots, c_k+_k 0_k)\\ t(d_1+_1 0_1,\ldots, d_k+_k 0_k) \end{pmatrix}\,. \] Since $\m A^2/\Delta$ is abelian, we can replace each $0_i$ in the second row by $a_i-_i b_i$. Therefore \[ \begin{pmatrix} t(a_1,\ldots, a_k)\\ t(a_1,\ldots, a_k) \end{pmatrix} \equiv_\Delta \begin{pmatrix} t(c_1,\ldots, c_k)\\ t(d_1+_1 a_1-_k b_1,\ldots, d_k+_k a_k -_k b_k) \end{pmatrix}\,. \] That is, $(a,a)\equiv_\Delta (c,d+a-b)$. The abelianness of $\m A$ implies that the diagonal is a $\Delta$-class, hence $c=d+a-b$. Thus $+$ is indeed a compatible operation. To prove that every congruence $\alpha$ of $\m A$ is a group-congruence it is sufficient to verify that if $(a,b)\in\alpha$, then $(a+c,b+c)\in\alpha$. Again, let $a=t(\wec{a})$, $b=t(\wec{b})$ and $c=t(\wec{c})$. Then \[ t(a_1+_1 0_1,\ldots, a_k+_k 0_k) \equiv_\alpha t(b_1+_1 0_1,\ldots, b_k+_k 0_k)\,. \] Since $\m A/\alpha$ is abelian, we can move each $0_i$ to $c_i$, proving~(1). Now suppose that ${\mathcal V}(\m A)$ is abelian. Recall that an algebra is called \textsl{Hamiltonian,} if every subalgebra is a block of some congruence. By the main result of~\cite{kiss-val}, every member of ${\mathcal V}(\m A)$ is Hamiltonian. Notice that $A^n=\hat t(U_1^n,\ldots,U_k^n)$ holds in $\m A^n$, where $\hat t$ is $t$ acting componentwise. Hence (1) applies to all powers of $\m A$, and the compatible group operation on~$\m A^n$ is the operation of the group $(A,+)^n$. Construct the free algebra $\m F$ of ${\mathcal V}(\m A)$ generated by $x$, $y$ and $z$ as a subalgebra of $\m A^n$, where $n=|A|^3$. By the Hamiltonian property, there exists a congruence $\alpha$ of $\m A^n$ such that $F$ is a class of $\alpha$. Item (1) of the theorem implies that $\alpha$ is a group congruence, so $F$ is a coset modulo a subgroup, hence $x-y+z\in F$. Therefore $x-y+z$ is a term operation of~$\m A$. \end{proof} We present two examples showing that no obvious weakening of the conditions in the previous theorem is possible. Our starting point is the example following Corollary~4.4 of \cite{kkv2}. \begin{exmp} Let $\BZ_2$ be the two-element field, let $A=\BZ_2^3$, denote by $+$ the (elementary abelian) group operation on~$A$, and consider the following matrices in $\BZ_2^{3\times 3}$: \[ F_1= \begin{pmatrix} 1&0&0\\ 0&1&0\\ 1&0&0 \end{pmatrix}\,, \qquad F_2= \begin{pmatrix} 1&0&0\\ 0&1&0\\ 0&0&0 \end{pmatrix}\,, \qquad G= \begin{pmatrix} 1&0&0\\ 1&1&0\\ 0&0&1 \end{pmatrix}\,. \] Define a binary operation $*$ on $A$ by $u*v=F_1u+F_2v$ (matrix-vector multiplication) and a unary operation $g$ on $A$ by $g(v)=Gv+(1,0,0)^T$, so \[ g: \begin{pmatrix} a\\ b\\ c \end{pmatrix} \mapsto \begin{pmatrix} a+1\\ a+b\\ c \end{pmatrix}\,. \] These are affine operations, so the algebra $\m A=\langle A;*,g\rangle$ is abelian, and $x-y+z$ is a compatible operation. Define three subgroups of $\langle A;+\rangle$ as follows: $B$ is the subgroup generated by $(0,0,1)^T$, $C$ is the subgroup generated by $(0,0,1)^T$ and $(0,1,0)^T$, finally $D$ is the subgroup generated by $(0,1,0)^T$. Let $\beta$, $\gamma$ and $\delta$ denote the corresponding congruences of $\langle A;+\rangle$. It is easy to check by hand or by computer that $\m A$ has only these three nontrivial proper congruences. We have $\beta\wedge\delta =0_A$ and $\beta\vee\delta =\gamma$. It can be verified also that $\mathsf{H}(\m A)$ is abelian. The type of $\langle 0_A,\beta\rangle$ is $\utyp$, while both $\langle\beta,\gamma\rangle$ and $\langle\gamma,1_A\rangle$ have type~$\atyp$. These latter quotients have the same minimal sets. There are four such type~$\atyp$ minimal sets, one of which is $U=\{(0,0,0)^T,(0,1,0)^T,(1,0,0)^T,(1,1,0)^T\}$. Thus the induced algebra $\m A|_U$ is affine (since $\m A$ is abelian). We also have that $U*U=A$, so the initial condition in Theorem~\ref{spreadaff} is satisfied. The algebra $\m A^2/\Delta$ is an $8$-element abelian algebra (although it has a nonabelian quotient, so $\mathsf{H}(\m A^2)$ is not abelian). Nevertheless, the abelianness of $\m A^2/\Delta$ and $\mathsf{H}(\m A)$ are sufficient to make the proof of statement~(1) of Theorem~\ref{spreadaff} work for~$\m A$ (which indeed has a compatible $+$ operation). But $\m A$ is not affine, so we cannot drop the assumption from statement~(2) in Theorem~\ref{spreadaff} that ${\mathcal V}(\m A)$ is abelian. \end{exmp} \begin{exmp} Now let $A$, $*$, $g$ be as in the preceding example, and let $h$ be the transposition of $A$ which switches $(1,0,0)^T$ and $(1,0,1)^T$ (and fixes all other elements of $A$). Let $\overline{\m a}=\langle A;*,g,h\rangle$. This algebra has only four congruences, the nontrivial proper ones are $\beta$ and $\gamma$ above. The quotients $\langle\beta,\gamma\rangle$ and $\langle\gamma,1_A\rangle$ still have type~$\atyp$, and they have the same minimal sets, but now there are $16$ of them. One of these is $U$ above, so the initial condition in Theorem~\ref{spreadaff} is satisfied again. It is also true that $\mathsf{H}(\overline{\m a})$ is abelian. However, $h$ is not linear, and $\overline{\m a}^2/\Delta$ is a $5$-element nonabelian algebra. Therefore it is not sufficient to assume in (1) of Theorem~\ref{spreadaff} that $\mathsf{H}(\m A)$ is abelian. \end{exmp} \section{Semisimple algebras}\label{fact} An algebra is called semisimple if it is isomorphic to a subdirect product of simple algebras, or equivalently, if its maximal congruences intersect to zero. \begin{thm}\label{simplefact} Let $\m A$ be a finite, semisimple, solvable algebra. If the growth rate of~$\m A$ is linear, then $\m A$ has a Maltsev polynomial, and so it is affine. \end{thm} (This establishes for semisimple abelian algebras the extra implication (iv)$\Rightarrow$(i) among the conditions of Theorem~\ref{basic}.) We shall need the following corollary of~Lemma~\ref{proj}. \begin{cor}\label{tame} Let $\m A$ be a finite solvable algebra, $\eta_1,\ldots,\eta_n$ maximal congruences of~$\m A$ and $\eta=\bigwedge_{i=1}^n \eta_i$. Suppose that \begin{enumerate} \item For each $1\le i\le n$, the $\lb\eta_i,1\rangle$-minimal sets are the same (that is, $M_{\m A}(\eta_i,1)=M_{\m A}(\eta_j,1)$ for every~$i$ and~$j$). \item $A/\eta$ has no nontrivial strongly abelian quotient algebras. \end{enumerate} Then $\lb\eta,1\rangle$ is a tame quotient, and if\/ $\theta\ge\eta$ is a coatom in $\mathop{}\mathopen{}\mathrm{Con}(\m A)$, then $M_{\m A}(\theta,1)=M_{\m A}(\eta_1,1)$. \end{cor} \begin{proof} Clearly, for every upper cover $\delta$ of~$\eta$, the quotient $\lb\eta,\delta\rangle$ is perspective to some $\lb\eta_i,1\rangle$ (and so has type~$\atyp$). Condition $(2)$ implies that for every maximal congruence~$\theta\ge\eta$ the quotient $\lb\theta,1\rangle$ also has type~$\atyp$. Therefore it is projective to some $\lb\eta_i,1\rangle$ by Lemma~\ref{proj}. Thus, if $U$~is an $\lb\eta_1,1\rangle$-minimal set, then it is minimal with respect to each such quotient $\lb\eta,\delta\rangle$ and $\lb\theta,1\rangle$, and also $U\in M_{\m A}(\eta,1)$. This implies that restriction to~$U$ is a $0,1$-separating map, so by Definition~2.6 of~\cite{hobby-mckenzie}, the quotient $\lb\eta,1\rangle$ is tame. \end{proof} \begin{proof}[Proof of Theorem~\ref{simplefact}] Our assumption that the growth rate of $\m a$ is linear implies, by Corollary~\ref{abelian_cor}~(1), that \begin{equation}\label{no_strab_q} \text{no power of $\m a^n$ of has a nontrivial strongly abelian homomorphic image.} \end{equation} In particular, if $\alpha_1,\ldots,\alpha_\ell$ is the list of all maximal congruences of~$\m A$, then for each~$i$, the solvability of $\m A$ implies that $\m a/\alpha_i$ is abelian, and (\ref{no_strab_q}) implies that it is not strongly abelian. Therefore, $\langle \alpha_i,1\rangle$ has type $\atyp$ for each~$i$. It follows also that $\m A$ is abelian, since it is a subdirect product of the abelian algebras $\m a/\alpha_i$. Call two maximal congruences $\alpha_i$ and~$\alpha_j$ of $\m a$ equivalent, if the $\lb\alpha_i,1\rangle$-minimal sets and the $\lb\alpha_j,1\rangle$-minimal sets are the same. This is an equivalence relation. For each equivalence class, consider the intersection $\beta_i$ of its members. Thus we get congruences $\beta_1,\ldots,\beta_m$ of $\m a$ such that their intersection is zero. In view of (\ref{no_strab_q}), Corollary~\ref{tame} shows that $\lb\beta_i,1\rangle$ is a tame quotient for every~$i$, and the set of all coatoms above~$\beta_i$ forms an equivalence class, for every~$i$. Denote $\m A/\beta_i$ by $\m B_i$. Then $\m A$ can be viewed as a subdirect subalgebra of $\m B_1\times\dots\times\m B_m$. Fix a minimal set $U_i$ corresponding to the coatoms whose intersection is $\beta_i$. This is a minimal set for the tame quotient $\lb\beta_i,1\rangle$. Since~$\m A$ is abelian, its type~$\atyp$ minimal sets are affine and E-minimal (see the remark at the beginning of Section~\ref{abvar}). This means that if $\alpha_i$ and $\alpha_j$ are not equivalent, then $\alpha_i$ collapses every $\lb\alpha_j,1\rangle$-minimal set (by which we mean that every $\lb\alpha_j,1\rangle$-minimal set is contained in a single $\alpha_i$-class). Therefore if $i\ne j$, then $\beta_i$ collapses~$U_j$. Hence $\beta_i$ restricts trivially to~$U_i$ (that is, $\beta_i|_{U_i}=0_{U_i}$), since $\beta_1\wedge\ldots\wedge\beta_m=0$. In fact, $U_i$ is a $\lb\beta_i,1\rangle$-trace, and therefore $U_i/\beta_i$ is polynomially isomorphic to a vector space over a finite field. Since $\m A$ is abelian and satisfies (\ref{no_strab_q}), Theorem~\ref{spreadnilp} implies that $\m A$ is a spread of its type~$\atyp$ minimal sets. Applying Lemma~\ref{proj} to $\m a$ and its maximal congruences $\alpha_1,\ldots,\alpha_\ell$ we see that every type $\atyp$ quotient of $\m a$ has the same minimal sets as $\langle\alpha_i,1\rangle$ for some $\alpha_i$. In a spread, each set can be replaced with a polynomially isomorphic one, therefore we can use the representatives $U_1,\ldots,U_m$. Thus there exists a polynomial~$t$ such that \begin{equation}\label{fullA} t(U_1,\ldots,U_1,U_2,\ldots,U_2,\ldots,U_m,\ldots,U_m)=A\,. \end{equation} Write this polynomial as $t(\wec{x}^1,\wec{x}^2,\ldots,\wec{x}^m)$, where $\wec{x}^i$ is the string of variables where $U_i$ occurs. Let $0_i\in U_i$ be fixed arbitrarily, and substitute $0_i$ to every variable in $\wec{x}^i$ for~$i\ne 1$. We get a polynomial \[ f(\wec{x}^1)=f(x_1^1,\ldots,x_n^1)= t(\wec{x}^1,\hat 0_2,\ldots,\hat 0_m) \] of~$\m A$. Let $T_1=f(U_1,\ldots,U_1)$. Each $\beta_j$ collapses~$T_1$ whenever $j\ne 1$, and therefore $\beta_1$ restricts trivially to~$T_1$ (using again that $\beta_1\wedge\ldots\wedge\beta_m=0$). Therefore, as $U_1$ is a $\lb\beta_1,1\rangle$-trace, for every $a\ne b\in T_1$ there exists a unary polynomial mapping $T_1$ to~$U_1$ that separates $a$ and~$b$. Thus we can apply Lemma~3.8 of~\cite{kkv1} to $S=U_1$, zero element $0_1$ and~$f$. The proof of this lemma shows that there exist \begin{enumerate} \item unary polynomials $g_1,\ldots,g_k$ mapping $T_1$ to~$U_1$, that satisfy $g_i(0_1)=0_1$, and \item $k$-ary polynomials $\ell_1,\ldots,\ell_n$ satisfying $\ell_i(\hat 0_1)=0_1$ and $\ell_i(U_1,\ldots,U_1)\subseteq U_1$ \end{enumerate} such that for $f'(\wec{y})= f\big(\ell_1(\wec{y}),\ldots,\ell_n(\wec{y})\big)$ we have \[ g_i\big(f'(y_1,\ldots,y_k)\big) = y_i \] when each $y_1,\ldots,y_k\in U_1$ and \begin{equation}\label{coord} f'\big(g_1(x),\ldots,g_k(x)\big) = x \end{equation} for every~$x\in T_1$. Thus $f'(U_1,\ldots,U_1)=T_1$. Note that $f'(\hat 0_1)=f(\hat 0_1)$ holds. We defined polynomials $g_i=g_i^1$ and $\ell_i=\ell_i^1$ above for the first block of variables $\wec{x}^1$ of the polynomial $t(\wec{x}^1,\ldots,\wec{x}^m)$. Do this in an analogous way for all other blocks of variables $\wec{x}^j$ to obtain unary polynomials~$g_i^j$ ($1\le i\le k_j$) and $k_j$-ary polynomials~$\ell_i^j$ ($1\le i\le n_j$), and the set~$T_j$. We shall substitute these polynomials into $t$ in the following way. Consider the $j$-th block $\wec{x}^j=(x_1^j,\ldots,x_{n_j}^j)$ of the variables of $t$. Then make the substitution \[ x_i^j \to \ell_i^j\big(g_1^j(y_j),\ldots,g_{k_j}^j(y_j)\big) \] for $1\le i\le n_j$ and $1\le j\le m$. This way, we get a polynomial $r(y_1,\ldots,y_m)$. Equation~(\ref{coord}) and its analogues ensure that \[ r(0_1,\ldots,0_{j-1},c,0_{j+1},\ldots,0_m)=c \] holds whenever $c\in T_j$. Thus if $c_j\in T_j$, then $r(c_1,\ldots,c_m)\equiv_{\beta_j}c_j$. Let $a\in A$. By equation~(\ref{fullA}) we have $a = t(\wec{a}^1,\ldots,\wec{a}^m)$ for appropriate vectors $\wec{a}^j$ such that each component of $\wec{a}^j$ is in~$U_j$. Replace each $\wec{a}^j$ by $\hat 0_j$ for $j\ge 2$, and call the resulting element~$c_1\in T_1$. Since each $U_j$ is contained in a $\beta_i$-class for every~$i\ne j$ we see that $c_1\equiv_{\beta_1} a$. Do the analogous substitutions for all other variables to obtain elements $c_i\in T_i$ such that $a\equiv_{\beta_i} c_i$. Then $r(c_1,\ldots,c_m)\equiv_{\beta_i}a$ for every~$i$, and since the $\beta_i$-s intersect to zero we have that $r(c_1,\ldots,c_m)=a$. We show that if $c_j'\in T_j$ are such that $r(c_1',\ldots,c_m')=a$, then $c_j=c_j'$ for every~$j$. Indeed, $c_j \equiv_{\beta_j} a=r(c_1',\ldots,c_m')\equiv_{\beta_j}c_j'$, and $\beta_j$ restricts trivially to~$T_j$. In other words, $r:T_1\times\dots\times T_m\to A$ is a bijection. Corollary~3.7 of~\cite{kkv1} states that $T_j$ is the range of an idempotent polynomial~$e_j$, and the induced algebra on $T_j$ is isomorphic to an algebra polynomially equivalent to a full matrix power of $\m A|_{U_j}$. Therefore $T_j$ has an induced Maltsev polynomial $d_j$. Let \[ d(x,y,z)=r(\ldots,d_j\big(e_j(x),e_j(y),e_j(z)\big),\ldots)\,. \] We prove that $d(x,x,z)=r(\ldots,e_j(z),\ldots)=z$. Let $z=r(z_1,\ldots,z_m)$, where $z_j\in T_j$. Then $z\equiv_{\beta_j} z_j$, so $e_j(z)\equiv_{\beta_j} e_j(z_j)=z_j$, and so $e_j(z)=z_j$, since $\beta_j$ restricts trivially to~$T_j$. Similarly, $d(x,z,z)=x$. Hence $d$ is a Maltsev polynomial of~$\m A$ and the proof of~ Theorem~\ref{simplefact} is complete. \end{proof} We actually proved that $\m A$ is the full direct product of the algebras~$\m B_i$, since $T_i$ and $B_i$ are in a bijective correspondence via factoring modulo~$\beta_i$. The following example shows that the direct product of two affine algebras need not have a Maltsev polynomial, even if its growth rate is linear, and so Theorem~\ref{simplefact} cannot be improved to say that if $\m A/\alpha_1$ and $\m A/\alpha_2$ have Maltsev polynomials, then so does $\m A/(\alpha_1\wedge \alpha_2)$. \begin{exmp}\label{abelian_spread_not_maltsev} Consider $V=\BZ_2^2$ as a vector space over the two-element field $\BZ_2^2$. Let \[ P= \begin{pmatrix} 1&0\\ 0&0 \end{pmatrix} \qquad\text{and}\qquad N= \begin{pmatrix} 0&0\\ 1&0 \end{pmatrix}\,. \] Define an operation $f$ on $V$ by \[ f(\wec{x},\wec{y})= P\wec{x}+N\wec{y}. \] Clearly, $P^2=P$, $PN=N^2=0$ and $NP=N$. Therefore $\{0,P,N\}$ is a semigroup under matrix multiplication, and the clone of all term operations of the algebra $\langle V;+,f,\wec{0}\rangle$ consists of all functions of the form \[ M_1\wec{x}_1+\ldots+M_n\wec{x}_n\,, \] where each $M_i$ is either $P$, $N$, or the identity matrix. If one deletes $+$ from the basic operations, then the identity matrix cannot occur as a coefficient (except for projections), and there can be at most one instance of $P$ and at most one instance of~$N$. It is left to the reader to check that $f(V,V)=V$. Next we define two algebras $\m B$ and $\m C$. The language contains the operation symbols $+$, $\oplus$, a $4$-ary $g$, and $0$. In the algebra~$\m B$ the underlying set is~$V$, $+$ interprets as the usual addition, $\oplus$ as the constant zero function, and $0$ as the zero vector~$\wec{0}$. In the algebra $\m C$, the only difference is that $+$ is interpreted as constant zero, and $\oplus$ as the usual addition of~$V$. Finally, \begin{align*} g^{\m B}(\wec{x},\wec{y},\wec{u},\wec{v}) &= P\wec{x}+N\wec{y}=f(\wec{x},\wec{y}),\\ g^{\m C}(\wec{x},\wec{y},\wec{u},\wec{v}) &= P\wec{u}\oplus N\wec{v}=f(\wec{u},\wec{v})\,. \end{align*} Both algebras $\m b$ and $\m c$ have Maltsev terms, namely $x-y+z$ for $\m b$ and $x\ominus y\oplus z$ for $\m c$. In addition, both algebras are abelian, hence affine. Now consider the algebra $\m A=\m B\times \m C$. It follows that $\m a$ is also abelian. Our goal is to show that $\m a$ has linear growth rate, but has no Maltsev polynomial. To see that the growth rate of $\m a$ is linear, observe first that $e_B(x)=x+(\wec{0},\wec{0})$ is an idempotent unary polynomial of $\m A$ mapping $A=B\times C$ to the subset $U=B\times \{\wec{0}\}$. Similarly, $e_C(x)=x\oplus (\wec{0},\wec{0})$ is an idempotent unary polynomial of $\m A$ mapping $A$ to the subset $W=\{\wec{0}\}\times C$. The induced algebras $\m A|_U$ and $\m A|_W$ are abelian and have Maltsev polynomials, hence they are affine algebras. Moreover, \[ g^{\m A}(U,U,W,W)=A. \] Thus $\m a$ has linear growth rate by Theorem~\ref{affine_prep}, Lemma~\ref{spreadaffrate} and Corollary~\ref{abelian_cor}~(2). It remains to prove that $\m A$ does not have a Maltsev polynomial. Suppose that $\m a$ has a Maltsev polynomial. Since $\m a$ is abelian, we get that $\m a$ is affine, and hence it has a Maltsev term $M(x,y,z)$. Since $\m b$, $\m c$ are affine, each one has a unique Maltsev term, so $M^{\m b}(x,y,z)=x-y+z$ and $M^{\m c}(x,y,z)=x\ominus y\oplus z$. Thus $M^{\m a}$ acts componentwise as $x-y+z$ in the $\m b$-coordinate and as $x\ominus y\oplus z$ in the $\m c$-coordinate. In particular, if $\wec{v}=(0,1)^T\in V$, then for the elements $(\wec{v},\wec{0}),(\wec{0},\wec{0}),(\wec{0},\wec{v})$ of $A$ we get that $M^{\m a}\bigl((\wec{v},\wec{0}),(\wec{0},\wec{0}),(\wec{0},\wec{v})\bigr) =(\wec{v},\wec{v})$. This implies that $\{(\wec{v},\wec{0}),(\wec{0},\wec{0}),(\wec{0},\wec{v})\}$ is not a subuniverse of $\m a$. On the other hand, it is not hard to check that the set $\{(\wec{v},\wec{0}),(\wec{0},\wec{0}),(\wec{0},\wec{v})\}$ is closed under all operations $+$, $\oplus$, $g$, $0$ of $\m a$. Indeed, closure under $+$ and $\oplus$ follows, because $+$ and $\oplus$ are constant $\wec{0}$ in one of the components and the usual addition of $V$ in the other, and $\{\wec{v},\wec{0}\}$ is closed under addition. Finally, closure under $g$ follows from the fact that we have $f(\wec{v},\wec{v})=f(\wec{v},\wec{0})= f(\wec{0},\wec{v})=\wec{0}$, since $P\wec{v}=N\wec{v}=\wec{0}$. This shows that $\{(\wec{v},\wec{0}),(\wec{0},\wec{0}),(\wec{0},\wec{v})\}$ is a subuniverse of $\m a$. The contradiction obtained proves that $\m a$ has no Maltsev polynomial. \end{exmp} In this example, $\m A$ is abelian, $d_{\m A}(n)\in O(n)$, and yet $\m A$ is not affine. This proves that, for abelian algebras, no one of the equivalent conditions (iii), (iv), (v) or (vi) implies any one of the equivalent conditions (i) or (ii). \section{Summary of results. Problems.}\label{summary_sec} We are now in a position to prove Theorem~\ref{basic}, which summarizes our main results. \begin{proof}[Proof of Theorem~\ref{basic}] [(i)$\Rightarrow$(iv)] A Maltsev polynomial is a $3$-ary, $0$-pointed, $2$-cube term. Hence this implication follows from Corollary~\ref{pointed_polynomial_cor} which implies that, if $\m a$ is an algebra with a $0$-pointed, $k$-cube polynomial, and $\m a^k$ is finitely generated, then $d_{\m a}(n)\in O(n^{k-1})$. [(i)$\Rightarrow$(ii)] The definition of a pointed cube polynomial generalizes that of a Maltsev polynomial. [(ii)$\Rightarrow$(v)] Theorem~\ref{pointed_polynomial} shows that if $\m a$ is an algebra with a $p$-pointed $k$-cube polynomial, and $p\geq 1$, then $d_{\m a}(n)$ is bounded above by a polynomial in $n$ if $\m a^{p+k-1}$ is finitely generated. The restriction on $\m a^{p+k-1}$ is satisfied when $\m a$ is finite. This proves that (ii) implies (v) when $p\geq 1$. The case $p=0$ is handled similarly using Corollary~\ref{pointed_polynomial_cor}. [(iii)$\Rightarrow$(iv)] A binary polynomial with a unit element is a $2$-ary, $1$-pointed, $2$-cube polynomial. Corollary~\ref{pointed_polynomial_cor} implies that the growth rate of any finite algebra that has such a polynomial lies in $O(n)$. The induced algebra on an $\langle\alpha,\beta\rangle$-minimal set of type $\atyp, \btyp, \ltyp$ or $\styp$ has such a binary polynomial: take $d(x,0,y)$ with $d$ a pseudo-Maltsev polynomial and $0$ in the body if the type is $\atyp$, and take $x\wedge y$ with $\wedge$ a pseudo-meet polynomial in the other cases. Therefore, by Lemma~\ref{spreadaffrate} of this paper, $d_{\m a}(n)\in O(n)$ whenever $\m a$ is a spread of minimal sets whose types are not $\utyp$. [(iv)$\Rightarrow$(v)] $O(n)\cap 2^{\Omega(n)}=\emptyset$. [(v)$\Rightarrow$(vi)] Theorem~\ref{first_bounds} implies that $d_{\m a}(n)\notin 2^{\Theta(n)}$ is equivalent to $d_{\m a}(n)\notin 2^{\Omega(n)}$ for finite algebras. Hence (v)$\Rightarrow$(vi) is just the contrapositive of Corollary~\ref{abelian_cor}~(1). [(vi)$\not\Rightarrow$(v)] Example~5.3.5 of \cite{kksz-A} describes finite implication algebras with exponential growth. These satisfy (vi), since no nontrivial implication algebra is strongly abelian, but do not satisfy (v). [(ii)$\not\Rightarrow$(iv), (v)$\not\Rightarrow$(iv)] Given $k\geq 2$, Theorem~5.3.1 of \cite{kksz-A} constructs a finite algebra with a cube polynomial whose $d$-function satisfies $d_{\m a}(n)\in \Theta(n^{k-1})$. When $k=3$ one has $d_{\m a}(n)\notin 2^{\Omega(n)}$, yet $d_{\m a}(n)\notin O(n)$. [(i)$\not\Rightarrow$(iii), (iv)$\not\Rightarrow$(iii)] If $\m a$ is a 2-element Boolean algebra, then $d_{\m a}(n)\in O(\log(n))$ (hence $d_{\m a}(n)\in O(n)$), and therefore (i) and (iv) hold. But (iii) does not hold, since $\m a$ has no type $\atyp$ minimal sets. [(iii)$\not\Rightarrow$(ii)] Example~\ref{abelian_spread_not_maltsev} describes an abelian algebra $\m c$ that is a spread of type $\atyp$ minimal sets, but does not have a Maltsev term. If $\m c$ had a pointed cube polynomial, then by Theorem~\ref{nilpotent} it would have Maltsev polynomial. But it is well known that an abelian algebra with a Maltsev polynomial has a Maltsev term, and $\m c$ does not have such a term. [(iv)$\not\Rightarrow$(ii)] According to Theorem~\ref{avoid}, any function that can be realized as the growth rate of a finite algebra can also be realized as the growth rate of a finite algebra that does not have a pointed cube polynomial. \end{proof} Recall the six growth-restricting conditions portrayed in this theorem: \begin{enumerate} \item[(i)] $\m a$ has a Maltsev polynomial. \item[(ii)] $\m a$ has a pointed cube polynomial. \item[(iii)] $\m A$ is a spread of its type~$\atyp$ minimal sets. \item[(iv)] $d_{\m a}(n)\in O(n)$. \item[(v)] $d_{\m a}(n)\notin 2^{\Omega(n)}$. \item[(vi)] No finite power $\m a^n$ has a nontrivial strongly abelian homomorphic image. \end{enumerate} We have shown that for arbitrary finite algebras, the following implications hold: \begin{center} \begin{picture}(200,70) \setlength{\unitlength}{1mm} \put(0,0){$(iii)$} \put(10,0){$\Longrightarrow$} \put(20,0){$(iv)$} \put(30,0){$\Longrightarrow$} \put(40.5,0){$(v)$} \put(50,0){$\Longrightarrow$} \put(60,0){$(vi)$.} \put(22,10){\rotatebox[origin=c]{270}{$\Longrightarrow$}} \put(21,20){$(i)$} \put(30,20){$\Longrightarrow$} \put(40,20){$(ii)$} \put(42,10){\rotatebox[origin=c]{270}{$\Longrightarrow$}} \end{picture} \end{center} If $\m a$ is a finite solvable algebra, then this can be strengthened to \begin{center} \begin{picture}(200,70) \setlength{\unitlength}{1mm} \put(0,0){$(iii)$} \put(10,0){$\Longrightarrow$} \put(20,0){$(iv)$} \put(30,0){$\Longrightarrow$} \put(40.5,0){$(v)$} \put(50,0){$\Longrightarrow$} \put(60,0){$(vi)$.} \put(10,10){\rotatebox[origin=c]{225}{$\Longrightarrow$}} \put(21,20){$(i)$} \put(30,20){$\Longrightarrow$} \put(40,20){$(ii)$} \put(30,10){\rotatebox[origin=c]{225}{$\Longrightarrow$}} \end{picture} \end{center} If $\m a$ is a finite left nilpotent algebra, then we have established that \begin{center} \begin{picture}(200,70) \setlength{\unitlength}{1mm} \put(0,0){$(iii)$} \put(10,0){$\Longleftrightarrow$} \put(20,0){$(iv)$} \put(30,0){$\Longleftrightarrow$} \put(40.5,0){$(v)$} \put(50,0){$\Longleftrightarrow$} \put(60,0){$(vi)$.} \put(10,10){\rotatebox[origin=c]{225}{$\Longrightarrow$}} \put(21,20){$(i)$} \put(30,20){$\Longleftrightarrow$} \put(40,20){$(ii)$} \put(30,10){\rotatebox[origin=c]{225}{$\Longrightarrow$}} \end{picture} \end{center} Finally, if $\m a$ is a semisimple abelian algebra or if it generates an abelian variety, then all six conditions are equivalent. On the other hand, we gave an example of an abelian algebra satisfying the properties in the bottom row but not satisfying those in the top row, so no other implications hold for finite abelian or nilpotent algebras. Now let us return to the ``solvability'' diagram. The example preceding Theorem~6.10 in \cite{kkv1}, which is a finite, solvable algebra which has no Maltsev polynomial, but has a binary polynomial with a unit element, shows that (ii)$\not\Rightarrow$(i) for finite solvable algebras. We have seen that no item on the bottom row implies any item on the top row for solvable algebras, so the true implications yet to be discovered can only be (ii)$\Rightarrow$(iii), or the reversal of some of the implications along the bottom row. This suggests some problems. \begin{prb} Does (ii)$\Rightarrow$(iii) hold for finite solvable algebras? \end{prb} \begin{prb} Which of the true implications (iii)$\Rightarrow$(iv)$\Rightarrow$(v)$\Rightarrow$(vi) can be reversed for finite solvable algebras? \end{prb} \bibliographystyle{plain}
\section{Introduction} One of the central issues in minimal surface theory is to understand the global behavior of the Gauss map. In the latter half of the twentieth century, Osserman \cite{Os1959, Os1964, Os1986} initiated a systematic study of the Gauss map and, in particular, proved that the image of the Gauss map of a nonflat complete minimal surface in Euclidean 3-space ${\R}^{3}$ must be dense in the unit 2-sphere ${\Si}^{2}$. Xavier \cite{Xa1981} then showed that the Gauss map can omit at most a finite number of values in ${\Si}^{2}$, and Fujimoto \cite{Fu1988} proved that the precise maximum for the number of omitted values possible is 4. Fujimoto also gave a curvature bound for a minimal surface when all of the multiple values of the Gauss map are totally ramified (\cite{Fu1992, Fu1993}). Here a value $\alpha$ of a map or function $g$ is said to be {\it totally ramified} if the equation $g=\alpha$ has no simple roots. Moreover, Fujimoto obtained a unicity theorem for the Gauss maps of nonflat complete minimal surfaces, which is analogous to the Nevanlinna unicity theorem (\cite{Ne1926}) for meromorphic functions on the complex plane $\C$ (\cite{Fu1993-2}). There exist several classes of immersed surfaces whose Gauss maps have these function-theoretical properties. For instance, Yu \cite{Yu1997} showed that the hyperbolic Gauss map of a nonflat complete constant mean curvature one surface in hyperbolic 3-space ${\H}^{3}$ can omit at most 4 values. The author and Nakajo \cite{KN2012} obtained that the maximal number of omitted values of the Lagrangian Gauss map of a weakly complete improper affine front in the affine 3-space ${\R}^{3}$ is 3, unless it is an elliptic paraboloid. As an application of this result, a simple proof of the parametric affine Bernstein theorem for an improper affine sphere in ${\R}^{3}$ was provided. Moreover the author \cite{Ka2013-2} gave similar results for flat fronts in ${\H}^{3}$. In \cite{Ka2013}, we revealed a geometric meaning for the maximal number of omitted values of their Gauss maps. To be precise, we gave a curvature bound for the conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ on an open Riemann surface $\Sigma$, where $\omega$ is a holomorphic 1-form and $g$ is a meromorphic function on $\Sigma$ (\cite[Theorem 2.1]{Ka2013}) and, as a corollary of the theorem, proved that the precise maximal number of omitted values of the nonconstant meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}$ is $m+2$ (\cite[Corollary 2.2, Proposition 2.4]{Ka2013}). Since the induced metric from ${\R}^{3}$ of a complete minimal surface is $ds^{2}=(1+|g|^{2})^{2}|\omega|^{2}$ (i.e. $m=2$), the maximum number of omitted values of the Gauss map $g$ of a nonflat complete minimal surface in ${\R}^{3}$ is $4\,(=2+2)$. On the other hand, for the Lagrangian Gauss map $\nu$ of a weakly complete improper affine front, because $\nu$ is meromorphic, $dG$ is holomorphic and the complete metric is $d{\tau}^{2}=(1+|\nu|^{2})|dG|^{2}$ (i.e. $m=1$), the maximal number of omitted values of the Lagrangian Gauss map of a weakly complete improper affine front in ${\R}^{3}$ is $3\,(=1+2)$, unless it is an elliptic paraboloid. The goal of this paper is to elucidate the geometric background of function-theoretic properties for the Gauss maps. The paper is organized as follows: In Section 2, we first give a curvature bound for the conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ on an open Riemann surface $\Sigma$ when all of the multiple values of the meromorphic function $g$ are totally ramified (Theorem \ref{thm-ramification}). This is a generalization of Theorem 2.1 in \cite{Ka2013}, and the proof is given in Section 3.1. As a corollary of this theorem, we give a ramification theorem for the meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}$ (Corollary \ref{cor-ramification}). We remark that this corresponds to the defect relation in Nevanlinna theory (see \cite{Ko2003}, \cite{NO1990}, \cite{NW2013}, \cite{Ru2001} for the details). Next, we provide two applications of the result. The first one is to show that the precise maximal number of omitted values of the nonconstant meromorphic function $g$ on $\Sigma$ with complete conformal metric $ds^{2}$ is $m+2$ (Corollary \ref{cor-exceptional}). The second one is to prove an analogue of a special case of the Ahlfors islands theorem \cite[Theorem B.2]{Be2000} for $g$ on $\Sigma$ with the complete conformal metric $ds^{2}$ (Corollaries \ref{cor-covering} and \ref{cor-Ahlfors-island}). The Ahlfors islands theorem has found various applications in complex dynamics; see \cite{Be2000} for an exposition. We also give a unicity theorem for the nonconstant meromorphic function $g$ on an open Riemann surface $\Sigma$ with the complete conformal metric $ds^{2}$ (Theorem \ref{thm-unicity}). This theorem is optimal in that for every even number $m$, there exist examples (Example \ref{exa-unicity}). The proof is given in Section 3.2. When $m=0$, all results coincide with the results for meromorphic functions on $\C$ (Remarks \ref{rem-geometry}, \ref{rem-Ahlfors-island} and \ref{rem-unicity-thm}). In Section 4, as applications of the main results, we show some function-theoretic properties for the Gauss maps of the following classes of surfaces: minimal surfaces in ${\R}^{3}$ (Section 4.1), constant mean curvature one surfaces in ${\H}^{3}$ (Section 4.2), maxfaces in ${\R}^{3}_{1}$ (Section 4.3), improper affine fronts in ${\R}^{3}$ (Section 4.4) and flat fronts in ${\H}^{3}$ (Section 4.5). In particular, we give their geometric background. Finally, the author would like to thank Professors Junjiro Noguchi, Wayne Rossman, Masaaki Umehara and Kotaro Yamada for their useful advice. The author also would like to express his thanks to Professors Ryoichi Kobayashi, Masatoshi Kokubu, Miyuki Koiso and Reiko Miyaoka for their constant encouragement. \section{Main results} \subsection{Curvature bound and its corollaries} We first give the following curvature bound for the conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ on an open Riemann surface $\Sigma$. This is more precise than Theorem 2.1 in \cite{Ka2013}. \begin{theorem}\label{thm-ramification} Let $\Sigma$ be an open Riemann surface with the conformal metric \begin{equation}\label{equ-metric} ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}, \end{equation} where $\omega$ is a holomorphic $1$-form, $g$ is a meromorphic function on $\Sigma$, and $m\in \N$. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that \begin{equation}\label{equ-ramification} \gamma= \displaystyle \sum_{j=1}^{q} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}> m+2. \end{equation} If $g$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on $m$, $\gamma$ and ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ we have \begin{equation}\label{equ-curvature} |K_{ds^{2}}(p)|^{1/2}\leq \dfrac{C}{d(p)}, \end{equation} where $K_{ds^{2}}(p)$ is the Gaussian curvature of the metric $ds^{2}$ at $p$ and $d(p)$ is the geodesic distance from $p$ to the boundary of $\Sigma$, that is, the infimum of the lengths of the divergent curves in $\Sigma$ emanating from $p$. \end{theorem} As a corollary of Theorem \ref{thm-ramification}, we give the following ramification theorem for the meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$. \begin{corollary}\label{cor-ramification} Let $\Sigma$ be an open Riemann surface with the conformal metric given by (\ref{equ-metric}). Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the metric $ds^{2}$ is complete and that the inequality (\ref{equ-ramification}) holds. If $g$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then $g$ must be constant. \end{corollary} \begin{proof} Since $ds^{2}$ is complete, we may set $d(p)=\infty$ for all $p\in \Sigma$. By virtue of Theorem \ref{thm-ramification}, $K_{ds^{2}}\equiv 0$ on $\Sigma$. On the other hand, the Gaussian curvature of the metric $ds^{2}$ is given by \begin{equation}\label{equ-Gaussian} K_{ds^{2}}=-\dfrac{2m|g'_{z}|^{2}}{(1+|g|^{2})^{m+2}|\hat{\omega}_{z}|^{2}}, \end{equation} where $\omega= \hat{\omega}_{z}dz$ and $g'_{z}=dg/dz$. Hence $K_{ds^{2}}\equiv 0$ if and only if $g$ is constant. \end{proof} \begin{remark}\label{rem-geometry} The geometric meaning of the ``2'' in ``$m+2$'' is the Euler number of the Riemann sphere. Indeed, if $m=0$ then the metric $ds^{2}=(1+|g|^{2})^{0}|\omega|^{2}=|\omega|^{2}$ is flat and complete on $\Sigma$. We thus may assume that $g$ is a meromorphic function on $\C$ because $g$ is replaced by $g\circ \pi$, where $\pi\colon \C\to \Sigma$ is a holomorphic universal covering map. On the other hand, Ahlfors \cite{Ah1935} and Chern \cite{Ch1960} showed that the least upper bound for the defect relation for meromorphic functions on $\C$ coincides with the Euler number of the Riemann sphere. Hence we get the conclusion. \end{remark} We next give two applications of Corollary \ref{cor-ramification}. The first one is to provide the precise maximal number of omitted values of the meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$. \begin{corollary}[{\cite[Corollary 2.2]{Ka2013}}]\label{cor-exceptional} Let $\Sigma$ be an open Riemann surface with the conformal metric given by (\ref{equ-metric}). If the metric $ds^{2}$ is complete and the meromorphic function $g$ is nonconstant, then $g$ can omit at most $m+2$ distinct values. \end{corollary} \begin{proof} By way of contradiction, assume that $g$ omits $m+3$ distinct values. In Corollary \ref{cor-ramification}, if $g$ does not take a value ${\alpha}_{j}$ $(j=1, \ldots, q)$, we may set ${\nu}_{j}=\infty$ in (\ref{equ-ramification}). Thus we can consider the case where $\gamma \geq m+3\, (>m+2)$. By virtue of Corollary \ref{cor-ramification}, the function $g$ is constant. This contradicts the assumption that $g$ is nonconstant. \end{proof} The number ``$m+2$'' is sharp because there exist examples in \cite[Proposition 2.4]{Ka2013}. The second one is to show an analogue of the Ahlfors islands theorem \cite[Theorem B.2]{Be2000} for the meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$. We first recall the notion of chordal distance between two distinct values in the Riemann sphere $\C\cup \{\infty \}$. For two distinct values $\alpha$, $\beta\in \C\cup \{\infty\}$, we set $$ |\alpha, \beta|:= \dfrac{|\alpha -\beta|}{\sqrt{1+|\alpha|^{2}}\sqrt{1+|\beta|^{2}}} $$ if $\alpha \not= \infty$ and $\beta \not= \infty$, and $|\alpha, \infty|=|\infty, \alpha| := 1/\sqrt{1+|\alpha|^{2}}$. We remark that, if we take $v_{1}$, $v_{2}\in {\Si}^{2}$ with $\alpha =\varpi (v_{1})$ and $\beta = \varpi (v_{2})$, we have that $|\alpha, \beta|$ is a half of the chordal distance between $v_{1}$ and $v_{2}$, where $\varpi$ denotes the stereographic projection of ${\Si}^{2}$ onto $\C\cup \{\infty \}$. We next explain the definition of an island of a meromorphic function on a Riemann surface. \begin{definition}\label{def-island} Let $\Sigma$ be a Riemann surface and $g\colon \Sigma\to \C\cup\{\infty\}$ a meromorphic function. Let $V\subset \C\cup\{\infty\}$ be a Jordan domain. A simply-connected component $U$ of $g^{-1}(V)$ with $\overline{U}\subset \Sigma$ is called an {\it island} of $g$ over $V$. Note that $g|_{U}\colon U\to V$ is a proper map. The degree of this map is called the {\it multiplicity} of the island $U$. An island of multiplicity one is called a {\it simple island}. \end{definition} When all islands of the meromorphic function $g$ with the complete conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ are small disks, we get the following result by applying Corollary \ref{cor-ramification}. \begin{corollary}\label{cor-covering} Let $\Sigma$ be an open Riemann surface with the conformal metric given by (\ref{equ-metric}). Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the metric $ds^{2}$ is complete and that the inequality (\ref{equ-ramification}) holds. Then there exists $\varepsilon > 0$ such that, if $g$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $g$ must be constant. \end{corollary} \begin{proof} If such an $\varepsilon$ does not exist, for any $\varepsilon$ we can find a nonconstant meromorphic function $g$ which has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$. However this implies that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, contradicting Corollary \ref{cor-ramification}. \end{proof} The important case of Corollary \ref{cor-covering} is the case where $q=2m+5$ and ${\nu}_{j}=2$ for each $j$ $(j=1, \ldots, q)$. This corresponds to the so-called five islands theorem in the Ahlfors theory of covering surfaces (\cite{Ah1935}, \cite[Chapter XIII]{Ne1970}). \begin{corollary}\label{cor-Ahlfors-island} Let $\Sigma$ be an open Riemann surface with the complete conformal metric given by (\ref{equ-metric}). Let ${\alpha}_{1}, \ldots, {\alpha}_{2m+5} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 2m+5)$. Then there exists $\varepsilon > 0$ such that, if $g$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$, then $g$ must be constant. \end{corollary} \begin{remark}\label{rem-Ahlfors-island} Theorem \ref{cor-Ahlfors-island} is valid for the case where $m=0$. In fact, by the same argument in Remark \ref{rem-geometry}, we can easily show that the theorem corresponds to a special case of the Ahlfors five islands theorem. \end{remark} \subsection{Unicity theorem} We give another type of function-theoretic property of the meromorphic function $g$ on $\Sigma$ with the complete conformal metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$. In \cite{Ne1926}, Nevanlinna showed that two nonconstant meromorphic functions on $\C$ coincides with each other if they have the same inverse images for five distinct values. We get the following analogue to this unicity theorem. \begin{theorem}\label{thm-unicity} Let $\Sigma$ be an open Riemann surface with the conformal metric \begin{equation}\label{equ-metric1} ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}, \end{equation} and $\widehat{\Sigma}$ be another open Riemann surface with the conformal metric \begin{equation}\label{equ-metric2} d{\hat{s}}^{2}=(1+|\hat{g}|^{2})^{m}|\hat{\omega}|^{2}, \end{equation} where $\omega$ and $\hat{\omega}$ are holomorphic $1$-forms, $g$ and $\hat{g}$ are nonconstant meromorphic functions on $\Sigma$ and $\widehat{\Sigma}$ respectively, and $m\in \N$. We assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Suppose that there exist $q$ distinct points ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ such that $g^{-1}({\alpha}_{j})=(\hat{g}\circ \Psi)^{-1}({\alpha}_{j})$ $(1\leq j\leq q)$. If $q \geq m+5 \,(=(m+4)+1)$ and either $ds^{2}$ or $d{\hat{s}}^{2}$ is complete, then $g\equiv \hat{g}\circ \Psi$. \end{theorem} \begin{remark}\label{rem-unicity-thm} When $m=0$, Theorem \ref{thm-unicity} coincides with the Nevanlinna unicity theorem. \end{remark} The maps $g$ and $\hat{g}\circ \Psi$ are said to share the value $\alpha$ (ignoring multiplicity) when $g^{-1}(\alpha)= (\hat{g}\circ \Psi)^{-1}(\alpha)$. Theorem \ref{thm-unicity} is optimal for an arbitrary even number $m\, (\geq 2)$ because there exist the following examples. \begin{example}\label{exa-unicity} For an arbitrary even number $m\, (\geq 2)$, we take $m/2$ distinct points ${\alpha}_{1}, \ldots, {\alpha}_{m/2}$ in $\C\backslash \{0, \pm 1 \}$. Let $\Sigma$ be either the complex plane punctured at $m+1$ distinct points $0$, ${\alpha}_{1}, \ldots, {\alpha}_{m/2}$, $1/{\alpha}_{1}, \ldots, 1/{\alpha}_{m/2}$ or the universal covering of that punctured plane. We set $$ \omega = \dfrac{dz}{z\prod_{i=1}^{m/2} (z-{\alpha}_{i})({\alpha}_{i}z-1)}, \quad g(z)=z\,, $$ and $$ \hat{\omega}\, (=\omega) = \dfrac{dz}{z\prod_{i=1}^{m/2} (z-{\alpha}_{i})({\alpha}_{i}z-1)}, \quad \hat{g}(z)=\dfrac{1}{z}. $$ We can easily show that the identity map $\Psi\colon \Sigma \to \Sigma$ is a conformal diffeomorphism and the metrics $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ and $d\hat{s}^{2}=(1+|\hat{g}|^{2})^{m}|\hat{\omega}|^{2}$ are complete. Then the maps $g$ and $\hat{g}$ share the $m+4$ distinct values $0,\, \infty,\, 1,\, -1,\, {\alpha}_{1},\,\ldots, {\alpha}_{m/2},\, 1/{\alpha}_{1},\, \ldots,\, 1/{\alpha}_{m/2}$ and $g\not\equiv \hat{g}\circ \Psi$. These show that the number $m+5$ in Theorem \ref{thm-unicity} cannot be replaced by $m+4$. \end{example} \section{Proof of main theorems} \subsection{Proof of Theorem \ref{thm-ramification}} Before proceeding to the proof of Theorem \ref{thm-ramification}, we recall two lemmas. \begin{lemma}[{\cite[Corollary 1.4.15]{Fu1993}}]\label{main-lem1} Let $g$ be a nonconstant meromorphic function on ${\triangle}_{R}=\{z\in \C; |z|< R\}$ $(0<R\leq \infty)$. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup\{\infty\}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that $$ \gamma =\displaystyle \sum_{j=1}^{q} \biggl(1-\dfrac{1}{{\nu}_{j}} \biggr)> 2. $$ If $g$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then, for arbitrary constants $\eta\geq 0$ and $\delta >0$ with $\gamma -2>\gamma(\eta +\delta)$, then there exists a positive constant $C'$, depending only on $\gamma$, $\eta$, $\delta$, and $L:={\min}_{i<j}|{\alpha}_{i}, {\alpha}_{j}|$, such that \begin{equation}\label{eq-lem-estimate} \dfrac{|g'|}{1+|g|^{2}}\dfrac{1}{({\prod}_{j=1}^{q}|g, {\alpha}_{j}|^{1-1/{\nu}_{j}})^{1-\eta-\delta}}\leq C'\dfrac{R}{R^{2}-|z|^{2}}. \end{equation} \end{lemma} \begin{lemma}[{\cite[Lemma 1.6.7]{Fu1993}}]\label{main-lem2} Let $d{\sigma}^{2}$ be a conformal flat metric on an open Riemann surface $\Sigma$. Then, for each point $p\in \Sigma$, there exists a local diffeomorphism $\Phi$ of a disk ${\Delta}_{R}=\{z\in \C\, ;\, |z|<{R}\}$ $(0<{R}\leq +\infty)$ onto an open neighborhood of $p$ with $\Phi (0)=p$ such that $\Phi$ is a local isometry, that is, the pull-back ${\Phi}^{\ast}(d{\sigma}^{2})$ is equal to the standard Euclidean metric $ds_{Euc}^{2}$ on ${\Delta}_{R}$ and, for a point $a_{0}$ with $|a_{0}|=1$, the $\Phi$-image ${\Gamma}_{a_{0}}$ of the curve $L_{a_{0}}=\{w:=a_{0}s\, ;\, 0<s<R \}$ is divergent in $\Sigma$. \end{lemma} \begin{proof}[{\it Proof of Theorem \ref{thm-ramification}}] For the proof of Theorem \ref{thm-ramification}, we may assume the following: \begin{enumerate} \item[(A)] For any proper subset $I$ in $\{1, 2, \ldots, q\}$, $$ \displaystyle \sum_{j\in I}\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}\leq m+2. $$ \item[(B)] There exists no set of positive integers $({\nu}^{\ast}_{1}, \ldots, {\nu}^{\ast}_{q})$ distinct with $({\nu}_{1}, \ldots, {\nu}_{q})$ satisfying the conditions \begin{equation}\label{equ-proof-assume} {\nu}^{\ast}_{j}\leq {\nu}_{j} \; (1\leq j\leq q), \quad \displaystyle \sum_{j=1}^{q} \biggl{(}1-\dfrac{1}{{\nu}^{\ast}_{j}} \biggr{)}> m+2. \end{equation} \end{enumerate} If there exists some proper subset $I$ in $\{1, 2, \ldots, q\}$ such that $$ \displaystyle \sum_{j\in I}\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}> m+2, $$ then the assumption in Theorem \ref{thm-ramification} for $\{{\alpha}_{j}\,;\, 1\leq j\leq q\}$ can be replace by the assumption for $\{{\alpha}_{j}\,;\, j\in I \}$. Moreover if there exist some $({\nu}^{\ast}_{1}, \ldots, {\nu}^{\ast}_{q})$ satisfying the conditions (\ref{equ-proof-assume}), then we may prove Theorem \ref{thm-ramification} after replacing each integer ${\nu}_{j}$ by ${\nu}^{\ast}_{j}$. \begin{lemma}\label{lem-finite} There exist only finite many sets of integers ${\nu}_{1}, \ldots, {\nu}_{q}$ with ${\nu}_{j}\geq 2$ which satisfy the conditions (A) and (B). \end{lemma} \begin{proof} We take positive integers ${\nu}_{1}, \ldots, {\nu}_{q}$ satisfying the conditions (A) and (B). We may assume that ${\nu}_{1}\leq \ldots \leq {\nu}_{q}$. Then, for the number $$ \displaystyle \gamma =\sum_{j=1}^{q}\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}, $$ we shall show that \begin{equation}\label{equ-proof-331} \gamma - (m+2) \leq \dfrac{1}{{\nu}_{q}({\nu}_{q}-1)}. \end{equation} In fact, we suppose that $\gamma - (m+2)> 1/{\nu}_{q}({\nu}_{q}-1)$. If ${\nu}_{q}=2$, then $$ \gamma > (m+1)+\dfrac{1}{2} $$ and $$ \displaystyle \sum_{j=1}^{q-1} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}> m+2, $$ which contradicts the assumption (A). Thus, ${\nu}_{q}\geq 3$. Here, if we set ${\nu}^{\ast}_{j}:={\nu}_{j}$ $(1\leq j\leq q-1)$ and ${\nu}^{\ast}_{q}:={\nu}_{q}-1$, then $$ \displaystyle \sum_{j=1}^{q}\biggl{(}1-\dfrac{1}{{\nu}^{\ast}_{j}} \biggr{)}= \sum_{j=1}^{q}\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}-\dfrac{1}{{\nu}_{q}({\nu}_{q}-1)}> m+2. $$ This contradicts the assumption (B). We have thus proved the inequality (\ref{equ-proof-331}). By virtue of (\ref{equ-proof-331}), $$ \gamma \leq m+2+\dfrac{1}{{\nu}_{q}({\nu}_{q}-1)}< m+2+\dfrac{1}{2}= m+\dfrac{5}{2}. $$ On the other hand, since ${\nu}_{j}\geq 2$ for all $j$, we have $$ \gamma =\displaystyle \sum_{j} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}\geq q\biggl{(}1-\dfrac{1}{{\nu}_{1}} \biggr{)}\geq \dfrac{q}{2}, $$ where $q\geq 5$. Hence we obtain that ${\nu}_{1}<2q/(2q-2m-5)$ and $q< 2m+5$. Now we consider the numbers ${\nu}_{1}, \ldots, {\nu}_{q}$ satisfying the conditions (A) and (B). We set $$ {\gamma}_{0}:= \displaystyle \sum_{j=1}^{k} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}. $$ Since $$ m+2 < \gamma = {\gamma}_{0}+ \displaystyle \sum_{j=k+1}^{q}\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}, $$ we have $$ {\gamma}_{0}+q-k-(m+2)=\gamma -(m+2)+\displaystyle \sum_{j=k+1}^{q}\dfrac{1}{{\nu}_{j}}> 0. $$ Then we take a number $N$ with ${\gamma}_{0}+q-k-(m+2)>{\eta}_{0}:=1/N(N-1)$. If ${\nu}_{q}\leq N$, then we have ${\nu}_{k+1}\leq N$. Otherwise, by the inequality (\ref{equ-proof-331}) and ${\nu}_{k}\leq {\nu}_{j}$ for $j=k+1, \ldots, q$, we have \begin{eqnarray} 0<{\gamma}_{0}+q-k-{\eta}_{0}-(m+2) &\leq& {\gamma}_{0}+q-k-(m+2)-\dfrac{1}{{\nu}_{q}({\nu}_{q}-1)} \nonumber \\ &\leq& \displaystyle \sum_{j=k+1}^{q}\dfrac{1}{{\nu}_{j}}\leq \dfrac{q-k}{{\nu}_{k+1}}. \nonumber \end{eqnarray} This gives $$ {\nu}_{k+1}\leq \dfrac{q-k}{{\gamma}_{0}-(m+2)+q-k-{\eta}_{0}}. $$ We thereby get that $$ {\nu}_{k+1}\leq \max\biggl{\{}N,\, \dfrac{q-k}{{\gamma}_{0}-(m+2)+q-k-{\eta}_{0}}\biggr{\}}. $$ Since the boundedness of ${\nu}_{1}$ has been already shown, by induction on $k$ $(=1, \ldots, q)$, we have completed the proof of the lemma. \end{proof} By Lemma \ref{lem-finite}, if we take the maximum $C_{0}$ in constants which are chosen for the finitely many possible cases of ${\nu}_{j}'s$ satisfying the conditions (A) and (B), then $C_{0}$ satisfies the desired inequality (\ref{equ-curvature}). Hence, for the proof of Theorem \ref{thm-ramification}, we shall show the existence of a constant satisfying (\ref{equ-curvature}) which may depend on the given data ${\nu}_{1}, \ldots, {\nu}_{q}$. We may assume that $m\not= 0$ and ${\alpha}_{q}=\infty$ after a suitable M\"obius transformation. We choose some $\delta$ such that $\gamma -(m+2)> 2\gamma\delta >0$ and set $m\not=0$ and \begin{equation}\label{equ-proof-211} \eta :=\dfrac{\gamma -(m+2)-2\gamma\delta}{\gamma}, \quad \lambda :=\dfrac{m}{m+\gamma\delta}. \end{equation} Then if we choose a sufficiently small positive number $\delta$ depending on $\gamma$ and $m$, for the constant ${\varepsilon}_{0}:= (\gamma -(m+2))/2m\gamma$ we have \begin{equation}\label{equ-proof-212} 0<\lambda <1, \quad \dfrac{{\varepsilon}_{0}\lambda}{1-\lambda}\biggl{(}=\dfrac{\gamma -(m+2)}{2\delta{\gamma}^{2}} \biggr{)}>1. \end{equation} Now we define a new metric \begin{equation}\label{equ-proof-213} \displaystyle d{\sigma}^{2}=|\hat{\omega}_{z}|^{2/(1-\lambda)} \biggl{(}\dfrac{1}{|g'_{z}|}\prod_{j=1}^{q-1}\biggl{(}\dfrac{|g-{\alpha}_{j}|}{\sqrt{1+|{\alpha}_{j}|^{2}}} \biggr{)}^{{\mu}_{j}(1-\eta -\delta)} \biggr{)}^{2\lambda /(1-\lambda)} |dz|^{2} \end{equation} on the set ${\Sigma}'=\{p\in \Sigma \,;\, g'_{z}\not= 0\; \text{and}\; g(z)\not= {\alpha}_{j}\; \text{for all}\; j\}$, where $\omega =\hat{\omega}_{z}dz$, $g'_{z}=dg/dz$ and ${\mu}_{j}=1-(1/{\nu}_{j})$. Take a point $p\in {\Sigma}'$. Since the metric $d{\sigma}^{2}$ is flat on ${\Sigma}'$, by Lemma \ref{main-lem2}, there exists a local isometry $\Phi$ satisfying $\Phi (0)=p$ from a disk ${\triangle}_{R}=\{ z\in \C\,;\, |z|<R \}$ $(0< R\leq +\infty)$ with the standard metric $ds^{2}_{Euc}$ onto an open neighborhood of $p$ in ${\Sigma}'$ with the metric $d{\sigma}^{2}$ such that for a point $a_{0}$ with $|a_{0}|=1$, the $\Phi$-image ${\Gamma}_{a_{0}}$ of the curve $L_{a_{0}}=\{w:=a_{0}s\,;\, 0<s<R \}$ is divergent in ${\Sigma}'$. For brevity, we denote the function $g\circ \Phi$ on ${\triangle}_{R}$ by $g$ in the following. By Lemma \ref{main-lem1}, we get that \begin{equation}\label{equ-proof-214} R\leq C'\dfrac{1+|g(0)|^{2}}{|g'_{z}(0)|} \displaystyle \prod_{j=1}^{q}|g(0), {\alpha}_{j}|^{{\mu}_{j}(1-\eta -\delta)} < +\infty. \end{equation} Hence $$ L_{d\sigma} ({\Gamma}_{a_{0}})=\int_{{\Gamma}_{a_{0}}} d\sigma =R < +\infty , $$ where $L_{d\sigma} ({\Gamma}_{a_{0}})$ denotes the length of ${\Gamma}_{a_{0}}$ with respect to the metric $d{\sigma}^{2}$. Now we prove that ${\Gamma}_{a_{0}}$ is divergent in $\Sigma$. Indeed, if not, then ${\Gamma}_{a_{0}}$ must tend to a point $p_{0}\in \Sigma\backslash {\Sigma}'$ where $g'_{z}(p_{0})= 0$ or $g(p_{0})={\alpha}_{j}$ for some $j$ because ${\Gamma}_{a_{0}}$ is divergent in ${\Sigma}'$ and $L_{d\sigma} ({\Gamma}_{a_{0}})< +\infty$. Taking a local complex coordinate $\zeta$ in a neighborhood of $p_{0}$ with $\zeta (p_{0})=0$, we can write the metric $d{\sigma}^{2}$ as $$ d{\sigma}^{2}= |\zeta|^{2k\lambda /(1-\lambda)} w|d\zeta|^{2}, $$ with some positive smooth function $w$ and some real number $k$. If $g-{\alpha}_{j}$ has a zero of order $l\,(\geq {\nu}_{j}\geq 2)$ at $p_{0}$ for some $1\leq j\leq q-1$, then $g'_{z}$ has a zero of order $l-1$ at $p_{0}$ and $\hat{\omega}_{z}(z_{0})\not= 0$. Then we obtain that \begin{eqnarray} k &=& -(l-1)+l\biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}(1-\eta -\delta) \nonumber \\ &=& \biggl{(}1-\dfrac{l}{{\nu}_{j}} \biggr{)}-\dfrac{l}{{\nu}_{j}}({\nu}_{j}-1)(\eta +\delta) \nonumber \\ &=& -(\eta +\delta)\leq -{\varepsilon}_{0}. \nonumber \end{eqnarray} For the case where $g$ has a pole of order $l\,(\geq {\nu}_{q}\geq 2)$, $g'_{z}$ has a pole of order $l+1$, $\hat{\omega}_{z}$ has a zero of order $ml$ at $p_{0}$ and each component $g-{\alpha}_{j}$ in the right side of (\ref{equ-proof-213}) has a pole of order $l$ at $p_{0}$. Using the identity ${\mu}_{1}+\cdots +{\mu}_{q-1}=\gamma -{\mu}_{q}$ and (\ref{equ-proof-211}), we get that \begin{eqnarray} k &=& \dfrac{ml}{\lambda} +(l+1)-l(\gamma -{\mu}_{q})(1-\eta -\delta) \nonumber \\ &=& l{\mu}_{q}(1-\eta -\delta) -(l-1)\leq -{\varepsilon}_{0}. \nonumber \end{eqnarray} Moreover, for the case where $g'_{z}(p_{0})= 0$ and $g(p_{0})\not= {\alpha}_{j}$ for all $j$, we see that $k\leq -1$. In any case, $k\lambda /(1-\lambda)\leq -1$ by (\ref{equ-proof-212}) and there exists a positive constant $\widetilde{C}$ such that $$ d{\sigma}\geq \widetilde{C}\dfrac{|d\zeta|}{|\zeta|} $$ in a neighborhood of $p_{0}$. Thus we obtain that $$ R=\int_{{\Gamma}_{a_{0}}} d\sigma \geq \widetilde{C}\int_{{\Gamma}_{a_{0}}} \dfrac{|d\zeta|}{|\zeta|}= +\infty, $$ which contradicts (\ref{equ-proof-214}). Since ${\Phi}^{\ast}d{\sigma}^{2}=|dz|^{2}$, we get by (\ref{equ-proof-213}) that \begin{equation}\label{equ-proof-215} |\hat{\omega}_{z}|=\displaystyle \biggl{(}|g'_{z}| \prod_{j=1}^{q-1}\biggl{(}\dfrac{\sqrt{1+|{\alpha}_{j}|^{2}}}{|g-{\alpha}_{j}|} \biggr{)}^{{\mu}_{j}(1-\eta -\delta)} \biggr{)}^{\lambda}. \end{equation} By Lemma \ref{main-lem1}, we obtain that \begin{eqnarray} {\Phi}^{\ast}ds &=& (1+|g|^{2})^{m/2}|\omega| \nonumber \\ &=& \biggl{(}|g'_{z}|(1+|g|^{2})^{m/2\lambda}\displaystyle \prod_{j=1}^{q-1}\biggl{(}\dfrac{\sqrt{1+|{\alpha}_{j}|^{2}}}{|g-{\alpha}_{j}|} \biggr{)}^{{\mu}_{j}(1-\eta -\delta)} \biggr{)}^{\lambda} |dz|\nonumber \\ &=& \biggl{(}\dfrac{|g'_{z}|}{1+|g|^{2}}\dfrac{1}{\prod_{j=1}^{q} |g, {\alpha}_{j}|^{\mu_{j}(1-\eta -\delta)}} \biggr{)}^{\lambda} |dz|\nonumber \\ &\leq &(C')^{\lambda}\biggl{(}\dfrac{R}{R^{2}-|z|^{2}} \biggr{)}^{\lambda}|dz|. \nonumber \end{eqnarray} Thus we have $$ d(p)\leq \int_{{\Gamma}_{a_{0}}} ds = \int_{L_{a_{0}}}{\Phi}^{\ast}ds \leq (C')^{\lambda} \int_{L_{a_{0}}} \biggl{(}\dfrac{R}{R^{2}-|z|^{2}} \biggr{)}^{\lambda}|dz|\leq (C')^{\lambda}\dfrac{R^{1-\lambda}}{1-\lambda}\,(<+\infty) $$ because $0<\lambda <1$. Moreover, by (\ref{equ-proof-214}), we get that $$ d(p)\leq \dfrac{(C')^{\lambda}}{1-\lambda}\biggl{(}\dfrac{1+|g(0)|^{2}}{|g'_{z}(0)|}\displaystyle \prod_{j=1}^{q} |g(0), {\alpha}_{j}|^{{\mu}_{j}(1-\eta -\delta)} \biggr{)}^{1-\lambda}. $$ On the other hand, the Gaussian curvature $K_{ds^{2}}$ of the metric $ds^{2}=(1+|g|^{2})^{m}|\omega|^{2}$ is given by $$ K_{ds^{2}}=-\dfrac{2m|g'_{z}|^{2}}{(1+|g|^{2})^{m+2}|\hat{\omega}_{z}|^{2}}. $$ Thus, by (\ref{equ-proof-215}), we also get that $$ |K_{ds^{2}}|^{1/2}=\sqrt{2m}\,\biggl{(}\dfrac{|g'_{z}|}{1+|g|^{2}} \biggr{)}^{1-\lambda}\biggl{(}\displaystyle \prod_{j=1}^{q}|g, {\alpha}_{j}|^{{\mu}_{j}(1-\eta -\delta)} \biggr{)}^{\lambda}. $$ Since $|g, {\alpha}_{j}|\leq 1$ for each $j$, we obtain that $$ |K_{ds^{2}}(p)|^{1/2}d(p)\leq \dfrac{\sqrt{2m}C'}{1-\lambda}=:C. $$ Hence we get the conclusion. \end{proof} \subsection{Proof of Theorem \ref{thm-unicity}} We review the following two lemmas used in the proof of Theorem \ref{thm-unicity}. \begin{lemma}[{\cite[Proposition 2.1]{Fu1993-2}}]\label{main-lem3} Let $g$ and $\hat{g}$ be mutually distinct nonconstant meromorphic function on a Riemann surface $\Sigma$. Let $q\in \N$ and ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C \cup \{\infty\}$ be distinct. Suppose that $q>4$ and $g^{-1}({\alpha}_{j})={\hat{g}}^{-1}({\alpha}_{j})$ $(1\leq j\leq q)$. For $b_{0}>0$ and $\varepsilon$ with $q-4> q\varepsilon >0$, we set that $$ \xi :=\displaystyle \biggl{(}\prod_{j=1}^{q} |g, {\alpha}_{j}| \log{\biggl{(}\dfrac{b_{0}}{|g, {\alpha}_{j}|^{2}} \biggr{)}}\biggr{)}^{-1+\varepsilon}, \quad \hat{\xi} := \displaystyle \biggl{(}\prod_{j=1}^{q} |\hat{g}, {\alpha}_{j}| \log{\biggl{(}\dfrac{b_{0}}{|\hat{g}, {\alpha}_{j}|^{2}} \biggr{)}}\biggr{)}^{-1+\varepsilon}, $$ and define that \begin{equation}\label{equ-proof-221} du^{2}:=\biggl{(}|g, \hat{g}|^{2}\xi\, \hat{\xi}\dfrac{|g'|}{1+|g|^{2}}\dfrac{|\hat{g}'|}{1+|\hat{g}|^{2}}\biggr{)}|dz|^{2} \end{equation} outside the set $E:=\bigcup_{j=1}^{q}g^{-1}({\alpha}_{j})$ and $du^{2}=0$ on $E$. Then for a suitably chosen $b_{0}$, $du^{2}$ is continuous on $\Sigma$ and has strictly negative curvature on the set $\{du^{2}\not= 0\}$. \end{lemma} \begin{lemma}[{\cite[Corollary 2.4]{Fu1993-2}}]\label{main-lem4} Let $g$ and $\hat{g}$ be a meromorphic function on ${\triangle}_{R}$ satisfying the same assumption as in Lemma \ref{main-lem3}. Then for the metric $du^{2}$ defined by (\ref{equ-proof-221}), there exists a constant $\widehat{C}>0$ such that $$ du^{2}\leq \widehat{C}\dfrac{R^{2}}{(R^{2}-|z|^{2})^{2}}|dz|^{2}. $$ \end{lemma} \begin{proof}[{\it Proof of Theorem \ref{thm-unicity}}] For brevity, we denote the function $\hat{g}\circ \Psi$ by $\hat{g}$ in the following. We assume that there exist $q$ distinct values ${\alpha}_{1}, \ldots, {\alpha}_{q}$ such that $g^{-1}({\alpha}_{j})={\hat{g}}^{-1}({\alpha}_{j})$ $(1\leq j\leq q)$, ${\alpha}_{q}=\infty$ after a suitable M\"obius transformation and $m\not= 0$. Moreover we assume that $q> m+4$, either $ds^{2}$ or $d\hat{s}^{2}$, say $ds^{2}$, is complete and $g\not\equiv \hat{g}$. Then the map $\Psi$ gives a biholomorphic isomorphism between $\Sigma$ and $\widehat{\Sigma}$. Thus, for each local complex coordinate $z$ defined on a simply connected open domain $U$, we can find a nonzero holomorphic function $h_{z}$ such that \begin{equation}\label{equ-proof-223} ds^{2}=|h_{z}|^{2}(1+|g|^{2})^{m/2}(1+|\hat{g}|^{2})^{m/2}|dz|^{2}. \end{equation} We take some $\eta$ with $q-(m+4)> q\eta > 0$ and set \begin{equation}\label{equ-proof-224} \lambda :=\dfrac{m}{q-4-q\eta}\: (< 1). \end{equation} Now we define a new metric \begin{equation}\label{equ-proof-225} d{\sigma}^{2}= |h_{z}|^{2/(1-\lambda)}\Biggl{(}\dfrac{\prod_{j=1}^{q-1}(|g-{\alpha}_{j}||\hat{g}-{\alpha}_{j}|)^{1-\eta}}{|g-\hat{g}|^{2}|g'_{z}||\hat{g}'_{z}|\prod_{j=1}^{q-1}(1+|{\alpha}_{j}|^{2})^{(1-\eta)/2}} \Biggr{)}^{\lambda /(1-\lambda)}|dz|^{2}, \end{equation} on the set ${\Sigma}'=\Sigma \backslash E'$, where $$ E'=\{z\in \Sigma \,;\, g'_{z}(z)=0,\, \hat{g}'_{z}(z)=0,\; \text{or}\; g(z)(=\hat{g}(z))={\alpha}_{j}\; \text{for some $j$} \}. $$ On the other hand, setting $\varepsilon := \eta /2$, we can define another pseudo-metric $du^{2}$ on $\Sigma$ given by (\ref{equ-proof-221}), which has strictly negative curvature on ${\Sigma}'$. Take a point $p\in {\Sigma}'$. Since the metric $d{\sigma}^{2}$ is flat on ${\Sigma}'$, by Lemma \ref{main-lem2}, there exists a local isometry $\Phi$ satisfying $\Phi (0)=p$ from a disk ${\triangle}_{R}=\{z\in \C\,;\, |z|<R \}$ $(0<R \leq +\infty)$ with the standard metric $ds_{Euc}^{2}$ onto an open neighborhood of $p$ in ${\Sigma}'$ with the metric $d{\sigma}^{2}$. We shall denote the functions $g\circ \Phi$ and $\hat{g}\circ \Phi (= \hat{g}\circ \Psi \circ \Phi)$ by, respectively, $g$ and $\hat{g}$ in the following. Moreover the pseudo-metric ${\Phi}^{\ast}du^{2}$ on ${\triangle}_{R}$ has strictly negative curvature. Since there exists no metric with strictly negative curvature on $\C$ (see \cite[Corollary 4.2.4]{Fu1993}), we get that $R< +\infty$. Furthermore, by the same argument as in the proof of Theorem \ref{thm-ramification}, we can choose a point $a_{0}$ with $|a_{0}|=1$ such that, for the curve $L_{a_{0}}=\{w:=a_{0}s\,;\, 0<s<R \}$, the $\Phi$-image ${\Gamma}_{a_{0}}$ tends to the boundary of ${\Sigma}'$ as $s$ tends to $R$. Here, if we suitably choose the constant $\eta$ in the definition (\ref{equ-proof-224}) of $\lambda$, then ${\Gamma}_{a_{0}}$ tends to the boundary of $\Sigma$. Since ${\Phi}^{\ast}d{\sigma}^{2}=|dz|^{2}$, we get by (\ref{equ-proof-225}) that $$ |h_{z}|^{2}=\biggl{(}\dfrac{|g-\hat{g}|^{2}|g'_{z}||\hat{g}'_{z}|\prod_{j=1}^{q-1}(1+|{\alpha}_{j}|^{2})^{(1-\eta)/2}}{\prod_{j=1}^{q-1}(|g-{\alpha}_{j}||\hat{g}-{\alpha}_{j}|)^{1-\eta}} \biggr{)}^{\lambda}. $$ By (\ref{equ-proof-223}), we have that \begin{eqnarray} ds^{2} &=& |h_{z}|^{2}(1+|g|^{2})^{m/2}(1+|\hat{g}|^{2})^{m/2}|dz|^{2} \nonumber \\ &=& \biggl{(}\dfrac{|g-\hat{g}||g'_{z}||\hat{g}'_{z}|(1+|g|^{2})^{m/2\lambda}(1+|\hat{g}|^{2})^{m/2\lambda}\prod_{j=1}^{q-1}(1+|{\alpha}_{j}|^{2})^{(1-\eta)/2}}{\prod_{j=1}^{q-1}(|g-{\alpha}_{j}||\hat{g}-{\alpha}_{j}|)^{1-\eta}} \biggr{)}^{\lambda}|dz|^{2} \nonumber \\ &=& \Biggl{(} {\mu}^{2}\prod_{j=1}^{q}(|g-{\alpha}_{j}||\hat{g}-{\alpha}_{j}|)^{\varepsilon}\biggl{(}\log{\dfrac{b_{0}}{|g, {\alpha}_{j}|^{2}}}\log{\dfrac{b_{0}}{|\hat{g}, {\alpha}_{j}|^{2}}}\biggr{)}^{1-\varepsilon} \Biggr{)}^{\lambda}|dz|^{2}, \nonumber \end{eqnarray} where $\mu$ is the function with $du^{2}={\mu}^{2}|dz|^{2}$. Since the function $x^{\varepsilon}\log^{1-\varepsilon}(b_{0}/x^{2})$ $(0< x\leq 1)$ is bounded, we obtain that $$ ds^{2}\leq C''\biggl{(}\dfrac{|g, \hat{g}|^{2}|g'_{z}||\hat{g}'_{z}|\xi\hat{\xi}}{(1+|g|^{2})(1+|\hat{g}|^{2})} \biggr{)}^{\lambda}|dz|^{2} $$ for some constant $C''$. By Lemma \ref{main-lem4}, we have that $$ ds^{2}\leq C'''\biggl{(}\dfrac{R}{R-|z|^{2}} \biggr{)}^{\lambda}|dz|^{2} $$ for some constant $C'''$. Thus we obtain that $$ \int_{{\Gamma}_{a_{0}}}ds \leq (C''')^{1/2}\int_{L_{a_{0}}}\biggl{(}\dfrac{R}{R^{2}-|z|^{2}} \biggr{)}^{\lambda /2}|dz|\leq C\dfrac{R^{(2-\lambda)/2}}{1-(\lambda /2)}< +\infty $$ because $0< \lambda < 1$. However it contradicts the assumption that the metric $ds^{2}$ is complete. Hence we have necessarily $g\equiv \hat{g}$. \end{proof} \section{Applications} In this section, we give several applications of our main results. \subsection{Gauss map of minimal surfaces in ${\R}^{3}$} We first recall some basic facts of minimal surfaces in Euclidean 3-space ${\R}^{3}$. Details can be found, for example, in \cite{Fu1993}, \cite{La1982} and \cite{Os1986}. Let $X=(x^{1}, x^{2}, x^{3})\colon \Sigma \to {\R}^{3}$ be an oriented minimal surface in ${\R}^{3}$. By associating a local complex coordinate $z=u+\sqrt{-1}v$ with each positive isothermal coordinate system $(u, v)$, $\Sigma$ is considered as a Riemann surface whose conformal metric is the induced metric $ds^{2}$ from ${\R}^{3}$. Then \begin{equation}\label{equ-411-Laplacian} {\triangle}_{ds^{2}}X=0 \end{equation} holds, that is, each coordinate function $x^{i}$ is harmonic. With respect to the local complex coordinate $z=u+\sqrt{-1}v$ of the surface, (\ref{equ-411-Laplacian}) is given by $$ \bar{\partial}\partial X = 0, $$ where $\partial =(\partial /\partial u -\sqrt{-1}\partial /\partial v)/2$, $\bar{\partial} =(\partial /\partial u +\sqrt{-1}\partial /\partial v)/2$. Hence each ${\phi}_{i}:=\partial x^{i}dz$ $(i=1, 2, 3)$ is a holomorphic 1-form on $\Sigma$. If we set that $$ \omega ={\phi}_{1}-\sqrt{-1}{\phi}_{2}, \quad g=\dfrac{{\phi}_{3}}{{\phi}_{1}-\sqrt{-1}{\phi}_{2}}, $$ then $\omega$ is a holomorphic 1-form and $g$ is a meromorphic function on $\Sigma$. Moreover the function $g$ coincides with the composition of the Gauss map and the stereographic projection from ${\mathbf{S}}^{2}$ onto $\C\cup \{\infty \}$, and the induced metric is given by \begin{equation}\label{equ-412-metric} ds^{2}=(1+|g|^{2})^{2}|\omega|^{2}. \end{equation} Applying Theorem \ref{thm-ramification} to the metric $ds^{2}$, we can get the Fujimoto curvature bound for a minimal surface in ${\R}^{3}$. \begin{theorem}[{\cite[Theorem C]{Fu1992}}, {\cite[Theorem 1.6.1]{Fu1993}}] \label{thm-4-minimal1} Let $X\colon \Sigma \to {\R}^{3}$ be an oriented minimal surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that \begin{equation}\label{equ-413-ramification} \gamma= \displaystyle \sum_{j=1}^{q} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}> 4\,(=2+2). \end{equation} If the Gauss map $g\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ the inequality (\ref{equ-curvature}) holds. \end{theorem} As a corollary of Theorem \ref{thm-4-minimal1}, we have the following ramification theorem for the Gauss map of a complete minimal surface in ${\R}^{3}$. \begin{corollary}[{\cite[Theorem 2]{Ru1993}}, {\cite[Corollary 3.4]{Fu1988}}]\label{thm-4-minimal2} Let $X\colon \Sigma \to {\R}^{3}$ be a complete minimal surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-413-ramification}) holds. If the Gauss map $g\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then $g$ must be constant, that is, $X(\Sigma)$ is a plane. In particular, the Gauss map of a nonflat complete minimal surface in ${\R}^{3}$ can omit at most $4\,(=2+2)$ values. \end{corollary} We remark that the author, Kobayashi and Miyaoka \cite{KKM2008} gave a similar result for the Gauss map of a special class of complete minimal surfaces (this class is called the pseudo-algebraic minimal surfaces). As an application of Corollary \ref{thm-4-minimal2}, we obtain the following analogue to the Ahlfors islands theorem. We remark that Klots and Sario \cite{KS1966} investigated the upper bound for the number of islands for the Gauss map of a minimal surface in ${\R}^{3}$. \begin{theorem}\label{thm-4-minimal3} Let $X\colon \Sigma \to {\R}^{3}$ be a complete minimal surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the inequality (\ref{equ-413-ramification}) holds. Then there exists $\varepsilon > 0$ such that, if the Gauss map $g$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $g$ must be constant, that is, $X(\Sigma)$ is a plane. \end{theorem} The important case of Theorem \ref{thm-4-minimal3} is the case where $q=9\,(=2\cdot 2+5)$ and ${\nu}_{j}=2$ for each $j$ $(j=1,\ldots, q)$. \begin{corollary}\label{thm-4-minimal4} Let $X\colon \Sigma \to {\R}^{3}$ be a complete minimal surface. Let ${\alpha}_{1}, \ldots, {\alpha}_{9} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 9)$. Then there exists $\varepsilon > 0$ such that, if the Gauss map $g$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , 9 \}$, then $g$ must be constant, that is, $X(\Sigma)$ is a plane. \end{corollary} Finally, by applying Theorem \ref{thm-unicity}, we provide the Fujimoto unicity theorem for the Gauss maps of complete minimal surfaces in ${\R}^{3}$. \begin{theorem}[{\cite[Theorem I]{Fu1993-2}}]\label{thm-4-minimal5} Let $X\colon \Sigma \to {\R}^{3}$ and $\widehat{X}\colon \widehat{\Sigma}\to {\R}^{3}$ be two nonflat minimal surfaces and assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Let $g\colon \Sigma \to \C\cup\{\infty \}$ and $\hat{g}\colon \widehat{\Sigma}\to \C\cup\{\infty \}$ be the Gauss maps of $X(\Sigma)$ and $\widehat{X}(\widehat{\Sigma})$, respectively. If $g\not\equiv \hat{g}\circ \Psi$ and either $X(\Sigma)$ or $\widehat{X}(\widehat{\Sigma})$ is complete, then $g$ and $\hat{g}\circ \Psi$ share at most $6\,(=2+4)$ distinct values. \end{theorem} \subsection{Hyperbolic Gauss map of constant mean curvature one surfaces in ${\H}^{3}$} We denote by ${\H}^{3}$ hyperbolic 3-space, that is, the simply connected Riemannian 3-manifold with constant sectional curvature $-1$, which is represented as $$ {\H}^{3}= SL(2, \C) / SU(2) =\{aa^{\ast} \,;\, a\in SL(2, \C) \} \quad (a^{\ast}:=\tr{\bar{a}}). $$ Then there exists the following representation formula for constant mean curvature one (CMC-1, for short) surfaces in ${\H}^{3}$ as an analogy of the Enneper-Weierstrass representation formula in minimal surface theory. \begin{theorem}[\cite{Br1987}, \cite{UY1993}]\label{thm-CMC-1} Let $\widetilde{\Sigma}$ be a simply connected Riemann surface with a base point $z_{0}\in \widetilde{\Sigma}$ and $(g, \omega)$ a pair consisting of a meromorphic function and a holomorphic 1-form on $\widetilde{\Sigma}$ such that \begin{equation}\label{equ-CMC-1} ds^{2}=(1+|g|^{2})^{2}|\omega|^{2} \end{equation} gives a (positive definite) Riemannian metric on $\widetilde{\Sigma}$. Take a holomorphic immersion $F=(F_{ij})\colon \widetilde{\Sigma}\to SL(2, \C)$ satisfying $F(z_{0})=id$ and \begin{equation}\label{equ-CMC-2} F^{-1}dF=\left( \begin{array}{cc} g & -g^{2} \\ 1 & -g \end{array} \right)\omega. \end{equation} Then $f\colon \widetilde{\Sigma}\to {\H}^{3}$ defined by \begin{equation}\label{equ-CMC-3} f=FF^{\ast} \end{equation} is a CMC-1 surface and the induced metric of $f$ is $ds^{2}$. Moreover the second fundamental form $h$ and the Hopf differential $Q$ of $f$ are given by $$ h=-Q-\overline{Q}+ds^{2}, \quad Q=\omega\, dg. $$ Conversely, for any CMC-1 surface $f\colon \widetilde{\Sigma}\to {\H}^{3}$, there exist a meromorphic function $g$ and a holomorphic 1-form $\omega$ on $\widetilde{\Sigma}$ such that the induced metric of $f$ is given by (\ref{equ-CMC-1}) and (\ref{equ-CMC-3}) holds, where the map $F\colon \widetilde{\Sigma}\to SL(2, \C)$ is a holomorphic null (``null'' means $\det{(F^{-1}dF)}=0$) immersion satisfying (\ref{equ-CMC-2}). \end{theorem} Following the terminology of \cite{UY1993}, $g$ is called a {\it secondary Gauss map} of $f$. The pair $(g, \omega)$ is called {\it Weierstrass data} of $f$. Let $f\colon \Sigma\to {\H}^{3}$ be a CMC-1 surface on a (not necessarily simply connected) Riemann surface $\Sigma$. Then the map $F$ is defined only on its universal covering surface $\widetilde{\Sigma}$. Thus the pair $(\omega, g)$ is not single-valued on $\Sigma$. However the {\it hyperbolic Gauss map} of $f$ defined by $$ G=\dfrac{dF_{11}}{dF_{21}}=\dfrac{dF_{12}}{dF_{22}}, \quad \text{where}\quad F(z)=\left( \begin{array}{cc} F_{11}(z) & F_{12}(z) \\ F_{21}(z) & F_{22}(z) \end{array} \right) $$ is a single-valued meromorphic function on $\Sigma$. By identifying the ideal boundary ${\Si}^{2}_{\infty}$ of ${\H}^{3}$ with the Riemann sphere $\C\cup\{\infty \}$, the geometric meaning of $G$ is given as follows (cf. \cite{Br1987}): The hyperbolic Gauss map $G$ sends each $p\in \Sigma$ to the point $G(p)$ at ${\Si}^{2}_{\infty}$ reached by the oriented normal geodesics emanating from the surface. The inverse matrix $F^{-1}$ is also a holomorphic null immersion and produce a new CMC-1 surface $f^{\sharp}:=F^{-1}(F^{-1})^{\ast}\colon \widetilde{\Sigma} \to {\H}^{3}$ which is called the {\it dual} of $f$ (\cite{UY1997}). Then the Weierstrass data $(g^{\sharp}, {\omega}^{\sharp})$, the Hopf differential $Q^{\sharp}$, and the hyperbolic Gauss map $G^{\sharp}$ of $f^{\sharp}$ are given by \begin{equation}\label{equ-CMC-4} g^{\sharp}=G, \quad {\omega}^{\sharp}=-\dfrac{Q}{dG}, \quad Q^{\sharp}=-Q, \quad G^{\sharp}= g. \end{equation} By Theorem \ref{thm-CMC-1} and (\ref{equ-CMC-4}) , the induced metric $ds^{2\sharp}$ of $f^{\sharp}$ is given by \begin{equation}\label{equ-CMC-5} ds^{2\sharp}=(1+|g^{\sharp}|^{2})^{2}|{\omega}^{\sharp}|^{2}=(1+|G|^{2})^{2}\biggl{|}\dfrac{Q}{dG}\biggr{|}^{2}. \end{equation} We call the metric $ds^{2\sharp}$ the {\it dual metric} of $f$. There exists the following relationship between the metric $ds^{2}$ and the dual metric $ds^{2\sharp}$. \begin{theorem}[\cite{UY1997}, \cite{Yu1997}]\label{thm-CMC-2} The metric $ds^{2}$ is complete (resp. nondegenerate) if and only if the dual metric $ds^{2\sharp}$ is complete (resp. nondegenerate). \end{theorem} Applying Theorem \ref{thm-ramification} to the dual metric $ds^{2\sharp}$, we get the following theorem. \begin{theorem}\label{thm-CMC-3} Let $f\colon \Sigma \to {\H}^{3}$ be a CMC-1 surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-413-ramification}) holds. If the hyperbolic Gauss map $G\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $G$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ we have $$ |K_{ds^{2\sharp}}(p)|^{1/2}\leq \dfrac{C}{d(p)}, $$ where $K_{ds^{2\sharp}}(p)$ is the Gaussian curvature of the metric $ds^{2\sharp}$ at $p$ and $d(p)$ is the geodesic distance from $p$ to the boundary of $\Sigma$. \end{theorem} Combining of Theorems \ref{thm-CMC-2} and \ref{thm-CMC-3}, we get the following ramification theorem for the hyperbolic Gauss map of CMC-1 surfaces in ${\H}^{3}$. \begin{corollary}[\cite{Ka2011}, \cite{Yu1997}]\label{thm-CMC-4} Let $f\colon \Sigma \to {\H}^{3}$ be a complete CMC-1 surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-413-ramification}) holds. If the hyperbolic Gauss map $G\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $G$ have multiplicity at least ${\nu}_{j}$, then $G$ must be constant, that is, $f(\Sigma)$ is a horosphere. In particular, if the hyperbolic Gauss map of a nonflat complete CMC-1 surface in ${\H}^{3}$ can omit at most $4\,(=2+2)$ values. \end{corollary} Moreover we obtain the following analogue to the Ahlfors islands theorem. \begin{theorem}\label{thm-CMC-5} Let $f\colon \Sigma \to {\H}^{3}$ be a complete CMC-1 surface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the inequality (\ref{equ-413-ramification}) holds. Then there exists $\varepsilon > 0$ such that, if the hyperbolic Gauss map $G$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $G$ must be constant, that is, $f(\Sigma)$ is a horosphere. \end{theorem} The important case of Theorem \ref{thm-CMC-5} is the case where $q=9\,(=2\cdot 2+5)$ and ${\nu}_{j}=2$ for each $j$ $(j=1, \ldots, q)$. \begin{corollary}\label{thm-CMC-6} Let $f\colon \Sigma \to {\H}^{3}$ be a complete CMC-1 surface. Let ${\alpha}_{1}, \ldots, {\alpha}_{9} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 9)$. Then there exists $\varepsilon > 0$ such that, if the hyperbolic Gauss map $G$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , 9 \}$, then $G$ must be constant, that is, $f(\Sigma)$ is a horosphere. \end{corollary} Finally, by applying Theorem \ref{thm-unicity}, we provide the following unicity theorem for the hyperbolic Gauss maps of complete CMC-1 surfaces in ${\H}^{3}$. \begin{theorem}\label{thm-CMC-7} Let $f\colon \Sigma \to {\H}^{3}$ and $\widehat{f}\colon \widehat{\Sigma}\to {\H}^{3}$ be two nonflat CMC-1 surfaces and assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Let $G\colon \Sigma \to \C\cup\{\infty \}$ and $\widehat{G}\colon \widehat{\Sigma}\to \C\cup\{\infty \}$ be the hyperbolic Gauss maps of $f(\Sigma)$ and $\widehat{f}(\widehat{\Sigma})$, respectively. If $G\not\equiv \hat{G}\circ \Psi$ and either $f(\Sigma)$ or $\hat{f}(\widehat{\Sigma})$ is complete, then $G$ and $\hat{G}\circ \Psi$ share at most $6\,(=2+4)$ distinct values. \end{theorem} \subsection{Lorentzian Gauss map of maxfaces in ${\R}^{3}_{1}$} Maximal surfaces in the Lorentz-Minkowski 3-space ${\R}^{3}_{1}$ are closely related to minimal surfaces in ${\R}^{3}$. We treat maximal surfaces with some admissible singularities, called {\it maxfaces}, as introduced by Umehara and Yamada \cite{UY2006}. We remark that maxfaces, non-branched generalized maximal surfaces in the sense of \cite{ER1992} and non-branched generalized maximal maps in the sense of \cite{IK2008} are all the same class of maximal surfaces. The Lorentz-Minkowski $3$-space ${\R}^{3}_{1}$ is the affine $3$-space ${\R}^{3}$ with the inner product $$ \langle \, , \, \rangle = -(dx^{1})^{2}+(dx^{2})^{2}+(dx^{3})^{2}, $$ where $(x^{1}, x^{2}, x^{3})$ is the canonical coordinate system of ${\R}^{3}$. We consider a fibration $$ p_{L}\colon {\C}^{3} \ni ({\zeta}^{1}, {\zeta}^{2}, {\zeta}^{3}) \mapsto \text{Re} (-\sqrt{-1}{\zeta}^{1}, {\zeta}^{2}, {\zeta}^{3}) \in {\R}^{3}_{1}. $$ The projection of null holomorphic immersions into ${\R}^{3}_{1}$ by $p_{L}$ gives maxfaces. Here, a holomorphic map $F=(F_{1}, F_{2}, F_{3})\colon \Sigma \to {\C}^{3}$ is said to be {\it null} if $\{(F_{1})'_{z}\}^{2}+\{(F_{2})'_{z}\}^{2}+\{(F_{3})'_{z}\}^{2}$ vanishes identically, where $'=d / dz$ denotes the derivative with respect to a local complex coordinate $z$ of $\Sigma$. For a maxface, an analogue of the Enneper-Weierstrass representation formula is known (see also \cite{Ko1983}). \begin{theorem}[{\cite[Theorem 2.6]{UY2006}}]\label{max-EW} Let $\Sigma$ be a Riemann surface and $(g, \omega)$ a pair consisting of a meromorphic function and a holomorphic $1$-form on $\Sigma$ such that \begin{equation}\label{max-nullmet} d{\sigma}^{2}:= (1+|g|^{2})^{2}|\omega|^{2} \end{equation} gives a (positive definite) Riemannian metric on $\Sigma$, and $|g|$ is not identically $1$. Assume that $$ \text{Re} \int_{\gamma} (-2g, 1+g^{2}, \sqrt{-1}(1-g^{2}))\, \omega = 0 $$ for all loops $\gamma$ in $\Sigma$. Then \begin{equation}\label{max-immer} f= \text{Re} \int^{z}_{z_{0}} (-2g, 1+g^{2}, \sqrt{-1}(1-g^{2}))\, \omega \end{equation} is well-defined on $\Sigma$ and gives a maxface in ${\R}^{3}_{1}$, where $z_{0}\in {\Sigma}$ is a base point. Moreover, all maxfaces are obtained in this manner. The induced metric $ds^{2}:=f^{\ast} \langle \, , \, \rangle$ is given by $$ ds^{2} = (1-|g|^{2})^{2}|\omega|^{2}, $$ and the point $p\in \Sigma$ is a singular point of $f$ if and only if $|g(p)|=1$. \end{theorem} We call $g$ the {\it Lorentzian Gauss map} of $f$. If $f$ has no singularities, then $g$ coincides with the composition of the Gauss map (i.e. (Lorentzian) unit normal vector) $n \colon \Sigma \to {\H}^{2}_{\pm}$ into the upper or lower connected component of the two-sheet hyperboloid ${\H}^{2}_{\pm}={\H}^{2}_{+}\cup {\H}^{2}_{-}$ in ${\R}^{3}_{1}$, where \begin{eqnarray} {\H}^{2}_{+} &:=& \{n=(n^{1}, n^{2}, n^{3})\in {\R}^{3}_{1}\, ; \, \langle n, n \rangle = -1, n^{1}> 0\}, \nonumber \\ {\H}^{2}_{-} &:=& \{n=(n^{1}, n^{2}, n^{3})\in {\R}^{3}_{1}\, ; \, \langle n, n \rangle = -1, n^{1}< 0\}, \nonumber \end{eqnarray} and the stereographic projection from the north pole $(1, 0, 0)$ of the hyperboloid onto the Riemann sphere $\C\cup \{\infty\}$ (see \cite[Section 1]{UY2006} for the details). A maxface is said to be {\it weakly complete} if the metric $d{\sigma}^{2}$ as in (\ref{max-nullmet}) is complete. We note that $(1/2)d{\sigma}^{2}$ coincides with the pull-back of the standard metric on ${\C}^{3}$ by the null holomorphic immersion of $f$ (see \cite[Section 2]{UY2006}). Applying Theorem \ref{thm-ramification} to the metric $d{\sigma}^{2}$, we can get the following curvature estimate. \begin{theorem}\label{thm-4-maximal1} Let $f\colon \Sigma \to {\R}^{3}_{1}$ be a maxface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-413-ramification}) holds. If the Lorentzian Gauss map $g\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ we have $$ |K_{d{\sigma}^{2}}(p)|^{1/2}\leq \dfrac{C}{d(p)}, $$ where $K_{d{\sigma}^{2}}(p)$ is the Gaussian curvature of the metric $d{\sigma}^{2}$ at $p$ and $d(p)$ is the geodesic distance from $p$ to the boundary of $\Sigma$. \end{theorem} As a corollary of Theorem \ref{thm-4-maximal1}, we have the following ramification theorem for the Lorentzian Gauss map of a weakly complete maxface in ${\R}^{3}_{1}$. \begin{corollary}\label{thm-4-maximal2} Let $f\colon \Sigma \to {\R}^{3}_{1}$ be a weakly complete maxface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-413-ramification}) holds. If the Lorentzian Gauss map $g\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $g$ have multiplicity at least ${\nu}_{j}$, then $g$ must be constant, that is, $f(\Sigma)$ is a plane. In particular, the Lorentzian Gauss map of a nonflat weakly complete maxface in ${\R}^{3}_{1}$ can omit at most $4\,(=2+2)$ values. \end{corollary} As an application of Corollary \ref{thm-4-maximal2}, we can give a simple proof of the Calabi-Bernstein theorem (\cite{Ca1970}, \cite{CY1976}) for a maximal space-like surface in ${\R}^{3}_{1}$ from the viewpoint of function-theoretic properties of the Lorenzian Gauss map. For the details, see \cite[Section 4.2]{Ka2013}. As another application of Corollary \ref{thm-4-maximal2}, we obtain the following analogue to the Ahlfors islands theorem. \begin{theorem}\label{thm-4-maximal3} Let $f\colon \Sigma \to {\R}^{3}_{1}$ be a weakly complete maxface. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the inequality (\ref{equ-413-ramification}) holds. Then there exists $\varepsilon > 0$ such that, if the Lorentzian Gauss map $g$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $g$ must be constant, that is, $f(\Sigma)$ is a plane. \end{theorem} The important case of Theorem \ref{thm-4-maximal3} is the case where $q=9\,(=2\cdot 2+5)$ and ${\nu}_{j}=2$ for each $j$ $(j=1,\ldots, q)$. \begin{corollary}\label{thm-4-maximal4} Let $f\colon \Sigma \to {\R}^{3}_{1}$ be a weakly complete maxface. Let ${\alpha}_{1}, \ldots, {\alpha}_{9} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 9)$. Then there exists $\varepsilon > 0$ such that, if the Lorentzian Gauss map $g$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , 9 \}$, then $g$ must be constant, that is, $f(\Sigma)$ is a plane. \end{corollary} Finally, by applying Theorem \ref{thm-unicity}, we provide the following unicity theorem for the Lorentzian Gauss maps of weakly complete maxfaces in ${\R}^{3}_{1}$. \begin{theorem}\label{thm-4-maximal5} Let $f\colon \Sigma \to {\R}^{3}_{1}$ and $\widehat{f}\colon \widehat{\Sigma}\to {\R}^{3}_{1}$ be two nonflat maxfaces and assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Let $g\colon \Sigma \to \C\cup\{\infty \}$ and $\hat{g}\colon \widehat{\Sigma}\to \C\cup\{\infty \}$ be the Lorentzian Gauss maps of $f(\Sigma)$ and $\widehat{f}(\widehat{\Sigma})$, respectively. If $g\not\equiv \hat{g}\circ \Psi$ and either $f(\Sigma)$ or $\widehat{f}(\widehat{\Sigma})$ is weakly complete, then $g$ and $\hat{g}\circ \Psi$ share at most $6\,(=2+4)$ distinct values. \end{theorem} \subsection{Lagrangian Gauss map of improper affine fronts in ${\R}^{3}$} Improper affine spheres in the affine $3$-space ${\R}^{3}$ also have similar properties to minimal surfaces in Euclidean $3$-space. Recently, Mart\'inez \cite{Ma2005} discovered the correspondence between improper affine spheres and smooth special Lagrangian immersions in the complex $2$-space ${\C}^{2}$ and introduced the notion of {\it improper affine fronts}, that is, a class of (locally strongly convex) improper affine spheres with some admissible singularities in ${\R}^{3}$. We remark that this class is called ``improper affine maps'' in \cite{Ma2005}, but we call this class ``improper affine fronts'' because all of improper affine maps are wave fronts in ${\R}^{3}$ (\cite{Na2009}, \cite{UY2011}). The differential geometry of wave fronts is discussed in \cite{SUY2009}. Moreover, Mart\'inez gave the following holomorphic representation for this class. \begin{theorem}[{\cite[Theorem 3]{Ma2005}}] Let $\Sigma$ be a Riemann surface and $(F, G)$ a pair of holomorphic functions on $\Sigma$ such that $\text{Re}(FdG)$ is exact and $|dF|^{2}+|dG|^{2}$ is positive definite. Then the induced map $\psi\colon \Sigma \to {\R}^{3}=\C\times {\R}$ given by $$ \psi :=\biggl{(}G+\overline{F}, \dfrac{|G|^{2}-|F|^{2}}{2}+\text{Re}\biggl{(} GF- 2\int FdG \biggr{)} \biggr{)} $$ is an improper affine front. Conversely, any improper affine front is given in this way. Moreover we set $x:= G+\overline{F}$ and $n:= \overline{F}-G$. Then $L_{\psi}:=x+\sqrt{-1}n\colon \Sigma \to {\C}^{2}$ is a special Lagrangian immersion whose induced metric $d{\tau}^{2}$ from ${\C}^{2}$ is given by $$ d{\tau}^{2}=2(|dF|^{2}+|dG|^{2}). $$ In addition, the affine metric $h$ of $\psi$ is expressed as $h:=|dG|^{2}-|dF|^{2}$ and the singular points of $\psi$ correspond to the points where $|dF|=|dG|$. \end{theorem} We remark that Nakajo \cite{Na2009} constructed a representation formula for indefinite improper affine spheres with some admissible singularities. The nontrivial part of the Gauss map of $L_{\psi}\colon \Sigma \to {\C}^{2}\simeq {\R}^{4}$ (see \cite{CM1987}) is the meromorphic function $\nu\colon \Sigma \to \C\cup\{\infty \}$ given by $$ \nu := \dfrac{dF}{dG}, $$ which is called the {\it Lagrangian Gauss map} of $\psi$. An improper affine front is said to be {\it weakly complete} if the induced metric $d{\tau}^{2}$ is complete. We remark that $$ d{\tau}^{2}=2(|dF|^{2}+|dG|^{2})=2(1+|\nu|^{2})|dG|^{2}. $$ Applying Theorem \ref{thm-ramification} to the metric $d{\tau}^{2}$, we can get the following theorem. This is a generalization of \cite[Theorem 4.6]{Ka2013}. \begin{theorem}\label{thm-4-improper1} Let $\psi\colon \Sigma \to {\R}^{3}$ be an improper affine front. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that \begin{equation}\label{equ-4-imramification} \gamma= \displaystyle \sum_{j=1}^{q} \biggl{(}1-\dfrac{1}{{\nu}_{j}} \biggr{)}> 3\,(=1+2). \end{equation} If the Lagrangian Gauss map $\nu\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $\nu$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ we have $$ |K_{d{\tau}^{2}}(p)|^{1/2}\leq \dfrac{C}{d(p)}, $$ where $K_{d{\tau}^{2}}(p)$ is the Gaussian curvature of the metric $d{\tau}^{2}$ at $p$ and $d(p)$ is the geodesic distance from $p$ to the boundary of $\Sigma$. \end{theorem} As a corollary of Theorem \ref{thm-4-improper1}, we have the following ramification theorem for the Lagrangian Gauss map of a weakly complete improper affine front in ${\R}^{3}$. \begin{corollary}[{\cite[Theorem 3.2]{KN2012}}]\label{thm-4-improper2} Let $\psi\colon \Sigma \to {\R}^{3}$ be a weakly complete improper affine front. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-4-imramification}) holds. If the Lagrangian Gauss map $\nu\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $\nu$ have multiplicity at least ${\nu}_{j}$, then $\nu$ must be constant, that is, $\psi(\Sigma)$ is an elliptic paraboloid. In particular, if the Lagrangian Gauss map of a weakly complete improper affine front in ${\R}^{3}$ is nonconstant, then it can omit at most $3\,(=1+2)$ values. \end{corollary} Since the singular points of $\psi$ correspond to the points where $|\nu|=1$, we can get a simple proof of the parametric affine Bernstein theorem (\cite{Ca1958}, \cite{Jo1954}) for an improper affine sphere from the viewpoint of function-theoretic properties of the Lagrangian Gauss map. For the details, see \cite[Corollary 3.6]{Ka2013}. As an application of Corollary \ref{thm-4-improper2}, we obtain the following analogue to the Ahlfors islands theorem. \begin{theorem}\label{thm-4-improper3} Let $\psi\colon \Sigma \to {\R}^{3}$ be a weakly complete improper affine front. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the inequality (\ref{equ-4-imramification}) holds. Then there exists $\varepsilon > 0$ such that, if the Lagrangian Gauss map $\nu$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $\nu$ must be constant, that is, $\psi(\Sigma)$ is an elliptic paraboloid. \end{theorem} The important case of Theorem \ref{thm-4-improper3} is the case where $q=7\,(=2\cdot 1+5)$ and ${\nu}_{j}=2$ for each $j$ $(j=1,\ldots, q)$. \begin{corollary}\label{thm-4-improper4} Let $\psi\colon \Sigma \to {\R}^{3}$ be a weakly complete improper affine front. Let ${\alpha}_{1}, \ldots, {\alpha}_{7} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 7)$. Then there exists $\varepsilon > 0$ such that, if the Lagrangian Gauss map $\nu$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , 7 \}$, then $\nu$ must be constant, that is, $\psi(\Sigma)$ is an elliptic paraboloid. \end{corollary} Finally, by applying Theorem \ref{thm-unicity}, we provide the following unicity theorem for the Lagrangian Gauss maps of weakly complete improper affine fronts in ${\R}^{3}$. \begin{theorem}\label{thm-4-improper5} Let $\psi\colon \Sigma \to {\R}^{3}$ and $\widehat{\psi}\colon \widehat{\Sigma}\to {\R}^{3}$ be two improper affine fronts and assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Let $\nu\colon \Sigma \to \C\cup\{\infty \}$ and $\hat{\nu}\colon \widehat{\Sigma}\to \C\cup\{\infty \}$ be the Lagrangian Gauss maps of $\psi(\Sigma)$ and $\widehat{\psi}(\widehat{\Sigma})$, respectively. Suppose that there exist $q$ distinct points ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ such that ${\nu}^{-1}({\alpha}_{j})=(\hat{\nu}\circ \Psi)^{-1}({\alpha}_{j})$ $(1\leq j\leq q)$. If $q \geq 6 \,(=(1+4)+1)$ and either $\psi (\Sigma)$ or $\widehat{\psi} (\widehat{\Sigma})$ is weakly complete, then either $\nu\equiv \hat{\nu}\circ \Psi$ or $\nu$ and $\hat{\nu}$ are both constant, that is, $\psi (\Sigma)$ and $\widehat{\psi} (\widehat{\Sigma})$ are both elliptic paraboloids. \end{theorem} \subsection{Ratio of canonical forms of flat fronts in ${\H}^{3}$} For a holomorphic Legendrian immersion $\Lc\colon \Sigma \to SL(2, \C)$ on a simply connected Riemann surface $\Sigma$, the projection $$ f:= \Lc{\Lc}^{\ast}\colon \Sigma \to {\H}^{3} $$ gives a {\it flat front} in ${\H}^{3}$. Here, flat fronts in ${\H}^{3}$ are flat surfaces in ${\H}^{3}$ with some admissible singularities (see \cite{KRUY2007}, \cite{KUY2004} for the definition of flat fronts in ${\H}^{3}$). We call $\Lc$ the {\it holomorphic lift} of $f$. Since $\Lc$ is a holomorphic Legendrian map, ${\Lc}^{-1}{d\Lc}$ is off-diagonal (see \cite{GMM2000}, \cite{KUY2003}, \cite{KUY2004}). If we set that $$ {\Lc}^{-1}{d\Lc} = \left( \begin{array}{cc} 0 & \theta \\ \omega & 0 \end{array} \right), $$ then the pull-back of the canonical Hermitian metric of $SL(2, \C)$ by $\Lc$ is represented as $$ ds^{2}_{\Lc}:=|\omega|^{2}+|\theta|^{2} $$ for holomorphic $1$-forms $\omega$ and $\theta$ on $\Sigma$. A flat front $f$ is said to be {\it weakly complete} if the metric $ds^{2}_{\Lc}$ is complete (\cite{KRUY2009, UY2011}). We define a meromorphic function on $\Sigma$ by the ratio of canonical forms $$ \rho := \dfrac{\theta}{\omega}. $$ Then a point $p\in \Sigma$ is a singular point of $f$ if and only if $|\rho (p)|=1$ (\cite{KRSUY2005}). We remark that $$ ds^{2}_{\Lc}=|\omega|^{2}+|\theta|^{2}=(1+|\rho|^{2})|\omega|^{2}. $$ Applying Theorem \ref{thm-ramification} to the metric $ds^{2}_{\Lc}$, we can get the following curvature estimate. This is a generalization of \cite[Theorem 4.8]{Ka2013}. \begin{theorem}\label{thm-4-flat1} Let $f\colon \Sigma \to {\H}^{3}$ be a flat front on a simply connected Riemann surface $\Sigma$. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-4-imramification}) holds. If the ratio of canonical forms $\rho\colon \Sigma \to \C\cup \{\infty \}$ satisfies the property that all ${\alpha}_{j}$-points of $\rho$ have multiplicity at least ${\nu}_{j}$, then there exists a positive constant $C$, depending on ${\alpha}_{1}, \ldots, {\alpha}_{q}$, but not the surface, such that for all $p\in \Sigma$ we have $$ |K_{ds^{2}_{\Lc}}(p)|^{1/2}\leq \dfrac{C}{d(p)}, $$ where $K_{ds^{2}_{\Lc}}(p)$ is the Gaussian curvature of the metric $ds^{2}_{\Lc}$ at $p$ and $d(p)$ is the geodesic distance from $p$ to the boundary of $\Sigma$. \end{theorem} If $\Sigma$ is not simply connected, then we consider that $\rho$ is a meromorphic function on its universal covering surface $\widetilde{\Sigma}$. As a corollary of Theorem \ref{thm-4-flat1}, we have the following ramification theorem for the ratio of canonical forms of a weakly complete improper affine front in ${\H}^{3}$. \begin{corollary}[{\cite[Theorem 3.2]{Ka2013-2}}]\label{thm-4-flat2} Let $f\colon \Sigma \to {\H}^{3}$ be a weakly complete flat front. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N\cup \{\infty \}$. Suppose that the inequality (\ref{equ-4-imramification}) holds. If the ratio of canonical forms $\rho$ satisfies the property that all ${\alpha}_{j}$-points of $\rho$ have multiplicity at least ${\nu}_{j}$, then $\rho$ must be constant, that is, $f(\Sigma)$ is a horosphere or a hyperbolic cylinder. In particular, if the ratio of canonical forms of a weakly complete flat front in ${\H}^{3}$ is nonconstant, then it can omit at most $3\,(=1+2)$ values. \end{corollary} As an application of Corollary \ref{thm-4-flat2}, we can obtain a simple proof of the classification (\cite{Sa1973}, \cite{VV1971}) of complete nonsingular flat surfaces in ${\H}^{3}$. For the details, see \cite[Corollary 4.5]{Ka2013-2}. As another application of Corollary \ref{thm-4-flat2}, we get the following analogue to the Ahlfors islands theorem. \begin{theorem}[{\cite[Corollary 4.2]{Ka2013-2}}]\label{thm-4-flat3} Let $f\colon \Sigma \to {\H}^{3}$ be a weakly complete flat front. Let $q\in \N$, ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ be distinct, $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq q)$ be pairwise disjoint and ${\nu}_{1}, \ldots, {\nu}_{q}\in \N$. Suppose that the inequality (\ref{equ-4-imramification}) holds. Then there exists $\varepsilon > 0$ such that, if the ratio of canonical forms $\rho$ has no island of multiplicity less than ${\nu}_{j}$ over $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , q\}$, then $\rho$ must be constant, that is, $f(\Sigma)$ is a horosphere or a hyperbolic cylinder. \end{theorem} The important case of Theorem \ref{thm-4-flat3} is the case where $q=7\,(=2\cdot 1+5)$ and ${\nu}_{j}=2$ for each $j$ $(j=1,\ldots, q)$. \begin{corollary}[{\cite[Corollary 4.3]{Ka2013-2}}]\label{thm-4-flat4} Let $f\colon \Sigma \to {\H}^{3}$ be a weakly complete flat front. Let ${\alpha}_{1}, \ldots, {\alpha}_{7} \in \C\cup \{\infty \}$ be distinct and $D_{j}({\alpha}_{j}, \varepsilon):=\{z\in \C\cup \{\infty \} \,;\, |z, {\alpha}_{j}|< \varepsilon \}$ $(1\leq j\leq 7)$. Then there exists $\varepsilon > 0$ such that, if the ratio of canonical forms $\rho$ has no simple island of over any of the small disks $D_{j}({\alpha}_{j}, \varepsilon)$ for all $j\in \{1, \ldots , 7 \}$, then $\nu$ must be constant, that is, $f(\Sigma)$ is a horosphere or a hyperbolic cylinder. \end{corollary} Finally, by applying Theorem \ref{thm-unicity}, we provide the following unicity theorem for the ratios of canonical forms of weakly complete flat fronts in ${\H}^{3}$. \begin{theorem}\label{thm-4-flat5} Let $f\colon \Sigma \to {\H}^{3}$ and $\widehat{f}\colon \widehat{\Sigma}\to {\R}^{3}$ be two flat fronts on simply connected Riemann surfaces and assume that there exists a conformal diffeomorphism $\Psi\colon \Sigma \to \widehat{\Sigma}$. Let $\rho\colon \Sigma \to \C\cup\{\infty \}$ and $\hat{\rho}\colon \widehat{\Sigma}\to \C\cup\{\infty \}$ be the ratio of canonical forms $f(\Sigma)$ and $\widehat{f}(\widehat{\Sigma})$, respectively. Suppose that there exist $q$ distinct points ${\alpha}_{1}, \ldots, {\alpha}_{q}\in \C\cup \{\infty \}$ such that ${\rho}^{-1}({\alpha}_{j})=(\hat{\rho}\circ \Psi)^{-1}({\alpha}_{j})$ $(1\leq j\leq q)$. If $q \geq 6 \,(=(1+4)+1)$ and either $f (\Sigma)$ or $\widehat{f} (\widehat{\Sigma})$ is weakly complete, then either $\rho\equiv \hat{\rho}\circ \Psi$ or $\rho$ and $\hat{\rho}$ are both constant. \end{theorem}
\section*{This is an unnumbered first-level section head} \section{Introduction and main results} We consider meromorphic solutions of the differential equation \begin{equation} \label{E:ode} y^{(n)}(z) + \sum_{\nu=0}^{n-1} f_\nu (z)\, y^{(\nu)}(z)=0, \end{equation}We apply freely the classical Nevanlinna Theory notation throughout this paper \cite{Ha75}, \cite{La93}. where the coefficients $f_\nu (z)\, \ 0\le \nu\le n-1$ are meromorphic in $\mathbb{C}$. We mention that Heittokangas, Korhonen and R\"atty\"a obtained sharper estimates for analytic solutions when the coefficients are analytic functions in \cite{HKR2004}. They also considered non-homogeneous equations in \cite{HKR2009}. For an up-to-date account on the growth of meromorphic solutions of algebraic differential equations with meromorphic coefficients, we refer the reader to Hayman \cite{Ha1996}. \medskip Bank asked, if as in the case when the (\ref{E:ode}) admits an entire solution \cite{Ba75}, an meromorphic solution of (\ref{E:ode}) can be estimated in terms of growth of Nevanlinna characteristics of the meromorphic coefficients alone. In general, he showed that this statement is not true in \cite{Ba76} by constructing an example for any given real-valued increasing function $\Phi(r)\uparrow +\infty$ on $(0,+\infty)$, then one can construct a first order linear differential equation with entire coefficient of zero Nevanlinna order such that the differential equation admits a meromorphic function $f$ with $T(r,\, f)>\Phi(r)$ as $r\to +\infty$. One would need some extra terms to bound the growth of meromorphic solutions. An example of such a result is given by Bank and Laine. \begin{theorem}[\cite{BL77}] Suppose that the coefficients of (\ref{E:ode}) are arbitrary meromorphic functions and that $y(z)$ is a meromorphic solution of (1.1). If \[ \Phi(r)=\max_{0 \leq i \leq n} \Big(\log r, T(r,f_i)\Big), \] then for any $\sigma > 1$, there exist positive constants $c, c_1$ and $r_0$, such that for $r \geq {r_0},$ \begin{equation} \label{E:BL-upper} T(r,y) \leq c \Big\{rN(\sigma r, y)+r^2 \exp \Big(c_1 J(\sigma r) \log (rJ(\sigma r)\Big) \Big\}, \end{equation} where \[ J(r)={\overline N} (r, 1/y) +\Phi (r). \] \end{theorem} \medskip We note that one needs counting function $\overline{N}(r,\, 1/y)$ of \textit{distinct zeros} in the $J(r)$ above as part of the upper bound in (\ref{E:BL-upper}). \medskip Since the equation (\ref{E:ode}) is linear, so one can deduce from the expression \[ \frac{f^{(n)}(z)}{f(z)} + \sum_{\nu=1}^{n-1} f_\nu (z) \frac{f^{(\nu)}(z)}{f(z)}+f_0(z)=0, \] that \[ \overline{N}(r,\, f)\le \sum_{\nu=0}^{n-1} \overline{N}(r,\, f_\nu) \le \sum_{\nu=0}^{n-1} {T}(r,\, f_\nu), \] indicating that the coefficients can only bound the \textit{distinct} poles of $f$. Indeed, the example constructed by Bank \cite{Ba76} mentioned above has poles of rapidly increasing multiplicities, that is $N(r,\, f)/\overline{N}(r,\, f)$ is unbounded. Hayman and the author \cite{ChiangHayman} showed that one can still bound the growth of a meromorphic solution $f$ of (\ref{E:ode}) in terms of the characteristic functions of coefficients alone if the solution $f$ has \textit{relatively few poles}. In particular, this means that $\delta(\infty,\, f)>0$. This follows from the following result. \medskip \begin{theorem}[\cite{ChiangHayman}]\label{T:ChiangHayman} Suppose that $0<\rho< r< R$ and suppose that the coefficients $f_\nu,\ 0\le\nu\le n-1$ of (\ref{E:ode}) are analytic on the path $\Gamma=\Gamma (\theta_0,\, \rho,\, t)$ defined by the segment \[ \Gamma_1: z=\tau e^{i\theta_0},\quad \rho\le \tau\le t, \] followed by the circle \[ \Gamma_2: z=te^{i\theta},\quad \theta_0\le\theta\le \theta_0+2\pi. \] We suppose that $y(z)$ is a solution of the equation (1.1) and define \[ K=2\max\biggl\{1, \sup_{0\le\nu\le n}|y^{(\nu)}(z_0)|\biggr\}, \] where $z_0=\rho e^{i\theta_0}$. We also define \[ C=C(f_\nu,\,\rho ,\, r,\, R)= (n+2)\exp\biggl\{ \frac{20 R}{R-r}\sum_{\nu=0}^n T(R,\, f_\nu)+\biggl(\sum_{\nu=0}^n p_\nu\biggr)\log\biggl(\frac{R}{\rho}\biggr) \biggr\}, \] where $p_\nu$ is the multiplicity of the order of pole of $f_\nu$ at $z=0$. Then we have for $|z|=t$, where $t$ is some number such that $r<t<\frac{1}{4}(3r+R)$, \[ |y^{(\nu)}(z)|< KC^\nu e^{(2\pi+1)CR},\quad 0\le \nu\le n. \] \end{theorem} \bigskip One can easily deduce when $R=+\infty$, and for a transcendental meromorphic $f$ with $\delta(\infty,\, f)>0$, then for $0<\varepsilon<\delta$, we have \[ T(r,\, y)\le \Big(\frac{1}{\delta-\varepsilon}\Big)\,(2\pi+1)\,R\, C. \] \medskip The main purpose of this paper is to give a shorter proof of a slightly different statement to Theorem \ref{T:ChiangHayman} and asymptotic results outside some exceptional sets using a different method. On the other hand, the original Theorem \ref{T:ChiangHayman} can deal with non-homogeneous (\ref{E:ode}), while our alternative can only deal with the (\ref{E:ode}). We prove \medskip \begin{theorem}\label{T:main} Let $y$ be a meromorphic solution to the differential equation (\ref{E:ode}) with meromorphic coefficients $f_\nu,\, \nu=0, \cdots, n-1$ in $|z|=r< R\le +\infty$ such that $f_\nu$ has a pole of order $q_\nu\ge 0\ (0\le \nu\le n-1)$. Given a constant $C>1$ and $0<\eta< 3\,e/2$ and $r=|z|$ is outside a union of discs centred at the poles of $y$ such that the sum of radii is not greater than $4\eta R$, then there is a $B=B(C)>1$ and a path \begin{equation} \Omega=\Omega(\theta,\, \rho,\, r) \end{equation} consists of the line segment \[ \Omega_1 : z =\tau e^{i\theta_0},\quad 0\le\rho\le \tau\le r \] followed by the circle \[ \Omega_2 : z =re ^{i\theta},\quad \theta_0\le \theta< \theta_0+2\pi, \] on which the coefficients $f_\nu$ are analytic and we have, for $z$ on $\Omega$, \[ \label{E:main-estimate} \sum_{j=0}^{n-1}|y^{(j)}(z)|\le K_1\, e^{(2\pi+1)\,D\, r}\le K_1\, e^{(2\pi+1)\,D\,R}, \] where \[ \label{E:initial} K_1= \sum_{j=0}^{n-1}|y^{(j)}(z_0)|,\qquad z_0=\rho\, e^{i\theta_0} \] and \begin{equation} \label{E:D} \begin{split} D: & =D(f_\nu,\, \rho,\, r,\, R;\, \eta,\, B,\, C)\\ &\quad =n\bigg\{1+ \Big(\frac{R^{H(\eta )(\frac{R+2\,er}{R-2\,er})}}{r}\Big)^q\,\exp\Big[B\,(1+H(\eta))\Big(\frac{R+2\,er}{R-2\,er}\Big)\, T(CR)\Big]\bigg\} \end{split} \end{equation} and where \[ \label{E:defn} T(r)=\max_{0\le\nu\le n-1} T(r,\, f_\nu),\quad q=\max_{0\le\nu\le n-1} {q_\nu}, \quad H(\eta)=2+\log\frac{3e}{2\eta}. \] \end{theorem} \bigskip For any $r^\prime$, we choose $r$ outside a union of discs centred at the zeros of $y$ such that the sum of radii is not greater than $4\eta R$ such that $r^\prime <r <R$ as described in the Theorem \ref{T:main}, we have \[ \begin{split} T(r^\prime,\, y) &\le N(r,\, y)+m(r,\, y)\\ &\le N(r,\, y)+ (2\pi+1)\, R\, D(f_\nu,\, \rho,\, r,\, R;\, \eta,\, B,\, C). \end{split} \] \bigskip This improves upon Bank and Laine's estimate mentioned above. Suppose that $\delta(\infty,\, y)>0$, we choose $r$ outside a union of discs centred at the zeros of $y$ such that the sum of radii is not greater than $4\eta R$ and sufficiently large such that $N(r,\, y)<(1-\delta+\varepsilon/2)\, T(r,\, y)$. Without loss of generality, we may also assume that $r$ is so chosen such that $|\log K_1|<\frac12\varepsilon T(r,\, y)$ so that $T(r,\, y) < (1-\delta+\varepsilon)\, T(r,\, y)+(2\pi+1)R\, D$. We can easily deduce \[ T(r^\prime,\, y)\le \Big(\frac{1}{\delta(\infty,\, y)-\varepsilon}\Big)\, (2\pi+1)R\, D. \] \bigskip \begin{theorem}\label{T:density} Let $y(z)$ be a meromorphic solution to equation (\ref{E:ode}), and we choose $0< \eta <(1+\log 2)/(16\,e^{5/2})<1$. Then there is a constant $B>1$ such that for every $\varepsilon>0$ be given, there is a $r(\varepsilon)>0$ we have \medskip \[ \label{E:density-bound} \log m(r,\, y^{(j)})\le 5B\big(1+H(\eta)\big)\, T(3\,e^2r)+[(5H(\eta)-1)q+1+\varepsilon]\, \log r \] \medskip $j=0,\, 1,\, \cdots, n-1$ holds for all $r>r(\varepsilon)$ except perhaps for a set of positive logarithmic density $\displaystyle\frac{16\eta e^{5/2}}{1+\log 2}$. \end{theorem} \bigskip This result is to be compared with the following density-type result also obtained previously in \cite{ChiangHayman}: \begin{theorem} Let $y(z)$ be a meromorphic solution of (\ref{E:ode}) such that the $f_\nu$ are not all constant, we have \begin{equation} \log m(r,\, y) < \Big(\sum_{\nu=0}^{n-1} T(r,\, f_\nu)\Big)\Bigg[(\log r) \log \Big(\sum_{\nu=0}^{n-1} T(r,\, f_\nu)\Big)\Bigg]^\sigma, \end{equation} where $\sigma>1$ is a constant, to hold outside an exceptional set of finite logarithmic measure. \end{theorem} \bigskip \section{Preliminaries} \label{S:pre} Let us write ${\bf y}(z)=(y_0,\, \cdots,\, y_{n-1})^T$ where $y_j(z),\, j=0,\, \cdots, n-1$ are complex functions of $z$. We define $\|{\bf y}\|=\sum_{j=0}^{n-1}|y_j|$. Suppose further that ${\bf A}=(a_{ij}(z))$ is a square matrix then we define $\|{\bf A}\|=\sum_{i,j}|a_{ij}|$. We note that \[ \left\|\int {\bf A}\, dt\right\|\le \int \left\|{\bf A}\right\|\, |dt| \] (see e.g., \cite[pp. 1--4]{Bellman}). \bigskip \begin{lemma}\cite[pp. 21--22]{Levin}\label{L:lower-bound} Let $R>0$ and $f(z)$ be analytic in $|z|\le 2\,e\, R$ with $f(0)=1$, and let $\eta$ be an arbitrary positive constant not exceeding $3\,e/2$. Then we have \begin{equation}\label{E:lower-bound} \log|f(z)| > -H(\eta) \log M(2\,e\,R,\, f) \end{equation} for all $z$ in $|z|\le R$ but outside a union of disks centred at the zeros of $f$ such that the sum of radii is not greater than $4\eta\, R$, where \[ H(\eta)=2+ \log \frac{3\,e}{2\,\eta}. \] \end{lemma} \medskip We also need the following quotient representation of meromorphic functions due to Miles \cite{Miles1972} and Rubel \cite[Chapter 14]{Rubel1996}. \medskip \begin{lemma}[\cite{Miles1972}]\label{L:Miles} Let $f$ be a meromorphic function in the plane, and let $C>1$ be a given constant, then there exist entire functions $f_1$ and $f_2$, and a constant $B=B(C)>0$ such that \[ f(z)=\frac{f_1(z)}{f_2(z)},\quad\quad \text{and}\quad\quad T(r,\, f_j)\le B\,T(C\,r,\, f), \] $j=1,\, 2$ and $r>0$. Here both the constants $B$ and $C$ are absolute constants, i.e., they are independent of the function $f$. \end{lemma} \bigskip \section{Proof of Theorem \ref{T:main}} We state and prove our main lemma that leads to the proof of the Theorem \ref{T:main}. \begin{lemma}\label{L:on-path} Let $y$ be a meromorphic solution to the differential equation (\ref{E:ode}) with meromorphic coefficients $f_\nu,\, \nu=0, \cdots, n-1$ in $|z|=r< R\le +\infty$. Suppose that the coefficients $f_\nu,\, \nu=0, \cdots, n-1$ are analytic on the path $\Omega=\Omega(\theta,\, \rho,\, r)$ as defined in the Theorem \ref{T:main}. Suppose $z_0=\rho e^{i\theta_0}$, then for all $z$ on $\Omega$, \begin{equation} \sum_{j=0}^{n-1}|y^{(j)}(z)|\le K_1\, \exp\bigg[\Big(\max_{\Omega}\sum_{\nu=0}^{n-1} (|f_\nu(z)|+1)\Big)\,(2\pi+1)\,r\bigg], \end{equation} where $K_1$ is given in (\ref{E:initial}). \end{lemma} \bigskip \begin{proof} It is well-known that equation (\ref{E:ode}) can be written in the matrix form \begin{equation} \label{E:matrix-eqn} {\bf F'}(z)={\bf A}(z)\, \textbf{F}(z), \end{equation} where ${\bf F}=(y,\, y',\, \cdots, y^{(n-1)})^T$, and \begin{equation} {\bf A}(z)= \left( \begin{matrix} 0 & 1 & 0 & \cdots & 0 \\ 0 & 0 & 1 & \cdots & 0 \\ \vdots & \vdots & \vdots & \ddots & \vdots\\ 0 & 0 & \cdots & \cdots & 1\\ -f_0 & -f_1 & \cdots & \cdots & -f_{n-1}\\ \end{matrix} \right). \end{equation} \bigskip A solution to the above matrix equation (\ref{E:matrix-eqn}) is given by \begin{equation} {\bf F}(z)={\bf F}(z_0)+\int_{z_0}^z {\bf A}(t)\, \mathbf{F}(t)\, dt. \end{equation} \bigskip We now apply Gronwall's inequality \cite[pp. 35--36]{Bellman} to (\ref{E:matrix-eqn}) to obtain \begin{equation}\label{E:matrix} \begin{split} \|{\bf F}(z)\| & \le \|{\bf F}(z_0)\|+\int_{z_0}^z \|{\bf A}(t)\|\,\|{\bf F}(t)\|\,|dt|\\ & \le \|{\bf F}(z_0)\|\,\exp\biggl(\int_{z_0}^z \|{\bf A}(t) \|\,|dt|\biggr)\\ & \le \|{\bf F}(z_0)\|\, \exp\bigg[\Big(\max_{\Omega}\sum_{\nu=0}^{n-1} |f_\nu(z)|+(n-1)\Big)\,(2\pi+1)\,r\bigg]\\ & < \|{\bf F}(z_0)\|\, \exp\bigg[\Big(\max_{\Omega}\sum_{\nu=0}^{n-1} (|f_\nu(z)|+1)\Big)\,(2\pi+1)\,r\bigg],\\ \end{split} \end{equation} where we have parametrized the path $\Omega$ with respect to arc length. Clearly the length of $\Omega$ is $(2\pi+1)\,r$ at most. This proves Lemma \ref{L:on-path}. \end{proof} \bigskip We are ready to prove the Theorem \ref{T:main}, which is a direct application of the Lemma \ref{L:on-path} and the two lemmas stated in \S\ref{S:pre} . \medskip \subsubsection*{Proof of the Theorem \ref{T:main}} Given $C>1$ be given. Then Miles' result in Lemma \ref{L:Miles} asserts that we can choose a $B>0$ such that we can write the coefficients in $f_\nu=f_{\nu,1}/f_{\nu, 2}$ from (\ref{E:ode}) such that \[ T(r,\, f_{\nu,\, j})\le B\, T(C\, r,\, f_\nu),\qquad j=1,\, 2;\quad 0\le \nu\le n-1 \] for $r>0$. We first assume that $f_{\nu, 2}(0)\not=0$ for $ 0\le \nu\le n-1$, then it is easy to see that we may assume that $f_{\nu, 2}(0)=1$ after dividing the numerator and the denominator by a suitable constant. Lemmas \ref{L:lower-bound} and \ref{L:Miles} with \begin{equation}\label{E:miles-1} \begin{split} \log|f_\nu| & =\log |f_{\nu,1}| + \log |f_{\nu, 2}|^{-1}\\ & \le \log M(r,\, f_{\nu,1})+ H(\eta)\log M(2\,e\,r,\, f_{\nu,2})\\ & \le \log^+ M(r,\, f_{\nu,1})+ H(\eta)\log^+ M(2\,e\,r,\, f_{\nu,2})\\ & \le \Big(\frac{R+r}{R-r}\Big)\, T(R,\, f_{\nu,1})+H(\eta)\,\Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(R,\, f_{\nu,\, 2})\\ & \le B\, \Big(\frac{R+r}{R-r}\Big)\,T(CR,\,f_{\nu})+BH(\eta)\,\Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(CR,\, f_{\nu})\\ & \le B\big(1+H(\eta)\big) \Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(CR, f_{\nu})\\ & \le B\big(1+H(\eta)\big)\, \Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(CR), \end{split} \end{equation} where \[ T(r)=\max_{0\le \nu\le n-1} T(r,\, f_\nu). \] \bigskip If, however, $f_{\nu, 2}$ has a zero of order $q_\nu$ at $z=0$, we consider \[ f_\nu=\frac {f_{\nu,1}/z^{q_\nu}}{f_{\nu,2}/z^{q_\nu}} \] in $0<\rho\le |z|$ in which the $f_{\nu,2}/z^{q_\nu}$ is clearly still analytic and not zero at the origin. We deduce from Lemmas \ref{L:lower-bound} and \ref{L:Miles} again that for $0\le\nu\le n-1$ \begin{equation}\label{E:miles-2} \begin{split} \log| f_\nu| & =\log |f_{\nu,1}|-q_\nu\log r + \log |f_{\nu, 2}/z^{q_\nu}|^{-1}\\ & \le \log M(r,\, f_{\nu,1})-q_\nu\log r + H(\eta)\log M\big(2\,e\,r,\, f_{\nu,2}/z^{q_\nu}\big)\\ & \le \log^+ M(r,\, f_{\nu,1})-q_\nu\log r+ H(\eta)\log^+ M\big(2\,e\,r,\, f_{\nu,2}/z^{q_\nu}\big)\\ & \le \Big(\frac{R+r}{R-r}\Big)\,T(R,\,f_{\nu,\, 1}) -q_\nu\log r +H(\eta)\,\Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(R,\, f_{\nu,2}/z^{q_\nu})\\ & \le \Big(\frac{R+r}{R-r}\Big)\,T(R,\,f_{\nu,\, 1}) -q_\nu\log r\\ & \qquad +H(\eta)\,\Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, \big(T(R,\, f_{\nu,2})+q_\nu\log R\big)\\ & \le B\Big(\frac{R+r}{R-r}\Big)\,T(CR,\,f_{\nu}) -q_\nu\log r \\ & \qquad +H(\eta)\,\Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, \big(B T(CR,\, f_{\nu})+q_\nu\log R\big)\\ & <B\big(1+H(\eta)\big)\, \Big(\frac{R+2\,e\,r}{R-2\,e\,r}\Big)\, T(CR,\, f_{\nu})+q_\nu\,\log\Big[\frac{R^{H(\eta)(\frac{R+2\,e\,r}{R-2\,e\,r})}}{r}\Big]. \end{split} \end{equation} Applying (\ref{E:lower-bound}) to (\ref{E:miles-1}) or (\ref{E:miles-2}) ($0\le\nu\le n-1$) and substituting them into (\ref{E:matrix}) completes the proof. \qed \bigskip \section{Proof of Theorem \ref{T:density}} We are now ready to prove Theorem \ref{T:density}. We choose $C=e$ in Theorem \ref{T:main}. Let $\alpha=2\, e$. We define annuli by \[ \Lambda_j=\bigl\{ z: \alpha^j \le |z| \le \alpha^{j+3/2}\bigr\},\quad\quad j=1,\, 2,\, \cdots. \] \medskip We take $R=3\,er$ in (\ref{E:D}) in Theorem \ref{T:main} and suppose $z$ belongs to $\Omega\cap\Lambda_j$ where $\Omega$ is defined in Theorem \ref{T:main}. That is, we have, for each $0\le j\le n-1$, \begin{equation} \begin{split} m(r,\, y^{(j)}) & \le (2\pi+1)\,r\,n\Bigg\{1+(3e)^{5H(\eta)q}r^{(5H(\eta)-1)q}\, \exp\Big[5B\big(1+H(\eta)\big)\, T(3\,e^2r)\Big]\Bigg\}\\ &\quad +\log K_1\\ &= \log K_1+(2\pi+1)\, r\, D(f_\nu,\, \rho,\, r,\, 3er,\, \eta,\, B,\, e), \end{split} \end{equation} where $D$ is given in (\ref{E:D}). Taking logarithm on both sides of the inequality once more yield the required estimate (\ref{E:density-bound}). \bigskip It remains to verify the size of the exceptional set of $r$, which follows from \medskip \begin{lemma}\sl Let $\eta$ and $H(\eta)$ be defined in Lemma \ref{L:lower-bound} and Theorem \ref{T:main} respectively. Then the estimate (\ref{E:density-bound}) for a meromorphic solution $y(z)$ is valid for all $r$ sufficiently large except on a set of positive logarithmic density at most $16\eta e^{5/2}/(1+\log 2)$. \end{lemma} \bigskip Let $E_j$ be the union of exceptional circles lying in $\Lambda_j$ and \[ E(r)=[1,\, r)\cap\bigl(\cup_{j=1}^\infty E_j\bigr). \] Let $q=\left[\frac{\log r}{\log\alpha}\right]$, then Lemma \ref{L:lower-bound} gives \[ \begin{split} \int_{E(r)}\frac{dt}{t} &\le\sum_{j=1}^q \int_{E_j}\frac{dt}{t}\\ & \le\sum_{j=1}^q \frac{4\eta(2\, er)}{\alpha^j}\\ & \le\sum_{j=1}^q \frac{4\eta(2e\alpha^{j+3/2})}{\alpha^j}\\ & \le\frac{\log r}{\log \alpha}(8\eta e\alpha^{3/2})\\ & \le \Big(\frac{16\,\eta e^{5/2}}{1+\log 2}\Big)\,\log r. \end{split} \] Thus \[ \limsup_{r\to\infty} \frac{1}{\log r}\int_{E(r)}\frac{dt}{t}\le \frac{16\eta e^{5/2}}{1+\log 2}<1. \] \medskip This completes the proof of the Lemma. This completes the proof of the theorem.\qed \subsection*{Acknowledgment} The author would like to acknowledge Ms. Xudan Luo for her careful reading of the manuscript. \medskip \bibliographystyle{amsplain}
\section{Introduction} \ifnum\sync=1{ \subsection*{SyncTable} Theorem \textcolor{red}{alb\_glue}: \ref{alb_glue}\\ Theorem \textcolor{red}{alberti\_rep\_prod}: \ref{alberti_rep_prod}\\ Lemma \textcolor{red}{lem:meas\_fix}: \ref{lem:meas_fix}\\ Theorem \textcolor{red}{thm:alb\_derivation} \ref{thm:alb_derivation}\\ Lemma \textcolor{red}{partialderivatives} \ref{partialderivatives}\\ Remark \textcolor{red}{rem:derivation\_extension} \ref{rem:derivation_extension}\\ Corollary \textcolor{red}{derbound} \ref{derbound}\\ Theorem \textcolor{red}{onedimapprox\_multi} \ref{onedimapprox_multi}\\ Lemma \textcolor{red}{lem:loc\_estimate\_dist} \ref{lem:loc_estimate_dist}\\ Lemma \textcolor{red}{lem:loc\_estimate\_fnull} \ref{lem:loc_estimate_fnull}\\ Corollary \textcolor{red}{cor:mu\_arb\_cone} \ref{cor:mu_arb_cone}\\ Lemma \textcolor{red}{biLip\_dis} \ref{biLip_dis}\\ Corollary \textcolor{red}{cor:assouad\_bound} \ref{cor:assouad_bound}\\ Theorem \textcolor{red}{thm:weak*density} \ref{thm:weak*density}\\ Lemma \textcolor{red}{borel\_rest} \ref{borel_rest}\\ Theorem \textcolor{red}{compact\_reduction} \ref{compact_reduction}\\ Lemma \textcolor{red}{lem:vector\_alberti} \ref{lem:vector_alberti}\\ }\fi \subsection*{Overview} The aim of this paper is to relate two structures associated with metric measure spaces: Alberti representations and Weaver derivations. The main results establish a correspondence between these objects that respects local invariants, such as the dimension and local norms. These results are then applied to further investigate the structure of differentiability spaces in the sense of Cheeger-Keith \cite{cheeger99,keith04}. Further applications will appear in a sequel to this work. \subsection*{Alberti representations} In this Subsection we give an informal account of Alberti representations. This tool has played an important r\^ole in the proof of the rank-one property of BV functions \cite{alberti_rank_one}, in understanding the structure of measures which satisfy the conclusion of the classical Rademacher's Theorem \cite{acp_proceedings,alberti_marchese}, and in describing the structure of measures in differentiability spaces \cite{bate-diff}. We start with a description in vague terms and refer the reader to \cite{bate-diff} and Subsection \ref{subsec:alberti} for further details. \par An \textbf{Alberti representation} of a Radon measure $\mu$ is a generalized Lebesgue decomposition of $\mu$ in terms of rectifiable measures supported on path fragments; a \textbf{path fragment} in $X$ is a biLipschitz map $\gamma:K\to X$ where $K\subset\real$ is compact with positive Lebesgue measure. Denoting the set of fragments in $X$ by $\frags(X)$,% \nomenclature[frags]{$\frags(X)$}{parametrized biLipchitz curve fragments in $X$}% \ an Alberti representation of $\mu$ takes the form: \begin{equation} \mu=\int_{\frags(X)}\nu_\gamma\,dP(\gamma) \end{equation} where $P$ is a regular Borel probability measure on $\frags(X)$ and $\nu_\gamma\ll\hmeas 1._\gamma$; here $\hmeas 1._\gamma$ denotes the $1$-dimensional Hausdorff measure on the image of $\gamma$. Note that in general it is necessary to work with path fragments instead of Lipschitz paths because the space $X$ on which $\mu$ is defined might lack any rectifiable curve. \begin{exa} A simple example of an Alberti representation is offered by Fubini's Theorem. Let $\hmeas n.$ denote the Lebesgue measure in $\real^n$ and, for $y\in[0,1]^{n-1}$, let \begin{equation} \begin{aligned} \psi(y):[0,1]&\to\frags([0,1]^n)\\ t&\mapsto y+te_{n}, \end{aligned} \end{equation} $e_n$ denoting the last vector in the standard orthonormal basis of $\real^n$. Then $P=\mpush\psi.\hmeas n-1.\mrest[0,1]^{n-1}$ is a regular Borel probability measure on $\frags([0,1]^n)$; if we let $\nu_\gamma=\hmeas 1._\gamma$, by Fubini's Theorem \begin{equation} \hmeas n.\mrest[0,1]^n=\int_{\frags([0,1]^n)}\nu_\gamma\,dP(\gamma) \end{equation} and so $(P,\nu)$ is an Alberti representation of $\hmeas n.\mrest[0,1]^n$. \end{exa} \subsection*{Weaver derivations} In this Subsection we give an informal account of Weaver derivations, hereafter simply called derivations, and refer the reader to \cite{weaver_book99,weaver00} and Subsection \ref{subsec:der_modules} for more details. Note that we need a less general setting than that of Weaver because we deal with Radon measures on separable metric spaces. \par We will denote by $\lipfun X.$% \nomenclature[lipfun]{$\lipfun X.$}{real-valued Lipschitz functions defined on $X$}% \ the space of real-valued Lipschitz functions defined on $X$ and by $\lipalg X.$% \nomenclature[lipfun]{$\lipalg X.$}{bounded real-valued Lipschitz functions defined on $X$}% \ the real algebra of real-valued bounded Lipschitz functions. The algebra $\lipalg X.$ is a dual Banach space\footnote{With a unique predual} and therefore has a well-defined weak* topology: a sequence $f_n\to f$ in the weak* topology if and only if the global Lipschitz constants of the $f_n$ are uniformly bounded and $f_n\to f$ pointwise. \par \textbf{Derivations} can be regarded as \textbf{measurable vector fields} in metric spaces allowing one to take \emph{partial derivatives} of Lipschitz functions once a background measure $\mu$ has been established. More precisely, a \textbf{derivation} is a weak* continuous bounded linear map $D:\lipalg X.\to L^\infty(\mu)$ which satisfies the product rule $D(fg)=fDg+gDf$. The set of derivations forms an $L^\infty(\mu)$-module $\wder\mu.$. \begin{exa} Many examples of derivations can be found in \cite[Sec.~5]{weaver00}. Considering $\real^n$, the partial derivatives $\frac{\partial}{\partial x^i}$ induce derivations with respect to the Lebesgue measure $\hmeas n.$% \nomenclature[measure]{$\hmeas m.$}{$m$-dimensional Hausdorff measure}% \nomenclature[measure]{$\mu\mrest A$}{restriction of the measure $\mu$ on $A$}% \nomenclature[derivation]{$\wder\mu.$}{module of derivations}% \ because of Rademacher's Theorem; this example generalizes to Lipschitz manifolds \cite[Subsec.~5.B]{weaver00}. Similarly, for an $m$-rectifiable set $M\subset\real^n$, $\wder{\hmeas m.}\mrest M.$ can be identified with the module of bounded measurable sections of approximate tangent spaces \cite[Thm.~38]{weaver00}. \end{exa} \subsection*{Matching Alberti representations and Derivations} We now start to describe the results obtained in this paper. Note that we will assume metric spaces to be separable. The first step is to use Alberti representations to construct derivations. We introduce the notion of \textbf{Lipschitz} and \textbf{biLipschitz} Alberti representations: an Alberti representation $\albrep.=(P,\nu)$ is called \textbf{$C$-Lipschitz (resp.~$(C,D)$-biLipschitz)} if $P$ gives full-measure to the set of $C$-Lipschitz (resp.~$(C,D)$-biLipschitz) fragments. In Subsection \ref{subsec:derivations_alberti} we show that (Theorem \ref{thm:alb_derivation}): \begin{itemize} \item To a Lipschitz Alberti representation $\albrep .$ of $\mu$, one can associate a derivation $D_{\albrep.}\in\wder\mu.$% \nomenclature[alberti]{$D_{\albrep.}$}{derivation associated to the Alberti representation $\albrep.$}% \ by using duality and by taking an average of derivatives along fragments \eqref{eq:derivation_alberti}. \item Denoting by $\Alb_{{\rm sub}}(\mu)$% \nomenclature[alberti]{$\Alb_{{\rm sub}}(\mu)$}{family of Lipschitz Alberti representations of measures of the form $\mu\mrest S$}% \ the set of Lipschitz Alberti representations of some measure of the form $\mu\mrest S$ for $S\subset X$ Borel, we obtain a map \begin{equation}\Der:\Alb_{{\rm sub}}(\mu)\to\wder\mu.. \end{equation}% \nomenclature[alberti]{$\Der$}{map associating to an Alberti representation $\albrep.$ a derivation $D_{\albrep.}$}% \end{itemize} \par Note that the previous construction produces a wealth of derivations; in fact, Subsection \ref{subsec:alberti} provides a standard criterion (Theorem \ref{alberti_rep_prod}) which allows to produce Alberti representations which are $(1,1+\epsi)$-biLipschitz: this is an improvement on the treatment in \cite{bate-diff} and will play an important r\^ole in the rest of the paper. We also point out that the choice of $\Alb_{{\rm sub}}(\mu)$ reflects the fact that $\wder\mu.$ depends only on the measure class of $\mu$: i.e.~if $\mu_1\ll\mu_2$ and $\mu_2\ll\mu_1$, $\wder\mu_1.=\wder\mu_2.$. \par We next relate the notion of algebraic independence in $\wder\mu.$ to a notion of independence for Alberti representations introduced in \cite{bate-diff}\footnote{We replace the constant cones in \cite{bate-diff} by Borel cones}: if $f:X\to\real^n$ is Lipschitz and $\cone$% \ is an \textbf{$n$-dimensional cone field}, i.e.~a Borel map on $X$ which takes values in the set of cones in $\real^n$ (see Definition \ref{defn:cone}), an Alberti representation $\albrep.=(P,\nu)$ is in the \textbf{$f$-direction of $\cone$} if, for $P$-a.e.~fragment $\gamma$ and $\lebmeas\mrest\dom\gamma$-a.e.~point $t$, $(f\circ\gamma)'(t)\in\cone(\gamma(t))$; cone fields $\{\cone_i\}_{i=1}^k$ are \textbf{independent} if for each $x\in X$, choosing $v_i\in\cone_i(x)\setminus\{0\}$, the $\{v_i\}_{i=1}^k$ are linearly independent. In Subsection \ref{subsec:derivations_alberti} we show that (Theorem \ref{thm:directional_cone}): \begin{itemize} \item If $\albrep.$ is in the $f$-direction of $\cone$, then $D_{\albrep.}f\in\cone$. \item Alberti representations $\{\albrep i.\}_{i=1}^k$ in the $f$-directions of independent cone fields $\{\cone_i\}_{i=1}^k$ generate independent derivations $\{D_{\albrep i.}\}_{i=1}^k$. \end{itemize} \par Furthermore, our construction relates to a notion of speed for Alberti representations introduced in \cite{bate-diff}\footnote{We allow the speed to be a Borel function}: if $f:X\to\real$ is Lipschitz we say that an Alberti representation $\albrep.=(P,\nu)$ \textbf{has $f$-speed $\ge\delta$} if, for $P$-a.e.~$\gamma$ and $\lebmeas\mrest\dom\gamma$-a.e.~point $t$, $(f\circ\gamma)'(t)\ge\delta\metdiff\gamma(t)$, where $\metdiff\gamma$% \nomenclature[frags]{$\metdiff$}{metric differential}% \ denotes the the metric differential of $\gamma$ (Definition \ref{def:met_diff}). Using an averaging process \eqref{eq:alberti_speed}, we associate an \textbf{effective speed} $\sigma_{\albrep.}$ to ${\albrep.}$ and show that: \begin{itemize} \item If $\albrep.$ has $f$-speed $\ge\delta$, then $D_{\albrep.}f\ge\sigma_{\albrep.}\delta$ (Theorem \ref{thm:directional_speed}). \end{itemize} \par We then address questions related to the injectivity and surjectivity of $\Der$. The map $\Der$ is very far from being injective: as an illustration of this fact, we show that (Lemma \ref{lem:domain_reduction}): \begin{itemize} \item Given a nondegenerate compact interval $I\subset\real$, an Alberti representation $\albrep.$ can be replaced by a new representation $\albrep.'$, whose probability is concentrated on fragments with domain inside $I$, and such that $D_{\albrep.}=D_{\albrep.'}$. \item Properties like the Lipschitz/biLipschitz constant, the speed and the direction are preserved in replacing $\albrep.$ by $\albrep.'$. \end{itemize} \par To study the surjectivity of $\Der$, we use derivations to produce Alberti representations, the intuition being that independent derivations can be used to produce Alberti representations in the directions of independent cone fields. The starting point is the observation that the independence of derivations $\{D_i\}_{i=1}^k\subset\wder\mu\mrest U.$, up to taking a Borel partition of $U$ and linear combinations of the $D_i$, is detected by \textbf{pseudodual} Lipschitz functions $\{g_i\}_{i=1}^k\subset\lipalg X.$ such that $D_ig_j=\delta_{i,j}\chi_U$% \nomenclature[measure]{$\chi_U$}{indicator function of $U$}% \ (by Corollary \ref{cor:pseudoduality}). To deal with Borel partitions, one is led to introduce the \textbf{restriction of $\albrep.=(P,\nu)$ to a Borel set $U$}: $\albrep.\mrest U=(P,\nu\mrest U)$ \cite[pg.~6]{bate-diff}.% \nomenclature[alberti]{$\albrep.\mrest U$}{restriction of $\albrep.$ to $U$}% \ In Subsection \ref{subsec:derivations_to_alberti} we show: \begin{itemize} \item If the $\{D_i\}_{i=1}^k\subset\wder\mu\mrest U.$ have pseudodual Lipschitz functions $\{g_i\}_{i=1}^k\subset\lipalg X.$, letting $g=(g_i)_{i=1}^k$, for any constant $k$-dimensional cone field $\cone$, it is possible to obtain a $(1,1+\epsi)$-biLipschitz Alberti representation of $\mu\mrest U$ in the $g$-direction of $\cone(w,\alpha)$ with \emph{almost optimal} \eqref{eq:speed_almost_optimal} $\langle w,g\rangle$-speed (Theorem \ref{derivation_alberti}). \item If the $\{D_i\}_{i=1}^k\subset\wder\mu\mrest U.$ are independent, passing to a Borel partition $U=\bigcup_\alpha U_\alpha$, there are Lipschitz functions $f_\alpha$ such that, for each $k$-dimensional cone field $\cone$, there is an Alberti representation $\albrep.$ of $\mu$ with $\albrep.\mrest U_\alpha$ in the $f_\alpha$-direction of $\cone$ (Corollary \ref{der-alb}). \item If $f:X\to\real^k$ and $\mu$ admits an Alberti representation in the $f$-direction of $k$ independent cone fields, then for each $k$-dimensional cone field $\cone$, $\mu$ admits an Alberti representation in the $f$-direction of $\cone$ (Corollary \ref{cor:mu_arb_cone}). \end{itemize} Corollary \ref{cor:mu_arb_cone} is saying that there cannot be \emph{gaps} in the directions accessible by curve fragments. This result was obtained, for differentiability spaces and constant cone fields, in \cite[Sec.~9]{bate-diff} using porosity techniques. The method employed here is more general and employs some ``soft'' Functional Analysis. \par In Subsection \ref{subsec:derivations_to_alberti} we finally show \begin{itemize} \item If $\wder\mu.$ is finitely generated, then $\Der$ is surjective (Theorem \ref{thm:fin_gen_surjectivity}). \item In the general case, $\Der(\Alb_{{\rm sub}}(\mu))$ is weak* dense in $\wder\mu.$ (Theorem \ref{thm:weak*density}). \end{itemize} \par The \textbf{dual module} $\wform\mu.$% \nomenclature[derivations]{$\wform\mu.$}{module of forms}% \ of $\wder\mu.$ can be regarded as the $L^\infty(\mu)$-module of differential forms because each $f\in\lipalg X.$ gives rise to a form $df\in\wform\mu.$. The modules $\wder\mu.$ and $\wform\mu.$ admit \textbf{local norms} $\locnorm\,\cdot\,,{\wder\mu.}.$ and $\locnorm\,\cdot\,,{\wform\mu.}.$,% \nomenclature[derivations]{$\locnorm\,\cdot\,,{\wder\mu.}.$}{local norm on $\wder\mu.$}% \nomenclature[derivations2]{$\locnorm\,\cdot\,,{\wform\mu.}.$}{local norm on $\wform\mu.$}% \ which can be thought of as families $\{\|\cdot\|_x\}_{x\in X}$ of pointwise norms from which one can reconstruct the global norms by taking the essential supremum. In Subsection \ref{subsec:weaver_norm} we obtain a geometric characterization of $\locnorm\,\cdot\,,{\wform\mu.}.$, which plays a central r\^ole in our characterization of differentiability spaces: \begin{itemize} \item For $U$ Borel, $f\in\lipalg X.$ and $\alpha>0$, if $\locnorm df,{\wform\mu\mrest U.}.\approx\alpha$, then there is an Alberti representation $\albrep.=(P,\nu)$ of $\mu\mrest U$ with $P$ concentrated on the fragments where $(f\circ\gamma)'\approx \alpha\metdiff\gamma$ (this is an informal restatement of Theorem \ref{thm:weaver_fnorm_char}). \end{itemize} \par This research was motivated by the work of Bate \cite{bate-diff} on differentiability spaces: Bate obtained an intrinsic characterization of differentiability spaces in terms of Alberti representations which implies that these spaces have a rich structure of curve fragments. We started our investigation with the vague question of what happens in spaces which, despite not being differentiability spaces, still have a rich curve structure. In particular, our results apply to quite a general class of metric measure spaces; examples are: \begin{itemize} \item Spaces $(X_{\rm lack},\mu_{\rm lack})$ which either lack any rectifiable curve or where $\mu_{\rm lack}$ does not admit any Alberti representation: in this case $\wder\mu_{\rm lack}.=\{0\}$. \item Spaces which are $k$-rectifiable $(X_{\text{$k$-rect}},\hmeas k.)$. \item Products $(X_{\rm lack}\times X_{\text{$k$-rect}}, \mu_{\rm lack}\times \hmeas k.)$ and quotients of such products, for example Laakso spaces and the non-doubling Laakso spaces of \cite{weaver00} and \cite{bate_speight}. \item Differentiability spaces. \item Carnot groups equipped with a Radon measure $\mu$: in this case $\wder\mu.$ is always finitely generated and the number of generators is at most the dimension of the horizontal distribution. \item Spaces which have rectifiable fragments in infinitely many directions, for example the Hilbert cubes in $l^p$ and $c_0$ considered in \cite{weaver00}. \end{itemize} \subsection*{Structure of differentiability spaces} We now describe the implications of the present work for the theory of differentiability spaces, first recalling the current knowledge about these spaces. For more details, we refer to Subsection \ref{subsec:diff_spaces}. \par We first recall notions of \textbf{finite dimensionality} for measures on metric spaces: \begin{itemize} \item A measure $\mu$ on $X$ \textbf{is doubling (with constant $C$)} if, for all pairs $(x,r)\in X\times (0,\diam X]$, \begin{equation}\label{eq:meas_doubling} \mu\left(\ball x,r/2.\right)\ge C\mu\left(\ball x,r.\right). \end{equation} \item If \eqref{eq:meas_doubling} holds for $\mu$-a.e.~$x$ for $r\in(0,r(x)]$, the measure $\mu$ is called \textbf{asymptotically doubling (with constant $C$)}. \item If there are disjoint Borel sets $X_\alpha$ with $\mu(X\setminus \bigcup_\alpha X_\alpha)=0$ and the measure $\mu\mrest X_\alpha$ is doubling on $X_\alpha$, $\mu$ is called \textbf{$\sigma$-asymptotically doubling}. \end{itemize} \par We now recall the definitions of \textbf{infinitesimal Lipschitz constants} and differentiability for Lipschitz functions. For a real-valued Lipschitz function $f$, the variation of $f$ at $x$ at scale $r$ is $\sup_{y\in\ball x,r.}|f(x)-f(y)|/{r}$; the lower and upper limits of the variation as $r\searrow0$ are denoted by $\smllip f(x)$ and $\biglip f(x)$\footnote{Sometimes called the ``small Lip'' lip and the ``big Lip'' Lip in the literature. Other terms used are pointwise lower and upper Lipschitz constants.}.% \nomenclature[loclip]{$\smllip f(x)$}{pointwise lower Lipschitz constant}% \nomenclature[loclip]{$\biglip f(x)$}{pointwise upper Lipschitz constant}% \ Following Cheeger \cite{cheeger99}, given $f\in\lipfun X.$ and $g:X\to\real^k$ Lipschitz, we say that $f$ \textbf{is differentiable at $x$ with respect to $g$ with derivative $(a_i)_{i=1}^k\in\real^k$}, if \begin{equation} \biglip\left(f-\sum_{i=1}^ka_ig_i\right)(x)=0. \end{equation} \par A differentiability space $(X,\mu)$ is a metric measure space for which there are an integer $N$ and countably many charts \begin{equation}\label{eq:chartsdef}\{(U_\alpha,\psi_\alpha)\}_\alpha\subset\left\{\text{Borel subsets of X}\right\}\times\left\{\text{Lipschitz maps $f:X\to\real^Q$, ($Q\le N$)}\right\} \end{equation} such that each $f\in\lipfun X.$ is $\mu$-a.e.~differentiable, with a unique derivative, on $U_\alpha$ with respect to the $\psi_\alpha$; the number of components $N_\alpha$ of $\psi_\alpha$ is called the dimension of the chart $(U_\alpha,\psi_\alpha)$ and the $l$-th component of the derivative, the partial derivative, is denoted by $\partial f/\partial\psi_\alpha^l$. We will \emph{require} in the definition of the charts that there is a constant $C_\alpha$ such that, for each $l$, \begin{equation} \label{eq:regchart} \|\partial f/\partial\psi_\alpha^l\|_{L^\infty(\mu\mrest U_\alpha)}\le C_\alpha\glip f., \end{equation} where $\glip f.$ denotes the global Lipschitz constant of $f$;% \nomenclature[globlip]{$\glip f.$}{global Lipschitz constant}% \ in this case one obtains linear bounded operators $\partial/\partial\psi_\alpha^l:\lipfun X.\to L^\infty(\mu\mrest U_\alpha)$ which are called \textbf{partial derivative operators}. The requirement \eqref{eq:regchart} is not assumed by all authors\footnote{for example, it is used in \cite{cheeger99} but not in \cite{keith04}} in the definition of a differentiability space; however, it can always be satisfied by partitioning the charts. For example, considering the differentiability space $(\real,\lebmeas)$ and following our terminology, one would not consider $((1,\infty),1/(|x|+1))$ a chart; however, on writing $(1,\infty)=\bigcup_{n=1}^\infty(n,n+1]$, one would consider the $\{((n,n+1],1/(|x|+1))\}_n$ charts. The smallest value of $N$ in \eqref{eq:chartsdef} is called the \textbf{differentiability dimension} of $(X,\mu)$. \par The notions of a differentiability space and of measurable charts that we have presented have originated in Cheeger's work \cite{cheeger99}. Later Keith \cite{keith04} clarified some aspects of this construction: a nice exposition can be found in \cite{kleiner_mackay}. \par We will now present two sufficient conditions for a metric measure space to be a differentiability space. A $(C,\tau, p)$-PI space is a doubling metric measure space with constant $C$ supporting a $p$-Poincar\'e inequality (\cite{heinonen_analysis}) with constant $\tau$. In \cite{cheeger99} Cheeger showed that: \begin{thm}\label{thm:cheeger} If $(X,\mu)$ is a $(C,\tau,p)$-PI-space, then it is a differentiability space with differentiability dimension $\le N(C,\tau)$. Moreover, for each Lipschitz function $f$, the equality $\smllip f = \biglip f$ holds $\mu$-a.e. \end{thm} In \cite{keith04} Keith was able to relax the assumptions of Cheeger: \begin{thm}\label{thm:keith} If $(X,\mu)$ is a metric measure space with a $C$-doubling measure $\mu$ and there is a constant $\tau>0$ such that, for each real-valued Lipschitz function $f$, \begin{equation}\label{eq_Lip_lip_ineq} \tau\smllip f(x)\ge\biglip f(x)\quad\text{\normalfont(for $\mu$-a.e.~$x$),} \end{equation} then $(X,\mu)$ is a differentiability space with differentiability dimension $\le N(C,\tau)$. \end{thm} The inequality \eqref{eq_Lip_lip_ineq} is sometimes called the \textbf{Lip-lip inequality}. \par Theorems \ref{thm:cheeger} and \ref{thm:keith} can be proved under the relaxed assumption that $\mu$ is asymptotically doubling. Moreover, in \cite{bate_speight} Bate and Speight showed: \begin{thm}\label{thm:bate_speight} If $(X,\mu)$ is a differentiability space, then $\mu$ is $\sigma$-asymptotically doubling. \end{thm} \begin{exa} Note that one cannot conclude that $\mu$ is asymptotically doubling, i.e.~that the local doubling constant is uniformly bounded. For example, consider a compact metric measure space with $\mu$ finite which is obtained by gluing together, along some geodesics, countably many (rescalings of) Laakso spaces with Hausdorff dimensions tending to $\infty$; this construction produces a differentiability space but the constant $C$ in \eqref{eq:meas_doubling} is not uniformly bounded. \end{exa} \par To deal with the case in which there is only a local bound on the differentiability dimension we introduce the following terminology: a \textbf{$\sigma$-differentiability space} $(X,\mu)$ is a complete separable metric measure space, such that, there are disjoint Borel sets $X_\alpha$ with $\mu(X\setminus \bigcup_\alpha X_\alpha)=0$ and such that $(X_\alpha, \mu\mrest X_\alpha)$ is a differentiability space. In \cite{bate-diff} Bate obtained the following characterization of $\sigma$-differentiability spaces: \begin{thm}\label{thm:bate-diff_char} A metric measure space $(X,\mu)$ is a $\sigma$-differentiability space if and only if: \begin{enumerate} \item The measure $\mu$ is $\sigma$-aymptotically doubling. \item There is a Borel map $\tau:X\to(0,\infty)$ such that, for each real-valued Lipschitz function $f$, the measure $\mu$ admits an Alberti representation with $f$-speed $\ge\biglip f/\tau$. \end{enumerate} \end{thm} Note that Theorem \ref{thm:bate-diff_char} implies that a weaker form of the Lip-lip inequality, where $\tau$ is allowed to be a Borel function, must hold in a $\sigma$-differentiability space. We significantly strengthen this result showing that, in a $\sigma$-differentiability space, the equality $\biglip f=\smllip f$ holds $\mu$-a.e. Our proof relies on the geometric characterization of $\locnorm\,\cdot\,,{\wform\mu.}.$; another proof will appear in \cite{cks_metric_diff}. \par The work \cite{bate-diff} is a partial extension to metric measure spaces of a structure theory for null sets in Euclidean spaces developed by Alberti-Cs\"ornyei-Preiss (we refer the reader to the expository papers \cite{acp_plane,acp_proceedings}; for the structure of measures we refer the reader to \cite{alberti_marchese}). In particular, Alberti-Cs\"ornyei-Preiss found a collection of sets which must be annihilated by measures satisfying the conclusion of the classical Rademacher's Theorem. \par Our characterization of $\sigma$-differentiability spaces uses a similar idea: we introduce the collection $\Gap(X)$ of Borel subsets of $X$: $S\in\Gap(X)$ if there are a Lipschitz function $f$ and constants $\alpha>\beta\ge0$, such that, for each Radon measure $\mu'$, \begin{equation} \locnorm df,{\wform\mu'.}.\le\beta<\alpha\le\biglip f\quad{\text{($\mu'$-a.e.~on $S$).}} \end{equation} In Subsection \ref{subsec:char_diff} we show that $(X,\mu)$ is a $\sigma$-differentiability space if and only if any of the following equivalent conditions holds: \begin{itemize} \item Every $S\in\Gap(X)$ is $\mu$-null (Theorem \ref{thm:diff_char1}). \item The measure $\mu$ is $\sigma$-asymptotically doubling and, for each real-valued Lipschitz function $f$, the following equalities hold $\mu$-a.e.: \begin{equation}\label{eq:intro_quant} \locnorm df,{\wform\mu.}.=\biglip f=\smllip f\quad\text{(Theorem \ref{thm:diff_char2})}. \end{equation} \end{itemize} \par There is also a connection between differentiability and derivations. We show that: \begin{thm}\label{thm:schioppa} A metric measure space $(X,\mu)$ is a differentiability space with dimension $\le N$ if and only if $\mu$ is $\sigma$-asymptotically doubling and there are a conformal factor $\lambda\in L^\infty(\mu)$ and derivations $\{D_i\}_{i=1}^N\subset{\rm Der}(\mu)$ such that, for each $f\in\lipalg X.$, \begin{equation}\label{revder} \lambda(x)\max_i|D_if(x)|\ge\biglip f(x)\quad\text{for $\mu$-a.e.~$x$.} \end{equation} \end{thm} In \cite{derivdiff}, motivated by \cite{gong11-revised}, the author showed that \eqref{revder} gives sufficient conditions for the existence of a differentiable structure\footnote{In \cite{derivdiff} the measure $\mu$ was assumed doubling. However, taking a Borel partition, it suffices to assume that $\mu$ is $\sigma$-asymptotically doubling.}; in \cite{derivdiff} the author also proved a partial converse: if $(X,\mu)$ is a differentiability space with $\mu$ doubling and if the partial derivative operators are derivations, then \eqref{revder} holds. Thus Theorem \ref{thm:schioppa} follows from \cite{derivdiff} because we provide two different proofs that the partial derivative operators are derivations: the first proof uses \cite{bate-diff} and can be found in Subsection \ref{subsec:derivation_differentiability}; the second proofs uses \eqref{eq:intro_quant} and can be found at the end of Subsection \ref{subsec:char_diff}. We include both proofs to illustrate the difference between Bate's characterization of $\sigma$-differentiability spaces and our quantitative description. In \cite{bate-diff} the existence of Alberti representations in independent directions is the \emph{starting point} of a characterization of $\sigma$-differentiability spaces. On the other hand, we obtain independent Alberti representations as a \emph{consequence} of the quantitative relation \eqref{eq:intro_quant}. \par An important consequence of Theorem \ref{thm:schioppa} is that in Theorems \ref{thm:cheeger} and \ref{thm:keith} one can replace the bound $N(C,\tau)$ by the Assouad dimension, removing the dependence on $\tau$ (Corollary \ref{cor:assouad_bound}). \subsection*{Technical tools}\label{subsec:technical-tools} In this Subsection we give an overview of four technical tools used in this paper. \par The first tool is an approximation scheme for Lipschitz functions in the weak* topology. The intuition is that if a set $S$ is $\frags(X,f,\delta)$-null (Definitions \ref{def:frag_nullity} and \ref{def:classes_frags}), i.e.~does not contain fragments where $(f\circ\gamma)'(t)\ge\delta\metdiff\gamma(t)$, then $f\in\lipalg X.$ can be approximated by Lipschitz functions which have Lipschitz constant at most $\delta$ in sufficiently small balls centred on $S'$, where $S'\subset S$ has full measure in $S$ . We prove an approximation scheme, Theorem \ref{onedimapprox_multi}, which takes into account also the direction of the fragments. We state here a simplified version which is sufficient for the results on differentiability spaces. \begin{thm}\label{onedimapprox} Let $X$ be a compact metric space, $f:X\to\real$ $L$-Lipschitz and $S\subset X$ compact. Let $\mu$ be a Radon measure on $X$. Assume that $S$ is $\frags(X,f,\delta)$-null. Then there are $\max(L,\delta)$-Lipschitz functions $g_k\xrightarrow{\text{w*}} f$ with $g_k$ $\mu$-a.e.~locally $\delta$-Lipschitz on $S$. \end{thm} \par The motivation to prove Theorem \ref{onedimapprox_multi} came from reading \cite[Subsec.~6.1]{bate-diff}: the author observed that Bate's construction can be used to produce an approximation scheme for Lipschitz functions with $\biglip f=0$ (\emph{flat}) on $S$. In the author's opinion, \cite[Subsec.~6.1]{bate-diff} is a metric space version of a construction sketched in \cite[Defn.~1.14]{acp_proceedings}. However, the approximation scheme based on \cite[Subsec.~6.1]{bate-diff} could be used only to prove part of the results presented here: the major obstacle is that the approximation works only for functions which are flat on $S$. The problem stems from the presence of a potential term $\delta\hmeas 1.$ in the $Q$-functional introduced by Bate. However, the discussion in \cite[Sec.~3]{acp_proceedings} provides evidence suggesting that it is possible to produce a finer approximation scheme. The starting point of \cite[Sec.~3]{acp_proceedings} is \cite[Thm.~2.4]{acp_proceedings}; the author found a proof of that result written in A.~Marchese's PhD~thesis\footnote{To appear in \cite{alberti_marchese}} and, starting from it, elaborated Theorem \ref{onedimapprox_multi}. \par The passage from Euclidean spaces to general metric spaces uses an \emph{abstract nonsense construction} of a cylinder which is a geodesic metric space containing the graph of the function to be approximated and extra degrees of freedom to deform it. The starting point, as in \cite[Subsec.~6.1]{bate-diff}, is a Kuratowski embedding in $l^\infty$. \par The second technical tool is a decomposition of $\wder\mu.$ into free modules. The problem stems from the fact that $L^\infty(\mu)$ is not an integral domain and thus the notion of linear independence of derivations behaves quite differently than in a vector space. In \cite{derivdiff} the author introduced the following concept of finite dimensionality: \begin{defn} The module $\wder\mu\mrest U.$ is said to have \textbf{index $\le N$} if any linearly independent (over $L^\infty(\mu\mrest U)$) set of derivations in it has cardinality at most $N$. If, for any $\mu$-measurable $U$, the module $\wder\mu\mrest U.$ has index $\le N$, we say that $\wder\mu.$ has \textbf{index locally bounded by $N$}. \end{defn} \par Under this assumption the author \cite{derivdiff} obtained the following decomposition result. \begin{thm}\label{thm:free_dec} Suppose that $\wder\mu.$ has index locally bounded by $N$. Then there is a Borel partition $X=\bigcup_{i=0}^N X_i$ such that, if $\mu(X_i)>0$, then $\wder\mu\mrest X_i.$ is free of rank $i$. A basis of $\wder\mu\mrest X_i.$ will be called a \textbf{local basis of derivations}. \end{thm} \par Theorem \ref{thm:free_dec} should compared with \cite[Thm.~10]{weaver00}; a posteriori the two approaches turn out to be equivalent; however, in the author's opinion, a local dimensional bound and a decomposition into free modules are more natural in the study of differentiability spaces. \par The third technical tool is to relate Alberti representations and \textbf{blow-ups} of metric spaces. A \textbf{blow-up\footnote{Sometimes called a tangent space / cone} of a metric space $X$ at a point $p$} is a (complete) pointed metric space $(Y,q)$ which is a pointed Gromov-Hausdorff limit of a sequence $(\frac{1}{t_n}X,p)$ where $t_n\searrow0$: the notation $\frac{1}{t_n}X$ means that the metric on $X$ is rescaled by $1/t_n$; the class of blow-ups of $X$ at $p$ is denoted by $\tang(X,p)$. A \textbf{blow-up of a Lipschitz function $f:X\to\real^Q$ at a point $p$} is is a triple $(Y,q,g)$ with $(\frac{1}{t_n}X,p)\to(Y,q)\in\tang(X,p)$, $g:X\to\real^Q$ Lipschitz and such that the rescalings $(f-f(p))/t_n:\frac{1}{t_n}X\to\real^Q$ converge to $g$; the class of blow-ups of $f$ at $p$ is denoted by $\tang(X,p,f)$. The general existence of blow-ups requires the notion of \textbf{ultralimits}: we simplified the treatment assuming that $X$ has \textbf{finite Assouad dimension} (\cite{tyson_mackay_conf}) because this assumption is not restrictive in the theory of differentiability spaces. \par In Subsection \ref{subsec:dimens-bounds-tang} we show that: \begin{itemize} \item If $f:X\to\real^N$ and $\mu$ admits Alberti representations in the $f$-directions of independent cone fields, for $\mu$-a.e.~$p$ all blow-ups $(Y,q,g)\in\tang(X,p,f)$ are such that $g:Y\to\real^N$ is surjective (Theorem \ref{alberti_blow_up}). \item If $X$ has Assouad dimension $D$, then $\wder\mu.$ has index locally bounded by $D$ (Corollary \ref{derbound}). \end{itemize} \par Note that Corollary \ref{derbound} improves \cite[Lem.~1.10]{gong11-revised} by giving an explicit bound \emph{equal} to the Assouad dimension. \par The motivation to prove Theorem \ref{alberti_blow_up} came from a constructon of separated nets in \cite[Sec.~8]{bate-diff} and the work of Keith \cite{keith04} which uses blow-ups to study spaces satisfying a Lip-lip inequality. \par The fourth tool is a construction of independent Lipschitz functions, Theorem \ref{lip_ind}. Lipschitz functions $\{\psi_1,\ldots,\psi_n\}$ are \textbf{independent} on a set $S$ if, for each $x\in S$, the map $\real^n\ni (a_i)\mapsto\biglip(\sum_{i=1}^na_i\psi_i)(x)$ is a norm. The property of being a differentiability space can be reformulated as a \textbf{finite dimensionality statement}: there is a uniform upper bound on the number of Lipschitz functions which are independent on a positive measure set. \par In the case of Euclidean spaces, instead of constructing independent functions on a set $S$, it is more natural to construct Lipschitz functions which are not differentiable on $S$. In the case of $\real$, there is a classical construction of Zahorski (\cite{zahorski_line}); for $\real^n$, a generalized construction is announced in \cite[Thm.~1.15]{acp_proceedings}; in the case in which one relaxes the conclusion to nondifferentiability $\mu$-a.e.~on $S$\footnote{``Nondifferentiability for measures''}, the construction is simplified as it can handled independently on different parts of $S$ using a truncation principle (Lemma \ref{lip_trunc}). Recently, Alberti-Marchese \cite{alberti_marchese} strengthnened this approach showing that the set of Lipschitz functions which are nondifferentiable on a large part of $S$ is \textbf{comeagre} in a suitable metric space of Lipschitz functions. \par In \cite[Sec.~4]{bate-diff} Bate produces a construction of nondifferentiable Lipschitz functions for measures on metric spaces; the approach is similar to the Euclidean case but requires the tool of \textbf{structured charts} introduced in \cite{bate_speight}. Theorem \ref{lip_ind} is a reformulation of this result in the language of independent functions; the author thinks this is useful because: [1] it is technically simpler avoiding a discussion of structured charts; [2] it fits more naturally with the approaches of Cheeger and Keith. The author wonders whether it is possible to do a construction of independent functions \emph{for sets} in metric spaces. \subsection*{Acknowledgements} The author wants to thank his advsisor, prof.~Kleiner, for reading this work and providing many stimulating questions. In particular, some of these questions motivated the author to prove Theorem \ref{thm:weak*density}. \renewcommand{\nomname}{List of Symbols} \printnomenclature[2cm] \section{Preliminaries}\label{sec:prelim} \subsection{Alberti representations}\label{subsec:alberti} The goal of this Subsection is to define Alberti representations precisely and prove Theorem \ref{alberti_rep_prod}, which can be regarded as a standard criterion to produce Alberti representations. The treatment has been a expanded a bit to address what seem to be a couple of little gaps in \cite{bate-diff}: \begin{enumerate} \item In \cite[Lem.~5.2]{bate-diff} Alberti representations are produced with $P$ a probability measure on $1$-rectifiable measures, instead of fragments (point addressed in Lemma \ref{biLip_dis}). \item In \cite[Sec.~6]{bate-diff} Alberti representations are produced with $P$ defined on $\frags(\bana)$ where $\bana$ is a Banach space containing $X$ (point addressed in Theorem \ref{compact_reduction}). \end{enumerate} We start by defining paths and fragments. Throughout this Section $X$ will denote a Polish metric space. \begin{defn} Given a metric space $Y$, let $\haus(Y)$ denote the topological space of compact sets with the Vietoris topology, which is induced by the Hausdorff distance. If $Y$ is compact, $\haus(Y)$ is compact and if $Y$ is Polish, $\haus(Y)$ is Polish. We introduce the set of path fragments: \begin{equation} \frags(X)=\left\{\gamma:K\to X:\text{$\gamma$ biLipschitz, $K\subset\real$ compact, $\lebmeas(K)>0$}\right\}; \end{equation} which is identified with a subspace of $\haus(\real\times X)$ via the map $\gamma\mapsto\graph\gamma$. Given a nondegenerate compact interval $I\subset\real$, we denote by $\frags(X,I)$ the subset of fragments $\gamma$ with $\dom\gamma\subset I$. \end{defn} \par In order to define Alberti representations precisely, we need to recall some facts from measure theory. Let $Z$ denote a locally compact metric space, $M(Z)$% \nomenclature[measure]{$M(Z)$}{finite Radon measures on $Z$}% \nomenclature[measure]{$P(Z)$}{finite Borel probability measures on $Z$}% \ the Banach space of finite (signed) Borel measures on $Z$ and $P(Z)\subset M(Z)$ the subset of probability measures. It might be useful to recall that finite Borel measures on metric spaces are regular and that a finite Borel measure on a Polish space is Radon \cite[Thm.~7.1.7]{bogachev_measure}. The Banach space $M(Z)$ is also a dual Banach space; in the definition of $M(Z)$-valued Borel maps we will consider on $M(Z)$ the weak* topology. In particular, given a metric space $Y$, a map $\psi:Y\to M(Z)$, $\psi$ is Borel if and only if for each $g\in C_c(Z)$\footnote{Real-valued continuous functions with compact support} the map $y\mapsto\int_Z g\,d\psi(y)$ is Borel (compare \cite[Rem.~1.1]{alberti_rank_one}). \par We will use the following result (\cite[Defn.~1.2]{alberti_rank_one}): \begin{lem}\label{integration_meas} Let $Y$ be a separable locally compact topological space, and $\lambda$ a locally finite Borel measure on $Y$. Let $\psi:Y\to M(Z)$ be Borel and, denoting by $\|\psi(y)\|_{M(Z)}$ the norm of $\psi(y)$, assume that \begin{equation} \int_Y\|\psi(y)\|_{M(Z)}\,d\lambda(y)<\infty; \end{equation} then the integral \begin{equation} \mu=\int_Y\psi(y)\,d\lambda(y) \end{equation} exists and defines an element of $M(Z)$. More precisely, for a Borel set $A\subset Z$, \begin{equation} \mu(A)=\int_Y\psi(y)(A)\,d\lambda(y). \end{equation} \end{lem} We can now state precisely the definition of an Alberti representation; note that condition (4) has been added for technical convenience (compare the proof of Lemma \ref{lem:meas_fix}). \begin{defn}\label{defn:Alberti_rep} Let $\mu$ be a Radon measure on a metric space $X$; an Alberti representation of $\mu$ is a pair $(P,\nu)$: \begin{enumerate} \item The measure $P$ is in $P(\frags(X))$. \item The map $\nu:\frags(X)\to M(X)$ is Borel and $\nu_\gamma\ll\hmeas 1._\gamma$, the one-dimensional Hausdorff measure associated to the image of $\gamma$. \item The measure $\mu$ can be represented as $\mu=\int_{\frags(X)}\nu_\gamma\,dP(\gamma)$. \item For each Borel set $A\subset X$ and for all real numbers $a\ge b$, the map $\gamma\mapsto\nu_\gamma\left(A\cap\gamma(\dom\gamma\cap[a,b])\right)$ is Borel. \end{enumerate} \end{defn} \par We now define precisely the notions of Euclidean cones and metric differential employed in the Introduction to introduce the notions of direction and speed. \begin{defn}\label{defn:cone} Let $\alpha\in(0,\pi/2)$, $w\in{\mathbb S}^{q-1}$; the \textbf{open cone} $\cone(w,\alpha)\subset\real^q$ with axis $w$ and opening angle $\alpha$ is: \begin{equation} \cone(w,\alpha)=\{u\in\real^q: \tan\alpha\langle w,u\rangle>\|\pi_w^\perp u\|_2\}. \end{equation}% \nomenclature[frags]{$\cone(w,\alpha)$}{cone / cone field of direction $w$ and angle $\alpha$}% The set of open cones is given the topology of ${\mathbb S}^{q-1}\times(0,\pi/2)$. \end{defn} \begin{rem}Note that if $u\in\cone(w,\alpha)\cap{\mathbb S}^{q-1}$, \begin{equation} \|u-w\|_2\le(1-\cos\alpha)+\sin\alpha. \end{equation} \end{rem} \par We recall the definition of \textbf{metric differential}, which is an adaptation of \cite[Defn.~4.1.2]{ambrosio_metric} (compare \cite[Sec.~3]{ambrosio-rectifiability}): \begin{defn}\label{def:met_diff} Given $\gamma\in\frags(X)$, the metric differential $\metdiff\gamma(t)$ of $\gamma$ at $t\in\dom\gamma$ is the limit \begin{equation} \lim_{\dom\gamma\ni t'\to t}\frac{\dist\gamma(t'),\gamma(t).}{|t'-t|} \end{equation} whenever it exists\footnote{We convene that if $t$ is an isolated point of $\dom\gamma$, the limit does not exist}. \end{defn} \par We will abbreviate a set of conditions on the Lipschitz/biLipschitz constant, speed and direction by \albcond. \par If $X$ is a metric space and $C\subset X$ with $\mu$ a Radon measure on $C$, there are a priori two different notions of Alberti representations $(P,\nu)$ depending on whether $P\in P(\frags(X))$ or $P\in P(\frags(C))$; in the former case we will say that the Alberti representation is defined on $\frags(X)$. We will prove Theorem \ref{compact_reduction} which produces an Alberti representation defined on $\frags(C)$ given an Alberti representation defined on $\frags(X)$ and preserves a set \albcond. \begin{thm}\label{compact_reduction} Let $C$ be compact subset of $X$. Suppose that the Radon measure $\mu$, with support contained in $C$, admits an Alberti representation $(P,\nu)$ defined on $\frags(X)$ and satisfying \albcond. Then $\mu$ admits an Alberti representation $(P',\nu')$ defined on $\frags(C)$ and satisfying \albcond. \end{thm} \par The proof of Theorem \ref{compact_reduction} requires some preparation. We start introducing some subsets of fragments. \begin{defn} Let $C\subset X$ and $n\in\natural$. We define: \begin{align} \frags(X,C)&=\left\{\gamma\in\frags(X): \text{$\gamma^{-1}(C)$ has positive Lebesgue measure}\right\}.\\ \begin{split} \frags_n(X)&=\bigl\{\gamma\in\frags(X):\text{$\dom\gamma\subset[-n,n]$, $\gamma$ is $\left(\frac{1}{n},n\right)$-biLipschitz,}\\ &\phantom{=\bigl\{\gamma\in\frags(X):\text{$\dom\gamma\subset[-n,n]$}} \text{and $\lebmeas(\dom\gamma)\ge\frac{1}{n}$}\bigr\}; \end{split}\\ \begin{split} \frags_n(X,C)&=\bigl\{\gamma\in\frags_n(X): \text{ $\lebmeas(\gamma^{-1}(C))\ge\frac{1}{n}$}\bigr\}. \end{split} \end{align} \end{defn} \begin{lem}\label{lem:frag_compact} The sets $\frags_n(X)$ and $\frags_n(X,C)$ are closed in $\haus(\real\times X)$. If $C\subset X$ is compact, then $\frags_n(C)$ is compact. \end{lem} \begin{proof} The subset of $\haus(\real\times X)$ consisting of graphs of $\left(\frac{1}{n},n\right)$-biLipschitz maps $\gamma:K\subset\real\to X$ is closed. Also the set of those compact sets $K\subset\real\times X$, whose projection $\pi_\real(K)$ on $\real$ satisfies $\pi_\real(K)\subset[-n,n]$, is closed. Consider a sequence of compact sets $\{K_k\}\subset\haus(\real)$ and assume that for each $k$ one has $\lebmeas(K_k)\ge\frac{1}{n}$; then, if the $K_k$ converge to $K$, we have $\lebmeas(K)\ge\frac{1}{n}$. We thus conclude that the set $\frags_n(X)$ is closed in $\haus(\real\times X)$. The compactness of $\frags_n(C)$ follows from Ascoli-Arzel\`a. We now show that $\frags_n(X,C)$ is a closed subset of $\frags_n(X)$; suppose that the sequence of fragments $\{\gamma_k\}\subset\frags_n(X,C)$ converges to $\gamma$; by passing to a subsequence we can also assume that the compact sets $\gamma_k^{-1}(C)$ converge to a compact set $K\subset\dom\gamma$. We then have $\lebmeas(K)\ge\frac{1}{n}$ and $\gamma(K)\subset C$, showing that $\frags_n(X,C)$ is closed. \end{proof} An immediate consequence of Lemma \ref{lem:frag_compact} is: \begin{lem}\label{frag_meas} The sets $\frags(X)$ and $\frags(X,C)$ are of class $F_\sigma$ in $\haus(\real\times X)$. If $C\subset X$ is compact, then $\frags(C)$ is of class $K_\sigma$ in $\haus(\real\times X)$. \end{lem} \par The key tool to prove Theorem \ref{compact_reduction} is the following Lemma. \begin{lem}\label{borel_rest} If $X$ is a complete separable metric space and if $C\subset X$ is compact, the restriction map \begin{equation} \begin{aligned} \frest C.:\frags(X,C)&\to\frags(C)\\ \gamma&\mapsto\gamma|\gamma^{-1}(C), \end{aligned} \end{equation} is Borel. \end{lem} \begin{proof} Note that $\frest C.$ maps $\frags_n(X,C)$ to $\frags_n(C)$ so by Lemma \ref{frag_meas} it suffices to show that the restriction $\frest C,n.=\frest C.|\frags_n(X,C)$ is Borel. By \cite[Lem.~6.2.5]{bogachev_measure} it suffices to show that for each $\psi\in C\left(\frags_n(C)\right)$\footnote{Set of real-valued continuous functions}, the map $\psi\circ\frest C,n.$ is Borel. \par If $(t,x)\in\real\times C$, let $d_{(t,x)}(K)$ denote the distance from the set $K$ to the point $(t,x)$. Note that $d_{(t,x)}$ is Lipschitz: for each $\epsi>0$ if $(t_1,x_1)\in K$ is a closest point to $(x,t)$, there is a point $(t_2,x_2)\in K'$ with \begin{equation} \dist {(t_1,x_1)},{(t_2,x_2)}.\le\hdist K,K'.+\epsi, \end{equation} which implies \begin{equation} d_{(t,x)}(K')\le d_{(t,x)}(K)+\hdist K,K'.. \end{equation} Note that these functions separate points in $\haus(\real\times X)$. As $\frags_n(C)$ is compact, by the Stone-Weierstra\ss\ Theorem \cite[Thm.~7.32]{rudin-principles} the unital subalgebra of $C\left(\frags_n(C)\right)$ generated by the functions $\{d_{(t,x)}\}$ is dense. In particular, any $\psi\in C\left(\frags_n(C)\right)$ is a uniform limit of polynomials in the $\{d_{(t,x)}\}$. Therefore, it suffices to show that $d_{(t,x)}\circ\frest C,n.$ is Borel. \par We show that $d_{(t,x)}\circ\frest C,n.$ is lower semicontinuous: assume that $\gamma_m\to\gamma$ in $\frags_n(X,C)$ and that along a subsequence $m_k$ one has \begin{equation} d_{(t,x)}(\frest C,n.(\gamma_{m_k}))\le\epsi. \end{equation} It is then possible to find points $t_{m_k}\in\dom\gamma_{m_k}$ with $\gamma_{m_k}(t_{m_k})=x_{m_k}\in C$ and $\dist {(t_{m_k},x_{m_k})}, {(t,x)}.\le\epsi$. By passing to a subsequence we can assume that $t_{m_k}\to\tilde t\in\dom\gamma$, $x_{m_k}\to\tilde x\in C$ and $\tilde x=\gamma(\tilde t)$. As $\dist {(\tilde t,\tilde x)},{(t,x)}.\le\epsi$, \begin{equation} d_{(t,x)}(\frest C,n.(\gamma))\le\liminf_{m\to\infty}d_{(t,x)}(\frest C,n.(\gamma_m)). \end{equation} \end{proof} \par The second step in the proof of Theorem \ref{compact_reduction} is a semplification of the description of Alberti representations which relies on the following map $\Psi$. \begin{lem} Let $X$ be a Polish space and consider the map \begin{equation} \begin{aligned} \Psi:\frags(X)&\to M(X)\\ \gamma&\mapsto\mpush\gamma.\left(\lebmeas\mrest\dom\gamma\right)=\Psi_\gamma; \end{aligned} \end{equation} then $\Psi$ is Borel and, if $A\subset X$ is Borel and $[a,b]\subset\real$, the map \begin{equation} \gamma\mapsto\Psi_\gamma\left(A\cap\gamma(\dom\gamma\cap[a,b])\right) \end{equation} is Borel. \end{lem} \begin{proof} It suffices to show that if $g$ is a real-valued continuous function on $X$, the map \begin{equation} \Psi_{g,a,b}:\gamma\mapsto\int_a^bg\circ\gamma(t)\chi_{\dom\gamma}(t)\,dt \end{equation} is Borel. Without loss of generality we can assume that $g$ is nonnegative, in which case we will show that $\Psi_{g,a,b}$ is upper semicontinuous. Assuming that $\gamma_n\to\gamma$ in $\frags(X)$, we have: \begin{equation}\label{eq:uppersem} \limsup_{n\to\infty}\left(g\circ\gamma_n\,\chi_{\dom\gamma_n}\right)\le g\circ\gamma\, \chi_{\dom\gamma}; \end{equation} in fact, either a point $t$ belongs to infinitely many of the $\dom\gamma_n$, in which case $t\in\dom\gamma$ and $\gamma_n(t)\to\gamma(t)$, or eventually the point $t$ does not belong to $\dom\gamma_n$, in which case the left hand side of \eqref{eq:uppersem} is $0$. By the reverse Fatou Lemma: \begin{equation} \limsup_{n\to\infty}\int_a^bg\circ\gamma_n\,\chi_{\dom\gamma_n}\,dt \le \int_a^bg\circ\gamma\, \chi_{\dom\gamma}\,dt, \end{equation} which is the upper semicontinuity of $\Psi_{g,a,b}$. \end{proof} \par We now obtain a semplification of the description of Alberti representations: one can assume that $\nu_\gamma=g\,\Psi_\gamma$, where $g$ is a Borel function on $X$. \begin{lem} Let $X$ be a compact metric space with a Radon measure $\mu$ admitting an Alberti representation $(P,\nu)$ satisfying \albcond. Then there is an Alberti representation $\albrep.'=(P',\nu')$ which satifies \albcond\ and with $\nu'_\gamma=g\,\Psi_\gamma$, where $g$ is Borel on $X$. \end{lem} \begin{proof} By Lemma \ref{frag_meas} we can find disjoint Borel sets $F_n\subset\frags_n(X)$ with $\frags(X)=\bigcup_nF_n$. Note that if $\gamma\in F_n$, then we have the bound $\|\Psi_\gamma\|\le 2n$. By Lemma \ref{integration_meas} the integral \begin{equation} \tilde\mu=\sum_n\int_{F_n}\frac{1}{2n}\Psi_\gamma\,dP(\gamma) \end{equation} defines a finite Radon measure on $X$ with total mass at most $1$. If $\tilde\mu(A)=0$, then for $P$-a.e.~$\gamma$, $\Psi_\gamma(A)=0$ which implies $\nu_\gamma(A)=0$ as $\hmeas 1._\gamma\ll\Psi_\gamma$. Introducing the measure \begin{equation} \tilde P = \sum_n\frac{1}{2n}P\mrest F_n, \end{equation} we obtain a finite Borel measure on $\frags(X)$ and we have \begin{equation} \tilde\mu=\int_{\frags(X)}\Psi_\gamma\,d\tilde P(\gamma). \end{equation} As $\tilde\mu\gg\mu$, by the Radon-Nikodym Theorem \cite[Thm.~6.10]{rudin-real}, there is a Borel function $\tilde g$ on $X$ with $\mu=\tilde g\,\tilde\mu$. Then \begin{equation} \mu=\int_{\frags(X)}\tilde g\,\Psi_\gamma\,d\tilde P \end{equation} and the result follows letting \begin{align} g&=\tilde g \tilde P (\frags(X));\\ \nu'_\gamma&=g\Psi_\gamma;\\ P'&=\frac{1}{\tilde P(\frags(X))}\tilde P. \end{align} From the way in which $\tilde P$ was obtained from $P$, we conclude that if the original Alberti representation $(P,\nu)$ satisfied \albcond, so does the new one $(P',\nu')$. \end{proof} \par We can now prove Theorem \ref{compact_reduction}. \begin{proof}[Proof of Theorem \ref{compact_reduction}] By the previous Lemma we can assume $\nu_\gamma=g\,\Psi_\gamma$. Note that as \begin{equation}\label{mu_on_C} \mu\left(X\setminus C\right)=0, \end{equation} for $P$-a.e.~$\gamma\in\frags(X)\setminus\frags(X,C)$, \begin{equation} g\,\Psi_\gamma=0. \end{equation} In particular, replacing $P$ by \begin{equation} \frac{P\mrest\frags(X,C)}{P\left(\frags(X,C)\right)}, \end{equation} we can assume that \begin{equation} P\left(\frags(X)\setminus\frags(X,C)\right)=0. \end{equation} Then for $A$ Borel, \begin{equation} \begin{aligned} \mu(A)&=\int_{\frags(X,C)}dP(\gamma)\int_Xg\chi_A\,d\Psi_\gamma & &\\ &=\int_{\frags(X,C)}dP(\gamma)\int_Xg\chi_A\chi_C\,d\Psi_\gamma &&\text{(by \eqref{mu_on_C})}\\ &=\int_{\frags(X,C)}dP(\gamma)\int_Xg\chi_A\,d\Psi(\frest C.(\gamma)) & &\\ &=\int_{\frags(C)}d\mpush{\frest C.}.P(\gamma)\int_Xg\chi_A\,d\Psi_\gamma &&\text{(by pushing forward)} \end{aligned} \end{equation} where in the last step we used that $\frest C.$ is Borel by Lemma \ref{borel_rest}. Letting $P'=\mpush{\frest C.}.P$ and $\nu'_\gamma=g\,\Psi_\gamma$, $(P',\nu')$ is an Alberti representation of $\mu$ defined on $\frags(C)$ (using again Lemma \ref{integration_meas}). From the way in which $P'$ was obtained from $P$, if $(P,\nu)$ satisfied \albcond, so does $(P',\nu')$. \end{proof} \par The following is a gluing principle for Alberti representations. \begin{thm}\label{alb_glue} Let $\mu$ be a Radon measure on $Z$ and $U\subset Z$ Borel. Assume that there are disjoint Borel sets $U_n\subset U$ isometrically embedded in some Polish spaces $Y_n$ ($U_n\hookrightarrow Y_n$) with $\mu\mrest U_n$ admitting an Alberti representation defined on $\frags(Y_n)$ and satisfying \albcond. Then $\mu\mrest U$ admits an Alberti representation defined on $\frags(Z)$ and satisfying \albcond. \end{thm} \begin{proof} Using that $\mu$ is Radon we can find disjoint compact subsets $\{C_\alpha\}$ such that for each $\alpha$ there is an $n_\alpha$ with $C_\alpha\subset U_{n_\alpha}$ and \begin{equation} \mu\left(U\setminus\bigcup_\alpha C_\alpha\right)=0. \end{equation} As $\mu\mrest U_{n_\alpha}$ admits an Alberti representation $(P_{n_\alpha},\nu_{n_\alpha})$ defined on $\frags(Y_{n_\alpha})$ and satisfiying \albcond, so does $\mu\mrest C_\alpha$ by letting $P'_\alpha=P_{n_\alpha}$ and $\nu'_\alpha=\nu_{n_\alpha}\mrest C_\alpha$. Then by Theorem \ref{compact_reduction}, $\mu\mrest C_\alpha$ admits an Alberti representation $(P_\alpha,\nu_\alpha)$ defined on $\frags(C_\alpha)$ and satisfying \albcond. Observing that the $\frags(C_\alpha)$ are disjoint Bo\-rel sub\-sets of $\frags(Z)$, we obtain an Alberti representation $(P,\nu)$ of $\mu\mrest U$ defined on $\frags(Z)$ and satisfing \albcond\ by letting \begin{align} P&=\sum_{\alpha=1}^\infty 2^{-\alpha}P_\alpha,\\ \nu&=\sum_{\alpha=1}^\infty 2^{\alpha}\nu_\alpha. \end{align} \end{proof} \par We now introduce the key concept to build Alberti representations in terms of nullity of a set with respect to a family of fragments. \begin{defn}\label{def:frag_nullity} If ${\mathcal G}\subset\frags(X)$, a Borel $S\subset X$ is said to be ${\mathcal G}$-null if for each $\gamma\in{\mathcal G}$, $\hmeas 1._\gamma(S)=0$. In case ${\mathcal G}$ is the family of fragments satisfying a set of conditions \albcond, we say that $S$ is \albcond-null. \end{defn} \par We now prove Lemma \ref{biLip_dis} which produces biLipchitz Alberti representations in Banach spaces. Throughout the remainder of this Subsection $\convgeo$ denotes a closed convex compact subset of $\bana^*$, where $\bana$ is a separable Banach space. This is motivated by the fact that any separable metric space $X$ can be isometrically embedded in $l^\infty$ via a Kuratowski embedding. The closed convex hull $\convgeo$ of $X$ in $\bana^*$ is compact by \cite[Thm.~3.25]{rudin-functional}. \begin{defn} For positive $\epsi$ and $\tau$, let $\QG\epsi,\tau.\subset\frags(\bana^*)$ denote the compact subset \begin{multline} \QG\epsi,\tau.=\left\{\gamma:[0,\tau]\to\convgeo:\forall t,s\in[0,\tau],\right. \\\left. |t-s|\le \|\gamma(t)-\gamma(s)\|_{\bana^*}\right.\left.\le(1+\epsi) |t-s|\right\}. \end{multline} \end{defn} \begin{lem} Let $\Psi:\QG\epsi,\tau.\to P(\convgeo)$ be defined by \begin{equation} \Psi_\gamma=\frac{1}{\tau}\mpush\gamma.\lebmeas\mrest[0,\tau]; \end{equation} then $\Psi$ is continuous. \end{lem} \begin{proof} Being $\QG\epsi,\tau.$ and $P(\convgeo)$ metric spaces, it suffices to check continuity for sequences. Assume that $\gamma_n\to\gamma$ in $\QG\epsi,\tau.$ and that $f$ is continuous with compact support in $\convgeo$. Then, as $f\circ\gamma_n\to f\circ\gamma$ uniformly, \begin{equation} \lim_{n\to\infty}\frac{1}{\tau}\int_0^\tau f\circ\gamma_n\,d\lebmeas= \frac{1}{\tau}\int_0^\tau f\circ\gamma\,d\lebmeas. \end{equation} \end{proof} \begin{lem}\label{biLip_dis} Let $\mathcal{G}\subset\QG\epsi,\tau.$ be compact and $\bar{H}\mathcal{G}$ denote the closed convex hull of $\Psi(\mathcal{G})$ in $M(\convgeo)$. Let $\mu$ be a Radon measure on $\convgeo$. Then \begin{equation} \mu=\mu_{\mathcal{G}}+\mu\mrest F_{\mathcal{G}}, \end{equation} with $\mu_{\mathcal{G}}\ll\rho_{\mathcal{G}}$ and $\rho_{\mathcal{G}}\in\bar{H}\mathcal{G}$; furthermore, there are a Borel regular probability measure $P_{\mathcal{G}}$ on $\mathcal{G}$ and a Borel function $f_{\mathcal{G}}$ on $\convgeo$ such that \begin{align} \rho_{\mathcal{G}}&=\int_{\mathcal{G}}\Psi_\gamma\,dP_{\mathcal{G}}\\ \label{gksrep} \mu_{\mathcal{G}}&=\int_{\mathcal{G}}f_{\mathcal{G}}\Psi_\gamma\,dP_{\mathcal{G}}. \end{align} The set $F_{\mathcal{G}}$ is a $\mathcal{G}$-null $F_\sigma$. \end{lem} \begin{proof} The space $M(\convgeo)$ in the weak* topology is a Fr\'echet space and so by \cite[Thm.~3.25]{rudin-functional} the closed convex hull of a compact subset is compact, so $\bar{H}\mathcal{G}$ is compact. By Rainwater's Lemma (\cite[Thm.~9.4.4]{rudin-complex-ball}, \cite{rainwater-note}) we can find $\mu_{\mathcal{G}}$, $\rho_{\mathcal{G}}\in \bar{H}\mathcal{G}$ and $F_{\mathcal{G}}$ such that: \begin{enumerate} \item The measure $\mu$ decomposes as $\mu=\mu_{\mathcal{G}}+\mu\mrest F_{\mathcal{G}}$. \item The measure $\mu_{\mathcal{G}}$ satisfies $\mu_{\mathcal{G}}\ll\rho_{\mathcal{G}}$. \item The set $F_{\mathcal{G}}$ is $\Psi(\mathcal{G})$-null in the sense that, for each $\gamma\in \mathcal{G}$, $\Psi_\gamma(F_{\mathcal{G}})=0$. \end{enumerate} Note that (3) implies that $F_{\mathcal{G}}$ is ${\mathcal{G}}$-null because for $\gamma\in{\mathcal{G}}$, $\Psi_\gamma$ and $\hmeas 1._\gamma$ are absolutely continuous one with respect to the other. By \cite[Thm.~3.28]{rudin-functional} there is a regular Borel probability measure $\pi_{\mathcal{G}}$ supported on $\Psi(\mathcal{G})$ and such that \begin{equation} \rho_{\mathcal{G}}=\int_{\Psi({\mathcal{G}})}\sigma\,d\pi_{\mathcal{G}}(\sigma). \end{equation} As $\Psi$ is continuous and ${\mathcal{G}}$ is compact, by \cite[Thm.~6.9.7]{bogachev_measure} there is a Borel $\tilde{\mathcal{G}}\subset{\mathcal{G}}$ with $\Psi(\tilde{\mathcal{G}})=\Psi({\mathcal{G}})$ and such that $\Psi$ is injective on $\tilde {\mathcal{G}}$ with a Borel inverse \begin{equation} \tilde\Psi^{-1}:\Psi(\tilde{\mathcal{G}})\to{\mathcal{G}}; \end{equation} in particular, \begin{equation} \begin{split} \rho_{\mathcal{G}}&=\int_{\Psi({\mathcal{G}})}\sigma\,d\pi_{\mathcal{G}}(\sigma)\\ &=\int_{\Psi(\tilde{\mathcal{G}})}\Psi(\tilde\Psi^{-1}(\sigma))\,d\pi_{\mathcal{G}}(\sigma)\\ &=\int_{{\mathcal{G}}}\Psi_\gamma\,d\underbrace{\left(\mpush\tilde\Psi^{-1}.\pi_{\mathcal{G}}\right)}_{P_{\mathcal{G}}}(\gamma). \end{split} \end{equation} If $f_{\mathcal{G}}$ is a Borel representative of the Radon-Nikodym derivative of $\mu_{\mathcal{G}}$ with respect to $\rho_{\mathcal{G}}$, \begin{equation} \mu_{\mathcal{G}}=\int_{{\mathcal{G}}}f_{\mathcal{G}}\Psi_\gamma\,dP_{\mathcal{G}}. \end{equation} \end{proof} The following theorem is the main tool we will use to produce Alberti representations. \begin{thm}\label{alberti_rep_prod} Let $X$ be a separable metric space and $\mu$ a Radon measure on $X$. Let $\Omega$ denote the set of fragments in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$\footnote{We require a strict inequality}. Then the following are equivalent: \begin{enumerate} \item The measure $\mu$ admits an Alberti representation in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$. \item For each $K\subset X$ compact and $\Omega$-null, $\mu(K)=0$. \item For each $\epsi>0$, $\mu$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$. \end{enumerate} \end{thm} \begin{proof} We show that (1) implies (2). Let $\albrep.=(P,\nu)$ with $P$ giving full measure to a Borel set $\Xi$ of fragments in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$. If $K\subset X$ is $\Omega$-null, it is $\Xi$-null, so \begin{equation} \mu(K)=\int_\Xi\underbrace{\nu_\gamma(K)}_{=0}\,dP(\gamma)=0. \end{equation} \par We show that (2) implies (3) first assuming that $X$ is compact and $v$, $\alpha$ and $\delta$ are constant. We rescale $f:X\to\real^q$ and $g$ to be $1$-Lipschitz, embedd $\graph (f,g)$ isometrically in $X\times\real^q\times\real=\bana^*$ with norm \begin{equation} \|(\xi,v,t)\|_{l^\infty\times\real^q\times\real}=\max(\|\xi\|_{l^\infty},\|v\|_2,|t|), \end{equation} and take the convex hull $\convgeo$ of the resulting set. \par Having chosen strictly decreasing sequences $\{\tau_n\}$, $\{\eta_n\}$ converging to $0$, we let \begin{multline} {\mathcal{G}}_n=\bigl\{\gamma\in\QG\epsi,\tau_n.:\text{$\forall t,s\in[0,\tau_n]$,}\\\text{$\sgn(t-s)\left(f\circ\gamma(t)-f\circ\gamma(s)\right)\in\bar\cone(v,\alpha-\eta_n)$}\\ \text{and $\sgn(t-s)\left( g\circ\gamma(t)-g\circ\gamma(s)\right)\ge(\delta+\eta_n)|t-s|$}\bigr\}, \end{multline} which is a compact subset of $\QG\epsi,\tau_n.$. We then apply Lemma \ref{biLip_dis} repeatedly; having obtained \begin{equation} \mu=\mu_{{\mathcal{G}}_1}+\mu\mrest F_{{\mathcal{G}}_1}, \end{equation} we apply Lemma \ref{biLip_dis} to $\mu\mrest F_{{\mathcal{G}}_1}$ to obtain a decomposition \begin{equation} \mu=\mu_{{\mathcal{G}}_1}+\mu_{{\mathcal{G}}_2}+\mu\mrest F_{{\mathcal{G}}_{1,2}} \end{equation} where $\mu_{{\mathcal{G}}_1}$ and $\mu_{{\mathcal{G}}_2}$ are concentrated on disjoint $G_\delta$ sets, and where $F_{{\mathcal{G}}_{1,2}}$ is an intersection of two $F_\sigma$ sets and is ${{\mathcal{G}}_1}$, ${{\mathcal{G}}_2}$-null. Iterating, \begin{equation} \mu=\sum_{n=1}^\infty\mu_{{\mathcal{G}}_n}+\mu\mrest F, \end{equation} where $F$ is an $F_{\sigma\delta}$ which is ${{\mathcal{G}}_n}$-null for every $n$. \par We now show that $\mu(F)=0$ using condition (2). The following observation (Ob1) will be used repeatedly: if $\gamma\in\frags(\convgeo)$ and the $\{K_\alpha\}\subset\dom\gamma$ are (countably many) compact sets with \begin{equation} \lebmeas\left(\dom\gamma\setminus\bigcup_\alpha K_\alpha\right)=0 \end{equation} and $\hmeas 1._{\gamma|K_\alpha}(F)=0$, then $\hmeas 1._\gamma(F)=0$. \par If $\gamma$ is a fragment in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$, by (Ob1) and Egorov and Lusin's Theorems \cite[Thms.~7.1.12 and 7.1.13]{bogachev_measure}, letting $I_\gamma=[\min\dom\gamma,\max\dom\gamma]$ we can assume that $\forall t,s\in\dom\gamma$ \begin{align} \|f\circ\gamma(t)-f\circ\gamma(s)-w(t-s)\|_2&\le\rho|t-s|;\label{conical_inclusion}\\ (l-\rho)|t-s|\le\|\gamma(t)-\gamma(s)\|_{\bana^*}&\le(L+\rho)|t-s|;\label{bound_metdiff}\\ \sgn(t-s)\left(g\circ\gamma(t)-g\circ\gamma(s)\right)&\ge(\delta+\eta-\rho)\|\gamma(t)-\gamma(s)\|_{\bana^*},\label{speed_bound} \end{align} where $\eta>0$, $w\in\cone(v,\alpha-\eta)$, $\frac{L}{l}<1+\epsi$ and $\rho$ is small enough to be chosen later. Moreover, by the Lebesgue's differentiation theorem we can also assume that $\forall t,s\in I_\gamma$ with at least one of them in $\dom\gamma$, \begin{equation} \lebmeas\left(\dom\gamma\cap[t,s]\right)\ge(1-\rho)|t-s|. \end{equation} Let $\{(u_n,v_n)\}$ be the components of $I_\gamma\setminus\dom\gamma$. On each component $(u_n,v_n)$ we extend $\gamma$ using the fact that $\convgeo$ is convex \begin{equation} \gamma(t)=\frac{t-u_n}{v_n-u_n}\gamma(v_n)+\frac{v_n-t}{v_n-u_n}\gamma(u_n); \end{equation} if $(s,t)\subset (u_n,v_n)$, \begin{equation} \gamma(t)-\gamma(s)=\frac{t-s}{v_n-u_n}(\gamma(v_n)-\gamma(u_n)) \end{equation} so that \eqref{bound_metdiff} remains true; in particular, the bound on the metric derivative implies that the extension is $(L+\rho)$-Lipschitz. Similarly, \begin{align} f\circ\gamma(t)-f\circ\gamma(s)&=\frac{t-s}{v_n-u_n}(f\circ\gamma(v_n)-f\circ\gamma(u_n))\\ g\circ\gamma(t)-g\circ\gamma(s)&=\frac{t-s}{v_n-u_n}(g\circ\gamma(v_n)-g\circ\gamma(u_n)) \end{align} so that \eqref{conical_inclusion} and \eqref{speed_bound} remain true. If $t,s\in I_\gamma\setminus\dom\gamma$ but in different components, there are $s',t'\in\dom\gamma$: \begin{equation} s\le s'\le t'\le t \end{equation} and each of $(s,s')$ and $(t',t)$ is contained in a component of $I_\gamma\setminus\dom\gamma$. In particular, \begin{align} |s-s'|&\le\rho|s'-t|\le\rho|s-t|;\\ |t'-t|&\le\rho|s-t'|\le\rho|s-t|. \end{align} Therefore, \eqref{conical_inclusion}, \eqref{bound_metdiff} and \eqref{speed_bound} generalize to \begin{equation} \|f\circ\gamma(t)-f\circ\gamma(s)-w(t-s)\|_2\le2\rho(L+\rho+\|v\|)|t-s|;\label{adj_conical_inclusion} \end{equation} \begin{multline}\label{adj_bound_metdiff} \left[(l-\rho)(1-2\rho)-2(L+\rho)\rho\right]|t-s|\le\|\gamma(t)-\gamma(s)\|_{\bana^*}\le\\ \left[L+\rho(1+2(L+\rho))\right]|t-s|; \end{multline} \begin{multline}\label{adj_speed_bound} \sgn(t-s)\left(g\circ\gamma(t)-g\circ\gamma(s)\right)\ge\Biggl(\delta+\eta-\rho \\ -{}2\rho \frac{L+\delta+\eta}{(l-\rho)(1-2\rho)-2(L+\rho)\rho} \Biggr)\|\gamma(t)-\gamma(s)\|_{\bana^*}. \end{multline} If $\eta'\in(0,\eta)$, for $\rho$ sufficiently small we have: \begin{align} \ball w,2\rho(L+\rho+\|v\|).&\in\cone(v,\alpha-\eta');\\ \frac{L+\rho(1+2(L+\rho))}{(l-\rho)(1-2\rho)-2(L+\rho)\rho}&\le1+\epsi;\\ \left(\delta+\eta-\rho-2\rho \frac{L+\delta+\eta}{(l-\rho)(1-2\rho)-2(L+\rho)\rho} \right)&\ge\delta+\eta'. \end{align} We can now reparametrize $\gamma$ to be $(1,1+\epsi)$-biLipschitz, subdivide the domain and choose $n$ sufficiently large so that $\gamma\in{{\mathcal{G}}_n}$. Thus, $\hmeas 1._\gamma(F)=0$. This implies that \begin{equation} \mu=\sum_{n=1}^\infty\mu_{{\mathcal{G}}_n}\quad\text{(on account of condition (2))}. \end{equation} \par As the measures $\mu_{{\mathcal{G}}_n}$ are concentrated on pairwise disjoint sets, by Theorem \ref{alb_glue} we obtain a representation $\albrep.'=(P',\nu')$ of $\mu$ in the $f$-direction of $\cone(v,\alpha)$ with $g$-speed $>\delta$ with $P'\in P(\frags(\convgeo))$; Theorem \ref{compact_reduction} gives a representation $\albrep.=(P,\nu)$ with $P\in P(\frags(X))$. \par The case in which $X$ is not compact and $v$, $\alpha$ and $\delta$ are not constant is treated by using Egorov and Lusin's Theorems to find Borel functions $(v_n,\alpha_n,\delta_n)$ such that: \begin{itemize} \item There are disjoint compacts $\{K_{n,j}\}_j$ with $\mu\left(X\setminus\bigcup_jK_{n,j}\right)=0$ and the $(v_n,\alpha_n,\delta_n)$ are constant on each $K_{n,j}$. \item The functions $\delta_n\searrow\delta$ pointwise. \item The cone fields $\cone(v_n,\alpha_n)\nearrow\cone(v,\alpha)$ pointwise. \end{itemize} \par One then applies the construction for constant cone fields and speeds on the $K_{n,j}$ recursively to obtain $\mu=\sum_{n,j}\mu_{n,j}$ where the measures $\{\mu_{n,j}\}_{n,j}$ have pairwise disjoint supports and each $\mu_{n,j}$ has an Alberti representation in the $f$-direction of $\cone(v_n,\alpha_n)$ with $g$-speed $>\delta_n$. These representations are then glued via Theorem \ref{alb_glue} to give a representation of $\mu$. \par That (3) implies (1) is immediate from the definitions. \end{proof} \subsection{Derivations and $L^\infty$-modules}\label{subsec:der_modules} In this Subsection we recall some facts about derivations and $L^\infty$-modules. Recall that the Banach space $L^\infty(\mu)$ is a real Banach algebra, and, in particular, a ring. We will denote by $L^\infty_+(\mu)$ the set of nonnegative elements $f$: \begin{equation} \mu\left\{x\in X:f(x)<0\right\}=0. \end{equation} The map absolute value \begin{align} L^\infty(\mu)&\xrightarrow{|\cdot|}L^\infty_+(\mu)\\ f&\mapsto|f|\notag \end{align} is sort of an $L^\infty$-valued seminorm: \begin{enumerate} \item For all $c_1,c_2\in\real$, and for all $f_1,f_2\in L^\infty(\mu)$, \begin{equation} |c_1f_1+c_2f_2|\le|c_1||f_1|+|c_2||f_2|. \end{equation} \item For all $f_1,f_2\in L^\infty(\mu)$, \begin{equation} |f_1f_2|=|f_1||f_2|. \end{equation} \end{enumerate} note also that $L^\infty_+(\mu)=\left\{f:f=|f|\right\}$. In particular, $\|f\|_{L^\infty(\mu)} =\|\,|f|\,\|_{L^\infty(\mu)}$. \par An \textbf{$L^\infty(\mu)$-module} $M$ is a Banach space $M$ which is also an $L^\infty(\mu)$-module (with the usual boundedness requirement on the $L^\infty(\mu)$-action). An algebraic submodule $M'\subset M$ is a called a \textbf{submodule} if it is closed. Algebraic concepts like direct sum, linear independence, free module and basis of a free module, extend immediately. Given $S\subset M$ the linear span of $S$ will be denoted by $\linspan_{L^\infty(\mu)}(S)$. The \textbf{dual $\dualmod M.$} of an $L^\infty(\mu)$-module $M$ is the $L^\infty(\mu)$-module of bounded module homomorphisms \begin{equation} f:M\to L^\infty(\mu). \end{equation} \par Among $L^\infty(\mu)$-modules a special r\^ole is played by \textbf{$L^\infty(\mu)$-normed modules}. \begin{defn}\label{defn:local_norm} An $L^\infty(\mu)$-module $M$ is said to be an \textbf{$L^\infty(\mu)$-normed module} if there is a map \begin{equation} |\cdot|_{M,{\text{loc}}}:M\to L^\infty_+(\mu) \end{equation} such that: \begin{enumerate} \item The function $|\cdot|_{M,{\text{loc}}}$ is a ``seminorm'': $\forall c_1,c_2\in\real$, $\forall m_1,m_2\in M$, \begin{equation} |c_1m_1+c_2m_2|\le|c_1||m_1|+|c_2||m_2|. \end{equation} \item For each $\lambda\in L^\infty(\mu)$ and each $m\in M$, \begin{equation} |\lambda m|_{M,{\text{loc}}}=|\lambda|\,|m|_{M,{\text{loc}}}. \end{equation} \item The local seminorm $|\cdot|_{M,{\text{loc}}}$ can be used to reconstruct the norm of any $m\in M$: \begin{equation} \|m\|_M=\|\,|m|_{M,{\text{loc}}}\,\|_{L^\infty(\mu)}. \end{equation} \end{enumerate} \end{defn} \begin{rem} Note that a submodule of an $L^\infty(\mu)$-normed module is an $L^\infty(\mu)$-normed module. The ring $L^\infty(\mu)$ has many idempotents $\chi_U\in L^\infty_+(\mu)$ (characteristic function of the $\mu$-measurable set $U$), which we will call projections as they give direct sum decompositions \begin{equation} L^\infty(\mu)=\chi_U\,L^\infty(\mu)\oplus (1-\chi_U)\,L^\infty(\mu), \end{equation} $\lambda\,L^\infty(\mu)$ denoting the ideal / submodule generated by $\lambda$. \end{rem} \par A simple and important criterion (\cite[Thm.~9]{phillips_cstarnorm}) for an $L^\infty(\mu)$-module $M$ to be an $L^\infty(\mu)$-normed module is the following: \begin{lem} An $L^\infty(\mu)$-module $M$ is an $L^\infty(\mu)$-normed module if and only if $\forall U$ $\mu$-measurable and $\forall m\in M$: \begin{equation} \|m\|_M=\max\left(\|\chi_Um\|_M,\|(1-\chi_U)m\|_M\right); \end{equation} in particular if the $U_\alpha$ are disjoint and $\mu\left(X\setminus\bigcup U_\alpha\right)=0$, \begin{equation} \|m\|_M=\sup_\alpha\|\chi_{U_\alpha}m\|_M. \end{equation} \end{lem} \begin{exa} While $L^\infty(\mu)$ is an $L^\infty(\mu)$-normed module, the previous Lemma implies that for $p\in[1,\infty)$, the $L^p(\mu)$ are not $L^\infty(\mu)$-normed modules. \end{exa} \begin{lem} If $M$ is an $L^\infty(\mu)$-module, $\dualmod M.$ is an $L^\infty(\mu)$-normed module. \end{lem} \begin{proof} Given $\xi\in\dualmod M.$ and $\epsi>0$, choose $m\in M$ with $\|m\|_M=1$ and \begin{equation} \|\xi\|_{\dualmod M.}\le \|\xi(m)\|_{L^\infty(\mu)}+\epsi. \end{equation} As $L^\infty(\mu)$ is an $L^\infty(\mu)$-normed module, $\forall U$ $\mu$-measurable, \begin{equation} \begin{split} \|\xi(m)\|_{L^\infty(\mu)}&=\max\left(\|\chi_U\xi(m)\|_{L^\infty(\mu)},\|(1-\chi_U)\xi(m)\|_{L^\infty(\mu)}\right)\\ &\le\max\left(\|\chi_U \xi\|_{\dualmod M.},\|(1-\chi_U)\xi\|_{\dualmod M.}\right). \end{split} \end{equation} \end{proof} \begin{rem}Actually, the previous argument shows that $\boundlin X,N.$ (bounded linear maps from the Banach space $X$ to $N$) is an $L^\infty(\mu)$-normed module whenever $N$ is too. \end{rem} \par For $L^\infty(\mu)$-normed modules there is an analogue of the Hahn-Banach Theorem \cite[Thm.~5]{weaver00}: \begin{lem}\label{lem:hahn-banach} Let $M'$ be an algebraic submodule of the $L^\infty(\mu)$-normed module $M$ and $\Phi'\in\dualmod M'.$ with norm $\le c$. Then there is a $\Phi\in\dualmod M.$ extending $\Phi'$ with norm $\le c$. \end{lem} We now introduce derivations, but first recall some facts about Lipschitz functions. We will denote by $\lipfun X.$ the space of real-valued Lipschitz functions defined on $X$ and by $\lipalg X.$ the real algebra of real-valued bounded Lipschitz functions defined on $X$. For $f\in\lipfun X.$ its global Lipschitz constant will be denoted by $\glip f.$. For $f\in\lipalg X.$ we define the norm \begin{equation} \|f\|_{\lipalg X.} = \max(\|f\|_\infty,\glip f.). \end{equation} This gives $(\lipalg X.,\|\cdot\|_{\lipalg X.})$ the structure of a Banach algebra, compare \cite[sec.~4.1]{weaver_book99}. An important property of $\lipalg X.$ is that it is a dual Banach space and it has a unique predual. For more information we refer the reader to \cite[chap.~2]{weaver_book99}. For the scope of the present work, the most useful topology on $\lipalg X.$ is the weak* topology. As we are assuming $X$ separable, by Krein-\v Smulian, on bounded subsets of $\lipalg X.$ the weak* topology will be metrizable; it turns out that a sequence $f_n\xrightarrow{\text{w*}}f$ if and only if $f_n\to f$ pointwise and if $\sup_n\glip f_n. <\infty$ \footnote{Note that this implies uniform convergence on compact subsets}. \par In the sequel, especially in Section \ref{sec:derivations_alberti}, we will sometimes need to consider functions which are Lipschitz with respect to a different (pseudo)-distance $d'$. If $f\in\lipfun X.$ is $L$-Lipschitz with respect to $d'$, we will say that $f$ is $(1,d')$-Lipschitz. \begin{defn} A \textbf{derivation $D:\lipalg X.\to L^\infty(\mu)$} is a weak* continuous, bounded linear map satisfying the product rule: \begin{equation} D(fg)=fDg+gDf. \end{equation} \end{defn} Note that the notion of derivation depends on the measure $\mu$ and that the product rule implies that $Df=0$ if $f$ is constant. The collection of all derivations $\wder\mu.$ is an $L^\infty(\mu)$-normed module. Moreover, any $D\in \wder\mu\mrest U.$ gives rise to an element of $\wder\mu.$ by extending $Df$ to be $0$ on the complement of $U$. In this way, $\wder\mu\mrest U.$ can be naturally identified with the submodule $\chi_U\, \wder\mu.$ of $\wder\mu.$. Derivations are local in the following sense \cite[Lem.~27]{weaver00}: \begin{lem}\label{lem:locality_derivations} If $U$ is $\mu$-measurable and if $f,g\in\lipalg X.$ agree on $U$, then $\forall D\in\wder\mu.$, $\chi_UDf=\chi_UDg$. \end{lem} \begin{rem} \label{rem:derivation_extension} This locality property allows to extend derivations to Lipschitz functions. Considering countably many disjoint compact sets $K_\alpha$ with uniformly bounded diametre and points $c_\alpha\in K_\alpha$, with $\mu\left(X\setminus\bigcup_\alpha K_\alpha\right)=0$, we define for $f\in\lipfun X.$ \begin{equation} Df=\sum_\alpha\chi_{K_\alpha}D\left(\min\left(\max\left(f-f(c_\alpha),-\max_{K_\alpha}|f-f(c_\alpha)|\right)\right), \max_{K_\alpha}|f -f(c_\alpha)|\right) \end{equation} which agrees with the previous definition for $f\in\lipalg X.$, using locality. Observe that we used that the norm of $f-f(c_\alpha)$ in $\lipalg K_\alpha.$ is bounded in terms of $\glip f.$ and the diametre of $K_\alpha$. Again using locality, it is possible to show that the extension does not depend on the choice of the $K_\alpha$ or the point $c_\alpha$. \end{rem} \begin{defn} The dual $L^\infty(\mu)$-normed module of $\wder\mu.$ will be denoted by $\wform\mu.$ and called \textbf{module of forms}. For $f\in\lipalg X.$ (or $\lipfun X.$) let $df\in\wform\mu.$ be defined by \begin{equation} \forall D\in\wder\mu.,\quad \langle df,D\rangle=Df; \end{equation} the map \begin{equation} \begin{aligned} d:\lipalg X.&\to\wform \mu.\\ f&\mapsto df \end{aligned} \end{equation} is linear, satisfies the product rule \begin{equation} d(fg)=fdg+gdf \end{equation} and is weak* continuous. \end{defn} The following result is also used repeatedly in the following. Proofs can be found in \cite{gong11-revised,derivdiff}. \begin{cor}\label{cor:pseudoduality} Let $\{D_i\}_{i=1}^N\subset\wder\mu\mrest U.$ be linearly independent, where $U$ is Borel. Then there are disjoint Borel subsets $V_\alpha\subset U$ with $\mu(U\setminus\bigcup_\alpha V_\alpha)=0$ such that, for each $\alpha$, there are $1$-Lipschitz functions $\{g_{i,\alpha}\}_{i=1}^N$ and derivations $\{D_{\alpha,i}\}_{i=1}^N$ in the linear span of $\{\chi_{V_\alpha}D_i\}_{i=1}^N$, satisfying \begin{equation}\label{eq:pseudoduality} D_{\alpha,i}g_{\alpha,i}=\delta_{ij}\chi_{V_\alpha}. \end{equation} \end{cor} \begin{rem} The space $\boundlin\lipalg X.,L^\infty(\mu).$ of bounded linear maps \begin{equation}\lipalg X.\to L^\infty(\mu) \end{equation} is a dual Banach space. In fact, let $B_1$ denote the unit ball in $\lipalg X.$ and consider the Banach space \begin{equation} l^1\left(B_1; L^1(\mu)\right)=\left\{f:S\to L^1(\mu):\sum_{s\in S}\left\|f(s)\right\|_{L^1(\mu)}<\infty\right\}; \end{equation} then $\boundlin\lipalg X.,L^\infty(\mu).$ can be identified with a subspace of the dual of the generalized $l^1$-space: $l^1\left(B_1;L^1(\mu)\right)$; on $\wder\mu.\subset\boundlin\lipalg X.,L^\infty(\mu).$ we can consider the corresponding weak* topology; in concrete terms, this is the coarsest topology making continuous all the seminorms $\{\nu_\xi\}_{l^1\left(B_1; L^1(\mu)\right)}$: \begin{equation} \nu_{\xi}(D)=\left|\sum_{f\in B_1}\int \xi(f)\,Df\,d\mu\right|; \end{equation} a basis for the weak* topology consists of the sets \begin{equation} \Omega(\xi_1,\ldots,\xi_k; D_0; \epsi)=\left\{D\in\wder\mu.: \nu_{\xi_i}(D-D_0)<\epsi\,(\forall i\in\left\{1,\ldots,k\right\})\right\} \end{equation} where $\left\{\xi_1,\ldots,\xi_k\right\}\subset l^1\left(B_1; L^1(\mu)\right)$, $D_0\in\wder\mu.$, $\epsi>0$. \end{rem} \subsection{Differentiability spaces}\label{subsec:diff_spaces} In this Subsection we recall some facts about differentiability spaces. We first recall the definition of the infinitesimal Lipschitz constants of $f\in\lipfun X.$. \begin{defn} For $f\in\lipfun X.$ we define the \textbf{lower and upper variations of $f$ at $x$ below scale $r$} by \begin{align} \biglip f(x,r)&=\sup_{s\le r}\sup_{y\in B(x,s)}\frac{|f(x)-f(y)|}{s}\\ \smllip f(x,r)&=\inf_{s\le r}\sup_{y\in B(x,s)}\frac{|f(x)-f(y)|}{s}. \end{align} The \textbf{(``big'' and ``small'') infinitesimal Lipschitz constants of $f$ at $x$} are defined by \begin{align} \biglip f(x)&=\inf_{r\ge0}\biglip f(x,r)\\ \smllip f(x)&=\sup_{r\ge0}\smllip f(x,r). \end{align} The functions $\biglip f(\cdot,r)$, $\smllip f(\cdot, r)$, $\biglip f$ and $\smllip f$ are Borel. \end{defn} \par The key notion for generalizing the classical Rademacher's Theorem to metric spaces is that of infinitesimal independence. \begin{defn}\label{def:infi_indep} Let $\{f_i\}_{i=1}^n\subset\lipfun X.$, then \begin{equation} \begin{aligned} \Phi_{x,\{f_i\}_{i=1}^n}:\real^n&\to[0,\infty)\\ (a_i)_{i=1}^n&\mapsto\biglip\left(\sum_{i=1}^na_if_i\right)(x) \end{aligned} \end{equation} defines a seminorm on $\real^n$; the functions $\{f_i\}_{i=1}^n$ are called \textbf{infinitesimally independent on a $\mu$-measurable $A\subset X$} if $\Phi_{x,\{f_i\}_{i=1}^n}$ is a norm for $\mu$-a.e.~$x\in A$. \end{defn} \par We can now define differentiability spaces. \begin{defn} A metric measure space $(X,\mu)$ is called a \textbf{differentiability space} if there is a uniform bound $N$ on the number of Lipschitz functions that can be inifinitesimally independent on a positive measure set. In this case there are (countably many) $\mu$-measurable $\{X_\alpha\}$ with $X=\bigcup_\alpha X_\alpha$ such that: \begin{enumerate} \item For each $\alpha$ there is $\{x_\alpha^j\}_{j=1}^{N_\alpha}\subset\lipfun X.$ such that $\forall f\in\lipfun X.$ there are unique $\left\{\frac{\partial f}{\partial x_\alpha^j}\right\}_{j=1}^{N_\alpha} \subset L^\infty(\mu\mrest X_\alpha)$ such that \begin{equation} \biglip\left(f-\sum_{j=1}^n\frac{\partial f}{\partial x_\alpha^j}(x)x_\alpha^j\right)(x)=0\quad\text{for $\mu$-a.e.~$x\in X_\alpha$}. \end{equation} \item The $N_\alpha$ are uniformly bounded by $N$ and the minimal value of $N$ is called \textbf{the differentiability dimension}. \end{enumerate} \par Each $(X_\alpha,\{x_\alpha^j\}_{j=1}^{N_\alpha})$ is called a \textbf{chart}, the $\{x_\alpha^j\}_{j=1}^{N_\alpha}$ are called \textbf{chart functions}, and the maps: \begin{equation} \begin{aligned} \frac{\partial }{\partial x_\alpha^j}:\lipfun X.&\to L^\infty(\mu\mrest X_\alpha)\\ f&\mapsto \frac{\partial f}{\partial x_\alpha^j} \end{aligned} \end{equation} are called \textbf{partial derivative operators}; the representatives $\frac{\partial f}{\partial x_\alpha^j}$ can be taken to be Borel if $X_\alpha$ is Borel. \par Note that differentiability spaces are usually assumed to be Polish. \end{defn} \section{Derivations and Alberti representations}\label{sec:derivations_alberti} \subsection{From Alberti representations to Derivations}\label{subsec:derivations_alberti} In this Subsection we associate derivations to Alberti representations; the fundamental construction is provided by Theorem \ref{thm:alb_derivation}; its proof uses the following Lemma to check the measurability of certain functions. \begin{lem}\label{lem:meas_fix} Suppose that $X$ is Polish and $\albrep.=(P,\nu)$ is an Alberti representation. Then for each $(f,g)\in\lipfun X.\times\bborel X.$ the map: \begin{equation} \label{eq:meas_fix_s1} \begin{aligned} H_{f,g}:\frags(X)&\to\real\\ \gamma&\mapsto\int_{\dom\gamma}(f\circ\gamma)'(t)\,(g\circ\gamma)(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t) \end{aligned} \end{equation} is Borel. \end{lem} \begin{proof} We start by showing that $H_{f,1}$ is Borel. Without loss of generality we can assume that $f$ is $1$-Lipschitz and just show that the restriction $H_{f,1}|\frags_n(X)$ is Borel. Recall that the set \begin{equation} \label{eq:meas_fix_p0} \lipalgb \real,n.=\left\{\psi\in\lipalg\real.: \glip\psi.\le n\right\} \end{equation} is closed in $\lipalg \real.$ and is a Polish space if we consider the weak* topology. Note that the set \begin{equation} \label{eq:meas_fix_p1} \text{\normalfont Ext}_n=\left\{(\gamma,\psi)\in\frags_n(X)\times\lipalgb \real,n.:\text{$\psi$ extends $f\circ\gamma$}\right\} \end{equation} is closed in $\frags_n(X)\times\lipalgb \real,n.$. For each $\gamma$, by taking a McShane extension of $f\circ\gamma$, one concludes that the section $(\text{\normalfont Ext}_n)_\gamma$ is non-empty. Moreover, by Ascoli-Arzel\`a each section $(\text{\normalfont Ext}_n)_\gamma$ is compact. By the Lusin-Novikov uniformization Theorem \cite[Thm.~18.10]{kechris_desc}, the set $\text{\normalfont Ext}_n$ admits a Borel uniformization and thus there is a Borel uniformizing function: \begin{equation} \label{eq:meas_fix_p2} F_n:\frags_n(X)\to\lipalgb \real,n. \end{equation} with $(\gamma,F_n(\gamma))\in\text{\normalfont Ext}_n$. To avoid a cumbersome notation for the value of $F_n(\gamma)$ at a point $t$, we will also write $F_{n,\gamma}$ to denote $F_n(\gamma)$. We now show that the function $G_{f,1}:\frags_n(X)\to\real$ defined by: \begin{equation} \label{eq:meas_fix_p3} G_{f,1}(\gamma)=\int_\real F_{n,\gamma}(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t) \end{equation} is Borel. For $k\in\natural$ we let $\Delta^k$ denote the collection of closed dyadic intervals in $\real$ of the form $\left[\frac{m}{2^k},\frac{m+1}{2^k}\right]$ for $m\in\zahlen$. Given an interval $I\in\Delta^k$ we denote by $a_I,b_I$ its extremes so that $[a_I,b_I]=I$. Consider the maps $G_{f,1;k}:\frags_n(X)\to\real$ defined by: \begin{equation} \label{eq:meas_fix_p4} G_{f,1;k}(\gamma)=\sum_{I\in\Delta^k}\int_I F_{n,\gamma}(a_I)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t); \end{equation} as the measure $\mpush\gamma^{-1}.\nu_\gamma$ is finite and $F_{n,\gamma}$ is Lipschitz, $\lim_{k\to\infty}G_{f,1;k}(\gamma)=G_{f,1}(\gamma)$. Note also that \begin{equation} \label{eq:meas_fix_p5} \int_IF_{n,\gamma}(a_I)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t)=F_{n,\gamma}(a_I)\,\nu_\gamma\left(\gamma(\dom\gamma\cap[a_I,b_I])\right); \end{equation} as the map $\gamma\mapsto F_{n,\gamma}(a_I)$ is Borel because it is the composition of $F_{n,\gamma}$ with the evalutation at the point $a_I$, and as the map $\gamma\mapsto\nu_\gamma\left(\gamma(\dom\gamma\cap[a_I,b_I])\right)$ is Borel by condition (4) in the Definition \ref{defn:Alberti_rep} of Alberti representations, we conclude that the maps $G_{f,1; k}$ and $G_{f,1}$ are Borel. We now fix $k\in\natural$ and note that also the map \begin{equation} \label{eq:meas_fix_p6} \gamma\mapsto\int_\real F_{n,\gamma}\left(t+\frac{1}{k}\right)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t) \end{equation} is Borel. We thus conclude that the function $H_{f,1; k}:\frags_n(X)\to\real$ defined by: \begin{equation} \label{eq:meas_fix_p7} H_{f,1;k}(\gamma)=\int_\real \frac{F_{n,\gamma}\left(t+\frac{1}{k}\right)-F_{n,\gamma}(t)}{1/k}\,d(\mpush\gamma^{-1}.\nu_\gamma)(t), \end{equation} is Borel. Note that for $\lebmeas$-a.e.~$t$ we have $\lim_{k\to\infty}\frac{F_{n,\gamma}\left(t+\frac{1}{k}\right)-F_{n,\gamma}(t)}{1/k}=F'_{n,\gamma}(t)$, which agrees with $(f\circ\gamma)'(t)$ for $\mpush\gamma^{-1}.\nu_\gamma$-a.e.~$t$. As $H_{f,1}(\gamma)=\lim_{k\to\infty}H_{f,1;k}(\gamma)$, we conclude that $H_{f,1}$ is Borel. For a Borel $A\subset X$, the previous argument can be applied to the Alberti representation $(P,\nu\mrest A)$ of $\mu\mrest A$ to conclude that $H_{f,\chi_A}$ is Borel. As each $g\in\bborel X.$ is a pointwise limit of uniformly bounded simple functions, we conclude that $H_{f,g}$ is Borel. \end{proof} \begin{thm}\label{thm:alb_derivation} If $\albrep.=(P,\nu)$ is a $C$-Lipschitz Alberti representation, the formula \begin{equation}\label{eq:derivation_alberti} \begin{split} \int_X gD_{\albrep.}f\,d\mu&=\int_{\frags(X)}dP(\gamma)\int_{\dom\gamma} (f\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t)\\ &=\int_{\frags(X)}dP(\gamma)\int_X g\partial_\gamma f\,d\nu_\gamma\quad(g\in L^1(\mu)\cap\bborel X.), \end{split} \end{equation} where \begin{equation} \partial_\gamma f(x)= \begin{cases} (f\circ\gamma)'(\gamma^{-1}(x))&\text{if $x\in\im\gamma$ and the derivative exists}\\ 0&\text{otherwise}, \end{cases} \end{equation} defines a derivation $D_{\albrep.}\in\wder\mu.$ with $\|D\|_{\wder\mu.}\le C$. \par Denoting by $\Alb(\mu)$ the set of Alberti representations of $\mu$ which are Lipschitz and letting \begin{equation} \Alb_{{\text{\normalfont sub}}}(\mu)=\bigcup\left\{\Alb(\mu\mrest S):\text{\normalfont $S\subset X$ Borel}\right\}, \end{equation} we obtain a map: \begin{equation} \begin{aligned} \Der:\Alb_{\text{\normalfont sub}}(\mu)&\to\wder\mu.\\ \albrep.&\mapsto D_{\albrep.}. \end{aligned} \end{equation} \end{thm} \begin{proof} Let $\albrep.=(P,\nu)$ be a $C$-Lipschitz Alberti representation of $\mu$; considering a $C$-Lipschitz fragment $\gamma$, from the estimate: \begin{equation} \left|(f\circ\gamma)'(t)\right|\le \glip f.\metdiff\gamma(t) \le C\glip f., \end{equation} we obtain \begin{equation} \left|\int_X gD_{\albrep.}f\,d\mu\right|\le C\glip f.\onenorm g.; \end{equation} in particular, we conclude that $D_{\albrep.}:\lipalg X.\to\elleinfty\mu.$ is a linear operator with norm bounded by $C$. \par That $D_{\albrep.}$ satisfies the product rule follows from a direct computation. \par We show that $D_{\albrep.}$ is weak* continuous; let $f_n\xrightarrow{\text{w*}} f$ in $\lipalg X.$, and consider a bounded Borel function $g\in L^1(\mu)$. Now, $\frac{d}{dt}$ is a derivation of $\real$ with respect to the Lebesgue measure; as $\mpush\gamma^{-1}.\nu_\gamma$ is absolutely continuous with respect to the Lebesgue measure, the function \begin{equation} \frac{d\mpush\gamma^{-1}.\nu_\gamma}{d{\rm Leb}}g \end{equation} is integrable, and so \begin{equation} \lim_{n\to\infty} \int_{\dom\gamma} (f_n\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t) = \int_{\dom\gamma} (f\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t). \end{equation} Assume that $\glip f_n.,\glip f.\le L$ and let \begin{align} h_n(\gamma)&=\int_{\dom\gamma} (f_n\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t),\\ h(\gamma)&=\int_{\dom\gamma} (f\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t),\\ H(\gamma)&=CL\int_{\dom\gamma}|g|\circ \gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t); \end{align} note that $h_n\to h$ pointwise, $|h_n|,|h|\le H$. The map $H$ is Borel by Definition \ref{defn:Alberti_rep} of Alberti representations, and the maps $h_n, h$ and $H$ are Borel by Lemma \ref{lem:meas_fix}; we thus conclude by the Lebesgue's dominated convergence Theorem that \begin{multline} \lim_{n\to\infty}\int_{\frags(X)}dP(\gamma)\int_{\dom\gamma} (f_n\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t)\\ =\int_{\frags(X)}dP(\gamma)\int_{\dom\gamma} (f\circ \gamma)'(t)g\circ\gamma(t)\,d(\mpush\gamma^{-1}.\nu_\gamma)(t), \end{multline} showing weak* continuity. \par The defition of $\Der$ is well-posed because $\wder\mu\mrest S.$ can be canonically identified with $\chi_S\wder\mu.$. \end{proof} \par We now describe what happens if $\albrep.$ is in the $f$-direction of some cone field. \begin{thm}\label{thm:directional_cone} Suppose $\albrep.=(P,\nu)\in\Alb(\mu)$ is in the $f$-direction of $\cone(v,\alpha)$; then \begin{equation} \label{eq:directional_cone} D_{\albrep.}f(x)\in\cone\left(v(x),\alpha(x)\right)\quad\text{(for $\mu$-a.e.~$x$).} \end{equation} In particular, if the Alberti representations $\{\albrep i.\}_{i=1}^k$ of $\mu$ are in the $f$-directions of independent cone fields $\{\cone(v_i,\alpha_i)\}_{i=1}^k$, the derivations $\{D_{\albrep i.}\}_{i=1}^k$ are independent. \end{thm} \begin{proof} Note that if $U\subset X$ is Borel, $D_{\albrep.\mrest U}=\chi_UD_{\albrep.}$; therefore we can assume that $\mu$ is finite with $\mu(X)>0$ and $\tan\alpha\in L^1(\mu)$. If $\varphi\in\bborel X.$% \nomenclature[Borel]{$\bborel X.$}{bounded real-valued Borel functions defined on $X$}% is nonnegative with $\int\varphi\,d\mu>0$, \begin{equation} \label{eq:directional_cone1} \begin{split} \int\varphi\tan\alpha\langle v,D_{\albrep.}f\rangle\,d\mu&=\int_{\frags(X)}dP(\gamma)\int \varphi\tan\alpha \langle v, \partial_\gamma f\rangle\,d\nu_\gamma\\ &>\int_{\frags(X)}dP(\gamma)\int \varphi\|\pi_v^\perp \partial_\gamma f\|_2\,d\nu_\gamma, \end{split} \end{equation} were we used that $\mu(X)>0$ and $\int\varphi\,d\mu>0$. On the other hand, $\forall\epsi>0$ there are countably many $(K_j,w_j)$ such that \begin{enumerate} \item The $K_j\subset X$ are disjoint compact sets and $\mu\left(X\setminus\bigcup_jK_j\right)=0$. \item Each $w_j$ is a unit vector field orthogonal to $v$ on $K_j$. \item The inequality $\|\pi_v^\perp D_{\albrep.}f\|_2\le \langle w_j,D_{\albrep.}f\rangle+\epsi$ holds on $K_j$. \end{enumerate} Thus \begin{equation} \label{eq:directional_cone2} \begin{split} \int \varphi \|\pi_v^\perp D_{\albrep.}f\|_2\,d\mu &\le\sum_j\int_{K_j} \varphi\langle w_j,D_{\albrep.}f\rangle\,d\mu+\max\varphi\,\epsi\,\mu(X)\\ &=\sum_j\int_{\frags(X)}dP(\gamma)\int_{K_j}\varphi\langle w_j,\partial_\gamma f\rangle\,d\nu_\gamma+\max\varphi\,\epsi\,\mu(X)\\ &\le\int_{\frags(X)}dP(\gamma)\int\varphi \|\pi_v^\perp \partial_\gamma f\|_2\,d\nu_\gamma +\max\varphi\,\epsi\,\mu(X); \end{split} \end{equation} letting $\epsi\searrow0$ and combining \eqref{eq:directional_cone1} and \eqref{eq:directional_cone2} \begin{equation}\label{eq:directional_cone3} \int \varphi\left(\tan\alpha\langle v,D_{\albrep.}f\rangle-\|\pi_v^\perp D_{\albrep.}f\|_2\right)\,d\mu>0, \end{equation} which implies the result. \end{proof} We now introduce the notion of \textbf{effective speed} of an Alberti representation. \begin{defn}\label{defn:alberti_speed} Let $\albrep.=(P,\nu)\in\Alb(\mu)$ an Alberti representation and define the measure \begin{equation}\label{eq:alberti_speed} \Sigma_{\albrep.}=\int_{\frags(X)} \metdiff\gamma\circ\gamma^{-1}\,\nu_\gamma\,dP(\gamma); \end{equation} an \textbf{effective speed of $\albrep.$} is a Borel representative $\sigma_{\albrep.}$ of the Radon-Nikodym derivative of $\Sigma_{\albrep.}$ with respect to $\mu$. \end{defn} We now describe what happens if one knows a lower bound of the $f$-speed of $\albrep.$. \begin{thm}\label{thm:directional_speed} If $\albrep.\in\Alb(\mu)$ has $f$-speed $\ge\delta$, then \begin{equation} \label{eq:directional_speed} D_{\albrep.}f(x)\ge\delta(x)\sigma_{\albrep.}(x)\quad\text{(for $\mu$-a.e.~$x$).} \end{equation} \end{thm} \begin{proof} Let $\varphi\in L^1(\mu)$ nonnegative; then \begin{equation} \begin{split} \int \varphi D_{\albrep.}f\,d\mu &= \int dP(\gamma) \int \varphi \partial_\gamma f\,d\nu_\gamma\\ &\ge\int dP(\gamma) \int \varphi\delta\,d\left(\metdiff\gamma\circ\gamma^{-1}\,\nu_\gamma\right)\\ &=\int \varphi\delta\sigma_{\albrep.}\,d\mu, \end{split} \end{equation} from which the result follows. \end{proof} As we deal with parametrized fragments, it is useful to know how Alberti representations are affected by affine reparametrizations of the fragments. \begin{lem}\label{lem:alberti_affine_action} For $a\in\real\setminus\{0\}$ and $b\in\real$ let $\tau_{a,b}:\real\to\real$ denote the homeomorphism \begin{equation} \tau_{a,b} (x)=ax+b; \end{equation} then \begin{equation} \begin{aligned} \pullb\tau_{a,b} .:\frags(X) &\to\frags(X)\\ \gamma&\mapsto\gamma\circ\tau_{a,b}, \end{aligned} \end{equation} is a homeomorphism.\par If $\mu\mrest A$ admits an Alberti representation $\albrep .=(P,\nu)$ then \begin{equation}\pullb\tau_{a,b}.\albrep.=\left(\mpush{(\pullb\tau_{a,b}.)}. P,\nu\circ(\pullb\tau_{a,b} .)^{-1}\right) \end{equation} is an Alberti representation and, moreover, if \begin{enumerate} \item The Alberti representation $\albrep.$ is $C$-Lipschitz ($(C,D)$-biLipschitz) then $\pullb\tau_{a,b}.\albrep.$ is $C|a|$-Lipschitz ($(C|a|,\allowbreak D|a|\-)$-biLipschitz). \item If $\albrep.$ has $f$-speed $\ge\delta$ then $\pullb\tau_{a,b}.\albrep.$ has $\sgn a\cdot f$-speed $\ge\delta$. \item If $\albrep.$ is in the $f$-direction of $\cone(v,\alpha)$, then $\pullb\tau_{a,b}.\albrep.$ is in the $f$-direction of $\cone(\sgn a\cdot v,\alpha)$. \end{enumerate} Finally, if $\albrep.$ is Lipschitz, $D_{\pullb\tau_{a,b} .\albrep.}=aD_{\albrep.}$. \end{lem} \begin{proof} Note that $\pullb\tau_{a,b}.$ is $\max(|a|,\frac{1}{|a|})$-Lipschitz and with inverse $\tau_{1/a,-b/a}$; if $B$ is Borel, from $\gamma\mapsto\nu_\gamma(B)$ being Borel, it follows that $\gamma\mapsto\nu({(\pullb\tau_{a,b}.)^{-1}\gamma})(B)$ is Borel. Then \begin{equation} \mu\mrest A(B)=\int_{\frags(X)}\nu(\gamma(B))\,dP(\gamma) =\int_{\frags(X)}\nu({(\pullb\tau_{a,b}.)^{-1}\gamma})(B)\,d\mpush{(\pullb\tau_{a,b}.)}.P(\gamma), \end{equation} which shows that $\pullb\tau_{a,b}.\albrep.$ is an Alberti representation of $\mu$. Points (1)--(3) follow observing that $\mpush{(\pullb\tau_{a,b}.)}.P$ is supported on $\pullb\tau_{a,b}.(\spt P)$. Finally, \begin{equation} \begin{split} \int D_{\pullb\tau_{a,b} .\albrep.}f\,g\,d\mu &= \int_{\frags(X)} d\mpush{(\pullb\tau_{a,b} .)}.P(\gamma) \int_{\dom\gamma}(f\circ\gamma)'(t) \,(g\circ\gamma)(t)\\ &\times d\left(\mpush\gamma^{-1}.\nu\left({(\pullb\tau_{a,b} .)^{-1}\gamma}\right)\right)(t)\\ &=\int_{\frags(X)} dP(\tilde\gamma) \int_{\dom\gamma}(f\circ\tilde\gamma\circ\tau_{a,b} )'(t) \,(g\circ\tilde\gamma\circ\tau_{a,b} )(t)\\ &\times d\left(\mpush(\tilde\gamma\circ\tau_{a,b} )^{-1}.\nu({ \tilde\gamma})\right)(t), \end{split} \end{equation} where $\gamma=\pullb\tau_{a,b}.\tilde\gamma$; then \begin{equation} \begin{split} \int D_{\pullb\tau_{a,b} .\albrep.}f\,g\,d\mu &= \int_{\frags(X)} dP(\tilde\gamma) \int_{\dom\gamma}a(f\circ\tilde\gamma)'\circ\tau_{a,b} (t) \,(g\circ\tilde\gamma\circ\tau_{a,b} )(t)\\ &\times\,d\left(\mpush\tau_{a,b} ^{-1}.\mpush\tilde\gamma^{-1}.\nu({ \tilde\gamma})\right)(t)\\ &=\int_{\frags(X)} dP(\tilde\gamma) \int_{\dom\tilde\gamma}a(f\circ\tilde\gamma)'(t) \,(g\circ\tilde\gamma)(t)d\left(\mpush\tilde\gamma^{-1}.\nu({\tilde\gamma})\right)(t)\\ &=\int aD_{\albrep.}f\,g\,d\mu. \end{split} \end{equation} \end{proof} To illustrate the lack of injectivity of $\Der$ we provide a useful reparametrization result which allows to use fragments with domain contained in a prescribed interval. \begin{lem}\label{lem:domain_reduction} Let $\albrep.\in\Alb(\mu\mrest S)$; then there is $ \albrep.'=(P',\nu')\in\Alb(\mu\mrest S)$ with \begin{equation}P'\left(\frags(X)\setminus\frags(X,I)\right)=0 \end{equation} and such that \begin{enumerate} \item We have the identity $D_{\albrep.'}=D_{\albrep.}$. \item If $\albrep.$ satisfies a set of conditions \albcond, so does $\albrep.'$. \end{enumerate} \end{lem} \begin{proof} We can assume that $\albrep.=(P,\nu)\in\Alb(\mu)$ and that $X$ is compact. Recall that $\haus(\real)$ denotes the set of nonempty compact subsets of $\real$ with the Vietoris topology. The maps: \begin{equation} \begin{aligned} \max,\min:\haus(\real)&\to\real\\ K&\mapsto\max_{x\in K} x,\min_{x\in K} x \end{aligned} \end{equation} are continuous. In particular, \begin{equation} \begin{aligned} \Delta_1:\haus(\real)&\to\haus(\real)\\ K&\mapsto K\cap[\min(K), \min(K)+\lebmeas(I)] \end{aligned} \end{equation} is continuous. For $n>1$ let $O_n$ denote the open set \begin{equation} O_n=\left\{K\in\haus(\real): \max(K)-\min(K)>(n-1)\lebmeas(I)\right\}, \end{equation} and $\Delta_n$ the continuous map \begin{equation} \begin{aligned} \Delta_n:O_n&\to\haus(X)\\ K&\mapsto\Delta_1\left(\overline{K\setminus\bigcup_{k=1}^{n-1}\Delta_k(K)}\right). \end{aligned} \end{equation} Note that the set \begin{equation} F_i=\left\{\gamma\in\frags(X):\lebmeas\left(\Delta_i(\dom\gamma)\right)>0\right\} \end{equation} is Borel and the map $R_i:F_i\to\frags(X,I)$ \begin{equation} \gamma\mapsto\pullb\tau_{1,\min(I)-\min(\dom\gamma)}.\gamma|_{\Delta_i(\dom\gamma)} \end{equation} is Borel. \par The following discussion applies only for those $i$ for which $P(F_i)>0$. We apply the Disintegration Theorem \cite[452O]{fremlin4} for \begin{equation} R_i:\left(F_i,\frac{P\mrest F_i}{P(F_i)}\right)\to\left(\frags(X,I),\underbrace{\frac{\mpush R_i.P\mrest F_i}{P(F_i)}}_{P_i}\right), \end{equation} which is, in the terminology of \cite{fremlin4}, an \textbf{inverse-measure-preserving} function. By \cite[434K(b)]{fremlin4} $\haus(\real\times X)$, being Polish, is a \textbf{Radon measure space} and, as $F_i$ is a Borel subset of $\haus(\real\times X)$, it is also a Radon measure space by \cite[434F(c)]{fremlin4}. On the other hand, $\left(\frags(X,I), P_i\right)$ is \textbf{strictly localizable} in the terminology of \cite{fremlin4} because the measure $P_i$ is finite. The Disintegration Theorem yields Radon probability measures $\{\pi_{\gamma}\}_{\gamma\in \frags(X,I)}$ on $\frags(X)$ such that \begin{align} \frac{P\mrest F_i}{P(F_i)}&=\int_{\frags(X,I)}\pi_{\gamma}\,dP_i(\gamma)\\ \pi_\gamma\left(R_i^{-1}\left(\{\gamma\}\right)\right)&=1\quad\text{(for $P_i$-a.e.~$\gamma$)}. \end{align} Consider the weakly measurable maps $\tilde\nu_i:\frags(X)\to M(X)$ \begin{equation} \tilde\nu_i(\gamma)= \begin{cases} 0&\text{if $\gamma\not\in F_i$,}\\ {P(F_i)}\nu_\gamma\chi_{\gamma\left(\Delta_i(\dom\gamma)\right)}&\text{otherwise;} \end{cases} \end{equation} and the measures \begin{equation} \mu_i=\frac{1}{P(F_i)}\int_{F_i}\tilde\nu_i(\gamma)\,dP(\gamma); \end{equation} if we let $\nu_i:\frags(X,I)\to M(X)$ be given by \begin{equation} \nu_i(\gamma)=\int_{\frags(X)}\tilde\nu_i(\tilde\gamma)\,d\pi_\gamma(\tilde\gamma), \end{equation} $\albrep i.=(P_i,\nu_i)$ is an Alberti representation of $\mu_i$ supported on the closure of $\im R_i$. In particular, if $\albrep.$ satisfies \albcond, so does $\albrep i.$. The derivation $D_i\in\wder\mu_i.$ given by \begin{equation} \int gD_if\,d\mu_i=\frac{1}{P(F_i)}\int_{F_i}dP(\gamma)\int g\partial_\gamma f\,\tilde\nu_i(\gamma) \end{equation} is naturally identified with an element of $\wder\mu.$ because $\mu_i\ll\mu$; in particular, $D_{\albrep.}=\sum_iD_i$. The calculation \begin{multline} \frac{1}{P(F_i)} \int_{F_i}dP(\tilde\gamma)\int_{\dom\tilde\gamma}\left(f\circ\tilde\gamma\right)'(t)\, \left(g\circ\tilde\gamma\right)(t)\,d\left(\mpush\tilde\gamma^{-1}.\tilde\nu_i(\tilde\gamma)\right)(t)\\ = \int_{\frags(X,I)}dP_i(\gamma)\int_{\frags(X)}d\pi_\gamma(\tilde\gamma) \int_{\dom\tilde\gamma}\left(f\circ\tilde\gamma\right)'(t)\, \left(g\circ\tilde\gamma\right)(t)\,d\left(\mpush\tilde\gamma^{-1}.\tilde\nu_i(\tilde\gamma)\right)(t)\\ = \int_{\frags(X,I)}dP_i(\gamma)\int_{\frags(X)}d\pi_\gamma(\tilde\gamma) \int_{\dom\gamma}\left(f\circ\gamma\right)'(t) \left(g\circ\gamma\right)(t)\,d\left(\mpush\gamma^{-1}.\tilde\nu_i(\tilde\gamma)\right)(t)\\ = \int_{\frags(X,I)}dP_i(\gamma) \int_{\dom\gamma}\left(f\circ\gamma\right)'(t) \left(g\circ\gamma\right)(t)\,d\left(\mpush\gamma^{-1}.\nu_i(\gamma)\right)(t)\\ =\int gD_{\albrep i.}f\,d\mu_i \end{multline} shows that $D_i=D_{\albrep i.}$. Let $P'$ the probability measure on $\frags(X,I)$ given by \begin{equation} P'=\sum_i2^{-i}P_i \end{equation} and $\varphi_i$ a Borel representative of the Radon-Nikodym derivative of $P_i$ with respect to $P'$; let \begin{equation} \begin{aligned} \nu':\frags(X,I)&\to M(X)\\ \gamma&\mapsto\sum_i\nu_i(\gamma)\varphi_i(\gamma); \end{aligned} \end{equation} then $\albrep.'=(P',\nu')$ is an Alberti representation of $\sum_i\mu_i=\mu$ and, if $\albrep.$ satisfies \albcond, so does $\albrep.'$. Moreover, \begin{equation} D_{\albrep.'}=\sum_iD_{\albrep i.}=\sum_iD_i=D_{\albrep.}. \end{equation} \end{proof} \subsection{From Derivations to Alberti representations}\label{subsec:derivations_to_alberti} The goal of this Subsection is to prove Theorems \ref{derivation_alberti}, \ref{thm:weak*density} and \ref{thm:fin_gen_surjectivity} and Corollaries \ref{der-alb} and \ref{cor:mu_arb_cone}. Recall that we deal with separable metric spaces. \begin{thm}\label{derivation_alberti} Let $X$ be a metric space and $\mu$ a Radon measure on $X$. Consider a Borel $V\subset X$, derivations $\{D_1,\cdots,D_k\}\subset\wder\mu.$ and a Lipschitz function $g:X\to\real^k$ such that $D_ig_j=\delta_{i,j}\chi_V$. Then for each $\epsi>0$, unit vector $w\in{\mathbb S}^{k-1}$, angle $\alpha\in(0,\pi/2)$ and speed parameter $\sigma\in(0,1)$, the measure $\mu\mrest V$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation in the $g$-direction of $\cone(w,\alpha)$ with $\langle w,g\rangle$-speed \begin{equation}\label{eq:speed_almost_optimal} \ge\frac{\sigma}{\locnorm D_w,{\wder\mu\mrest V.}.+(1-\sigma)}. \end{equation} \end{thm} The proof of Theorem \ref{derivation_alberti} relies on an approximation scheme for Lipschitz functions, Theorem \ref{onedimapprox_multi}. We state the relevant definitions and the approximation scheme here, and defer the proof to Subsection \ref{subsec:an-appr-scheme}. We define the following classes of fragments: \begin{defn}\label{def:classes_frags} For $\delta>0$ and $f:X\to\real$ Lipschitz we define: \begin{equation} \frags(X,f,\delta)=\left\{\gamma\in\frags(X): \text{$(f\circ\gamma)'(t)\ge\delta\metdiff\gamma(t)$ for $\lebmeas$-a.e.~$t\in\dom\gamma$}\right\}; \end{equation} For $\delta>0$, $f:X\to\real^q$ Lipschitz, $w\in{\mathbb{S}}^{q-1}$ and $\alpha\in(0,\pi/2)$, we define: \begin{multline} \frags(X,f,\delta,w,\alpha)=\bigl\{\gamma\in\frags(X,\langle w,f\rangle,\delta):\text{$(f\circ\gamma)'(t)\in\cone(w,\alpha)$} \\ \text{for $\lebmeas$-a.e.~$t\in\dom\gamma$}\bigr\}. \end{multline} \end{defn} \begin{defn}\label{def:locally_lipschitz} Let $f:X\to\real$ be a Lipschitz function, $S\subset X$ Borel and $\mu$ a Radon measure on $X$. We say that $f$ is locally $\delta$-Lipschitz on $S$ if for each $x\in S$ there is an $r>0$ such that the restriction $f|\ball x,r.$ is $\delta$-Lipschitz. Finally, we say that $f$ is $\mu$-a.e.~locally $\delta$-Lipschitz on $S$ if for $\mu$-a.e.~$x\in S$ there is an $r>0$ such that the restriction $f|\ball x,r.$ is $\delta$-Lipschitz. \end{defn} \begin{thm}\label{onedimapprox_multi} Let $X$ be a compact metric space, $f:X\to\real^q$ $L$-Lipschitz and $S\subset X$ compact. Let $\mu$ be a Radon measure on $X$. Assume that $S$ is $\frags(X,f,\delta,w,\alpha)$-null. Denoting by $d_{\delta,\alpha}$, $\tilde d_{\delta,\alpha}$ the distances \begin{equation} \begin{aligned} d_{\delta,\alpha}(x,y)&=\delta\dist x,y.+\cot\alpha\|\pi_w^\perp f(x)-\pi_w^\perp f(y)\|_2,\\ \tilde d_{\delta,\alpha}&=\max(\sqrt{q}L\dist\cdot,\cdot.,d_{\delta,\alpha}), \end{aligned} \end{equation} there are $(1,\tilde d_{\delta,\alpha})$-Lipschitz functions $g_k\xrightarrow{\text{w*}} \langle w,f\rangle$ with $g_k$ $\mu$-a.e.~locally $(1,d_{\delta,\alpha})$-Lipschitz on $S$. \end{thm} The proof of Theorem \ref{derivation_alberti} relies on the following technical Lemmas \ref{lem:loc_estimate_dist} and \ref{lem:loc_estimate_fnull}. \begin{lem}\label{lem:loc_estimate_dist} Let $X$ be a separable metric space and $\mu$ a Radon measure on $X$. Consider a Borel $S\subset X$, a Lipschitz function $f:X\to\real$, a derivation $D\in\wder\mu.$, and a (pseudo)-distance function $d'$ on $X$ satisfying $d'\le Cd$, where $d$ denotes the metric on $X$. Assume that $f|S$, regarded as a function defined on the metric space $(S,d)$, is $\mu\mrest S$-a.e.~locally $(1,d')$-Lipschitz and that, for some $c\ge0$, \begin{equation} \chi_S\left|Dd'(\cdot,p)\right|\le c\quad(\forall p\in S); \end{equation} then \begin{equation}\label{eq:loc_estimate_dist}\chi_S|Df|\le c. \end{equation} \end{lem} \begin{proof} As $\mu$ is Radon, it suffices to show \eqref{eq:loc_estimate_dist} when $S$ is replaced by a compact subset $K$. Fix a ball $B\subset X$ centred on $K$ such that $f$ is $(1,d')$-Lipschitz in $\bar B\cap K$. Let $\mathscr{N}_\eta$ a finite $\eta$-net in $\bar B\cap K$ (which is compact) and $f_\eta$ the McShane extension of $f|\mathscr{N}$ to $X$: \begin{equation} f_\eta(x)=\min_{p\in\mathscr{N_\eta}}\left(f(p)+d'(x,p)\right); \end{equation} note that $f_\eta$ is $(C,d)$-Lipschitz. Let \begin{equation} \tilde f_\eta(x)=\min\left(\max\left(f_\eta(x),-\sup_{\bar B}|f_\eta|\right),\sup_{\bar B}|f_\eta|\right), \end{equation} so that $\tilde f_\eta\in\lipalg X.$ and $\tilde f_\eta=f_\eta$ on $\bar B$. Choose a sequence $\eta_n\searrow0$. As $X$ is separable, the predual of $X$ is separable (compare the description of the predual using \emph{molecules} given in \cite[Subec.~2.2]{weaver_book99}) and thus the weak* topology on $\lipalg X.$ is metrizable on bounded subsets of $\lipalg X.$; by Banach-Alaoglu, the sequence $\tilde f_{\eta_n}$ has a convergent subsequence $\tilde f_{\eta_n}\xrightarrow{\text{w*}} \tilde f$ in $\lipalg X.$\footnote{This argument is a version of Ascoli-Arzel\`a in disguise}. Note that $\tilde f = f$ on $\bar B\cap K$. For a fixed $\eta$ there are finitely many closed sets $C_p\subset\bar B\cap K$ ($p\in\mathscr{N}$) which cover $\bar B\cap K$ and such that, for each $ x\in C_p$, \begin{equation} f_\eta(x)=f(p)+d'(x,p); \end{equation} this implies, by Lemma \ref{lem:locality_derivations}, \begin{equation} \chi_{C_p\cap K}|D\tilde f_\eta|=\chi_{C_p\cap K}\left|Dd'(\cdot,p)\right|\le c; \end{equation} thus $\left\|\chi_K D\tilde f_\eta\right\|_{L^\infty(\mu\mrest\bar B)}\le c$. Using lower semicontinuity of the norm under weak* convergence and Lemma \ref{lem:locality_derivations}, it follows that \begin{equation}\left\|\chi_K D\tilde f\right\|_{L^\infty(\mu\mrest\bar B)}=\left\|\chi_K Df\right\|_{L^\infty(\mu\mrest\bar B)}\le c \end{equation} which implies \eqref{eq:loc_estimate_dist}. \end{proof} \begin{lem}\label{lem:loc_estimate_fnull} Let $X$ be a separable metric space and $\mu$ a Radon measure on $X$. Consider a Lipschitz function $g:X\to\real^k$ , a derivation $D\in\wder\mu.$ and a Borel $S\subset X$. Assume that $S$ is $\frags(X,\allowbreak g,\allowbreak\delta,\allowbreak w,\alpha)$-null, where $w\in{\mathbb S}^{k-1}$ and $\alpha\in(0,\pi/2)$, and that, for each $u\in {\mathbb S}^{k-1}$ orthogonal to $w$, one has \begin{equation}\chi_S\left|D\langle u,g\rangle\right|\le\epsi\locnorm D,{\wder\mu.}.; \end{equation} then one has \begin{equation} \label{eq:loc_estimate_fnull1} \chi_S\left|D\langle w,g\rangle\right|\le\left(\delta+(k-1)\epsi\cot\alpha\right)\locnorm D,{\wder\mu.}.. \end{equation} \par In particular, if the derivations $\{D_1,\cdots,D_k\}\subset\wder\mu.$ satisfy $D_ig_j=\delta_{i,j}$ $\mu$-a.e., letting $D_w=\sum_{i=1}^kw_iD_i$, one has \begin{equation}\label{eq:loc_estimate_fnull2} \chi_S\left|D_w\langle w,g\rangle\right|\le\delta\locnorm D_w,{\wder\mu\mrest S.}.. \end{equation} \end{lem} \begin{proof} The second part, \eqref{eq:loc_estimate_fnull2}, follows from the first, \eqref{eq:loc_estimate_fnull1} as $D_w$ annihilates $\langle u,g\rangle$ for $u$ orthogonal to $w$. As $\mu$ is Radon, it suffices to show \eqref{eq:loc_estimate_fnull1} for $S$ replaced by a compact subset $K$. \par Consider the metric space $(K,d)$ and choose an orthonormal basis $\mathscr{B}$ of the plane orthogonal to $w$; let \begin{equation} \begin{aligned} d'(x,y)&=\delta\dist x,y.+\cot\alpha\sum_{u\in\mathscr{B}}|\langle u,g(x)-g(y)\rangle|,\\ \tilde d'(x,y)&=\max(\sqrt{k}\glip g.d(x,y),d'(x,y)). \end{aligned} \end{equation} As $K$ is $\frags(K,g,\delta,w,\alpha)$-null, by Theorem \ref{onedimapprox_multi} there are functions $f_n$ \begin{enumerate} \item Which are $(1,\tilde d')$-Lipschitz and $\mu\mrest K$-a.e.~locally $(1,d')$-Lipschitz on $K$. \item Which converge weak* to $\langle w,g\rangle$ in $\lipalg K.$. \end{enumerate} Reasoning as in the proof of Lemma \ref{lem:loc_estimate_dist}, we can take extensions $\tilde f_n\in\lipalg X.$ of the $f_n$ with $\tilde f_n\xrightarrow{\text{w*}} \tilde f$\footnote{After passing to a subsequence} and $\tilde f = f$ on $K$. Let $p\in K$; as $d(\cdot, p)$ is $1$-Lipschitz, \begin{equation} \label{eq:D_dist1} |Dd(\cdot,p)|\le\locnorm D,{\wder\mu.}.; \end{equation} on the other hand, by \cite[Lem.~7.2.2]{weaver_book99}\footnote{This is a consequence of weak* continuity and the classical approximation of functions in $\lipalg [0,1].$ by polynomials} \begin{equation} \label{eq:D_dist2} \left|\,D\left|\left\langle u,g(\cdot)-g(p)\right\rangle\right|\,\right|\le \left|D\left\langle u,g(\cdot)-g(p)\right\rangle\right|; \end{equation} but by hypothesis, \begin{equation} \label{eq:D_dist3} \chi_K\left|D\left\langle u,g(\cdot)-g(p)\right\rangle\right|\le\epsi\locnorm D,{\wder\mu.}.; \end{equation} putting together \eqref{eq:D_dist1}, \eqref{eq:D_dist2} and \eqref{eq:D_dist3}: \begin{equation} \chi_K\left|Dd'(\cdot,p)\right|\le\left(\delta+(k-1)\epsi\cot\alpha\right)\locnorm D,{\wder\mu.}.. \end{equation} By Lemma \ref{lem:loc_estimate_dist} applied to the $\tilde f_n$, \begin{equation} \chi_K\left|D\tilde f_n\right|\le\left(\delta+(k-1)\epsi\cot\alpha\right)\locnorm D,{\wder\mu.}.; \end{equation} by lower semicontinuity of the norm under weak* convergence, \begin{equation} \chi_K\left|D\tilde f\right|\le\left(\delta+(k-1)\epsi\cot\alpha\right)\locnorm D,{\wder\mu.}.; \end{equation} finally by Lemma \ref{lem:locality_derivations}, \begin{equation} \chi_K\left|D f\right|\le\left(\delta+(k-1)\epsi\cot\alpha\right)\locnorm D,{\wder\mu.}.. \end{equation} \end{proof} \begin{proof}[Proof of Theorem \ref{derivation_alberti}] By Theorem \ref{alb_glue} and by Lusin's Theorem we can replace $V$ by a compact subset $K$ on which, for some $\eta_0,\eta_1>0$, \begin{align} \|\chi_KD_w\|_{\wder\mu.}&\le\inf_{x\in K}\locnorm D_w,{\wder\mu.}.(x)+\eta_0,\\ \sup_{x\in K}\frac{\sigma}{\locnorm D_w,{\wder\mu.}.(x)+(1-\sigma)}&\le\frac{1}{\inf_{x\in K}\locnorm D_w,{\wder\mu.}.(x)+\eta_0}-\eta_1; \end{align} we let $\delta_{\eta_1}=\frac{1}{\|\chi_KD_w\|_{\wder\mu.}}-\eta_1$ and $\alpha'\in(0,\alpha)$. We want to show that if a Borel $S\subset K$ is $\frags(K,g,\delta_{\eta_1},w,\alpha')$-null, then it is $\mu$-null. Taking $X=K$ in Lemma \ref{lem:loc_estimate_fnull}, we get \begin{equation} \begin{split} \chi_S=\chi_SD_w\langle w,g\rangle&\le\delta_{\eta_1}\locnorm D_w,{\wder\mu.}.\\ &\le\delta_{\eta_1}\|\chi_KD_w\|_{\wder\mu.}\\ &\le1-\eta_1\|\chi_KD_w\|_{\wder\mu.}\\ &<1, \end{split} \end{equation} which implies $\mu(S)=0$. By Theorem \ref{alberti_rep_prod} the measure $\mu\mrest K$ admits an Alberti representation in the $g$-direction of $\cone(v,\alpha)$ with \begin{equation} \begin{split} \text{$\langle w,g\rangle$-speed}&\ge\frac{1}{\|\chi_KD_w\|_{\wder\mu.}}-\eta_1\\ &\ge\frac{1}{\inf_{x\in K}\locnorm D_w,{\wder\mu.}.(x)+\eta_0}-\eta_1\\ &\ge\sup_{x\in K}\frac{\sigma}{\locnorm D_w,{\wder\mu.}.(x)+(1-\sigma)}. \end{split} \end{equation} \end{proof} \begin{cor}\label{der-alb} Let $V$ be Borel and assume that $\wder\mu\mrest V.$ contains $k$-independent de\-ri\-va\-tions. Then there is a Borel partition $V=\bigcup_{\alpha}V_\alpha$ and, for each $\alpha$, there is a $1$-Lipschitz function $f_\alpha:V\to\real^k$ such that, for each $\epsi>0$ and for all Borel maps $w:X\to{\mathbb S}^{k-1}$ and $\theta:X\to(0,\pi/2)$, the measure $\mu$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation $\albrep.$ with $\albrep.\mrest V_\alpha$ in the $f_\alpha$-direction of the cone field $\cone(w,\theta)$. \par In particular, the measure $\mu\mrest V_\alpha$ admits Alberti representations in the $f_\alpha$-directions of independent cone fields $\cone_1,\cdots,\cone_k$. \end{cor} \begin{proof Applying Corollary \ref{cor:pseudoduality} one finds disjoint Borel sets $\{V_\alpha\}$ with $V=\bigcup_\alpha V_\alpha$ and such that, for each $\alpha$, there are $1$-Lipschitz functions $f_\alpha:X\to\real^k$ and derivations $\{D_{\alpha,1},\cdots,D_{\alpha,k}\}\subset\wder\mu.$ with \begin{equation} D_{\alpha,i}f_{\alpha,j}=\chi_{V_\alpha}\delta_{i,j}. \end{equation} In the case in which $w$ and $\theta$ are constant, the result follows applying Theorem \ref{derivation_alberti}. If $w$ and $\theta$ are not constant, one can find disjoint compact sets $\{C_{\alpha,\beta}\}_{\beta}$: \begin{enumerate} \item Each $C_{\alpha,\beta}$ is a subset of $V_\alpha$. \item We have the identity $\mu\left(V_\alpha\setminus\bigcup_\beta C_{\alpha,\beta}\right)=0$. \item For each $\beta$ there are a unit vector $w_\beta\in{\mathbb S^{k-1}}$ and an angle $\theta_\beta\in(0,\pi/2)$ with $\cone(w_\beta,\theta_\beta)\subset \cone(w(x),\theta(x))$ for each $x\in C_{\alpha,\beta}$. \end{enumerate} One then applies Theorem \ref{alb_glue} to glue together the Alberti representations of the measures $\mu\mrest C_{\alpha,\beta}$ in the $f_\alpha$-directions of the cone fields $\cone(w_\beta,\theta_\beta)$. \end{proof} \begin{cor}\label{cor:mu_arb_cone} Let $f:X\to\real^k$ be Lipschitz and $\mu$ a Radon measure on $X$ admitting Alberti representations $\albrep 1.,\cdots,\albrep k.$ in the $f$-directions of independent cone fields $\cone_1,\cdots,\cone_k$. Then, for each $\epsi>0$ and for all Borel maps $w:X\to{\mathbb S}^{k-1}$ and $\theta:X\to(0,\pi/2)$, the measure $\mu$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation in the $f$-direction of $\cone(w,\theta)$. \end{cor} \begin{proof By Theorem \ref{alberti_rep_prod} it is possible to assume that the Alberti representations are biLipschitz. According to Theorem \ref{thm:directional_cone} the corresponding derivations $D_{\albrep i.}$ are independent. By the gluing principle for Alberti representations, Theorem \ref{alb_glue}, it suffices to prove the statement for $w$ and $\theta$ constant and to show that there is a Borel partition $\spt\mu=\bigcup_\alpha U_\alpha$ with each $\mu\mrest U_\alpha$ admitting an Alberti representation in the $f$-direction of $\cone(w,\theta)$. As the cone fields $\cone_i$ are independent, the matrix $\left(D_{\albrep i.}f_j\right)_{i,j=1}^k$ is invertible $\mu$-a.e. This allows to pass to a Borel partition $\spt\mu=\bigcup_\alpha U_\alpha$ such that, for each $U_\alpha$, there are derivations $D_{\alpha,i}$ with $D_{\alpha,i}f_j=\delta_i^j$ $\mu\mrest U_\alpha$-a.e. This partition is necessary because the inverse of $\left(D_{\albrep i.}f_j\right)_{i,j=1}^k$ does not need to have its entries in $L^\infty(\mu)$. However, there are disjoint Borel sets $U_\alpha$ with $\mu(\spt\mu\setminus\bigcup_\alpha U_\alpha)=0$ and such that on each $U_\alpha$ the inverse of $\left(D_{\albrep i.}f_j\right)_{i,j=1}^k$ has its entries in $L^\infty(\mu)$. In fact, the entries of $\left(D_{\albrep i.}f_j\right)_{i,j=1}^k$ are in $L^\infty(\mu)$ and one needs a lower bound on its determinant on each $U_\alpha$, for example letting \begin{equation} U_n=\left\{x:\left|\det\left(D_{\albrep i.}f_j\right)_{i,j=1}^k\right|\in\left[ \frac{1}{n+1},\frac{1}{n}\right)\right\}. \end{equation} The result now follows by applying Theorem \ref{derivation_alberti}. \end{proof} \begin{thm}\label{thm:weak*density} The set $\Der(\Alb_{{\rm sub}}(\mu))$ is weak* dense in $\wder\mu.$. \end{thm} The proof of Theorem \ref{thm:weak*density} relies on the following technical Lemma. \begin{lem}\label{lem:vector_alberti} Consider a $1$-Lipschitz function $F:X\to\real^N$, a derivation $D_0\in\wder\mu.$, a compact $K\subset X$ and a vector $V=\lambda w$ where $\lambda>0$ and $w\in{\mathbb S}^{N-1}$; suppose that for $0<s_1<s_2$ and $\epsi>0$ \begin{equation} \begin{aligned} \locnorm D_0,{\wder\mu.}.(x)&\in(s_1,s_2)\quad(\forall x\in K)\\ \sup_{x\in K}\left\|D_0F(x)-V\right\|_{2}&<\epsi, \end{aligned} \end{equation} then, for each $\alpha\in(0,\frac{\pi}{2})$, the measure $\mu\mrest K$ admits a $C$-Lipschitz Alberti representation $\albrep.'$ with \begin{equation}\label{eq:vector_alberti} \begin{aligned} \left\|D_{\albrep.'}\right\|_{\wder\mu\mrest K.}\le C&\le (1+\epsi)\left( (1+\epsi)s_2+\epsi(N-1)\frac{s_2}{s_1}\cot\alpha+2\epsi \right),\\ \left\|D_{\albrep.'}F-V\right\|_2&\le\left(\epsi+\sin\alpha+1-\cos\alpha\right) \left( (1+\epsi)s_2+\epsi(N-1)\frac{s_2}{s_1}\cot\alpha+2\epsi \right). \end{aligned} \end{equation} \end{lem} \begin{proof} Note that if $u\perp w$, one has \begin{equation} \chi_K\left|D_0\langle u,F\rangle\right|=\chi_K\left|\langle u,D_0F-V\rangle\right|<\epsi; \end{equation} supposing that $S\subset K$ is Borel and $\frags(X,F,\delta,w,\alpha)$-null, Lemma \ref{lem:loc_estimate_fnull} implies \begin{equation} \chi_S\left|D_0\langle w,F\rangle\right|\le\left(\delta+(N-1)\frac{\epsi\cot\alpha}{s_1}\right)\locnorm D_0,{\wder\mu.}.; \end{equation} on the other hand, \begin{equation} \chi_K\left|D_0\langle w,F\rangle\right|\ge\chi_K\langle w,V\rangle- \chi_K\left|\langle w,D_0F-V\rangle\right|\ge\chi_K(\lambda-\epsi). \end{equation} This implies \begin{equation} \delta\ge\frac{\lambda-\epsi}{s_2}-(N-1)\frac{\epsi\cot\alpha}{s_1}; \end{equation} thus the measure $\mu\mrest K$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation $\albrep.=(P,\nu)$ in the $F$-direction of $\cone(w,\alpha)$ with $\langle w,F\rangle$-speed $\ge\delta_0$ where \begin{equation} \delta_0=\frac{\lambda-2\epsi}{s_2}-(N-1)\frac{\epsi\cot\alpha}{s_1}. \end{equation} Thus for $P$-a.e.~$\gamma$ and for $\mpush\gamma^{-1}.\nu_\gamma$-a.e.~$t$, one has \begin{equation} \begin{split} \left\|(F\circ\gamma)'(t)-w\right\|_2 & \le \left\|(F\circ\gamma)'(t) - \frac{(F\circ\gamma)'(t)} {\left\|(F\circ\gamma)'(t)\right\|_2}\right\|_2 + \left\|\frac{(F\circ\gamma)'(t)} {\left\|(F\circ\gamma)'(t)\right\|_2} -w\right\|_2\\ &\le\left|\, \left\|(F\circ\gamma)'(t)\right\|_2 -1\right| + \sin\alpha + 1-\cos\alpha\\ &\le\epsi+\sin\alpha+1-\cos\alpha; \end{split} \end{equation} note that this implies \begin{equation} \left\|D_{\albrep.}F-w\right\|_2\le\epsi+\sin\alpha+1-\cos\alpha; \end{equation} on the other hand, from \begin{equation} \delta_0\le\left\langle w,(F\circ\gamma)'(t)\right\rangle\le1+\epsi, \end{equation} one gets \begin{equation} \lambda\le(1+\epsi)s_2+\epsi(N-1)\frac{s_2}{s_1}\cot\alpha+2\epsi. \end{equation} If $\albrep.'=\pullb\tau_{\lambda,0}.\albrep.$ Lemma \ref{lem:alberti_affine_action} implies \eqref{eq:vector_alberti}. \end{proof} \begin{proof}[Proof of Theorem \ref{thm:weak*density}] It suffices to show that $\forall\Omega(\xi_1,\ldots,\xi_k; D_0; \epsi)$ there is $\albrep.\in\Alb_{{\rm sub}}(\mu)$: \begin{equation} \label{eq:weak*density1} D_{\albrep.}\in\Omega(\xi_1,\ldots,\xi_k; D_0; \epsi). \end{equation} Note that as the $\xi_i$ satisfy $\sum_{f\in B_1}\|\xi_i(f)\|_{L^1(\mu)}<\infty$, there are $\left\{(f_i,g_i)\right\}_{i=1}^{N_1}\subset\lipalg X.\times L^1(\mu)$ and $\epsi_1>0$ such that if \begin{align} \label{eq:weak*density2a} \|D\|_{\wder\mu.}&\le2\|D_0\|_{\wder\mu.},\\ \label{eq:weak*density2b} \left|\int g_i(D-D_0)f_i\,d\mu\right|&<\epsi_1\quad(\forall i),\\ \end{align} then $D\in\Omega(\xi_1,\ldots,\xi_k; D_0; \epsi)$. Let $F=(f_i)_{i=1}^{N_1}$ and, for $D\in\wder\mu.$, $DF=(Df_i)_{i=1}^{N_1}$; note that by possibly shrinking $\epsi_1$ we can assume that $F:X\to\real^{N_1}$ is $1$-Lipschitz. Moreover, as we need to produce an Alberti representation for a Borel subset of $S\subset X$ where $\locnorm D_0,{\wder\mu.}.>0$ and $\|D_0F\|_2>0$, we can assume $S=X$. Using Egorov's Theorem we find countably many $\{(K_j,\epsi_{2,j})\}$ with $K_j\subset X$ compact and $\epsi_{2,j}\in(0,\infty)$ such that if \eqref{eq:weak*density2a} holds and \begin{equation} \label{eq:weak*density3} \sup_{x\in K_j}\left\|DF(x_j)-D_0F(x_j)\right\|_{2}<\epsi_{2,j}\quad(\forall j), \end{equation} then \eqref{eq:weak*density2b} holds. Note that we can also assume that \begin{equation} \begin{aligned} \inf_{x\in K_j}\locnorm D_0,{\wder\mu.}.(x)&>0\\ \inf_{x\in K_j}\|D_0F(x)\|_2&>0. \end{aligned} \end{equation} Using again Egorov's Theorem, we can further partition the $K_j$ so that there are $(W_j,\epsi_{3,j},\alpha_j)\in \left(\real^{N_1}\setminus\{0\}\right)\times(0,\frac{\epsi_{2,j}}{2})\times(0,\frac{\pi}{2})$ such that: \begin{enumerate} \item The supremum $\sup_{x\in K_j}\left\|D_0F(x_j)-W_j\right\|_2<\epsi_{3,j}$. \item For each $ x\in K_j$ $\locnorm D_0,{\wder\mu.}.\in(s_{1,j},s_{2,j})\subset(0,\infty)$ with $\frac{s_{2,j}}{s_{1,j}}\le(1+\epsi_{3,j})$. \item We have the inequality \begin{equation}(1+\epsi_{3,j}) \left( (1+\epsi_{3,j})s_{2,j}+\epsi_{3,j}(N_1-1)\frac{s_{2,j}}{s_{1,j}}\cot\alpha_j+2\epsi_{3,j} \right)\le 2\|D_0\|_{\wder\mu.}. \end{equation} \item We have the inequality $2\|D_0\|_{\wder\mu.}(\epsi_{3,j}+\sin\alpha_j+1-\cos\alpha_j)<\frac{\epsi_{2,j}}{2}$. \end{enumerate} Using Lemma \ref{lem:vector_alberti} we obtain an Alberti representation $\albrep j.$ of $\mu\mrest K_j$ such that \eqref{eq:weak*density2a} and \eqref{eq:weak*density3} hold for $D_{\albrep j.}$. Using the gluing principle Theorem \ref{alb_glue} we obtain $\albrep .\in\Alb_{{\rm sub}}(\mu)$ such that \ref{eq:weak*density1} holds. \end{proof} \begin{thm}\label{thm:fin_gen_surjectivity} Suppose that $\wder\mu.$ is finitely generated and $D\in\wder\mu.$; let \begin{equation} S=\left\{x\in X:\locnorm D,{\wder\mu.}.>0\right\}; \end{equation} then $D\in\Der(\Alb(\mu\mrest S))$. \end{thm} The proof of Theorem \ref{thm:fin_gen_surjectivity} relies on the following technical Lemma. \begin{lem}\label{lem:derivation_idempotent} Suppose $\albrep.\in\Alb(\mu\mrest S)$ and $U\subset X$ Borel. Then there is $\albrep.'\in\Alb(\mu\mrest S)$ with $D_{\albrep.'}=\chi_UD_{\albrep.}$. Moreover, \begin{enumerate} \item If $\albrep.$ is $C$-Lipschitz ($(C,D)$-biLipschitz), so is $\albrep.'$. \item If $\albrep.$ has $f$-speed $\ge\delta$ or is in the $f$-direction of $\cone(w,\alpha)$, so does $\albrep.'\mrest U$. \end{enumerate} \end{lem} \begin{proof} Without loss of generality, $\albrep.=(P,\nu)\in\Alb(\mu)$. Using Lemma \ref{lem:domain_reduction}, we find an Alberti representation $\albrep 1.=(P_1,\nu_1)\in\Alb(\mu\mrest U^c)$ with \begin{equation} P_1\left(\frags(X,[0,1])\right)=1 \end{equation} and $D_{\albrep 1.}=\chi_{U^c}D_{\albrep.}$. Moreover, if $\albrep.$ satisfies a set of conditions \albcond, so does $\albrep 1.$. Similarly, combinining Lemmas \ref{lem:domain_reduction} and \ref{lem:alberti_affine_action}, we can find $\albrep 2.=(P_2,\nu_2)\in\Alb(\mu\mrest U^c)$, which is an Alberti representation of $\albrep.\mrest U^c$ with \begin{equation} P_2\left(\frags(X,[2,3])\right)=1 \end{equation} and \begin{equation}D_{\albrep 2.}=D_{\pullb\tau_{-1,0}.\albrep 1.}=-\chi_{U^c}D_{\albrep.}; \end{equation} note that conditions on the Lipschitz or biLipschitz constants satisfied by $\albrep.$ are also satisfied by $\albrep 2.$. Using Lemma \ref{lem:domain_reduction} we find an Alberti representation $\albrep 3.=(P_3,\nu_3)\in\Alb(\mu\mrest U)$ with \begin{equation} P_3\left(\frags(X,[4,5])\right)=1 \end{equation} and $D_{\albrep 3.}=\chi_UD_{\albrep.}$; note that if $\albrep.$ satisfies \albcond, so does $\albrep 3.$. We now let \begin{align} P'&=\frac{1}{4}(P_1+P_2)+\frac{1}{2}P_3;\\ \nu'(\gamma) &= \begin{cases} 2\nu_1(\gamma) &\text{if $\gamma\in\frags(X,[0,1])$}\\ 2\nu_2(\gamma) &\text{if $\gamma\in\frags(X,[2,3])$}\\ 2\nu_3(\gamma) &\text{if $\gamma\in\frags(X,[4,5])$;} \end{cases} \end{align} note that $\albrep.'=(P',\nu')\in\Alb(\mu)$ and \begin{equation} D_{\albrep.'}=\frac{1}{2}D_{\albrep 1.}+\frac{1}{2}D_{\albrep 2.}+D_{\albrep 3.}=\chi_UD_{\albrep.}. \end{equation} \end{proof} \begin{proof}[Proof of Theorem \ref{thm:fin_gen_surjectivity}] Without loss of generality we can assume that $S=X$, $\wder\mu.$ is free on $\{D_{\albrep i.}\}_{i=1}^k$ where $\albrep i.\in\Alb(\mu)$ and \begin{equation} D=\sum_{i=1}^k\lambda_iD_{\albrep i.} \end{equation} where the $\{\lambda_i\}_{i=1}^k\subset\bborel X.$ are nonnegative. The result now follows from the following Claims: \begin{description} \item[(CL1)] If $\{\albrep j.\}_{j=1}^m\subset\Alb(\mu)$, $\exists\albrep.'\in\Alb(\mu)$ with \begin{equation} D_{\albrep.'}=\sum_{j=1}^mD_{\albrep j.}. \end{equation} \item[(CL2)] If $\lambda\in\bborel X.$ is nonnegative and $\albrep.\in\Alb(\mu)$, then $\exists\albrep.'\in\Alb(\mu)$ with \begin{equation} D_{\albrep.'}=\lambda D_{\albrep.}. \end{equation} \end{description} \par To show (CL1), Lemma \ref{lem:domain_reduction} gives $\{\albrep j.'=(P_j,\nu_j)\}_{j=1}^m$ with $D_{\albrep j.'}=D_{\albrep j.}$ and \begin{equation} P_j\left(\frags(X,[2(j-1),2(j-1)+1])\right)=1; \end{equation} if we let \begin{align} P_\oplus&=\frac{1}{m}\sum_{j=1}^mP_j\\ \nu_\oplus(\gamma)&= \begin{cases} \nu_j(\gamma)&\text{if $\gamma\in\frags(X,[2(j-1),2(j-1)+1])$}\\ 0&\text{otherwise,} \end{cases} \end{align} $\albrep\oplus.=(P_\oplus,\nu_\oplus)\in\Alb(\mu)$ and \begin{equation} D_{\albrep\oplus.}=\frac{1}{m}\sum_{j=1}^mD_{\albrep j.}; \end{equation} thus it suffices to take $\albrep.'=\pullb\tau_{m,0}.\albrep\oplus.$. To show (CL2), let $2^{<\natural}$ denote the set of finite strings on $\{0,1\}$: \begin{equation} 2^{<\natural}=\left\{ \alpha=(a_1,\ldots,a_m): m\in\natural, a_i\in\{0,1\} \right\}; \end{equation} the length of $\alpha\in 2^{<\natural}$ will be denoted by $|\alpha|$, and the $k$-th entry by $\alpha(k)$. Suppose that the range of $\lambda$ lies in $[0, M)$ and let \begin{equation} \begin{aligned} \Delta:2^{<\natural}&\to[0,M)\\ \alpha&\mapsto\sum_{k=1}^{|\alpha|}\frac{\alpha(k)}{2^k}M; \end{aligned} \end{equation} let $I(\alpha)$ denote the unique subinterval $[a,b)$ of the dyadic subdivision of $[0,M)$ in $2^{|\alpha|}$ intervals which contains $\alpha$. If we let $U_\alpha$ denote the Borel set \begin{equation} U_\alpha=\left\{x\in X: \lambda(x)\in I(\alpha)\right\}, \end{equation} we can write \begin{equation} \lambda=\sum_{n=1}^\infty\frac{M}{2^n}\sum_{|\alpha|=n}\alpha(n)\chi_{U_\alpha}. \end{equation} By Lemma \ref{lem:derivation_idempotent} we can find $\albrep n.=(P_n,\nu_n)\in\Alb(\mu)$\footnote{With a uniform bound on the Lipschitz constant} with \begin{equation} D_{\albrep n.}=M\sum_{|\alpha|=n}\alpha(n)\chi_{U_\alpha}D_{\albrep .} \end{equation} and \begin{equation} P_n\left( \frags(X,[2(n-1),2(n-1)+1])\right)=1; \end{equation} if we let \begin{align} P'&=\sum_n2^{-n}P_n\\ \nu'(\gamma)&= \begin{cases} \nu_n(\gamma)&\text{if $\gamma\in\frags(X,[2(n-1),2(n-1)+1])$}\\ 0&\text{otherwise}, \end{cases} \end{align} $\albrep.'=(P',\nu')\in\Alb(\mu)$ and $D_{\albrep .'}=\lambda D_{\albrep.}$. \end{proof} \subsection{Geometric characterization of $\locnorm\,\cdot\,,{\wform\mu.}.$}\label{subsec:weaver_norm} In this subsection we prove Theorem \ref{thm:weaver_fnorm_char} which gives an intrinsic and geometric characterization of the Weaver norm on $\wder\mu.$. We also show that the inequality $\|Df\|_{L^\infty(\mu)}\le\|D\|_{\wder\mu.}\glip f.$ localizes (Lemma \ref{lem:local_lip_der}): $\glip f.$ can be replaced by $\biglip f$. \begin{thm}\label{thm:weaver_fnorm_char} Let $X$ be a compact metric space, $f:X\to\real$ Lipschitz, $S\subset X$ Borel, $\mu$ a Radon measure on $X$. Then $\locnorm df,{\wform\mu\mrest S.}.\le\alpha$ if and only if for each $\epsi>0$ there is an $S_\epsi\subset S$ which is Borel, $\frags(X,f,\alpha+\epsi)$-null, and satisfies $\mu(S\setminus S_\epsi)=0$. In particular, \begin{equation} \|df\|_{\wform\mu.}=\sup\left\{\|Df\|_{L^\infty(\mu)}:\text{$D\in\Der(\Alb_{{\rm sub}}(\mu))$ and $\|D\|_{\wder\mu.}\le1$}\right\}. \end{equation} \end{thm} \begin{proof} Necessity is proven by contrapositive. Assume that for some $\epsi>0$ $S$ does not contain a full $\mu$-measure subset which is $\frags(X,f,\alpha+\epsi)$-null . Inspection of the proof of \ref{alberti_rep_prod} shows that $\forall\eta>0$ \begin{equation} \mu=\mu\mrest G_\eta+\mu\mrest F_\eta, \end{equation} where $G_\eta$, $F_\eta$ are complementary, $F_\eta$ being a $\frags(X,f,\alpha+\epsi)$-null $F_{\sigma\delta}$ and $\mu\mrest G_\eta$ admitting a $(1,1+\eta)$-biLipschitz Alberti representation with $f$-speed $\ge\alpha+\epsi$. By assumption, $F_\eta\cap S$ cannot be a full $\mu$-measure subset of $S$, so $\mu(G_\eta\cap S)>0$. If $D_\eta$ denotes the derivation associated to that Alberti representation, \begin{equation} \|D_\eta\|_{\wder\mu.}\le 1+\eta \end{equation} and \begin{equation} \chi_{G_\eta}D_\eta f\ge\alpha+\epsi \end{equation} so that \begin{equation} \locnorm df,{\mu\mrest G_\eta}.\ge\frac{\alpha+\epsi}{1+\eta}>\alpha \end{equation} for $\eta$ sufficiently small. \par Sufficiency is proven via the approximation scheme Theorem \ref{onedimapprox}, reducing to the case in which $S$ is compact. By replacing $S$ by $\bigcap_nS_{\frac{1}{n}}$, we can assume that $S$ is, $\forall n$, $\frags(X,f,\alpha+{\frac{1}{n}})$-null; then $\exists g_n$ $\max\left(\glip f.,\alpha+{\frac{1}{n}}\right)$-Lipschitz and $\mu$-a.e.~$(\alpha+{\frac{1}{n}})$-Lipschitz on $S$ with $\|g_n-f\|_\infty\le{\frac{1}{n}}$. Thus, $g_n\xrightarrow{\text{w*}} f$ so that $dg_n\xrightarrow{\text{w*}}df$ and by lower semicontinuity, \begin{equation} \|df\|_{L^\infty(\mu\mrest S)}\le\liminf_{n\to\infty}\|dg_{{n}}\|_{L^\infty(\mu\mrest S)}\le\alpha. \end{equation} \end{proof} \par We also present a proof of the following useful result. \begin{lem}\label{lem:local_lip_der} Let $X$ be a metric space, $\mu$ a Radon measure on $X$, $f\in\lipalg X.$ and $D\in\wder\mu.$; then \begin{equation} |Df|\le\locnorm D,{\wder\mu.}.\biglip f. \end{equation} \end{lem} \begin{proof} For all $\epsi_0,\epsi_1>0$, using Egorov and Lusin's Theorems, there are triples $(K_\alpha,\lambda_\alpha,r_\alpha)$ such that: \begin{enumerate} \item The $K_\alpha$ are disjoint compact with $\mu\left(X\setminus\bigcup_\alpha K_\alpha\right)=0$ and $\mu(K_\alpha)>0$. \item The constants $\lambda_\alpha$ are nonnegative. \item The constants $r_\alpha$ are positive and $K_\alpha$ has diametre less than $r_\alpha$. \item For each $x\in K_\alpha$ and each $r\in(0,r_\alpha]$, \begin{equation} \biglip f(x,r)\in(\lambda_\alpha-\epsi_0,\lambda_\alpha). \end{equation} \item The local norm $\locnorm D,{\wder\mu\mrest K_\alpha.}.$ lies in $(\|D\|_{\wder\mu\mrest K_\alpha.}-\epsi_1,\|D\|_{\wder\mu\mrest K_\alpha.}]$. \end{enumerate} By point (3) $f$ is $(\lambda_\alpha)$-Lipschitz on $K_\alpha$; let $F$ denote a McShane extension of $f|K_\alpha$ with $\glip F.\le\lambda_\alpha$. Then \begin{equation} \|Df\|_{L^\infty(\mu\mrest K_\alpha)}=\|DF\|_{L^\infty(\mu\mrest K_\alpha)}\le\lambda_\alpha\|D\|_{\wder\mu\mrest K_\alpha.}; \end{equation} thus \begin{equation} \begin{split} \chi_{K_\alpha}|Df|&\le\lambda_\alpha\left(\locnorm D,{\wder\mu\mrest K_\alpha.}.+\epsi_1\right)\\ &<(\biglip f+\epsi_0)\left(\locnorm D,{\wder\mu\mrest K_\alpha.}.+\epsi_1\right). \end{split} \end{equation} As $\sum_\alpha\chi_{K_\alpha}=1$ in $L^\infty(\mu)$\footnote{The convergence of the series is in the weak* sense}, \begin{equation} |Df|<(\biglip f+\epsi_0)\left(\locnorm D,{\wder\mu.}.+\epsi_1\right). \end{equation} Letting $\epsi_0,\epsi_1\searrow0$ completes the proof. \end{proof} \section{Structure of differentiability spaces}\label{sec:app_diff_spaces} \subsection{Differentiability spaces and derivations}\label{subsec:derivation_differentiability} In this Subsection we make a first application of derivations to study differentiability spaces. The goal is to prove Lemma \ref{partialderivatives} and Corollary \ref{cor:assouad_bound}. We give a first proof of Lemma \ref{partialderivatives} which depends on the results in \cite{bate-diff}; another proof, which relies on Theorem \ref{thm:diff_char1} and \cite{derivdiff}, can be found at the end of Subsection \ref{subsec:char_diff}. \begin{lem}\label{partialderivatives} Let $(U,x)$ be a differentiability chart, with $U$ Borel, in a differentiability space $(X,\mu)$. Then the partial derivative operators $\frac{\partial}{\partial x^j}$ are derivations. \end{lem} \begin{proof} Let $n$ be the dimension of the chart $(U,x)$; we first show that there are disjoint Borel sets $\{U_\alpha\}$ with $U_\alpha\subset U$ and $\mu\left(U\setminus\bigcup_\alpha U_\alpha\right)=0$, and such that, for each $\alpha$, it is possible to find $g_{\alpha,k}^i\in L^\infty(\mu\mrest U_\alpha)$ ($i,k=1,\cdots,n)$ and derivations associated to Alberti representations $\{\albrep i.\}_{i=1}^n$, such that \begin{equation}\label{derrep} \sum_{i=1}^ng_{\alpha,k}^iD_{\albrep i.\mrest U_\alpha}x^j=\delta_k^j\chi_{U_\alpha}; \end{equation} in fact, because of \cite[Lem.~6.1]{derivdiff}, in a differentiability space derivations are completely determined by their action on the chart functions and so \eqref{derrep} represents $\chi_{U_\alpha}\frac{\partial}{\partial x^k}$. Note that even though in \cite[Sec.~6]{derivdiff} the measure $\mu$ is assumed doubling, this is not restrictive by Theorem \ref{thm:bate_speight}. By \cite[Thm.~6.6]{bate-diff} we know that $\mu\mrest U$ admits Alberti representations in the $x$-directions of indipendent cone fields $\{\cone_l\}_{l=1}^n$. Using Corollary \ref{der-alb}, we obtain $(1,3/2)$-biLipschitz Alberti representations $\{\albrep i.\}_{i=1}^n$ in the $x$-directions of cone fields $\cone(e_i,\theta)$ where $\{e_i\}_{i=1}^n$ is the standard Euclidean basis and $\theta$ is sufficiently small to ensure that the cone fields $\cone(e_i,\theta)$ are independent. In particular, letting $M=(D_{\albrep i.}x^j)_{i,j=1}^n$, and choosing Borel representatives for the entries of $M$, the bounded Borel function $\det M$ is different from zero on a $\mu$-full measure subset. In particular, one can choose disjoint Borel subsets $U_\alpha\subset U$ with $\mu\left(U\setminus\bigcup_\alpha U_\alpha\right)=0$, and such that, on each $U_\alpha$ the determinant $\det M$ is uniformly bounded from below by $\delta_\alpha>0$. Inverting the matrix $M$ one concludes that there are $g_{\alpha,k}^i\in L^\infty(\mu\mrest U_\alpha)$ such that \eqref{derrep} holds. \par The definition of chart implies that the partial derivative operators $\frac{\partial}{\partial x^j}$ are bounded linear maps from $\lipalg X.\to L^\infty(\mu\mrest U)$ that satisfy the product rule. In order to show weak* continuity, suppose that $f_t\xrightarrow{\text{w*}} f$ in $\lipalg X.$ and that $h\in L^1(\mu\mrest U)$. For each $\epsi>0$ there are finitely many $\{U_{\alpha_s}\}_{s=1}^{N(\epsi)}$ such that \begin{align}\label{eq:partialderivatives_p1} \sup_t\left|\int_{U\setminus\bigcup_{s=1}^{N(\epsi)}U_{\alpha_s}}h\frac{\partial{f_t}}{\partial x^j}\,d\mu\right|&\le\epsi; \\ \label{eq:partialderivatives_p2} \left|\int_{U\setminus\bigcup_{s=1}^{N(\epsi)}U_{\alpha_s}}h\frac{\partial{f}}{\partial x^j}\,d\mu\right|&\le\epsi. \end{align} Combining equations \eqref{eq:partialderivatives_p1} and \eqref{eq:partialderivatives_p2} with the fact that $\sum_{s=1}^{N(\epsi)}\chi_{U_{\alpha,s}}\frac{\partial}{\partial x^j}$ is a derivation, we obtain \begin{equation} \lim_{t\to\infty}\int_Uh\frac{\partial{f_t}}{\partial x^j}\,d\mu=\int_Uh\frac{\partial{f}}{\partial x^j}\,d\mu, \end{equation} which shows that $\frac{\partial}{\partial x^j}$ is weak* continuous. \end{proof} \par We now prove the dimensional bound Corollary \ref{cor:assouad_bound} which significantly strengthens previous bounds on differentiability dimensions. In fact, we remove from Theorems \ref{thm:cheeger}, \ref{thm:keith} and \ref{thm:bate-diff_char} the dipendence on $\tau$ from the bound on the differentiability dimension. \begin{cor}\label{cor:assouad_bound} Suppose $(X,\mu)$ is a $\sigma$-differentiability space with $\mu$ doubling, then $(X,\mu)$ is a differentiability space and the dimension of the mea\-surable differentiable structure is at most the Assouad dimension of $X$. \end{cor} \begin{proof} It follows by applying Lemma \ref{derbound} and Lemma \ref{partialderivatives}. \end{proof} \subsection{Characterization of differentiability spaces}\label{subsec:char_diff} In this Subsection we obtain a new characterization of differentiability spaces. The goal is to prove Theorems \ref{thm:diff_char1} and \ref{thm:diff_char2}. Throughout this section the metric space $X$ is assumed to be Polish. \begin{thm}\label{thm:diff_char1} The metric measure space $(X,\mu)$ is a $\sigma$-differentiability space if and only if \begin{equation}\mu(S)=0\quad\text{\normalfont($\forall S\in\Gap(X)$).} \end{equation} \end{thm} \begin{thm}\label{thm:diff_char2} The metric measure space $(X,\mu)$ is a $\sigma$-differentiability space if and only if $\mu$ is $\sigma$-asympto\-ti\-cal\-ly doubling and one of the following equivalent conditions holds: \begin{enumerate} \item For each $f\in\lipfun X.$ \begin{equation}\label{eq:diff=Lip_gl} \locnorm df,{\wform\mu.}.(x)=\biglip f(x)\quad\text{\normalfont(for $\mu$-a.e.~$x$)}. \end{equation} \item For each $f\in\lipfun X.$, denoting by $S_f$ the Borel set \begin{equation} \left\{x\in X:\biglip f(x)>0\right\}, \end{equation} for all $\epsi,\sigma\in(0,1)$ the measure $\mu\mrest S_f$ admits a $(1,1+\epsi)$-biLipschitz Alberti representation with $f$-speed $\ge\sigma\biglip f$. \item For each $f\in\lipfun X.$, \begin{equation} \biglip f(x)=\smllip f(x)\quad\text{\normalfont(for $\mu$-a.e.~$x$)}. \end{equation} \end{enumerate} \end{thm} The proofs of Theorems \ref{thm:diff_char1} and \ref{thm:diff_char2} require some preparation. We will use Theorem \ref{lip_ind}, the proof of which is deferred to Subsection \ref{subsec:constr-indep-lipsch}. \begin{thm}\label{lip_ind} Let $S\subset X$ be a Borel set and assume that for some $ \delta_0\in(0,1]$ and for each $m\in\natural$ there is an $L$-Lipschitz function $f_m$ such that: \begin{enumerate} \item There is a $\rho_m\in(0,\frac{1}{m})$ such that if $B$ is a ball of radius $\rho_m$ centred at some point of $S$, the Lipschitz constant of the restriction $f|_B$ is at most $\frac{1}{m}$. \item For each $x\in S$ there is $y\in \ball x, \frac{1}{m}.$ such that \begin{equation} |f_m(x)-f_m(y)|\ge\delta_0 d(x,y)>0. \end{equation} \end{enumerate} \par Then for all $(M,\alpha)\in\natural\times(0,\min\left( \frac{1}{2}\sqrt{\frac{\delta_0 L}{1+\delta_0}},\frac{1}{2}\delta_0\right))$ there are a Borel subset $S'\subset S$ and Lipschitz functions $\{\psi_0,\ldots,\psi_{M-1}\}$ such that: \begin{itemize} \item The set $S\setminus S'$ is $\mu$-null. \item The $\psi_i$ have Lipschitz constant bounded by \begin{equation} 3\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right). \end{equation} \item For each $x\in S'$: \begin{equation} \biglip\left(\sum_{i=0}^{M-1}\lambda_i\psi_i\right)(x)\ge \left(\delta_0-\frac{\alpha^2/L}{1-\alpha^2/L}-\alpha\right)\max_{i=0,\cdots,M-1}|\lambda_i|. \end{equation} \end{itemize} \end{thm} \par As a first step, we prove a Lemma that produces \emph{flat functions} as in the hypothesis of Theorem \ref{lip_ind}. \begin{lem}\label{lem:gap_flat_func} Let $X$ be a compact metric space with finite Assouad dimension, $f:X\to\real$ Lipschitz, $S\subset X$ Borel, $\mu$ a Radon measure on $X$. Suppose that there are $\alpha$ and $\beta$ such that $\alpha>\beta\ge0$ and \begin{equation} \inf_{x\in S}\biglip f(x)\ge\alpha>\beta\ge\locnorm df,{\wform\mu\mrest S.}.; \end{equation} then for each $(m,\delta,\gamma,\epsi)\in\natural\times(0,1)\times(0,\alpha-\beta)\times(0,1)$, there is a triple $(K,r_m,h)$: \begin{enumerate} \item The set $K$ is compact with $\mu(S\setminus K)\le\epsi$. \item The function $h$ is $3\glip f.$-Lipschitz and $r_m>0$. \item For each $ x\in K$ there is $y$: \begin{equation} 0<\dist x,y.\le\frac{1}{m}\quad\text{and}\quad \left|h(x)-h(y)\right|\ge\gamma\dist x,y.. \end{equation} \item For each $ x\in K$, $\glip h|{\ball x,r_m.}.\le\delta$. \end{enumerate} \par In the case in which $\beta=0$ the assumption on the finite Assouad dimension is not needed. \end{lem} \begin{proof} Without loss of generality we can assume that $S$ is compact and, for each $\eta_0>0$, $\frags(X,f,\beta+\eta_0)$-null. Thus, by Theorem \ref{onedimapprox} there are functions $g_n:X\to\real$ which are \begin{equation} L_0=\max\left(\glip f.,\beta+\eta_0\right)\text{-Lipschitz,} \end{equation} and such that $g_n\xrightarrow{\text{w*}}f$ and $g_n$ is $\mu$-a.e.~locally $(\beta+\eta_0)$-Lipschitz on $S$. \par The part of the argument starting here is the only one that requires that the Assouad dimension of the space is finite and is not needed if $\beta=0$. Note that $dg_n\xrightarrow{\text{w*}}df$ in $\wform\mu.$. By finiteness of the Assouad dimension, by Lemma \ref{derbound} and Theorem \ref{thm:free_dec}, there are finitely many disjoint Borel $\{X_\alpha\}$ with $\mu(X_\alpha)>0$, $\mu\left(X\setminus\bigcup_\alpha X_\alpha\right)=0$ and $\wder\mu\mrest X_\alpha.$ free of rank $N_\alpha$. Note that \begin{equation} \wder\mu.=\bigoplus_\alpha\wder\mu\mrest X_\alpha. \end{equation} and Lemma \ref{lem:hahn-banach} implies that $\wform\mu\mrest X_\alpha.$ is also free of rank $N_\alpha$. In particular, we have \begin{equation} \wform\mu.=\bigoplus_\alpha\wform\mu\mrest X_\alpha. \end{equation} and for each $\alpha$ one can find a basis of $\wder\mu\mrest X_\alpha.$, $\{D_{\alpha,i}\}_{i=1}^{N_\alpha}$, and a constant $C_\alpha>0$: for each $\omega\in\wform\mu\mrest X_\alpha.$, \begin{equation} \locnorm\omega,{\wform\mu\mrest X_\alpha.}.\le C_\alpha\left(\sum_{i=1}^{N_\alpha} \left|\langle D_{\alpha,i},\omega\rangle\right|^2\right)^{1/2}. \end{equation} We let $N=\max_\alpha N_\alpha$, $C=\max_\alpha C_\alpha$ and $\{e_i\}_{i=1}^{N_\alpha}$ the standard basis of $\real^{N_\alpha}$. Consider the map \begin{equation} \begin{aligned} \iota:\wform\mu.&\to L^2(\mu,\real^N)\\ \omega&\mapsto\sum_\alpha\sum_{i=1}^{N_\alpha}\chi_{X_\alpha}\langle D_{\alpha,i},\omega\rangle e_i; \end{aligned} \end{equation} then \begin{equation} \locnorm\omega,{\wform\mu.}.\le C|\iota(\omega)|_{l^2}\in L^\infty(\mu) \end{equation} and $\iota(dg_n)\xrightarrow{\text{w}-L^2}\iota(df)$\footnote{Weak convergence in $L^2(\mu,\real^N)$}. By Mazur's Lemma there are tail-convex combinations \begin{equation} \hat g_n=\sum_{k=n}^{M(n)}t_kg_k \end{equation} with $\iota(\hat g_n)\xrightarrow{L^2}\iota(df)$. Note that the $\hat g_n$ are $L_0$-Lipschitz and $\mu$-a.e.~locally $(\beta+\eta_0)$-Lipschitz on $S$. By passing to a subsequence we can assume that $\iota(\hat g_n)\to\iota(df)$ $\mu$-a.e. Using Egorov's Theorem, for each $\eta_1,\eta_2>0$ there are a compact $K_0\subset S$ and $n_0\in\natural$: \begin{align} \mu\left(S\setminus K_0\right)&\le\eta_1,\\ \sup_{x\in K_0}\left|\iota(\hat g_{n_0})(x)-\iota(df)(x)\right|&\le\frac{\eta_2}{C}, \end{align} which implies \begin{equation} \locnorm d\hat g_{n_0}-df,{\wform\mu\mrest K_0.}.\le\eta_2. \end{equation} This is the end of the part of the argument that requires that the Assouad dimension of the space is finite. \par As $\biglip$ behaves like a seminorm, \begin{equation} \inf_{x\in S}\biglip\left(f-\hat g_{n_0}\right)(x)\ge\inf_{x\in S}\biglip f(x)-\sup_{x\in S}\biglip \hat g_{n_0}(x)\ge\alpha-\beta-\eta_0; \end{equation} using the Borel measurability of $\biglip\left(f-\hat g_{n_0}\right)$ and Egorov and Lusin's Theorems, for all positive $\eta_3$, $\eta_4$ and for each natural $m$, there are a compact $K_1\subset K_0$, and constants $\rho_0\ge\rho_1>0$ such that: \begin{enumerate} \item We have the inequalities: \begin{align} \rho_1\le\rho_0&<\frac{1}{m},\\ \mu(K_0\setminus K_1)&\le\eta_3. \end{align} \item For each $ x\in K_1$ there is $ y\in X$: $\dist x,y.\in[\rho_1,\rho_0]$ and \begin{equation}\label{eq:low_var_bound} \left|\left(f-\hat g_{n_0}\right)(x)-\left(f-\hat g_{n_0}\right)(y)\right|\ge(\alpha-\beta-\eta_0-\eta_4)\dist x,y.. \end{equation} \end{enumerate} Note that $f-\hat g_{n_0}$ is \begin{equation} L_1=\left(\glip f.+L_0\right)\text{-Lipschitz}. \end{equation} Note also that for each $\eta_5>0$ the compact $K_0$ is $\frags(X,f-\hat g_{n_0},\eta_2+\eta_5)$-null. Applying again Theorem \ref{onedimapprox} we can find functions $h_n:X\to\real$ which are \begin{equation} L_2=\max(L_1,\eta_2+\eta_5)\text{-Lipschitz,} \end{equation} with $h_n\xrightarrow{\text{w*}}f-\hat g_{n_0}$ and which are $\mu$-a.e.~locally $(\eta_2+\eta_5)$-Lipschitz on $K_0$. Thus, for all $\eta_6,\eta_7>0$, there are $n_1\in\natural$, a compact $K_2\subset K_1$ and constants $\rho_1\ge\rho_2>0$: \begin{align} \mu(K_1\setminus K_2)&\le\eta_6,\\ \left\|h_{n_1}-\left(f-\hat g_{n_0}\right)\right\|_\infty&\le\eta_7\rho_1, \end{align} and for each $x\in K_2$, $\glip h|{\ball x,\rho_2.}.$ does not exceed $\eta_2+\eta_5$. Let $x\in K_2$; then there is $y\in X$ with $\dist x,y.\in[\rho_1,\rho_0]$ such that \eqref{eq:low_var_bound} holds. Therefore, \begin{equation} \begin{split} \left|h_{n_1}(x)-h_{n_1}(y)\right|&\ge\left|\left(f-\hat g_{n_0}\right)(x)-\left(f-\hat g_{n_0}\right)(y)\right| -2\left\|h_{n_1}-\left(f-\hat g_{n_0}\right)\right\|_\infty\\ &\ge(\alpha-\beta-\eta_0-\eta_4-2\eta_7)\dist x,y.. \end{split} \end{equation} We now require the constants $\eta_i$ to satisfy \begin{align} \eta_1+\eta_3+\eta_6&\le\epsi\\ \eta_2+\eta_5&\le\min(\delta,\glip f.)\\ \gamma&\le(\alpha-\beta-\eta_0-\eta_4-2\eta_7)\\ \beta+\eta_0&\le2\beta; \end{align} as a consequence, $L_2\le 3\glip f.$. We let $K=K_2$, $h=h_{n_1}$ and $r_m=\rho_2$. \end{proof} To see how \textbf{porosity} interacts with differentiability we will need to consider the pointwise Lipschitz constants of a function with respect to a given subset. \begin{defn} Let $Y\subset X$, $f:X\to\real$ Lipschitz. For $y\in Y$, the \textbf{pointwise upper and lower Lipschitz constants of $f|_Y$ at $y$} will be denoted by $\biglip_Y f(y)$ and $\smllip_Y f(y)$. Note that \begin{align} \label{eq:local_Lip} \biglip_Y f(y)&\le\biglip f(y)\\ \label{eq:local_lip} \smllip_Y f(y)&\le\smllip f(y). \end{align} \end{defn} We now introduce the classes of subsets used to characterize differentiability. \begin{defn} Given a complete metric space $X$ and a Radon measure $\mu$ on $X$, we define by $\Gap(\mu,X)$ the class of Borel subsets $V\subset X$ such that there are $\alpha>\beta\ge0$ and a Lipschitz function $f:X\to\real$ with \begin{equation} \inf_{x\in V}\biglip f(x)\ge\alpha\quad\text{and}\quad\locnorm df,{\wform\mu\mrest V.}.\le\beta. \end{equation} Note that the inequality involving $\locnorm df,{\wform\mu\mrest V.}.$ holds in $L^\infty(\mu\mrest V)$. We define by $\Gap_0(\mu,X)$ the class of Borel subsets $V\subset X$ such that there are $\alpha>0$ and a Lipschitz function $f:X\to\real$ with \begin{equation} \inf_{x\in V}\biglip f(x)\ge\alpha\quad\text{and}\quad\locnorm df,{\wform\mu\mrest V.}.=0. \end{equation} \par Note that $\Gap_0(\mu,X)\subset\Gap(\mu,X)$ and if $Y\subset X$ is Borel, applying McShane's Lemma and \eqref{eq:local_Lip}, \begin{align} \Gap(\mu\mrest Y,Y)&\subset\Gap(\mu,X)\\ \Gap_0(\mu\mrest Y,Y)&\subset\Gap_0(\mu,X). \end{align} We finally let \begin{align} \Gap(X)&=\bigcap_{\text{$\mu$ Radon}}\Gap(\mu,X)\\ \Gap_0(X)&=\bigcap_{\text{$\mu$ Radon}}\Gap_0(\mu,X). \end{align} Note that $\Gap_0(X)\subset\Gap(X)$ and if $Y\subset X$ is Borel, \begin{align} \Gap(Y)&\subset\Gap(X)\\ \Gap_0(Y)&\subset\Gap_0(X). \end{align} \end{defn} Even though we found useful to work with $\Gap(\mu, X)$ sets, the following Lemma shows that one can just work with $\Gap(X)$ sets. \begin{lem} \label{lem:gap_incl} If $S\in\Gap(\mu,X)$ with $\mu(S)>0$, then there is a Borel $S'\subset S$ with $\mu\left(S\setminus S'\right)=0$ and $S'\in\Gap(X)$. The same conclusion holds replacing $\Gap$ with $\Gap_0$. \end{lem} \begin{proof} Let $(f,\alpha,\beta)$ as in the defining property of $\Gap$. By Theorem \ref{thm:weaver_fnorm_char}, for $n$ sufficiently large, there is a $\frags(X,f,\beta+\frac{1}{n})$-null Borel $S_n\subset S$ such that $\mu(S\setminus S_n)=0$. Let $S'=\bigcap_n S_n$ so that $S'$ is a full $\mu$-measure Borel subset of $S$ which is $\frags(X,f,\beta+\frac{1}{n})$-null for each $n$. Thus, Theorem \ref{thm:weaver_fnorm_char} implies that \begin{equation} \locnorm df,{\wform\nu\mrest S'.}.\le\beta \end{equation} for all Radon measures $\nu$. As \begin{equation} \inf_{x\in S'}\biglip f(x)\ge\alpha, \end{equation} $S'\in\Gap(X)$. \end{proof} \par We now use Lemma \ref{lem:gap_flat_func} to construct independent Lipschitz functions. \begin{lem} \label{lem:ind_lip} Let $X$ be a compact metric space with finite Assouad dimension, $f:X\to\real$ Lipschitz, $S\in\Gap(\mu,X)$ with $\mu(S)>0$. Then for all $\epsi>0$ and $M\in\natural$, there is a Borel $S'\subset S$ with $\mu\left(S\setminus S'\right)\le\epsi$ and there are $1$-Lipschitz functions $\left\{\psi_0,\cdots,\psi_{M-1}\right\}$ which are infinitesimally independent on $S'$. \par In the case in which $S\in\Gap_0(\mu,X)$ the assumption on the Assouad dimension is not needed. \end{lem} \begin{proof} Let $(f,\alpha,\beta)$ as in the defining property of $\Gap$. By Lemma \ref{lem:gap_flat_func}, for each $m\in\natural$ and $\epsi_1^{(m)}>0$, there are $(K_m,\rho_m,h_m)$: \begin{enumerate} \item The set $K_m$ is compact with $\mu\left(S\setminus K_m\right)\le\epsi_1^{(m)}$. \item The function $h_m$ is $3\glip f.$-Lipschitz and $\rho_m>0$. \item For each $ x\in K_m$ there is $ y$: \begin{equation} 0<\dist x,y.\le\frac{1}{m}\quad\text{and}\quad \left|h_m(x)-h_m(y)\right|\ge\frac{\alpha}{2}\dist x,y.. \end{equation} \item For each $ x\in K_m$, $\glip h_m|{\ball x,\rho_m.}.\le\frac{1}{m}$. \end{enumerate} Let $K=\bigcap_m K_m$ so that \begin{equation} \mu\left(S\setminus K\right)\le\sum_m\epsi_1^{(m)}; \end{equation} choosing the $\epsi_1^{(m)}$ sufficiently small, we can ensure $ \mu\left(S\setminus K\right)\le\epsi$. Applying Theorem \ref{lip_ind} to $K$ we construct the functions $\left\{\psi_0,\cdots,\psi_{M-1}\right\}$. \end{proof} \par We now recall some facts about porosity and refer the reader to the survey \cite{zajicek_por_surv} for more information. \begin{defn} For $Y\subset X$ and $c>0$ we say that $Y$ is \textbf{$c$-porous at $y$} if there is a sequence $y_n\to y$ with \begin{equation} 0<c\dist y_n,y.<\setdist\{y_n\},Y.. \end{equation} For $y\in Y$ we let \begin{equation} W_c(y,Y)=\left\{y'\in Y: 0<c\dist y' ,y.<\setdist\{y'\},Y.\right\}; \end{equation} thus $c$-porosity of $Y$ at $y$ is equivalent to: \begin{equation} W_c(y,Y)\cap\ball y,r.\ne\emptyset\quad(\forall r>0). \end{equation} If $Y$ is $c$-porous at each $y\in Y$, $Y$ is called \textbf{$c$-porous}. \end{defn} \begin{rem}Given a $c$-porous set $Y$, one might wonder if there is also a Borel porous subset of $X$. By \cite[Lem.~1.4]{preiss_porosity} it follows that if $Y$ is $c$-porous, for each $\epsi>0$ there is $\tilde Y$, a $G_\delta$ set, with $Y\subset\tilde Y$ and $\tilde Y$ $(c-\epsi)$-porous. \end{rem} \begin{rem}Porosity is related to the relationship between $\biglip_Yf$ and $\biglip f$. For example, assume that $Y$ is $c$-porous at $y$ and let $f=\setdist Y,\{\cdot\}.$. Then \begin{equation}\label{eq:por_Lip_lbound} \biglip f(y)\ge\limsup_{n\to\infty}\frac{\setdist\{y_n\},Y.}{\dist y_n,y.}>c; \end{equation} but $\biglip_Yf(y)=0$ as $f$ is identically zero on $Y$. On the other hand, if for all $c\in(0,\frac{1}{2})$ the set $Y$ is not $c$-porous at $y\in Y$, then, for each Lipschitz function $f$, \begin{equation} \label{eq:npor_Lip_equal} \biglip_Yf(y)=\biglip f(y). \end{equation} To see this, choose $x_n\to y$, with $\dist x_n,y.>0$ and \begin{equation} \biglip f(y)=\lim_{n\to\infty}\frac{|f(y)-f(x_n)|}{\dist y,x_n.}; \end{equation} as $Y$ is not $c$-porous at $y$, $\exists r_c>0$ such that \begin{equation} \setdist{\ball y,r.},Y.\le cr\quad(\forall r\in(0,r_c]). \end{equation} So for $n$ large enough we find $Y\ni y_n\ne y$ with \begin{equation} \dist x_n,y_n.\le c\dist x_n,y.; \end{equation} this implies \begin{equation} \begin{split} \biglip_Yf(y)&\ge\limsup_{n\to\infty}\frac{|f(y)-f(y_n)|}{\dist y,y_n.}\\ &\ge\limsup_{n\to\infty}\frac{|f(y)-f(x_n)|-c\glip f.\dist y,x_n.}{(1+c)\dist x_n,y.}\\ &\ge\frac{1}{1+c}\left(\biglip f(y)-c\glip f.\right). \end{split} \end{equation} Letting $c\to0$ we obtain \eqref{eq:npor_Lip_equal}. \end{rem} We now show that porous sets are of class $\Gap_0(X)$. \begin{lem}\label{lem:por_gap} If $S\subset X$ is $c$-porous and Borel, then $S\in\Gap_0(X)$. \end{lem} \begin{proof} Let $\nu$ be a Radon measure on $X$. Let $f=\setdist S,\{\cdot\}.$. From \eqref{eq:por_Lip_lbound} it follows that \begin{equation} \inf_{x\in S}\biglip f(x)\ge c. \end{equation} If $\nu(S)=0$ we have, trivially, $\locnorm df,{\wform\nu\mrest S.}.=0$. If $\nu(S)>0$ we note that $\forall D\in\wder\nu.$, by locality of derivations, as $f=0$ on $S$, $\chi_SDf=0$. This implies \begin{equation} \locnorm df,{\wform\nu\mrest S.}.=0. \end{equation} \end{proof} \par We need the following technical Lemma because our approximation schemes are designed to work in compact spaces. \begin{lem}\label{lem:porous_completion} If $K\subset X$\footnote{$X$ is assumed complete} is compact and $c$-porous, then there is a compact $Y$ with $K\subset Y\subset X$ and such that $K$ is $\frac{2c}{3}$-porous in $Y$. \end{lem} \begin{proof} For $m\in\natural$ define \begin{equation} \begin{aligned} \psi_m:K&\to\left(0,\frac{1}{m}\right]\\ x&\mapsto\sup\left\{\dist x,y.:y\in W_c(x,K)\cap\ball x,\frac{1}{m}.\right\}; \end{aligned} \end{equation} note that $\psi_m$ is lower-semicontinuous as $\{x\in K:\psi_m(x)>r\}$ is open. In fact, choose $y\in W_c(x,K)$ with $\dist x,y.>r$. For $x'\in K$ sufficiently close to $x$, $\dist x',y.>r$ and $y\in W_c(x',K)$. Let $r_m=\min_K\psi_m$. As $K$ is compact, we can choose a finite $\frac{r_m}{3}$-net $N_m\subset K$. For $x\in N_m$ choose \begin{equation} w_m(x)\in W_c(x,K)\cap\ball x,\frac{1}{m}. \end{equation} with $\dist x,w_m(x).>\frac{2r_m}{3}$. Let $W_m=\bigcup_{x\in N_m}\{w_m(x)\}$. Let $x'\in K$; choose $x\in N_m$ with $\dist x,x'.\le\frac{r_m}{3}$. Then \begin{equation} \dist x',w_m(x).>\frac{r_m}{3} \end{equation} and \begin{equation} \begin{split} \dist w_m(x),x'.&\le\dist x,w_m(x).+\dist x,x'.\\ &<\dist x,w_m(x).+\frac{1}{2}\dist x,w_m(x).\\ &=\frac{3}{2}\dist x,w_m(x).. \end{split} \end{equation} This implies that \begin{equation} \frac{2c}{3}\dist w_m(x),x'.<c\dist x,w_m(x).<\dist w_m(x),K.; \end{equation} thus $w_m(x)\in W_{\frac{2c}{3}}(x',K)$ and $K$ is $\frac{2c}{3}$-porous in \begin{equation} Z=K\cup\bigcup_mW_m. \end{equation} We let $Y$ be the closure of $Z$ in $X$ and show that $Y$ is compact by showing that $Z$ is totally bounded. Let $\epsi>0$. By compactness of $K$, finitely many balls $B_1,\cdots, B_{M_\epsi}$ cover $K$; As $B_1\cup\cdots\cup B_{M_\epsi}$ is an open neighbhourhood of $K$ and as $W_m$ lies in a $\frac{1}{m}$-neighbourhood of $K$, there is $m_0$ such that $m\ge m_0$ implies \begin{equation} W_m\subset B_1\cup\cdots\cup B_{M_\epsi}; \end{equation} note that $\bigcup_{m<m_0}W_m$ is finite so finitely many balls of radius $\epsi$ are needed to cover it. \end{proof} \par We now prove Theorem \ref{thm:diff_char1}. \begin{proof}[Proof of Theorem \ref{thm:diff_char1}] We first prove necessity. Suppose that $S$ is a Borel $c$-porous subset of $X$ with $\mu(S)>0$. Then there is a compact $K\subset S$ with $\mu(K)>0$ and, by Lemma \ref{lem:porous_completion}, a compact $Y\supset K$ in which $K$ is $\frac{2c}{3}$-porous. By Lemma \ref{lem:por_gap} $K\in\Gap_0(Y)$ and so, by Lemma \ref{lem:ind_lip} and as $(Y,\mu\mrest Y)$ is a $\sigma$-differentiability space, $\mu(K)=0$, yielding a contradiction. This implies that $\mu(S)=0$. As observed in \cite{bate_speight}, \cite[Thm.~3.6]{preiss_porosity} implies that if $\mu$ annihilates porous sets, then $\mu$ is asymptotically doubling. We can therefore find disjoint compact sets $K_\alpha$ with $\mu(K_\alpha)>0$, $\mu(X\setminus\bigcup_\alpha K_\alpha)=0$ and $\mu\mrest K_\alpha$ doubling on $K_\alpha$. \par We now show that on a full $\mu$-measure Borel subset of $K_\alpha$, the set $K_\alpha$ is not $c$-porous for any $c\in(0,\frac{1}{2})$. Let \begin{equation} P_\alpha=\left\{x\in K_\alpha:\exists c\in\left(0,\frac{1}{2}\right):\text{$K_\alpha$ is $c$-porous at $x$}\right\}; \end{equation} note that \begin{equation} O_{c,\alpha}(r)=\left\{x\in X:\exists y\in\ball x,r.:\setdist K_\alpha,\{y\}.>c\dist x,y.>0\right\} \end{equation} is open; then \begin{equation} P_\alpha=\bigcup_{c\in\rational\cap(0,\frac{1}{2})}\bigcap_{n\in\natural}\bigcup_{r\in(0,\frac{1}{n})}O_{c,\alpha}(r)\cap K_\alpha \end{equation} is a $G_{\delta\sigma}$\footnote{$K_\alpha$, being closed, is a $G_\delta$} and is a countable union of the porous sets \begin{equation} \bigcap_{n\in\natural}\bigcup_{r\in(0,\frac{1}{n})}O_{c,\alpha}(r)\cap K_\alpha\quad\text{($c\in\rational\cap(0,\frac{1}{2})$).} \end{equation} Thus $\mu(P_\alpha)=0$. \par Let $S\in\Gap(X)$ and let $(f,\alpha,\beta)$ be as in the defining property of $\Gap$. Then \eqref{eq:npor_Lip_equal} implies that if $S\cap(K_\alpha\setminus P_\alpha)\ne\emptyset$, then $S\cap(K_\alpha\setminus P_\alpha)\in\Gap(K_\alpha)$. But by Lemma \ref{lem:ind_lip}, $(K_\alpha,\mu\mrest K_\alpha)$ being a $\sigma$-differentiability space, $\mu(S\cap K_\alpha)=0$ so that $\mu(S)=0$. \par We now prove sufficiency. If $S$ is Borel and $c$-porous, $S\in\Gap_0(X)$ by Lemma \ref{lem:por_gap}, so $S$ is $\mu$-null by hypothesis. Then, as in the necessity argument, $\mu$ is $\sigma$-asymptotically doubling and we find disjoint compact sets $K_\alpha$ with $\mu(K_\alpha)>0$, $\mu(X\setminus\bigcup_\alpha K_\alpha)=0$ and $\mu\mrest K_\alpha$ doubling on $K_\alpha$; by Corollary \ref{derbound} and Theorem \ref{thm:free_dec} we can partition the $K_\alpha$ and assume that the module $\wder\mu\mrest K_\alpha.$ is free of rank $N_\alpha$. As in the necessity argument, $P_\alpha$ is $\mu$-null so it suffices to show, by \eqref{eq:npor_Lip_equal}, that each $(K_\alpha, \mu\mrest K_\alpha)$ is a differentiability space in order to conclude that $(X,\mu)$ is a $\sigma$-differentiability space. \par Let $S\in\Gap(\mu\mrest K_\alpha,K_\alpha)\subset\Gap(\mu,X)$; by Lemma \ref{lem:gap_incl} it contains a $\mu$-full measure Borel subset of type $\Gap(X)$, so $\mu\mrest K_\alpha(S)=0$. By definition of $\Gap$ sets, this implies that for every $f:K_\alpha\to\real$ \begin{equation}\label{eq:diff=Lip_loc} \locnorm df,{\wform\mu\mrest K_\alpha.}.(x)=\biglip f(x)\quad\text{(for $\mu\mrest K_\alpha$-a.e.~$x$);} \end{equation} by Corollary \ref{cor:pseudoduality} up to furthering partitioning the $K_\alpha$, we can assume that there is a basis $\{D_{\alpha,i}\}_{i=1}^{N_\alpha}$ of $\wder\mu\mrest K_\alpha.$ and there are $1$-Lipschitz functions $\{g_{\alpha,j}\}_{j=1}^{N_\alpha}$ with \begin{equation} D_{\alpha,i}g_{\alpha,j}=\delta_{ij}\chi_{K_\alpha}. \end{equation} Because of \eqref{eq:diff=Lip_loc}, for each $\epsi>0$ there are disjoint Borel $\{U_{\alpha,\beta}\}$ which are subsets of $K_\alpha$, which satisfy $\mu\left(K_\alpha\setminus\bigcup_\beta U_{\alpha,\beta}\right)=0$, and such that there are derivations $\{D_{\alpha,\beta}\}$ with \begin{align} \locnorm D_{\alpha,\beta},{\wder\mu\mrest U_{\alpha,\beta}.}.&=1\\ D_{\alpha,\beta} f&\ge\biglip f-\epsi\quad\text{on $U_{\alpha,\beta}$;} \end{align} moreover, there are $\{\lambda_{i,\alpha,\beta}\}_{i=1}^{N_\alpha}\subset\bborel K_\alpha.$ with \begin{equation} D_{\alpha,\beta}=\sum_{i=1}^{N_\alpha}\lambda_{i,\alpha,\beta}D_{\alpha,i}; \end{equation} evaluating on the $g_{\alpha,j}$ gives \begin{equation} \|\lambda_{i,\alpha,\beta}\|_{L^\infty(\mu\mrest K_\alpha)}\le1, \end{equation} so that \begin{equation} |D_{\alpha,\beta}f|\le N_\alpha \max_{i=1,\cdots,N_\alpha}|D_{\alpha,i}f|. \end{equation} Thus \begin{equation}\label{eq:reverproof} \biglip f(x)\le N_\alpha\max_{i=1,\cdots,N_\alpha}|D_{\alpha,i}f|(x)\quad\text{(for $\mu\mrest K_\alpha$-a.e.~$x$),} \end{equation} and \cite[Thm.~2.148]{derivdiff} shows that $(K_\alpha, \mu\mrest K_\alpha)$ is a differentiability space. \end{proof} \par We now prove Theorem \ref{thm:diff_char2}. \begin{proof}[Proof of Theorem \ref{thm:diff_char2}] The characterization in terms of (1) is immediate because of Theorem \ref{thm:diff_char1}, in particular of \eqref{eq:diff=Lip_loc}. The characterization in terms of (2) follows from (1) and Theorem \ref{thm:weaver_fnorm_char}. The characterization in terms of (3) is an immediate consequence of (2). \end{proof} \par We now give an alternative proof of Theorem \ref{partialderivatives} which relies on Theorem \ref{thm:diff_char1}. \begin{proof}[Alternative proof of Theorem \ref{partialderivatives}] Let $(U,x)$ be an $n$-dimensional chart for the differentiability space $(X,\mu)$; then \eqref{eq:reverproof} shows that it is possible to find derivations $\{D_i\}_{i=1}^n\subset\wder\mu\mrest U.$ such that, for each $f\in\lipalg X.$, \begin{equation} \label{eq:altproof} \biglip f(x)\le n\max_{i=1,\cdots,n}|D_{i}f|(x) \end{equation} holds for $\mu\mrest U$-a.e.~$x$. We can then apply \cite[Lem.~6.20]{derivdiff} to conclude that the partial derivative operators $\frac{\partial}{\partial x^j}$ are derivations; note that even though in \cite{derivdiff} the measure $\mu$ is assumed doubling, this property of $\mu$ is used only to ensure that the Lebesgue's Differentiation Theorem holds for the measure $\mu$. However, the Lebesgue Differentiation Theorem holds for asymptotically doubling measures and so it is possible to apply the results in \cite{derivdiff}. \end{proof} \section{Technical tools}\label{sec:technical-tools} \subsection{An approximation scheme}\label{subsec:an-appr-scheme} The goal of this Subsection is to prove Theorem \ref{onedimapprox_multi}, whose proof requires some preparation. The first step is the geometric construction of a cylinder. For this we need to compare the notion of nullity for fragments in different spaces. We introduce the following auxiliary definition: \begin{defn}\label{def:classes_of_fragments_approx} For $\delta>0$, metric spaces $X$ and $W$, Lipschitz functions $G:X\to W$ and $f:X\to\real^q$, we define \begin{multline} \frags(X,f,G,\delta,w,\alpha)=\Biggl\{\gamma\in\frags(X): \text{$(\langle w,f\rangle\circ\gamma)'(t)\ge\delta\metdiff G\circ\gamma(t)$}\\ \text{and $(f\circ\gamma)'(t)\in\bar\cone(w,\alpha)$ for $\lebmeas$-a.e. $t\in \dom\gamma$}\Biggr\}. \end{multline} \end{defn} \begin{lem}\label{nullembedd_multi} Assume that $\psi:X\to Y$ is an isometric embedding and $S\subset X$. Then $S$ is $\frags(X,f,G,\delta,w,\alpha)$-null if and only if $\psi(S)$ is \begin{equation} \frags(Y,\tilde f,\tilde G,\delta,w,\alpha)\text{-null} \end{equation} where $\tilde f$ and $\tilde G$ are any Lipschitz extensions of $f\circ\psi^{-1}$, $G\circ\psi^{-1}$. \end{lem} \begin{proof} Necessity is proven by contrapositive, assuming that there is a \begin{equation} \gamma\in\frags(Y,\tilde f,\tilde G,\delta,w,\alpha) \end{equation} with $\hmeas 1.(\gamma\cap\psi(S))>0$. It is then possible to find a compact $K'\subset\dom\gamma$ with $\gamma(K')\subset \psi(S)$ and $\hmeas 1.\left((\gamma|K')\cap\psi(S)\right)>0$. Let $\tilde\gamma=\psi^{-1}\circ (\gamma|K')$. Then \begin{align} (f\circ\tilde\gamma)'(t)&=(\tilde f\circ\gamma)'(t)\\ \metdiff\tilde\gamma(t)&=\metdiff\gamma(t)\\ \metdiff\tilde G\circ\gamma(t)&=\metdiff G\circ\tilde\gamma(t) \end{align} at any $t\in K'$ which is a Lebesgue density point for $K'$. In particular, \begin{equation} \tilde\gamma\in\frags(X,f,G,\delta,w,\alpha) \end{equation} and $\hmeas 1.(\tilde\gamma\cap S)>0$. \par Sufficiency is proven by observing that the previous part of the argument allows to identify $\frags(X,f,G,\delta,w,\alpha)$ with $\frags(\psi(X),\tilde f,\tilde G,\delta,w,\alpha)$ via $\gamma\mapsto\psi\circ\gamma$ because the metric differential and the derivative of a Lipschitz function along a fragment are determined, at a point $t$, by the behaviour of the fragment on a subset for which $t$ is a Lebesgue density point. Sufficiency then follows because $\frags(X,f,G,\delta,w,\alpha)\subset\frags(Y,\tilde f,\tilde G,\delta,w,\alpha)$. \end{proof} \par We now introduce the definition of cylinder. \begin{defn} Let $X$ be a compact metric space and $M>0$. The cylinder $\Cyl(X,M)$ is the compact metric space $X\times[0,M]$ with metric \begin{equation} \dist {(x_1,t_1)},{(x_2,t_2)}.=\max\left(\dist x_1,x_2.,|t_1-t_2|\right). \end{equation} The projection on the base $X$ will be denoted by $\pi$ and the projection on the axis $[0,M]$ by $\tau$. \end{defn} \begin{rem} Note that if $X$ is geodesic, $\Cyl(X,M)$ is geodesic. If \begin{equation} \gamma:[0,\dist x_1,x_2.]\to X \end{equation} denotes a unit speed geodesic joining $x_1$ to $x_2$, then \begin{align} \sigma:[0,\dist {(x_1,t_1)},{(x_2,t_2)}.]&\to X\\ s&\mapsto\left(\gamma(\frac{s\dist x_1,x_2.}{\dist {(x_1,t_1)},{(x_2,t_2)}.}),t_1+ \frac{s(t_2-t_1)}{\dist {(x_1,t_1)},{(x_2,t_2)}.}\right) \end{align} is a unit speed geodesic joining $(x_1,t_1)$ to $(x_2,t_2)$. \end{rem} \par We now reduce the general approximation problem to the cylindrical case. \begin{lem}\label{con_red_multi} If $f:X\to\real^q$ is Lipschitz and $S\subset X$ is compact and \begin{equation}\frags(X,f,\delta,w,\alpha)\text{-null}, \end{equation} we can assume that: \begin{enumerate} \item The space $X$ is the cylinder $\Cyl(Y,M)$ for $Y=Z\times Q$, where $Z$ is a compact geodesic metric space, $Q$ is a compact rectangle in $\real^{q-1}$ and $M\le\diam X$. \item If $\tilde\tau:Y\to Q$ is the projection, $(\pi_w^\perp\circ f,\langle w,f\rangle)=(\tilde\tau,\tau)$. \item If $\pi_Z:Y\to Z$ is the projection, $S$ is $\frags(\Cyl(Y,M),(\tilde\tau,\tau),\pi_Z,\delta,e_q,\alpha)$-null. \end{enumerate} \end{lem} \par Points of $\Cyl(Y,M)$ will be denoted by $(z,v,t)$ where $z\in Z$, $v\in Q$ and $t\in[0,M]$. \begin{proof} Considering a Kuratowski embedding of $X$ in $l^\infty$ we can assume that $X$ lies in $l^\infty$. Let $Z$ denote the closed convex hull of $X$ in $l^\infty$ which is compact (\cite[Thm.~3.25]{rudin-functional}). Moreover, postcomposing $f$ with an orthogonal transformation $A\in O(q)$, we can assume that the vector $w=e_q$, where $\{e_1,\cdots,e_q\}$ is the standard Euclidean basis of $\real^q$. Note that this does not affect the global Lipschitz constant of $f$ with respect to the $l^2$-norm $\|\cdot\|_2$. \par Replace $f$ by taking a McShane extension on $Z$ of each component $f_i$: this does not affect the $\frags(X,f,\delta,w,\alpha)$-nullity of $S$ even though it might increase the global Lipschitz constant of $f$ by a factor $\sqrt{q}$. Postcomposing $f$ with an affine transformation $\lambda x+b$, where $(\lambda,b)\in (0,\infty)\times\real^q$, we can assume that: \begin{itemize} \item The function $f$ is $1$-Lipschitz with respect to the norm $\|\cdot\|_2$. \item The minimum of $f_i$ is $0$. \end{itemize} Note that postcomposing affects $\delta$ but not $\alpha$; furthermore, considering a geodesic from a point where the minimum is attained to a point where the maximum is attained, all the values in $[0,\max f_i]$ are assumed by $f_i$. \par Let $M=\max f_q\le\diam Z=\diam X$ and \begin{equation} Q=[0,\max f_1]\times\cdots\times[0,\max f_{N-1}]\subset\real^{q-1} \end{equation} with the metric induced by the norm $\|\cdot\|_2$. Let $Y=Z\times Q$ with the metric \begin{equation} \dist {(z_1,v_1)},{(z_2,v_2)}.=\max\left(\dist z_1,z_2.,\|v_1-v_2\|_2\right). \end{equation} Let $\tilde\tau$ denote the projection on the $Q$ factor. The metric space $Y$ is geodesic because if \begin{equation} \gamma:[0,\dist z_1,z_2.]\to Z \end{equation} denotes a unit speed geodesic joining $z_1$ to $z_2$, then \begin{align} \sigma:[0,\dist {(z_1,v_1)},{(z_2,v_2)}.]&\to X\\ s&\mapsto\left(\gamma(\frac{s\dist z_1,z_2.}{\dist {(z_1,v_1)},{(z_2,v_2)}.}),v_1+ \frac{s(v_2-v_1)}{\dist {(z_1,v_1)},{(z_2,v_2)}.}\right) \end{align} is a unit speed geodesic joining $(z_1,v_1)$ to $(z_2,v_2)$. We will replace $X$ by $Y$. \par Considering the isometric embedding \begin{equation} \begin{aligned} \psi:Z&\to \Cyl(Y,M)\\ z&\mapsto((z,f_1(z),\cdots,f_{q-1}(z)),f_q(z)), \end{aligned} \end{equation} we conclude that $\tilde\tau$ extends $(f_1,\cdots,f_{q-1})\circ\tilde\psi^{-1}$ and $\tau$ extends $f_q\circ\psi^{-1}$; noting that for $\gamma\in\frags(Z)$ \begin{equation} \metdiff\gamma=\metdiff\left(\pi_Z\circ\psi\right)\circ\gamma, \end{equation} we conclude from Lemma \ref{nullembedd_multi} that $\psi(S)$ is $\frags(\Cyl(Y,M), (\tilde\tau,\tau),\pi_Z,\delta,e_q,\alpha)$-null. The Radon measure $\mu$ on $X$ is replaced by the pushforward $\mpush\psi.\mu$. \end{proof} \par The next step is to cover $S$ by \textbf{thin} strips. \begin{defn} Given a Lipschitz function $f:Y\to[0,M]$ and $h>0$ we define the \textbf{open strip of width $h$ above $f$} by \begin{equation} \strip f,h.=\left\{(y,t)\in\Cyl(Y,M): t\in(f(y),f(y)+h)\right\}; \end{equation} note that $\strip f,h.$ is an open set. The \textbf{lower and upper hypersurfaces bounding $\strip f,h.$} are the closed sets: \begin{align} \partial_-\strip f,h.&=\left\{(y,t)\in\Cyl(Y,M): t=f(y)\right\};\\ \partial_+\strip f,h.&=\left\{(y,t)\in\Cyl(Y,M): t=f(y)+h\right\}. \end{align} \end{defn} \begin{lem}\label{strip_cov_multi} Assume that the compact $S\subset\Cyl(Y,M)$ is $\frags(\Cyl(Y,M),\allowbreak(\tilde\tau,\allowbreak\tau),\allowbreak\pi_Z, \allowbreak\delta,\allowbreak e_q,\allowbreak\alpha)$-null; then for each $n\in\natural$ the set $S$ can be covered by $M(n)$ open strips \begin{equation}\left\{\strip f_i,2\frac{\delta+\cot\alpha+1}{n}.\right\}_{i=1}^{M(n)} \end{equation} where the $f_i$ are $1$-Lipschitz with respect to the distance \begin{equation} d_{\delta,\alpha}((z_1,v_1,t_1),(z_2,v_2,t_2))=\delta\max\left(\dist z_1,z_2.,|t_1-t_2|\right)+\cot\alpha\|v_1- v_2\|_2, \end{equation}and \begin{equation} \lim_{n\to\infty}\frac{M(n)}{n}=0. \end{equation} \end{lem} \begin{proof} Let ${\mathscr{N}}_Y$ be a $\frac{1}{n}$-net in $Y$: a maximal set of points in $Y$ which are separated by a distance $\ge\frac{1}{n}$. By compactness of $Y$, the net ${\mathscr{N}}_Y$ is finite. Let ${\mathscr{N}}_{[0,M]}$ be a $\frac{1}{n}$-net in $[0,M]$, which is finite by compactness. Then ${\mathscr{N}}_{\Cyl(Y,M)}={\mathscr{N}}_Y\times {\mathscr{N}}_{[0,M]}$ is a $\frac{1}{n}$-net in $\Cyl(Y,M)$. Note that the open balls of radius $\frac{1}{n}$ centred on the points ${\mathscr{N}}_{\Cyl(Y,M)}$ cover $\Cyl(Y,M)$. We now consider the set of those points in ${\mathscr{N}}_{\Cyl(Y,M)}$ such that the corresponding balls intersect $S$: \begin{equation} {\mathscr{N}}_{\Cyl(Y,M)}^S=\left\{(y,t)\in{\mathscr{N}}_{\Cyl(Y,M)}: \ball {(y,t)},\frac{1}{n}.\cap S\ne\emptyset\right\}. \end{equation} \par On ${\mathscr{N}}_{\Cyl(Y,M)}^S$ define the partial order $\preceq$: \begin{multline} (z_1,v_1,t_1)\preceq(z_2,v_2,t_2)\Longleftrightarrow t_2-t_1\ge\delta\dist z_1,z_2.\\\quad\text{and}\quad(t_2-t_1)\tan\alpha\ge\|v_1-v_2\|_2; \end{multline} the binary relation $\preceq$ is reflexive; it is antisymmetric because if $(z_1,v_1,t_1)\preceq(z_2,v_2,t_2)$ and $(z_2,v_2,t_2)\preceq(z_1,v_1,t_1)$, then $\dist z_1,z_2.=0$, $\|v_1-v_2\|_2=0$ and $t_2-t_1=0$. Transitivity follows from the triangle inequality: \begin{align} t_2-t_1&\ge\delta\dist z_1,z_2.&&\text{and}&(t_2-t_1)\tan\alpha&\ge\|v_1-v_2\|_2\\ t_3-t_2&\ge\delta\dist z_2,z_3.&&\text{and}&(t_3-t_2)\tan\alpha&\ge\|v_2-v_3\|_2 \end{align} imply \begin{align} t_3-t_1&\ge\delta\dist z_1,z_2.+\delta\dist z_2,z_3.\ge\delta\dist z_1,z_3.,\\ (t_3-t_1)\tan\alpha&\ge\|v_1-v_2\|_2+\|v_2-v_3\|_2\ge\|v_1-v_3\|_2. \end{align} \par Let $M(n)$ denote the length of a maximal chain in $({\mathscr{N}}_{\Cyl(Y,M)}^S,\preceq)$\footnote{Note that the nets depend on $n$}. If \begin{equation} \lim_{n\to\infty}\frac{M(n)}{n}\ne0, \end{equation} there are naturals $n\to\infty$ with $M(n)\ge Cn$ for $C>0$. Let $\{z_i^{(n)},v_i^{(n)},t_i^{(n)})\}_{i=1}^{M(n)}$ denote a maximal chain with the $t_i^{(n)}$ in increasing order. We want to costruct a biLipschitz path $\gamma_n:[0,M]\to\Cyl(Y,M)$: \begin{itemize} \item On $[t_i^{(n)},t_{i+1}^{(n)}]$, $\gamma_n$ is defined as a constant speed geodesic joining $(z_i^{(n)},\allowbreak v_i^{(n)},\allowbreak t_i^{(n)})$ to $(z_{i+1}^{(n)},v_{i+1}^{(n)},t_{i+1}^{(n)})$. \item On $[0,t_1^{(n)}]$, $\gamma_n$ is defined as a constant speed geodesic from $(z_1^{(n)},v_1^{(n)},0)$ to $(z_1^{(n)},v_1^{(n)},t_1^{(n)})$. \item On $[t_{M(n)}^{(n)},M]$, $\gamma_n$ is defined as a constant speed geodesic from $(z_{M(n)}^{(n)},\allowbreak v_{M(n)}^{(n)},\allowbreak t_{M(n)}^{(n)})$ to $(z_{M(n)}^{(n)},v_{M(n)}^{(n)},M)$. \end{itemize} Note that \begin{equation} 1\le\frac{\dist{(z_i^{(n)},v_i^{(n)},t_i^{(n)})},{(z_{i+1}^{(n)}, v_{i+1}^{(n)},t_{i+1}^{(n)})}.} {t_{i+1}^{(n)}-t_i^{(n)}}\le\max\left(\frac{1}{\delta},\tan\alpha\right); \end{equation} this implies that $\metdiff\gamma_n\in[1,\max\left(\frac{1}{\delta},\tan\alpha\right)]$. Moreover, $(\tau\circ\gamma_n)'=1$ $\lebmeas$-a.e.~implying that $\gamma_n$ is $(1,\max\left(\frac{1}{\delta},\tan\alpha\right))$-biLipschitz. Furthermore, as $\pi_Z\circ\gamma_n$ is $\frac{1}{\delta}$-Lipschitz and $\tilde\tau\circ\gamma_n$ is $\tan\alpha$-Lipschitz, $\gamma_n\in\pathsp(\Cyl(Y,M),(\tilde\tau,\tau),\pi_Z,\delta,e_q,\alpha)$. By Ascoli-Arzel\`a we can assume that the $\gamma_n$ converge uniformly to a \begin{equation} \left(1,\max\left(\frac{1}{\delta},\tan\alpha\right)\right)\text{-biLipschitz}\;\gamma:[0,M]\to\Cyl(Y,M). \end{equation} Note that $(\tau\circ\gamma)'=1$ $\lebmeas$-a.e., that $\pi_Z\circ\gamma$ is $\frac{1}{\delta}$-Lipschitz and that $\tilde\tau\circ\gamma$ is $\tan\alpha$-Lipschitz; this implies that that $\gamma\in\pathsp(\Cyl(Y,M),(\tilde\tau,\tau),\pi_Z,\delta,e_q,\alpha)$. Moreover, \begin{equation} \gamma_n\left([t_i^{(n)}-\frac{\min(\delta,\cot\alpha)}{4n},t_i^{(n)}+\frac{\min(\delta,\cot\alpha)}{4n}]\right)\subset\ball {(z_i^{(n)},v_i^{(n)},t_i^{(n)})},\frac{1}{n}. \end{equation} and the intervals $[t_i^{(n)}-\frac{\min(\delta,\cot\alpha)}{4n},t_i^{(n)}+\frac{\min(\delta,\cot\alpha)}{4n}]$ are disjoint\footnote{For $i=1,M(n)$ one might have to consider one-sided intervals}; in particular, there is a compact $K_n\subset[0,M]$ with \begin{equation} \lebmeas(K_n)\ge\frac{\min(\delta,\cot\alpha) (M(n)-2)}{2n}\ge \left(C-\frac{1}{n}\right)\frac{\min(\delta,\cot\alpha)}{2} \end{equation} on which $\dist \gamma_n,S.\le \frac{3}{2n}$. We can pass to a subsquence such that the $K_n$ converge to a compact $K$. Regularity of the Lebesgue measure will imply that \begin{equation} \lebmeas(K)\ge\limsup_{n\to\infty}\lebmeas(K_n) \end{equation} because any $\epsi$-neighbourhood of $K$ will eventually contain the $K_n$. But on $K$ $\dist\gamma,S.=0$, contradicting that $S$ is $\frags(\Cyl(Y,M),(\tilde\tau,\tau),\pi_Z,\delta,e_q,\alpha)$-null. \par By Mirsky's Lemma (dual to Dilworth's Lemma, \cite{mirsky_dil}), there are $M(n)$ antichains ${\mathscr{A}}_1,\cdots,{\mathscr{A}}_{M(n)}$ covering ${\mathscr{N}}_{\Cyl(Y,M)}^S$. Recall that in an antichain no two elements are comparable with respect to the order. So if $(z,v,t),(z',v',t')\in{\mathscr{A}}_p$, either \begin{equation} |t-t'|\le\delta\dist z,z'.\quad\text{or}\quad|t-t'|\le\cot\alpha\|v-v'\|_2; \end{equation} in particular, ${\mathscr{A}}_p$ can be regarded as the graph of a function $g_p:\pi_Z\times\tilde\tau({\mathscr{A}}_p)\to\real$ which is $1$-Lipschitz with respect to the distance $d_{\delta,\alpha}$; these functions can be extended to $1$-Lipschitz functions $g_p:Y\to\real$. If $(z_i^{(n)},v_i^{(n)},t_i^{(n)})\in{\mathscr{A}}_p$ and if $(z,v,t)\in \ball (z_i^{(n)},v_i^{(n)},t_i^{(n)}),\frac{1}{n}.$, then \begin{equation} \begin{split} |g_p(z,v)-t|&\le|g_p(z,v)-g_p(z_i^{(n)},v_i^{(n)})| +\underbrace{|g_p(z_i^{(n)},v_i^{(n)})-t_i^{(n)}|}_{=0}+|t_i^{(n)}-t|\\ &\le\delta\dist z,z_i^{(n)}.+\cot\alpha\|v-v_i^{(n)}\|_2+|t_i^{(n)}-t|\\ &<\frac{\delta+\cot\alpha+1}{n}; \end{split} \end{equation} in particular, \begin{equation} S\subset\bigcup_{p=1}^{M(n)}\strip g_p-\frac{\delta+\cot\alpha+1}{n},2\frac{\delta+\cot\alpha+1}{n}.. \end{equation} \end{proof} \par We now rearrange the strips in \emph{increasing order}. \begin{lem} Given open strips $\{\strip f_i,h.\}_{i=1}^{N}$ ($h>0$) where the $f_i$ are $1$-Lipschitz with respect to the distance $d_{\delta,\alpha}$, there are $1$-Lipschitz functions (with respect to $d_{\delta,\alpha}$) $\{\tilde f_i\}_{i=1}^N$ such that \begin{equation} \tilde f_1\le\tilde f_2\le\cdots\le\tilde f_N, \end{equation} and \begin{equation} \bigcup_{i=1}^N\strip f_i,h. =\bigcup_{i=1}^N\strip \tilde f_i,h.. \end{equation} \end{lem} \begin{proof} We argue by induction on $N$, for $N=1$ there being nothing to prove. We now assume the result true for $N$ and prove it for $N+1$. By the inductive hypothesis we can assume that the first $N$ strips are defined by functions \begin{equation} f_1\le f_2\le \cdots\le f_N; \end{equation} the sorting is now done with respect to $f_{N+1}$. Let $F_1=\min(f_1,f_{N+1})$ and $G_1=\max(f_1,f_{N+1})$. For consistency we will also let $G_0=f_{N+1}$ and for $1<i\le N$ we will define $F_i=\min(f_i,G_{i-1})$ and $G_i=\max(f_i,G_{i-1})$. We will finally let $F_{N+1}=G_N$. We will show that the desired strips are defined by the functions $F_i$. Note that taking maxima and minima of $1$-Lipschitz functions produces $1$-Lipschitz functions. We show that $F_i(x)$ is nondecreasing in $i$ for $i\in\{1,\cdots,N+1\}$. In fact, \begin{equation} F_i=\min(f_i,G_{i-1})\le\min(f_{i+1},G_{i-1})\le\min(f_{i+1},G_{i})=F_{i+1}. \end{equation} Similarly, \begin{equation} F_N=\min(f_N,G_{N-1})\le G_N=F_{N+1}. \end{equation} We want to show that \begin{equation} \strip F_i,h.\subset\bigcup_{i=1}^{N+1}\strip f_i,h.; \end{equation} assume $(y,t)\in\strip F_i,h.$. If $F_i(y)=f_i(y)$ then $(y,t)\in\strip f_i,h.$. If this is not the case, then $F_i(y)=G_{i-1}(y)$ (recall that $G_0=f_{N+1}$ so that if $i-1=0$ we have $(y,t)\in\strip f_{N+1},h.$). If $G_{i-1}(y)=f_{i-1}(y)$ then $(y,t)\in\strip f_{i-1},h.$. Otherwise $G_{i-1}(y)=G_{i-2}(y)$ and one continues to argue in a similar way till one either stops at some $G_{i-k}(y)=f_{i-k}(y)$ or at $G_0(y)$. We want to show that \begin{equation} \strip f_i,h.\subset\bigcup_{i=1}^{N+1}\strip F_i,h.; \end{equation} the argument is similar to the one above but it is better to distinguish between the cases $i\in\{1,\cdots,N\}$ and $i=N+1$. Assume that $1\le i\le N$ and $(y,t)\in\strip f_i,h.$. From the definition of $F_i$ and $G_i$ we see that either $f_i(y)=F_i(y)$ or $f_i(y)=G_i(y)$. In the first case $(y,t)\in\strip F_i,h.$. In the second case one observes that either $G_i(y)=F_{i+1}(y)$ or $G_i(y)=G_{i+1}(y)$. In the first case $(y,t)\in\strip F_{i+1},h.$. Otherwise one continues to argue in a similar way till one either stops at some $G_{i+k}(y)=F_{i+k}(y)$ or at $G_{N}(y)=F_{N+1}(y)$. If $(y,t)\in\strip f_{N+1},h.$ then either $f_{N+1}(y)=F_1(y)$ or $f_{N+1}(y)=G_1(y)$. In the first case $(y,t)\in\strip F_1,h.$. In the second case one keeps arguing as above. \end{proof} \par The next step is to make the strips disjoint to induce an order structure on the set of strips. The cost to pay is to make the strips slightly bigger and allow to cover a full measure subset of $S$, but not necessarily the whole of $S$. \begin{lem}\label{strip_cov_disj_multi} Assume that $S\subset\bigcup_{i=1}^N\strip f_i,h.$ ($h>0$) where the functions $f_i$ are $(1,d_{\delta,\alpha})$-Lipschitz. Then there are $(1,d_{\delta,\alpha})$-Lipschitz functions $\{g_i\}_{i=1}^N$ and $\lambda_i\in(1,\frac{3}{2})$ such that \begin{equation} g_1\le g_2\le\cdots\le g_N, \end{equation} \begin{equation} \strip g_i,\lambda_i h.\cap\strip g_j,\lambda_j h.=\emptyset\quad\text{(for $i\ne j$),} \end{equation} and \begin{equation} \mu\left(S\setminus\bigcup_{i=1}^N\strip g_i,\lambda_i h.\right)=0. \end{equation} \end{lem} \begin{proof} Because of the previous Lemma we can assume that for each $y$ the map $i\mapsto f_i(y)$ is nondecreasing in $i$. As $\mu$ is a finite measure and $(1,\frac{3}{2})$ is uncountable, it is possible to find $\lambda_1\in(1,\frac{3}{2})$ such that \begin{equation} \mu\left(\partial_+\strip f_1,\lambda_1 h.\right)=0. \end{equation} Let $g_1=f_1$ so that $\strip f_1,h.\subset\strip g_1,\lambda_1 h.$. For $j\in\{2,\cdots,N\}$ we let $g_j=\max(g_{j-1}+\lambda_{j-1}h,f_j)$ and choose $\lambda_j\in(1,\frac{3}{2})$ such that \begin{equation} \mu\left(\partial_+\strip g_j,\lambda_j h.\right)=0; \end{equation} this is possible because $\mu$ is a finite measure and $(1,\frac{3}{2})$ is uncountable. We want now to show that \begin{equation}\label{esscover} \strip f_j,h.\subset\bigcup_{i=1}^j\strip g_i,\lambda_ih.\cup\bigcup_{i=1}^{j-1} \partial_+\strip g_i,\lambda_ih.. \end{equation} If $f_j(y)\ge g_{j-1}(y)+\lambda_{j-1}h$ then $g_j(y)=f_j(y)$ so that \begin{equation} \left(f_j(y),f_j(y)+h\right)\subset\left(g_{j}(y),g_{j}(y)+\lambda_{j}h\right); \end{equation} otherwise $f_j(y)< g_{j-1}(y)+\lambda_{j-1}h$ which implies $g_j(y)=g_{j-1}(y)+\lambda_{j-1}h$. If $f_j(y)\ge g_{j-1}(y)$ we conclude that \begin{multline} \left(f_j(y),f_j(y)+h\right)\subset\left(g_{j-1}(y),g_{j-1}(y)+\lambda_{j-1} h\right)\cup \left\{g_{j-1}(y)+\lambda_{j-1} h\right\}\\\cup\left(g_j(y),g_j(y)+\lambda_j h\right); \end{multline} if $f_j(y)<g_{j-1}(y)$ then $f_{j-1}(y)<g_{j-1}(y)$ which implies $g_{j-1}(y)=g_{j-2}(y)+\lambda_{j-2}h$. If $f_j(y)\ge g_{j-2}(y)$ we get, similarly to the previous equation, \begin{multline} \left(f_j(y),f_j(y)+h\right)\subset\left(g_{j-2}(y),g_{j-2}(y)+\lambda_{j-2} h\right)\\\cup \left\{g_{j-2}(y)+\lambda_{j-2} h\right\}\cup\left(g_{j-1}(y),g_{j-1}(y)+\lambda_{j-1} h\right); \end{multline} one can continue to argue inductively by decreasing the indices till \begin{equation} g_k(y)\le f_j(y)<g_k(y)+\lambda_{k}h=g_{k+1}(y); \end{equation} in that case, \begin{multline} \left(f_j(y),f_j(y)+h\right)\subset\left(g_{k}(y),g_{k}(y)+\lambda_{k} h\right)\cup \left\{g_{k}(y)+\lambda_{k} h\right\}\\\cup\left(g_{k+1}(y),g_{k+1}(y)+\lambda_{k+1} h\right); \end{multline} this argument shows that \eqref{esscover} holds. The result now follows, noting also that by definition of the functions $\{g_j\}$, for each $y$ the map $j\mapsto g_j(y)$ is nondecreasing in $j$. \end{proof} \par We can now prove Theorem \ref{onedimapprox_multi}. \begin{proof}[Proof of Theorem \ref{onedimapprox_multi}] By Lemma \ref{con_red_multi} we can assume that we are in the cylindrical case so that we will approximate $\tau$; in particular, we have to show that we will obtain approximations which are globally $1$-Lipschitz with respect to the distance: \begin{equation} D\left((z_1,v_1,t_1),(z_2,v_2,t_2)\right)=\max\left(|t_1-t_2|,d_{\delta,\alpha}\left((z_1,v_1),(z_2,v_2)\right)\right). \end{equation} By Lemma \ref{strip_cov_disj_multi} we can ``$\mu$-essentially'' cover $S$ by $M(n)$ disjoint strips \begin{equation}\left\{\strip f_i,2\lambda_i\frac{\delta+\cot\alpha+1}{n}.\right\}_{i=1}^{M(n)}, \end{equation} where the functions $f_i$ are $1$-Lipschitz with respect to the distance $d_{\delta,\alpha}$. One can define a natural total order relation between the hypersurfaces bounding the open strips: \begin{equation} \begin{split} &\partial_-\strip f_i,2\lambda_i\frac{\delta+\cot\alpha+1}{n}.\prec \partial_+\strip f_i,2\lambda_i\frac{\delta+\cot\alpha+1}{n}.\\ \prec&\partial_-\strip f_{i+1},2\lambda_{i+1}\frac{\delta+\cot\alpha+1}{n}.\prec \partial_+\strip f_{i+1},2\lambda_{i+1}\frac{\delta+\cot\alpha+1}{n}.. \end{split} \end{equation} \par The approximations are defined integrating over lines issuing from the base of the cylinder: \begin{equation} \tau_n(z,v,t)=\int_0^t\chi_{{\mathcal{T}}_n^c}(z,v,s)\,ds; \end{equation} given points $(z_1,v_1,t_1),(z_2,v_2,t_2)\in\Cyl(Y,M)$, we say that they are separated by a hypersurface $\partial_{\pm}\strip f_j,2\lambda_j\frac{\delta+\cot\alpha+1}{n}.$ if \begin{equation} \left(f_j(y_1)+\underbrace{2\lambda_j\frac{\delta+\cot\alpha+1}{n}}_{\text{omitted for $-$}}-t_1\right)\left(f_j(y_2) \underbrace{+2\lambda_j\frac{\delta+\cot\alpha+1}{n}}_{\text{omitted for $-$}}-t_2\right)\le0. \end{equation} If the two points are not separated by a hypersurface, there are 4 possibilities: \begin{enumerate} \item They lie in a strip $\strip f_j,2\lambda_j\frac{\delta+\cot\alpha+1}{n}.$. \item They lie between two strips $\strip f_j,2\lambda_j\frac{\delta+\cot\alpha+1}{n}.$, $\strip f_{j+1},2\lambda_{j+1}\frac{\delta+\cot\alpha+1}{n}.$ in the sense that \begin{equation} t_i-f_j(z_i,v_i)+2\lambda_j\frac{\delta+\cot\alpha+1}{n}>0\;\text{and}\;t_i-f_{j+1}(z_i,v_i)<0 \;\text{for $i=1,2$}. \end{equation} \item They both lie below the first strip: \begin{equation} t_i-f_1(z_i,v_i)<0\quad\text{for $i=1,2$}. \end{equation} \item They both lie above the last strip: \begin{equation} t_i-f_{M(n)}(z_i,v_i)-2\lambda_{M(n)}\frac{\delta+\cot\alpha+1}{n}>0\quad\text{for $i=1,2$}. \end{equation} \end{enumerate} We define the constants $\eta_j$: \begin{equation} \eta_j=\sum_{i=1}^{j-1}2\lambda_i\frac{\delta+\cot\alpha+1}{n}. \end{equation} In case (1) we have \begin{equation} \tau_n(z_i,v_i,t_i)=f_j(y_i)-\eta_{j-1} \end{equation} which implies \begin{equation} \left|\tau_n(z_1,v_1,t_1)-\tau_n(z_2,v_2,t_2)\right|\le d_{\delta,\alpha}\left((z_1,v_1),(z_2,v_2)\right). \end{equation} This also implies that the function $\tau_n$ is $\mu$-a.e.~locally $1$-Lipschitz on $S$ with respect to the distance $d_{\delta,\alpha}$. In case (2) we have \begin{equation} \tau_n(z_i,v_i,t_i)=t_i-\eta_j \end{equation} which implies \begin{equation} \left|\tau_n(z_1,v_1,t_1)-\tau_n(z_2,v_2,t_2)\right|\le|t_1-t_2|. \end{equation} Cases (3) and (4) are treated like case (2). \par Suppose now that the points $(z_1,v_1,t_1),(z_2,v_2,t_2)\in\Cyl(Y,M)$ are separated by an hypersurface. We can choose the hypersurface minimal with respect to the order $\prec$ and there are two possibilities: \begin{enumerate} \item They are separated by a hypersurface $\partial_-\strip f_j,2\lambda_j\frac{\delta+\cot\alpha+1}{n}.$. \item They are separated by a hypersurface $\partial_+\strip f_j,2\lambda_j\frac{\delta+\cot\alpha+1}{n}.$. \end{enumerate} We consider just case (1) and assume that $t_1\le f_j(z_1,v_1)$ and $t_2\ge f_j(z_2,v_2)$. The assumption of minimality on the separating strip implies that \begin{align} \tau_n(z_1,v_1,t_1)&=t_1-\eta_{j-1};\\ \tau_n(z_2,v_2,t_2)&\in[f_j(z_2,v_2)-\eta_{j-1},t_2-\eta_{j-1}]; \end{align} in particular, \begin{equation} \tau_n(z_2,v_2,t_2)\le t_2-\eta_{j-1}=t_2-t_1+t_1-\eta_{j-1}\le|t_2-t_1|+\tau_n(z_1,v_1,t_1); \end{equation} moreover, \begin{equation} \begin{split} \tau_n(z_1,v_1,t_1)=t_1-\eta_{j-1}&\le f_j(z_1,v_1)-\eta_{j-1}\\&=f_j(z_2,v_2)-\eta_{j-1}+f_j(z_1,v_1)-f_j(z_2,v_2) \\&\le\tau_n(z_2,v_2,t_2)+d_{\delta,\alpha}(z_1,z_2). \end{split} \end{equation} This shows that the function $\tau_n$ is $1$-Lipschitz with respect to the distance $D$. \par We finally observe that \begin{equation} \|\tau-\tau_n\|_\infty\le\eta_{M(n)}\le 3(1+\delta+\cot\alpha)\frac{M(n)}{n}=O(1/n); \end{equation} choosing a sequence $n_k\to\infty$ and letting $g_k=\tau_{n_k}$ we conclude that $g_k\xrightarrow{\text{w*}}\tau$. \end{proof} \subsection{Dimensional bounds and tangent cones}\label{subsec:dimens-bounds-tang} The goal of this Subsection is to prove Theorem \ref{alberti_blow_up} and Corollary \ref{derbound}. We first recall the definition of blow-ups of Lipschitz functions on metric spaces, following the approach of isometrically embedding all the pointed metric spaces in a common proper metric space \cite[Sec.~2.2]{keith-modulus}. \par \begin{defn}\label{defn_blow_up} A \textbf{blow-up of a metric space $X$ at a point $p$} is a complete pointed metric space $(Y,q)$ such that there is a sequence $(t_n>0)_n$ with $t_n\to0$ and $\left(\frac{1}{t_n}X,p\right)\to(Y,q)$ in the Gromov-Hausdorff sense. The class of blow-ups at $p$ is denoted by $\tang(X,p)$. \end{defn} \par To show the existence of blow-ups the following notion of finite dimensionality for metric spaces is useful. \begin{defn}\label{defn:assouad_dim} A metric space $X$ is \textbf{doubling} if there is a constant $C$ such that every set of diametre $\le N$ can be covered by at most $C$ sets of diametre $\le N/2$. By induction it follows that $X$ admits a covering function $C(\epsi)=C\epsi^{-D}$ where any set of diametre $\le N$ can be covered by at most $C(\epsi)$ sets of diametre $\le\epsi N$. The minimal exponent $D$ is called the \textbf{Assouad dimension} of $X$. \end{defn} \begin{rem}\label{rem:selection} If $X$ is doubling, given a sequence $(t_n>0)_n$ with $t_n\to0$, it is possible to find a subsequence $(t_{n_k})$ such that the sequence $\left(\frac{1}{t_{n_k}}X,p\right)$ converges in the Gromov-Hausdorff sense. In this case the spaces $\left(\frac{1}{t_{n_k}}X,p\right)$ and $(Y,q)$ can be isometrically embedded into a proper metric space $(Z,z)$\footnote{Mapping basepoints to basepoints} so that, for each $R>0$, \begin{align} \lim_{k\to\infty} \sup_{y\in \ball z,R.\cap Y}\setdist \frac{1}{t_{n_k}}X,\{y\}.&=0\\ \lim_{k\to\infty}\sup_{x\in \ball z,R.\cap \frac{1}{t_{n_k}}X}\setdist Y,\{x\}.&=0. \end{align} \par In particular, any point $q'\in Y$ can be \textbf{approximated} by a sequence $p'_{n_k}\in\frac{1}{t_{n_k}}X$ such that $p'_{n_k}\to q'$ in $Z$. \end{rem} \par We now define blow-ups of Lipschitz functions. \begin{defn}\label{defn:blow_up_lip} A \textbf{blow-up of a Lipschitz function $f:X\to\real^Q$ at a point $p$} is a triple $(Y,q,g)$ where: \begin{enumerate} \item The space $(Y,q)\in\tang(X,p)$ and is a limit realized by a sequence $(t_n)_n$ of scaling factors. \item The function $g:Y\to\real^Q$ is Lipschitz with $g(q)=0$. \item If $p'_n$ approximates $q'$, \begin{equation} \lim_{n\to\infty}\frac{f(p'_n)-f(p)}{t_n}=g(q'). \end{equation} \end{enumerate} The class of all blow-ups of $f$ at $p$ will be denoted by $\tang(X,p,f)$. By an Ascoli-Arzel\`a argument, if $X$ is doubling, $\tang(X,p,f)\ne\emptyset$. \end{defn} We now prove Theorem \ref{alberti_blow_up}: the assumption on completenss of $\mu$ is required to ensure that Suslin sets are $\mu$-measurable. Note that any Radon measure can be extended so that sets in the $\sigma$-algebra generated by Suslin sets are measurable. \begin{thm}\label{alberti_blow_up} Let $\mu$ be a complete Radon measure on a metric space $X$ which has finite Assouad dimension $D$. Consider a Lipschitz function $f:X\to \real^N$, points $\{v_i\}_{i=1}^N\subset{\mathbb S}^{N-1}$, and constants $\{\alpha_i\}_{i=1}^N\subset(0,\pi/2)$ and $\delta>0$. Suppose that $\mu$ admits Lipschitz Alberti representations $\{\albrep i.\}_{i=1}^N$ such that: \begin{enumerate} \item The Alberti representation $\albrep i.$ is in the $f$-direction of $\cone(v_i,\alpha_i)$ with $\langle v_i, f\rangle$-speed $\ge\delta$. \item For some $\theta>0$ the cone fields $\cone(v_i,\alpha_i+\theta)$ are independent. \end{enumerate} Then there is a $\mu$-full measure Borel subset $U\subset X$ such that for each $p\in U$ and for each blow up $(Y,q,g)\in\tang (X,p, f)$, the function $g:Y\to\real^N$ is surjective. \end{thm} \begin{proof} We assume that the Alberti representations $\{\albrep i.\}_{i=1}^N$ are $C$-Lipschitz. Given a $C$-Lipschitz $\albrep.$ in the $f$-direction of a cone field $\cone(v,\alpha)$ with $\langle v,f\rangle$-speed $\ge\delta$, we define the set $\frags(p,R,\epsi,\albrep.)$ of fragments $\gamma$ that satisfy the following conditions at $p$: \begin{enumerate} \item The fragment $\gamma$ is a $C$-Lipschitz path fragment in the $f$-direction of $\cone(v,\alpha)$ with $f$-speed $\ge\delta$, and such that $\lebmeas(\dom\gamma)>0$ and $0$ is a density point of $\dom \gamma$. \item The metric differential $\metdiff\gamma(0)$ and the derivative $(f\circ\gamma)'(0)\in\cone(v,\alpha)$ exist, $\gamma(0)=p$ and \begin{equation} \left|(f\circ\gamma)'(0)\right|\ge\delta\metdiff\gamma(0). \end{equation} \item For each $ r\in(0,R)$ one has $\lebmeas(\dom\gamma\cap\ball 0,r.)\ge 2r(1-\epsi)$. \item For each $ t\in\dom\gamma\cap\ball 0,r.$ one has \begin{equation} \left|f(\gamma(t))-f(p)-(f\circ\gamma)'(0)t\right|\le\epsi |t|. \end{equation} \item For all $ t,s\in\dom\gamma\cap\ball 0,r.$ one has\footnote{This is the approximate continuity of the metric differential (\cite[Thm.~3.3]{ambrosio-rectifiability})} \begin{equation} \left|\dist\gamma(t),\gamma(s).-\metdiff\gamma(0)|t-s|\right|\le \epsi(|t|+|s|). \end{equation} \end{enumerate} Let $F(R_1,\epsi_1,\albrep 1.)=\{p\in U:\frags(p,R_1,\epsi_1,\albrep 1.) \ne\emptyset\}$; this set is Suslin and, as $\mu$ admits the Alberti representation $\albrep 1.$, as $R_1\searrow0$ the sets \{$F(R_1,\epsi_1,\albrep 1.)\}$ increase to a full measure subset of $U$. By the Jankoff measurable selection principle \cite[Thm.~6.9.1]{bogachev_measure} there is a selection function $\Gamma(R_1,\epsi_1,\albrep 1.)$ associating to $p\in F(R_1,\epsi_1,\albrep 1.)$ a fragment satisfying the conditions (1)--(5) above. This choice is measurable in the sense that \begin{align} \varphi(R_1,\epsi_1,\albrep 1.)(p)&=(f\circ \Gamma(R_1,\epsi_1,\albrep 1.))'(0)\\ \psi(R_1,\epsi_1,\albrep 1.)(p)&=\metdiff \Gamma(R_1,\epsi_1,\albrep 1.)(0) \end{align} are measurable functions\footnote{With respect to the $\sigma$-algebra generated by Suslin sets}. By Lusin's Theorem \cite[Thm.~7.1.13]{bogachev_measure} there are compact sets \begin{equation} F_c(R_1,\epsi_1,\albrep 1.;\tau_1)\subset F(R_1,\epsi_1,\albrep 1.) \end{equation} suct that \begin{enumerate} \item for all $p,q\in F_c(R_1,\epsi_1,\albrep 1.;\tau_1)$ with $\dist p,q.\le C\tau_1$ one has \begin{align} \left|\varphi(R_1,\epsi_1,\albrep 1.)(p)-\varphi(R_1,\epsi_1,\albrep 1.)(q)\right|&<\epsi_1;\\ \left|\psi(R_1,\epsi_1,\albrep 1.)(p)-\psi(R_1,\epsi_1,\albrep 1.)(q)\right|&<\epsi_1; \end{align} \item as $\tau_1\to 0$ \begin{equation} \mu(F(R_1,\epsi_1,\albrep 1.)\setminus F_c(R_1,\epsi_1,\albrep 1.;\tau_1))\to0. \end{equation} \end{enumerate} The construction proceeds inductively in the following way: assuming for $k<N$ that \begin{equation}F(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.), \end{equation} \begin{equation} \Gamma(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.), \end{equation} and \begin{equation} F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k) \end{equation} have been constructed, let \begin{equation} F(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.) \end{equation} be the set of those \begin{equation} p\in F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k)\cap F(R_{k+1},\epsi_{k+1},\albrep k+1.) \end{equation} such that $\exists\gamma\in\frags(p,R_{k+1},\epsi_{k+1},\albrep k+1.)$: \begin{multline}\label{rec_dens} \forall r\le R_{k+1}\quad \lebmeas(\gamma^{-1}(F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k))\cap\ball 0,r.)\\\ge 2r(1-\epsi_{k+1}). \end{multline} These sets are Suslin measurable and as $R_{k+1}\to0$, increase to a full measure subset of $F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k)$. By Jankoff measurable selection principle there is a selection function $\Gamma(R_{k+1},\epsi_{k+1},\albrep k+1.)$ associating to \begin{equation} p\in F(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.) \end{equation} a fragment satisfying the conditions above. This choice is measurable in the sense that \begin{align} \varphi(R_{k+1},\epsi_{k+1},\albrep k+1.)(p)&=(f\circ \Gamma(R_{k+1},\epsi_{k+1},\albrep k+1.))'(0)\\ \psi(R_{k+1},\epsi_{k+1},\albrep k+1.)(p)&=\metdiff \Gamma(R_{k+1},\epsi_{k+1},\albrep k+1.)(0) \end{align} are measurable functions. By Lusin's theorem there are compact sets \begin{multline} F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.;\tau_{k+1})\\\subset F(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.): \end{multline} $\forall\dist p,q.\le C\tau_{k+1}$, \begin{align} \left|\varphi(R_{k+1},\epsi_{k+1},\albrep {k+1}.)(p)-\varphi(R_{k+1},\epsi_{k+1},\albrep {k+1}.)(q)\right|&<\epsi_{k+1};\\ \left|\psi(R_{k+1},\epsi_{k+1},\albrep {k+1}.)(p)-\psi(R_{k+1},\epsi_{k+1},\albrep {k+1}.)(q)\right|&<\epsi_{k+1}; \end{align} as $\tau_{k+1}\to0$ we can assume that \begin{equation} F_c(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.;\tau_{k+1}) \end{equation} increases to a full measure subset of \begin{equation} F(R_1,\epsi_1,\albrep 1.;\tau_1;\cdots;R_k,\epsi_k,\albrep k.;\tau_k;R_{k+1},\epsi_{k+1},\albrep k+1.). \end{equation} Having fixed $c>0$, it is possible to choose $(R^{(i)}_h), (\tau^{(i)}_h)$ such that: \begin{itemize} \item The sequences $(R^{(i)}_h), (\tau^{(i)}_h)$ are decreasing in both $h\in\natural$ and $i\in\{1,\allowbreak\cdots,\allowbreak N\}$. \item One has that $\lim_{h\to\infty}R^{(i)}_h=\lim_{h\to\infty}\tau^{(i)}_h=0$. \item Letting \begin{equation} F_h=F_c(R^{(1)}_h,\frac{1}{h},\albrep 1.;\tau^{(1)}_h;\cdots; R^{(N)}_h,\frac{N}{h},\albrep N.;\tau^{(N)}_h), \end{equation} the Borel set $V=\bigcap_hF_h$ satisfies $\mu\left(X\setminus V\right)<c$. \end{itemize} \par We will simplify the notation for selectors and derivatives writing $\Gamma^{(i)}_h$, $\varphi^{(i)}_h$, $\psi^{(i)}_h$. Let $p\in V$ and \begin{equation} \left(\frac{1}{t_h}X,p,\frac{f-f(p)}{t_h}\right)\to(Y,q,g); \end{equation} consider a subsequence $t_h$ such that $\lim_{h\to\infty}\frac{\tau^{(N)}_h}{t_h}=\infty$. By passing to further subsequences we can assume that: \begin{itemize} \item The $\varphi^{(i)}_h(p)$ converge to $w_i\in\cone(v_i,\alpha_i+\theta)\setminus\ball 0,\delta.$, implying that the $\{w_i\}_{i=1}^N$ are independent. \item The $\psi^{(i)}_h(p)$ converge to $\delta_i\in [\delta, C]$. \end{itemize} In particular, the functions \begin{equation} \Gamma^{(N)}_h(p):\dom \Gamma^{(N)}_h(p)\to X \end{equation} are $C$-Lipschitz and we define \begin{equation} \tilde\Gamma^{(N)}_h(p):\frac {1}{t_h}\dom\Gamma^{(N)}_h(p)\to\frac{1}{t_h} X \end{equation} by \begin{equation} \tilde\Gamma^{(N)}_h(p)(s)=\Gamma^{(N)}_h(p)(t_h\cdot s). \end{equation} From property (3) we obtain \begin{equation} \lebmeas(\dom\tilde\Gamma^{(N)}_h \cap\ball 0,\frac{\tau^{(N)}_h}{t_h}.)\ge 2\frac{\tau^{(N)}_h}{t_h}(1-\frac{1}{h}). \end{equation} The functions $\tilde\Gamma^{(N)}_h(p)$ are still $C$-Lipschitz with respect to the rescaled metrics and $\tilde\Gamma^{(N)}_h(p)(0)=p$ and by a variant of Ascoli-Arzel\`a we get a $C$-Lipschitz \begin{equation} \tilde\Gamma^{(N)}_\infty:\real\to Y \end{equation} with $\tilde\Gamma^{(N)}_\infty(0)=p$. For $s\in\real$ we choose $s_h\in\dom \tilde\Gamma^{(N)}_h(p)$ converging to $s$ and observe that by (4) \begin{equation}\label{der_conv} \left|\frac{f\circ\tilde\Gamma^{(N)}_h(p)(s_h)-f(p)}{t_h}-\varphi^{(N)}_h(p)\cdot s_h\right|\le\frac{|s_h|}{h} \end{equation} which implies \begin{equation} g\circ\tilde\Gamma^{(N)}_\infty(s)=w_N\cdot s. \end{equation} A similar argument involving the metric derivative and $\psi^{(N)}_h(p)$ shows that \begin{equation} \dist\tilde\Gamma^{(N)}_\infty(s),\tilde\Gamma^{(N)}_\infty(s').=\delta_N|s-s'|. \end{equation} If $s\in\real$, because of \eqref{rec_dens}, there are $s_h\in \dom\tilde\Gamma^{(N)}_h$ converging to $s$ such that $\Gamma^{(N-1)}_h\left(\Gamma^{(N)}_h(p)(s_h)\right)$ is defined. Let \begin{equation} \tilde\Gamma^{(N-1)}_h:\frac {1}{t_h}\dom \Gamma^{(N-1)}_h\left(\Gamma^{(N)}_h(p)(s_h)\right) \to\frac{1}{t_h} X \end{equation} be defined by \begin{equation} \tilde\Gamma^{(N-1)}_h(\sigma)=\Gamma^{(N-1)}_h\left(\Gamma^{(N)}_h(p)(s_h)\right)(t_h\cdot \sigma). \end{equation} These functions are $C$-Lipschitz with respect to the rescaled metrics and a variant of Ascoli-Arzel\`a yields a $C$-Lipschitz \begin{equation} \tilde\Gamma^{(N-1)}_\infty:\real\to Y \end{equation} with $\tilde\Gamma^{(N-1)}_\infty(0)=\tilde\Gamma^{(N)}_\infty(s)$. There is an analogue of \eqref{der_conv} where $\varphi^{(N)}_h(p)$ is replaced by $\varphi^{(N-1)}_h\left(\Gamma^{(N)}_h(p)(s_h)\right)$ but for $h$ sufficiently large \begin{equation} \left|\varphi^{(N-1)}_h\left(\Gamma^{(N)}_h(p)(s_h)\right)- \varphi^{(N-1)}_h(p)\right|<\frac{1}{h} \end{equation} so \begin{equation} g\circ \tilde\Gamma^{(N-1)}_\infty(\sigma)-g\circ \tilde\Gamma^{(N-1)}_\infty(0)=w_{N-1}\cdot\sigma. \end{equation} A similar argument involving the metric derivative shows that \begin{equation} \dist \tilde\Gamma^{(N-1)}_\infty(s),\tilde\Gamma^{(N-1)}_\infty(s').=\delta_{N-1}\left|s-s'\right|. \end{equation} Continuing inductively we conclude that $\exists\tilde\Gamma:\real^N\to Y$: $\tilde\Gamma(0)=q$ and $\forall(0\le k\le N-1)\forall s\in\real^k$ $\tilde\gamma(t)=\tilde\Gamma((0,t,s))$ is a $\delta_{N-k}$-constant speed geodesic with \begin{equation} g\circ\gamma(t)-g\circ\gamma(0)=w_{N-k}\cdot t. \end{equation} Being the $w_i$ independent, $g$ is surjective. \end{proof} \par We need a Lemma relating the Assouad dimension of a space to that of a blow-up \cite[Prop.~6.1.5]{tyson_mackay_conf}: \begin{lem}\label{assouad_tang} If $X$ has Assouad dimension $\le D$ and if $(Y,q)\in\tang(X,p)$, then $Y$ has Assouad dimension $\le D$. \end{lem} \par The following Lemma provides a lower bound on the Assouad dimension. \begin{lem}\label{assouad_surj} If $Y$ has Assouad dimension $\le D$ (or Hausdorff dimension $\le D$) and if there is a surjective Lipschitz map $g:Y\to\real^N$, then $D\ge N$. \end{lem} \begin{proof} The argument in \cite[Subsec.~8.7]{heinonen_analysis} shows that the Assouad dimension of $Y$ is at least its Hausdorff dimension. Now Lipschitz maps do not increase the Hausdorff dimension so the Hausdorff dimension of $Y$ is at least the Hausdorff dimension of $g(Y)=\real^N$. \end{proof} \par We can now prove Corollary \ref{derbound}. \begin{cor}\label{derbound} If $X$ is either a metric space with Assouad dimension $D$ and if $\mu$ is a Radon measure, then $\wder\mu.$ has index locally bounded by $D$. \end{cor} \begin{proof} We can reduce to the hypothesis of Theorem \ref{alberti_blow_up} because of Corollary \ref{der-alb}; we then apply Lemmas \ref{assouad_tang}, \ref{assouad_surj}. \end{proof} \begin{rem} Note that using ultralimits one can replace in Corollary \ref{derbound} the assumption on the Assouad dimension with a uniform upper bound $D$ on the Hausdorff dimension of the blow-ups of $X$. \end{rem} \subsection{Construction of independent Lipschitz functions}\label{subsec:constr-indep-lipsch} The goal of this Subsection is to prove Theorem \ref{lip_ind}. We will use the following truncation Lemma (\cite[Lem.~4.1]{bate-diff}). \begin{lem}\label{lip_trunc} Let $\epsi\in(0,h/4)$ and assume that $S\subset X$ is Borel and $f:X\to\real$ is $L$-Lipschitz. Then there is an $L$-Lipschitz function $g$, constructed from $f$, such that: \begin{enumerate} \item The function $g$ satisfies $0\le g\le h$. \item The function $g$ is supported in $\ball S,\frac{2h}{L}.$. \item If $x,y\in\ball S,\frac{h}{L}.$, \begin{equation} |f(x)-f(y)|\ge|g(x)-g(y)|. \end{equation} \item There is a Borel subset $S'\subset S$ with \begin{equation} \mu(S')\ge\left(1-\frac{4\epsi}{h}\right)\mu(S) \end{equation} such that if $x\in S$ and $\dist x,y.\le\epsi/L$, then \begin{equation} |f(x)-f(y)|=|g(x)-g(y)|. \end{equation} \end{enumerate} \end{lem} \begin{proof}[Proof of Theorem \ref{lip_ind}] Choose $m_1$ such that $\frac{1}{m_1}<\frac{\alpha^2}{2^5L}$ and let $h_1=\frac{\alpha^2}{4}$ and $\epsi_1=\frac{L}{m_1}$. We use Lemma \ref{lip_trunc} to find $g_1$ and $S_1\subset S$ such that the following holds: \begin{enumerate} \item We have the inequalities $\mu(S_1)\ge\left(1-\frac{4\epsi_1}{L}\right)\mu(S)\ge\frac{1}{2}\mu(S)$. \item The function $g_1$ is supported in $\ball S,\frac{2h_1}{L}.=\ball S,\frac{\alpha^2}{ 2L}.$ and for each $x,y\in\ball S,\frac{h_1}{L}.$ \begin{equation} |f_{m_1}(x)-f_{m_1}(y)|\ge|g_1(x)-g_1(y)|. \end{equation} \item If $B$ is a ball of radius $\rho_{m_1}$ centred at some point of $S$, the Lipschitz constant of $g_1|B$ is at most $\frac{1}{m_1}$. \item As $\frac{\epsi_1}{L}=\frac{1}{m_1}$, for each $x\in S_1$ there is $x_1\in \ball x,\frac{1}{m_1}.$ with \begin{equation} \left|g_1(x)-g_1(x_1)\right|\ge\delta_0\dist x,x_1.>0. \end{equation} \end{enumerate} We define $g_{k+1}$ inductively in the following way. Having chosen $m_{k+1}$ such that \begin{equation} \frac{1}{m_{k+1}}<\frac{\alpha^2}{2^{(k+1)+4}L}\rho_{m_k}, \end{equation} we let $h_{k+1}=\frac{\alpha^2}{4}\rho_{m_k}$ and $\epsi_{k+1}=\frac{L}{m_{k+1} }$. We use Lemma \ref{lip_trunc} to find $g_{k+1}$ and $S_{k+1}\subset S$ such that the following holds: \begin{enumerate} \item We have the inequalities $\mu(S_{k+1})\ge\left(1-\frac{4\epsi_{k+1}}{L}\right)\mu(S)\ge \left(1-\frac{\rho_{m_k}}{2^{k+1}}\right)\mu(S)$. \item The function $g_{k+1}$ is supported in $\ball S,\frac{2h_{k+1}}{ L}.=\ball S,\frac{\alpha^2\rho_{m_k}}{2L}.$ and for each $x,y\in \ball S,\frac{h_{k+1}}{L}.$, \begin{equation} |f_{m_{k+1}}(x)-f_{m_{k+1}}(y)|\ge|g_{k+1}(x)-g_{k+1}(y)|. \end{equation} \item If $B$ is a ball of radius $\rho_{m_{k+1}}$ centred at some point of $S$, the Lipschitz constant of $g_{k+1}|B$ is at most $\frac{1}{m_{k+1}}$. \item As $\frac{\epsi_{k+1}}{L}=\frac{1}{m_{k+1}}$, for each $x\in S_{k+1}$ there is $x_{k+1}\in \ball x,\frac{1}{m_{k+1}}.$ with \begin{equation} \left|g_{k+1}(x)-g_{k+1}(x_{k+1})\right|\ge\delta_0\dist x,x_{k+1}.>0. \end{equation} \end{enumerate} Note that from the definition of $m_k$ one can verify by induction that \begin{equation} \frac{1}{m_k}\le\left(\frac{\alpha^2}{L}\right)^k 2^{-\left(\frac{k(k+1)} {2}+4k\right)}; \end{equation} note also that \begin{align} h_{s+1}&=\frac{\alpha^2}{4}\rho_{m_s}\le\frac{\alpha^2}{4}\frac{1}{m_s},\\ h_{s+k}&=\frac{\alpha^2}{4}\left(\frac{\alpha^2}{L}\right)^{k-1} 2^{-(s+5)}\rho_{m_s}\text{ (here $k>1$)}; \end{align} so that \begin{equation} \sum_{k\ge s+1}h_k\le\frac{\alpha^2}{4}\left( 1+2^{-(s+5)}\frac{\alpha^2/L}{1-\alpha^2/L}\right)\rho_{m_s}. \end{equation} This implies that if $n_k\nearrow\infty$, then $\varphi=\sum_k g_{n_k}<\infty$ defines a continuous function. We show that $\varphi$ is Lipschitz with Lipschitz constant \begin{equation} 3\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right). \end{equation} We want to bound $|\varphi(x)-\varphi(y)|$. In the first case, we assume $x\in S$. If $y\ne x$ assume there is some $s$ such that $\dist x,y.\in(\frac{\alpha}{2}\rho_{m_s},\rho_{m_s}]$. For $t\le s$ $y\in\ball x,\rho_{m_t}.$ so \begin{equation} \left|g_{t}(x)-g_{t}(x_{t})\right|\le\frac{1}{m_t}\dist x,y.; \end{equation} in particular, from the estimate for $\frac{1}{m_k}$ we deduce that \begin{equation} \sum_{k}\frac{1}{m_k}\le\frac{\alpha^2/L}{1-\alpha^2/L}, \end{equation} so that \begin{equation} \sum_{t\le s}\left|g_{t}(x)-g_{t}(x_{t})\right|\le \frac{\alpha^2/L}{1-\alpha^2/L}\dist x,y. . \end{equation} For the second estimate we use the bound on $h_{t}$: \begin{equation}\label{tail_bound} \begin{split} \sum_{t\ge s+1} \left|g_{t}(x)-g_{t}(x_{t})\right|&\le 2\sum_{t\ge s+1}h_t\\& \le\frac{\alpha^2}{2}\left( 1+2^{-(s+5)}\frac{\alpha^2/L}{1-\alpha^2/L}\right)\rho_{m_s}\\ &\le\alpha\left( 1+2^{-(s+5)}\frac{\alpha^2/L}{1-\alpha^2/L}\right)\dist x,y.. \end{split} \end{equation} We therefore get the bound \begin{equation} \left|\varphi(x)-\varphi(y)\right|\le\left(\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right)\dist x,y.. \end{equation} The other possibility is that $\dist x,y.\in(\rho_{m_{s+1}},\frac{\alpha}{2} \rho_{m_s}]$ for some $s$ or $\dist x,y.>\rho_{m_1}$. In this case we can estimate as above except that for $g_{s+1}$ we must use the full Lipschitz constant $L$. We therefore have: \begin{equation} \left|\varphi(x)-\varphi(y)\right|\le\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right)\dist x,y.. \end{equation} In the second case $x,y\not\in S$. We use that the functions $g_s$ are supported on a neighbourhood of $S$. Without loss of generality we can assume $\dist x,S.\le\dist y,S.$. Now, for some $s$ we have $\dist x,S.\in[\frac{\alpha^2}{2L}\rho_{m_{s+1}},\frac{\alpha^2}{2L}\rho_{ m_s})$ (for $s=0$ we let $\rho_{m_0}=1$) or $\dist x,S.\ge\frac{\alpha^2}{2L}$. So for $t\ge s+2$ we have that $g_t(x)=g_t(y)=0$ and in the last case $\varphi(x)=\varphi(y)=0$ so there is nothing to prove. We assume that $\dist x,y.<\frac{\alpha^2}{2L}\rho_{m_s}$. Then the hypothesis on $\alpha$ implies that $\alpha^2/L<1$ so that $x,y$ belong to a ball centred on $S$ of radius at most $\rho_{m_s}$. For $t\le s$ the Lipschitz constant of $g_t$ is then at most $\frac{1}{m_t}$ implying \begin{equation} \sum_{t\le s}\left|g_t(x)-g_t(y)\right|\le\frac{\alpha^2/L}{1-\alpha^2/L} \dist x,y.. \end{equation} We therefore have: \begin{equation} \left|\varphi(x)-\varphi(y)\right|\le\left(L+\frac{\alpha^2/L}{1-\alpha^2/L}\right)\dist x,y.. \end{equation} We assume $\dist x,y.\ge\frac{\alpha^2}{2L}\rho_{m_s}$. We can find $\tilde x\in S$ such that $\dist x,\tilde x.\le\frac{\alpha^2}{2L}\rho_{m_s}$ so that \begin{align} \left|\varphi(x)-\varphi(\tilde x)\right|&\le\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right)\dist x,\tilde x.\\ \left|\varphi(y)-\varphi(\tilde x)\right|&\le\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right)\dist y,\tilde x.. \end{align} Note now that $\dist \tilde x,y.\le2\dist x,y.$ and $\dist \tilde x,x. \le\dist x,y.$. Therefore: \begin{equation} \left|\varphi(x)-\varphi(y)\right|\le 3\left(L+\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right)\dist x,y.. \end{equation} We now pass to the construction of $M$ independent functions. We let \begin{equation} \psi_j=\sum_{\text{$k\equiv j\bmod{M}$}}g_k \end{equation} and \begin{equation} \tilde S_j=\bigcap_{k}\bigcup_{\text{$s\ge k$ and $s\equiv j\bmod{M}$}} S_s \end{equation} which is a full measure Borel subset of $S$. In particular $S'=\bigcap_j \tilde S_j$ is a full measure Borel subset of $S$. Let $\tilde x\in S'$ and assume that $|\lambda_0|\ge\max_{0\le i\le M-1}|\lambda_i|$. For each $k$ there is an $n_k\ge k$ with $x\in S_{n_k}$ and $n_k\equiv0 \bmod{M}$. We can then find a point $x_{n_k}$ such that \begin{equation} \left| g_{n_k}(x)-g_{n_k}(x_{n_k})\right|\ge\delta_0\dist x,x_{n_k}.>0. \end{equation} Then \begin{equation} \left|\sum_{j=0}^{M-1}\lambda_j\psi_j(x)-\lambda_j\psi_j(x_{n_k}) \right|\ge|\lambda_0|\left(\delta_0\dist x,x_{n_k}.-\sum_{t\ne n_k} \left|g_t(x)-g_t(x_{n_k})\right|\right). \end{equation} The terms for $t>n_k$ can be bound using \eqref{tail_bound} to get: \begin{equation} \sum_{t> n_k} \left|g_t(x)-g_t(x_{n_k})\right| \le\alpha\left( 1+2^{-(n_k+5)}\frac{\alpha^2/L}{1-\alpha^2/L}\right)\dist x,x_{n_k}.. \end{equation} Note that for $t<n_k$, $\dist x,x_k.<\rho_{m_t}$ so we have the following bound on the terms for $t<n_k$: \begin{equation} \sum_{t< n_k} \left|g_t(x)-g_t(x_{n_k})\right|\le\frac{\alpha^2/L}{1-\alpha^2/L} \dist x,x_{n_k}.. \end{equation} Letting $k\nearrow\infty$ we conclude that \begin{equation} \biglip\left(\sum_{i=0}^{M-1}\lambda_i\psi_i\right)(x)\ge \left(\delta_0-\frac{\alpha^2/L}{1-\alpha^2/L}-\alpha \right)|\lambda_0|. \end{equation} \end{proof} \begin{rem} Consider what happens to $\varphi_\alpha$ as $\alpha\searrow0$: we get a full measure Borel subset $S'\subset S$ with \begin{equation} \inf_{x\in S'}\biglip\varphi_\alpha(x)\ge \left(\delta_0-\frac{\alpha^2/L}{1-\alpha^2/L}-\alpha \right) \end{equation} and if $x\in S'$ and $r\in(\frac{\alpha}{2}\rho_{m_s},\rho_{m_s}]$, \begin{equation} \varlip\varphi_\alpha(x,r)\le \left(\alpha+ \frac{\alpha^2/L}{1-\alpha^2/L}(1+2^{-6}\alpha)\right); \end{equation} in particular, on a full measure subset of $S$ the Lip-lip inequality \eqref{eq_Lip_lip_ineq} is violated. \end{rem} \bibliographystyle{alpha}
\section{Introduction} As users migrate information to cloud storage, the burden of reliability moves to the cloud provider. Thus many cloud vendors such as Amazon \cite{dynamo} and Azure \cite{bradpaper} use multiple loosely consistent replicas of user information because of the high overhead of keeping replicas synchronized at all times. Further, users often retain copies of their information on laptops, tablets, phones and Personal Digital Assistants (PDAs); these devices are often disconnected from cloud storage and thus can diverge from the corresponding copies in the cloud. The situation naturally grows even more complicated when multiple users have access to information, because the number of replicas can increase with the number of users. Periodically, however copies of information objects must be synchronized or {\em reconciled}. One can also view the need for reconciliation at a higher level, such as for loosely consistent replicas of large databases that may be used for availability by information providers. This paper focuses on the basic problem of set reconciliation. In the 2-party setting, two parties $A_1$ and $A_2$ respectively have (usually very similar) sets $S_1$ and $S_2$, and want to reconcile so both have $S_1 \cup S_2$. Our major contribution is to extend the recent approach to set reconciliation for two parties using Invertible Bloom Lookup Tables (IBLTs) to the multi-party setting, where there are three or more parties holding sets $S_1,S_2,S_3,\ldots,S_n$, and the goal is for all parties to obtain $\cup_i S_i$. This could of course be done by pairwise reconciliations, but we seek more efficient methods. We first extend the IBLT approach, showing that in the multi-part setting we can reconcile using messages of size $O(|\cup_i S_i - \cap_i S_i|)$. This generalizes results from the two-party setting, where the information theoretic goal has been to send information close to the size of the set difference, rather than sending information proportional to the size of the sets, as generally the set difference is very small compared to the set sizes. Our approach has other advantages, including that one does not need to know the number of parties in advance. Our main approach here, related to network coding, is to think of the information stored in the IBLT as corresponding to vectors over a suitable finite field instead of the binary vectors used in previous work. We further show that our methodology allows using further network coding techniques in conjunction with IBLTs, providing additional efficiency in terms of network utilization. By connecting reconciliation with network coding, we can provide more efficient reconciliation methods that apply to a number of natural distributed computing problems. For example, using recent results from gossip algorithms, we show that multi-party set reconciliation over a network with $n$ nodes can be done in $O(\phi^{-1} \log n)$ rounds of communication with IBLTs, where $\phi$ is the conductance of the network. While our work can be seen as a specific example of a linear sketch that has a natural affinity to the network coding approach, we believe it suggests that other linear sketch-based data structures may also find expanded use by combining them with ideas from network coding. \subsection{Potential Applications and Related Work} The use of IBLTs for distributed synchronization has already been proposed for specific applications. For example, recently the Bitcoin community has considered using IBLTs for scalable synchronization of transactions\footnote{See also \url{http://www.reddit.com/r/Bitcoin/comments/2hchs0/scaling_bitcoin_gavin_begins_work_on_invertible/} for further discussion.} \cite{github,CCN}. Multi-party variations would be potentially useful in the Bitcoin setting, where multiple parties may need to track transactions. In the setting of data centers, as shown in the survey of Bailis and Kingsbury~\cite{Bailis:2014:NR:2639988.2655736}, network errors abound in cost-effective large-scale environments. Thus, even if we attempt to keep multiple copies of data synchronized, the synchronization will periodically fail along with the network, leaving the problem of reconciling the differences. As related work, we note that a different generalization of the set reconciliation problems, to settings where a certain type of \emph{approximate} reconciliation is desired, was recently considered in the database community~\cite{Chen:2014:RSR:2588555.2610528}. However, that work focuses on the setting of two parties, leaving the question of scaling to many parties open. Another related model for problems on distributed data is that of \emph{distributed tracking} (see e.g.~\cite{Huang:2012:RAT:2213556.2213596}). Our problem differs from distributed tracking in two respects: we focus on exact computation (with an arbitrarily small error probability), and we focus on periodic, as opposed to continuous, computation of the joint function. Several further specific applications for set difference structures are given in \cite{setdifference}, including peer-to-peer transactions, deduplication, partition healing, and synchronizing parallel activations (e.g., of independent crawlers of a search engine). They also discuss why logging as an alternative may have disadvantages in multiple contexts; an example they provide is for ``hot'' data items that are written often and may therefore be in the log multiple times. We refer the reader to this paper for more information on these examples. Multi-party variations of IBLT-based synchronization methods could enhance their desirability in these applications when multiple parties naturally arise. In synchronizing parallel activations, for instance, several agents in a distributed system could be gathering information into local databases in a redundant fashion for near-optimal accuracy in the collection process, and then need to reconcile these local databases into a synchronized whole. \subsection{Background} We briefly summarize known results for the historically common setting of two parties with direct communication. Consider two parties $A_1$ and $A_2$ with sets $S_1$ and $S_2$ of keys from a universe $U$. An important value is the size of the set difference between $S_1$ and $S_2$, denoted by $d = |S_1-S_2|+|S_2-S_1| = |(S_1 \cup S_2) - (S_1 \cap S_2)|$. In this setting there are communication-efficient algorithms when $d$, or a good approximate upper bound for $d$, is known. Hence, in some algorithms for set reconciliation, there are two phases: in a first phase a bound on $d$ is obtained, which then drives the second phase of the algorithm, where reconciliation occurs. See \cite{minsky2003set} for further discussion on this point. One previous approach to set reconciliation uses characteristic polynomials, in a manner reminiscent of Reed-Solomon codes \cite{minsky2003set}. Treating keys as values, $A_1$ considers the characteristic polynomial $$\chi_{S_1}(Z) = \prod_{x_i \in S_1} (Z-x_i),$$ and similarly $A_2$ considers $\chi_{S_2}(Z)$. Observe that in the rational function $$\frac{\chi_{S_1}(Z)}{\chi_{S_2}(Z)},$$ common terms cancel out, leaving a rational function in $Z$ where the sums of the degrees of the numerator and denominator is at most $d$. The rational function can be determined through interpolation by evaluating the function at $d+1$ points over a suitably large field; hence, if $A_1$ sends the value of $\chi_{S_1}(Z)$ at $d+1$ points and $A_2$ sends the value of $\chi_{S_2}(Z)$ at the same (pre-chosen) $d+1$ points, then the two parties can reconcile their set difference. If the range can be embedded in a field of size $q$, the total number of bits sent in each direction would be approximately $(d+1) \lceil \log_2 q \rceil$. Note that this takes $O(d^3)$ operations using standard Gaussian elimination techniques. Similar ideas underlie similar results by Juels and Sudan \cite{JS}; these ideas can be extended to use other codes, such as BCH codes, with various computational tradeoffs \cite{DORS}. Recent methods for set reconciliation have centered on using randomized data structures, such as the Invertible Bloom Filter or the related but somewhat more general Invertible Bloom Lookup Table (IBLT) \cite{IBF,setdifference,IBLT}. For our purposes, the Invertible Bloom Lookup Table is a randomized data structure storing a set of keys that supports insertion and deletion operations, as well as a listing operation that lists the keys in the structure. We review the use of IBLTs for 2-party set reconciliation in Section~\ref{sec:review}. The main effect of using IBLTs is that one can give up a small constant factor in the data transmitted to obtain speed and simplicity that is desirable for many implementations. 2-party reconciliation using IBLTs can generally be done in linear time, using primarily hashing and XOR operations. As we show in this paper, the use of IBLTs can also be extended to multi-party reconciliation. While the theory set reconciliation among two parties has been widely studied, there appears to be no substantial prior work (that we are aware of) specifically examining the theory of multi-party reconciliation schemes, although the question of how to implement them was raised in \cite{setreconc}.\footnote{Some preliminary results for multi-party settings, also using the IBLT framework but based on pairwise reconciliations, were provided to us by Goyal and Varghese \cite{GV}. Also, after the appearance of this work on the arxiv, multi-party set reconciliation using characteristic polynomials and repeated pairwise reconciliations was examined in \cite{BoralM}; their conclusion is that is while it is possible, it currently seems much less efficient than the methods considered here.} Special cases, such as rumor spreading (see, e.g., \cite{chierichettialmost,DebMedard,demers,giakkoupis,haeupler,shah}), where one (or more) parties have a single key to share with everyone, have been studied, however. In more practical work, reconciliation among multiple parties has been studied, often in the context of distributed data distribution, using such techniques as erasure coding and Bloom filters to enhance performance or reduce the overall amount of data transferred (e.g., \cite{bcmr,bullet,lcsr,setdifference}). However, these works are also based on pairwise reconciliation, and some do not attempt to achieve data transmission proportional to the size of the set difference, which is our goal here. The work closest to ours is \cite{setdifference}, which also uses Invertible Bloom Filters, but is focused on pairwise reconciliation. \section{Review: The 2-Party Setting} \label{sec:review} We review 2-party set reconciliation, using the framework of the Invertible Bloom Filter/ Invertible Bloom Lookup Table (IBLT) \cite{IBF,setdifference,IBLT}. More concretely, we first describe an IBLT and its use for set reconciliation. IBLTs store keys, which here we will think of as fixed-length bit strings. An IBLT is designed with respect to a threshold number of keys, $t$, so that listing will be successful with high probability if the actual number of keys in the structure at the time of a listing operation is less than or equal to $t$. An IBLT ${\cal B}$ consists of a lookup table $T$ of $m$ cells initialized with all entries set to 0, where $m$ is $O(t)$, and the constant factor in the order notation is generally small (between 1 and 2) depending on the parameters chosen. Like a standard Bloom filter, an IBLT uses a set of $k$ random hash functions, $h_1$, $h_2$, $\ldots$, $h_k$, to determine where keys are stored.\footnote{To obtain structures where $m/t$ is very close to 1, one must use irregular IBLTs, where different keys utilize a different number of hash functions. We simplify our description here and use regular IBLTs, where the same number of hash functions are used for each key. See \cite{IBLT,Rink} for more discussion.} For simplicity, we assume random hash functions here, and for technical reasons we assume that the hashes of each key yield distinct locations; hence the $k$ hashes yield a uniform subset of $k$ distinct cells from the $m \choose k$ possibilities. Alternatively, this could easily be accomplished, for instance, by splitting the table into $k$ subtables and having the $i$th hash function choose a location independently and uniformly at random in the $i$th subtable, which would not asymptotically change the thresholds \cite{BWZ}. The IBLT uses a hash function $H$ that maps keys to hash values in a large range of size $q$ (where $q$ will be chosen later to bound the probability of error). The IBLT uses a hash function $H$ that maps keys to hash values in a large range of size $q$ (where $q$ is a power of 2 that will be chosen later to bound the probability of error). For the purpose of this paper we assume that $H$ is a fully random hash function\footnote{We note that it is possible to show that $H(x) = (a^x \bmod p) \bmod q$, where $1<a<p-1$ is a random positive integer and $p>q^2$, has the needed properties when all keys are smaller than $q$.}. Each key $x$ is placed into cells $T[h_1(x)]$, $T[h_2(x)], \dots, T[h_k(x)]$, where again $T$ is the lookup table that represents the IBLT, as follows. Each cell $T[i]$ contains an ordered pair $$(\textsf{keySum},\textsf{keyhashSum}) \enspace .$$ The \textsf{keySum} field is the XOR of all the keys that have been mapped to this cell, and hence must be the size of the keys (in bits). The \textsf{keyhashSum} field is the XOR of all the hash values $H(x)$ that have been mapped to this cell, and hence must be the size of the hash $H$ of the keys (in bits). Note that insertion and deletion is the same operation, as a deletion operation reverses an insertion. Hence it is possible to delete a key without it first being inserted. \paragraph{Set reconciliation using IBLTs} The above structure yields a set reconciliation algorithm. Consider two parties Angel and Buffy (or $A_1$ and $A_2$). Angel places his keys into an IBLT, as does Buffy. They are assumed to share the hash functions $h_i$ and $H$ according to some prior arrangement. They transfer their corresponding IBLTs, and each then deletes their own keys from the transferred IBLTs. Each IBLT then contains the set difference, and the set difference is recovered using the listing process. Alternatively, since deleting and inserting both correspond to XOR operations, we can say that the parties take the {\em sum} of the IBLTs, by which we simply mean that for each field in each cell, the corresponding values are summed via the bitwise XOR operation. As long as the set difference size $d$ is less than the threshold $t$, recovery will occur with high probability (in $t$). As it will help us subsequently, we describe how the listing process functions. Listing the contents of IBLTs uses a ``peeling process''. Peeling here corresponds to finding a cell with exactly one key contained in it, after the insertion/deletion steps by Angel and Buffy have effectively removed keys that appear in both sets. To find a cell with one key, we check the \textsf{keySum} field using \textsf{keyhashSum}. That is, if the \textsf{keySum} field contains a value $z$, we check whether \textsf{keyhashSum} contains the value $H(z)$. If $z$ is the actual key contained in the cell, then $H(z)$ will indeed appear in the \textsf{keyhashSum} field. If the \textsf{keySum} field contains the XOR of several keys, then (under our assumption of random hash values for $H$) there will be a false positive with probability only $1/q$ where $q$ is the size of the range of hash values for $H$. Let us temporarily assume that there are no false positives. Once we have a cell with a single recoverable key $z$, we can remove $z$ from the structure by computing $h_i(z)$ for all $i$ and deleting $z$ from the corresponding cells using exclusive-or operations to update the \textsf{keySum} and \textsf{keyhashSum} fields. Removing a key from the structure may yield new keys that can be recovered. This peeling process has been used in a variety of contexts, such as in erasure-corrected codes \cite{LMSS}. The peeling process may also fail simply because at some point there may not be any available cell with only a single recoverable key. It can be shown that recovery occurs with high probability, assuming a suitably-sized IBLT is used \cite{IBLT,LMSS,Molloy}. Specifically, the process of peeling corresponds to finding what is known as the 2-core -- the maximal subgraph where all vertices are adjacent to at least two hyperedges -- on a hypergraph where cells correspond to vertices and each key $x$ corresponds to the hyperedge $\{h_1(x),\dots,h_k(x)\}$. When $k$ is a constant, the peeling process yields an empty $k$-core with high probability whenever the table size $m$ satisfies $m > (c_k + \epsilon)t$ for a constant threshold coefficient $c_k$ and constant $\epsilon > 0$. As noted in \cite{IBLT}, the threshold coefficients, given in Table~\ref{tab:thr}, are close to 1. (Again, they can be made closer to 1 if desired using irregular hypergraph constructions \cite{IBLT,Rink}.) \begin{table}[ht] \begin{center} \begin{tabular}{c|ccccc} $k$ & 3 & 4 & 5 & 6 & 7\\\hline $c_k$ & 1.222 & 1.295 & 1.425 & 1.570 & 1.721 \\ \end{tabular} \end{center} \caption{Threshold coefficients for the 2-core rounded to four decimal places.} \label{tab:thr} \end{table} The following theorem, paraphrased from \cite{IBLT}, provides the probabilistic bounds on the failure probability of the peeling process. \begin{theorem} \label{thm:thm1} As long as we choose $m > (c_k + \epsilon) t$ for some $\epsilon > 0$, the listing operation (not counting the separate probability of false positives from \textsf{keyhashSum}) fails with probability $O(t^{-k+2})$. \end{theorem} We now bound the running time for peeling and error probability from false positives from the \textsf{keyhashSum} field. Given an IBLT to peel, we can start by taking a pass over the $O(t)$ cells of the IBLT to find cells where the \textsf{keySum} field contains a value $z$ and the \textsf{keyhashSum} contains the value $H(z)$. We keep a list of such cells and start the peeling with these cells. As we peel a cell we update the \textsf{keySum} and \textsf{keyhashSum} fields of other cells. As we proceed through the list, we might encounter cells that have already been peeled --- these are simply ignored. Also, while peeling we test cells that we delete keys from to see if now the \textsf{keySum} and \textsf{keyhashSum} fields match, in which case the cell can be added to the list of cells to peel. Overall this process clearly takes $O(t)$ time (constant time per peeling operation), and there are $O(t)$ times that we compare \textsf{keySum} and \textsf{keyhashSum} fields within a cell, each of which can yield a false positive with probability $1/q$. Under a worst-case assumption that any such error would cause a listing failure, this gives a total error probability of at most $O(t/q)$ caused by a false positive in the \textsf{keyhashSum} field. We can choose $q$ according to our desired error bound. \paragraph{Adapting to the set difference size} If we have an upper bound on $d$, we can use this upper bound as the value $t$ for the IBLT, and apply Theorem~\ref{thm:thm1}. We henceforth assume that an upper bound within a (small) constant factor of $d$ is available throughout this work, as finding an upper bound is essentially an orthogonal problem. Without an upper bound on $d$, some additional work may need to be done, as explored in for \cite{setdifference,minsky2003set}; we summarize these results and offer other alternatives as they apply here. Both of \cite{setdifference,minsky2003set} suggest approaches that correspond to repeated doubling; if the IBLT size is not sufficient, so that decoding is unsuccessful, then start over with larger IBLTs. Another option is to double the IBLT size by adding one additional lower-order bit to each hash value. Then it suffices to send every odd-numbered cell of the IBLT arrays, since the even-numbered cells can be found by subtraction from the old IBLT. In this way, the total number of cells transmitted is the same as if the final IBLT had been transmitted initially. If the set difference is small but still a constant fraction of the union of the set sizes, then using min-wise independence or related techniques \cite{minwise,cohen} to approximate the set difference may be suitable. Finally, one might incrementally improve the IBLT. If each hash function is assigned its own subarea of cells (as is often how Bloom filters are implemented to allow parallel lookups into the structure \cite{BM}, and as noted in \cite{IBLT} is advantageous for IBLTs), then the parties can incrementally add another hash function by sending additional information corresponding to the subarea for the additional hash function. \paragraph{Keeping count} Finally, the IBLT can also contain an optional \textsf{count} field, which gives a count for the number of keys in a cell. We can increase the count by 1 on insertion, and decrease it by 1 on deletion. With a count field, a cell can contain a recoverable key only if the count is 1 or $-1$. (From Angel's point of view, after deleting his keys from Buffy's IBLT, when the \textsf{count} is $-1$, it could correspond to a cell containing a key of his that Buffy does not have.) Note, however, a \textsf{count} of 1 or $-1$ does not necessarily correspond to a cell with a recoverable key. For example, if Angel has two keys Buffy does not hold that hash to the same cell, and Buffy has one key that Angel does not have that hashes to the same cell, then after taking the difference of the IBLTs the \textsf{count} will be 1 but there will be as sum of three keys in the cell. The \textsf{count} field can be useful for implementation and assists by acting like the sum of another trivial ``hash'' value for the key (all keys hash to 1), but it does not replace the \textsf{keyhashSum} field. \paragraph{Abstraction} At an abstract level, we can view the 2-party setting as follows. We desire a linear sketch (over an appropriate field) taken on sets of keys with the following properties. Let $f(S)$ denote the sketch of the set $S$. We desire \begin{itemize} \item $f((S_1 \cup S_2) - (S_1 \cap S_2)) = f(S_1) \oplus f(S_2)$; \item a set $X$ can be efficiently extracted from $f(X)$ under suitable conditions, which here means that $X$ is sufficiently small; \item the size of the sketch is small, which here means that that if we want to recover $(S_1 \cup S_2) - (S_1 \cap S_2)$ then the sketch size is $O(|(S_1 \cup S_2) - (S_1 \cap S_2)|)$. \end{itemize} We have focused our description on IBLTs as it is a sketch with the required properties. For multi-party reconciliation, we now show that IBLTs can be extended by working over an appropriate field to ensure that with multiple sets $S_1,S_2,\ldots,S_n$, we have $f(\cup_i S_i - \cap_i S_i) = \sum_i f(S)$, while still maintaining suitable extraction and size properties. \section{The 3-Party Setting} We now describe the extension of IBLTs to provide set reconciliation for three parties. Starting with the 3-party setting allows us to demonstrate the key ideas behind this approach and explore its capabilities; we then examine how these extensions can be used beyond three parties. To start each of our three parties -- Angel, Buffy, and Cordelia (or $A_1$, $A_2$, and $A_3$) -- inserts keys and hashes of keys into the IBLT. (For now, we will not use a \textsf{count} field.) However, in this setting both the keys and the hashes of keys are mapped injectively to values in $(\mathbb{F}_3)^b$ for an appropriate $b$. The particular way of mapping to $(\mathbb{F}_3)^b$ does not matter --- we could interpret the key and hash values as number base 3 (at the cost of converting to base 3), or (at the cost of some space) we could interpret the vector of bits of the key or hash value as a vector in $(\mathbb{F}_3)^b$. Now, instead of using XOR in our insertion operation -- which is equivalent to treating keys as $b$-bit elements of $(\mathbb{F}_2)^b$ -- we move to $(\mathbb{F}_3)^b$ (treating keys as sequences of $b$ trits; similarly, hash values are sequences of trits, perhaps of a different length). The three IBLT data structures are combined by summing each of the \textsf{keySum} fields for each cell, as well as summing the \textsf{keyhashSum} fields for each cell, where the sums are sums of elements in $(\mathbb{F}_3)^b$. To begin, let us ignore the issues of the underlying network and assume that all parties obtain all 3 IBLTs. Any key that appears in all three sets is canceled out of \textsf{keySum} by the summation, and similarly the summation of the three hashes of the key is canceled out of \textsf{keyhashSum}. Hence the number of keys existing in the IBLT after this cancellation is $|\cup_i S_i - \cap_i S_i|$. If a key $x$ is found in the \textsf{keySum} field and a matching $H(x)$ is in the \textsf{keyhashSum} field, the key is recovered and removed by subtracting $x$ and $H(x)$ (or equivalently, adding in $2x$ and $2H(x)$ in the appropriate fields). However, some of these keys may appear duplicated in the IBLT; for example, if Angel and Buffy have a key $x$ but Cordelia does not, then the sum of the 3 IBLTs may have a cell containing the value $2x$ in the \textsf{keySum} field and $2H(x)$ in the \textsf{keyhashSum} field. We therefore must further modify our method of recovery. If we see a value $z$ in the \textsf{keySum} field, we must check whether the value $H(z)$ appears in the \textsf{keyhashSum} field, but we must also check whether $2H(z/2)$ appears in the \textsf{keyhashSum} field, in which case we treat it is a verification of $z/2$ as the key. This increases our error rate due to false positives from \textsf{keyhashSum} by at most a factor of two, which is still $O(t/q)$. We note that we could reduce this error rate by instead keeping a count field, which would tell us which one of the two cases above may apply. Also, we note that when removing this key from the IBLT, each cell it is hashed to would then have to remove two copies of the key. \begin{figure*}[t] \begin{minipage}[t]{0.2\linewidth} \centering \includegraphics[scale=0.3]{Net1C.pdf} \end{minipage} \hspace{0.15cm} \begin{minipage}[t]{0.2\linewidth} \centering \includegraphics[scale=0.3]{Net2C.pdf} \end{minipage} \hspace{0.15cm} \begin{minipage}[t]{0.2\linewidth} \centering \includegraphics[scale=0.3]{Net3C.pdf} \end{minipage} \hspace{0.15cm} \begin{minipage}[t]{0.2\linewidth} \centering \includegraphics[scale=0.3]{Net4C.pdf} \end{minipage} \caption{Set reconciliation with network coding. The relay $R$ receives an IBLT from each party, and sends back to each party the sum of the other parties' IBLTs. Each party just needs to send and receive one message, improving on pairwise set reconciliation protocols. Alternatively, in the subfigure on the right, in a wireless setting the relay $R$ can broadcast the sum of all the IBLTs to all parties (denoted by dotted lines), reducing the number of messages further.} \label{fig:netc} \end{figure*} We emphasize that despite this difference, the IBLT listing process works in the same manner, and in particular has the same threshold size for successful listing. This is because a key is recovered exactly when a key is the only key hashed to a cell; the multiplicity of that key within the cell does not affect the listing process. As the IBLT recovery works in the same manner as in the 2-party case, we have the following theorem, based on Theorem~\ref{thm:thm1}. \begin{theorem} \label{thm:thm2} Consider a 3-party reconciliation using IBLTs with $k$ hash functions and a range of $q$ values in the \textsf{keyhashSum} field. As long as we choose $m > (c_k + \epsilon) t$ for some $\epsilon > 0$, and $t \geq |\cup_i S_i - \cap_i S_i|$, the 3-way reconciliation protocol fails with probability $O(t^{-k+2} + t/q)$. \end{theorem} The 3-party protocol uses $m$ cells, which contain the key and a hash. As long as the keys are sufficiently large so that the hash is relatively small compared to the keys, the constant factor from overhead will be small. For 64-bit keys and 32-bit hashes, for example, the total overhead should be less than a factor of 2 for $k =3,4$, and less than a factor of 3 for $k$ up to 7. In many settings, keys or associated stored values can be significantly longer than 64 bits and the overhead will be much smaller. \subsection{Useful Extensions} \label{sec:usefulext} Note that from this process each party can determine the number of other parties (1 or 2) that hold a key they do not possess. It is not hard to add some additional information so that each party can determine which other party holds the key. For example, we could add a 3-bit \textsf{IDs} field to each cell; each time $A_i$ adds (or removes) a key from the IBLT it would toggle the $i$th bit. Hence the $i$th bit of the \textsf{IDs} field would record the parity of the number of keys that $A_i$ has added to the cell. Recall that when we recover a key from the cell, it should be the only key in that cell; hence, the bits set to 1 in the \textsf{IDs} field correspond to the parties that hold that key. (We note that, in fact, having a modulo 3 counter, and having $A_i$ add $i$ to that counter when adding a key, would in fact suffice in this case; the details are left to the reader. Our description here generalizes more readily.) Another reasonable question one might ask is if one party drops out of the protocol, can the remaining two parties still reconcile their sets. The answer is yes. Suppose Cordelia does not participate, so that only Angel and Buffy swap IBLTs. In this case, when combining IBLTs, Angel adds his own IBLT twice (or simply multiples every entry in his IBLT by 2 initially), and similarly for Buffy. That is, each participating party can simply act as though they were two parties with the same set. This guarantees that any key shared by both parties appears 3 times in the IBLT and is canceled; the IBLT listing can be done as above. Note that a participating party must know the number of other participating parties, potentially requiring dropping parties to signal their dropping out in some way. \subsection{Combining with Network Coding} We have thus far assumed that all participating parties get all IBLTs, and hence it may not be clear that this approach is significantly advantageous when compared to simply performing pairwise reconciliations. However, significant advantages become clearer when we consider the transmission of IBLTs over a network. Because IBLTs are linear sketches, based solely on addition, linear network coding methods can be applied. Specifically, suppose $A_1$, $A_2$, and $A_3$ are communicating over a network via a relay $R$. We emphasize that this is a simple setting for illustrative purposes. In Figure~\ref{fig:netc} we let $I(A_i)$ represent the IBLT for $A_i$, and similarly let $I(A_i,A_j)$ be the sum of the IBLTs for $A_i$ and $A_j$, and so on. If we sent $I(A_1)$ to $A_2$ and $A_3$, even if the relay duplicates the IBLT it will have to cross 3 links, and similarly for the other two IBLTs. Hence, the total transmission cost will be 9 IBLTs worth of data. However, suppose as shown in Figure~\ref{fig:netc} that all the Bloom filters $I(A_1),I(A_2)$, and $I(A_3)$ are sent to the relay $R$, and $R$ then takes sums to send $I(A_2,A_3)$ to $A_1$, and similarly for the other parties. Now only 6 IBLTs worth of data need to be sent, saving 1/3 of the transmission cost. This savings is entirely similar to standard network coding techniques. Indeed, in the wireless setting, we could instead have the relay $R$ broadcast the joint IBLT $I(A_1,A_2,A_3)$ of all parties to all the parties, reducing the number of messages down to four. This approach is similar to the now well-known approach of using simple XOR-based network coding in wireless networks \cite{xors}. We note that, in this relay setting, instead of using an IBLT with entries over $(\mathbb{F}_3)^b$, we could build up a joint IBLT using standard IBLTs by having the relay do more work. Given IBLTs $I(A_1)$ and $I(A_2)$, the relay could determine the set difference from the pair of IBLTs, and correspondingly add elements to one IBLT to create an IBLT for the union of the sets. Then on the arrival of $I(A_3)$ the relay could again determine the set difference between $A_1 \cup A_2$ and $A_3$ and use this to build an IBLT for $A_1 \cup A_2 \cup A_3$. This repeated decoding and encoding approach is used in \cite{BoralM}. However, this approach requires much more work from the relay, namely a full decoding for each newly received IBLT, which should be avoided in many settings. Our work shows, for the first time, that such additional work can naturally be avoided. \section{Generalizing to $n$ Parties} We now consider the generalization to $n$ parties. We use a field of characteristic $p$, where $p \geq n$, and we assume keys are mapped injectively (in an arbitrary way) into $(\mathbb{F}_p)^b$, for some $b$. For convenience we simply take $p$ to be a prime here, so that keys are mapped to vectors of non-negative integers smaller than $p$, and sums are computed modulo $p$. (And similarly for hash values, for a possibly different $b$). For simplicity, the reader might think about $b=1$, so keys and hash values are mapped injectively to integers modulo $p$. For convenience we will assume multiplication and division modulo $p$ can be done in constant time; those who object to this assumption may add an appropriate $O(\log^2 p)$ factor to the time bounds (although better bounds may be possible with specially chosen primes). Let $v_i$ denote the representation of $I(A_i)$ as a vector as above. For efficiency and to reduce the probability of a false positive when using the \textsf{keyhashSum} to verify the key value in the cell, we keep a counter modulo $p$ in the \textsf{count} field to track the count of the number of (copies of) keys hashed to a cell by all parties We first state a general result that may not appear directly relevant at first blush. However, this form is useful in that it can be applied to more specific situations, including not only the straightforward generalization to $n$ parties, where we consider a sum of $n$ IBLTs, but also situations (motivated by randomized network coding) in which we are given two different linear combinations of IBLTs, and need to do set reconciliation. The condition under which set reconciliation succeeds is somewhat technical, and it is possible that some set $X$ of keys cannot be recovered. But as we will see, for the linear combinations of interest (such as random linear combinations) the probability that $X\ne\emptyset$ can be made small; in some settings (e.g., when a sum of IBLTs is obtained as in from a relay) the probability will be 0. \begin{theorem} \label{thm:mostgen} Consider an $n$-party reconciliation using IBLTs with $k$ hash functions and a range of $q$ values in the \textsf{keyhashSum} field. Suppose we know two linear combinations (over $(\mathbb{F}_p)^b$) $L_1 = \sum_{i=1}^n \alpha_i v_i$ and $L_2 = \sum_{i=1}^n \beta_i v_i$, as well as the sums of coefficients $\alpha = \sum_i \alpha_i \bmod p$ and $\beta = \sum_i \beta_i \bmod p$, where $\alpha_i \neq 0 \bmod p$ for all $i$, and $\beta \neq 0 \bmod p$. Let $X= \{ x\in \cup_i S_i \; | \; \sum_i (\alpha_i + \gamma \beta_i) 1_{x\in S_i} \bmod p = 0\}$, where $\gamma = -\alpha \beta^{-1} \bmod p$, and $1_E$ is an indicator random variable for the event $E$. As long as we choose $m > (c_k + \epsilon) t$ for some $\epsilon > 0$, and $t \geq |\cup_i S_i - \cap_i S_i|$, then from $L_1$ and $L_2$ we can determine all keys from $(\cup_i S_i - \cap_i S_i) - X$, with probability $1-O(t^{-k+2} + t/q)$. \end{theorem} \begin{proof} If a key is present in all sets $S_i$ then it appears with coefficient $\alpha$ in $L_1$, and similarly it will appear with coefficient $\beta$ in $L_2$. From these we can form the combination of $L_1 + \gamma L_2$, where as stated $\gamma = -\alpha \beta^{-1} \bmod p$. The coefficient of a key that is present in all sets $S_i$ is then 0 modulo $p$ and therefore the key is not present in the IBLT $L_1 + \gamma L_2$. Unfortunately, the same is true for any key that appears exactly in sets $S_j$ for $j \in T$ when $T$ has the property that $\sum_{i \in T} (\alpha_i + \gamma \beta_i) = 0 \bmod p$. All other keys can be found with the given probability using the IBLT recovery process. Note we assume the use of a \textsf{count} field that tracks the weighted multiplicity of the number of keys hashed to a cell modulo $p$ (that is, the sum of the coefficients of the keys hashed to that cell), so that only a single possible key value must be tested in a cell at any time. It is this use of the \textsf{count} field that limits the additional failure probability to $O(t/q)$. We also remark that in the case where $\alpha = 0$, the theorem also holds, but in fact there is no need for the second linear combination $L_2$ (as $\gamma = 0$). In this case, a key present in all sets has a multiplicity that is $0 \bmod p$, and $X$ corresponds those keys $x$ that appear exactly in sets $S_j$ for $j \in T$ where $T$ has the property that $\sum_{i \in T} \alpha_i = 0 \bmod p$. {\hspace*{\fill}\rule{6pt}{6pt}\smallskip} \end{proof} To see how Theorem~\ref{thm:mostgen} can be used, we first consider the basic case, where each party simply wants to find $\cup_i S_i - \cap_i S_i$. We first provide the argument without using the theorem, and then see how the theorem immediately implies the result. Let us assume that each party $A_i$ obtains the sum of all IBLTs $Z = \sum_i v_i$, and of course each party $A_i$ also has $S_i$ and hence $v_i$. If $p > n$, then after $A_i$ obtains the sum $Z$ of all IBLTs, it computes the sum $Z+(p-n)v_i$, essentially acting as $p-n+1$ parties with the same set. As before, this means that the contributions of keys appearing in all sets will cancel out. Each key not appearing in all sets will have an associated multiplicity of less than $n$ in $Z$ and hence less than $p$ in $Z+(p-n)v_i$ (regardless of whether the key is in $S_i$ or not). The algorithm can then for each cell examine the \textsf{count} $a$, the \textsf{keySum}~$y$, and the \textsf{keyhashSum} $z$ to determine if $H(a^{-1}y)=a^{-1} z \bmod p$, which is true when the count $a$ corresponds to a single key. Hence, as before, each key not in all sets can be recovered using the IBLT recovery process with probability $1-O(t^{-k+2} + t/q)$. (For small $n$ one might choose not to use a \textsf{count} field; one could avoid keeping the \textsf{count} field and test all possible \textsf{count} values, that is try all values of $a$ from 1 to $p-1$. Or one can be slightly smarter; if $A_i$ contains a single $x \in S_i$ that hashes to that cell, then $A_i$ need only test the value of $a$ that satisfies $ax = y$, and if $A_i$ has $x \in S_i$ that hashes to that cell, then only $a$ values from $1$ to $n-1$ need to be tested.) Alternatively, we see that this matches the setting of Theorem~\ref{thm:mostgen}, with $L_1 = Z$, $L_2 = v_i$, and $\gamma = -n$. Computing $Z-nv_i = Z+(p-n)v_i \bmod p$ we see that the set $X$ must be empty, because as we argued above any key not in all sets has a count that is non-zero modulo $p$ in $Z+(p-n)v_i$. The result follows. \begin{figure*}[t] \begin{minipage}[t]{0.95\linewidth} \centering \includegraphics[scale=0.6]{Net5C.pdf} \end{minipage} \caption{Set reconciliation example when the parties are connected by a communication tree, a common case in network communications.} \label{fig:tree} \end{figure*} \subsection{Extensions} Several of our extensions from Section~\ref{sec:usefulext} hold in this framework as well. For example, with an $n$-bit \textsf{IDs} field one can track the parity of the number of keys in a cell for each of the $n$ parties, and thereby determine the parties that hold a recovered key. Also, if some parties do not participate, each participating party can simply add in additional copies of its own IBLT to arrange for cancellation if all participating parties hold a key. All that needs to be known for recovery is the number of participating parties. In the case of $n$ parties connected through a relay node, we find that by passing IBLTs through the relay, we can arrange for $n$-party set reconciliation using $2n$ messages on a wired network and $n+1$ messages on a wireless network, where the relay broadcasts the sum of the IBLTs to all parties. This improves over the simplistic natural approach of using $n(n-1)$ point-to-point messages to compute all pairwise differences. However, it should be noted, the pairwise difference messages may in fact be smaller in size, since pairwise set differences may be smaller than the total set difference. \section{Network Coding and IBLTs} At a high level, our work thus far suggests that, by working over a suitable finite field, IBLTs can be naturally plugged in to linear network coding schemes to provide efficient set reconciliation mechanisms, where the message size corresponds (up to constant factors) to the generalized set difference. We believe this correspondence is indeed quite general, and while the broad nature of use and applications of network coding make it difficult to turn this statement into a theorem, we provide some sample applications. \subsection{Set Reconciliation on Trees} We first consider set reconciliation among $n$ parties, connected by a communication network that is a {\em rooted tree} with $E$ edges, known in advance. The parties are at the leaves of the tree. At each time step, each node in the tree can send a message to each of its neighbors. Let $P$ be the length of the longest path from a leaf to a root of the tree. We assume in what follows that keys are sufficiently large so that other overhead (e.g., the \textsf{keyhashSum} field) at most affects the total size by a constant factor, and that an upper bound on $|\cup_i S_i - \cap_i S_i|$ within a constant factor is known. We claim the following. \begin{theorem} Set reconciliation among $n$ parties on a rooted tree can be performed in $2P$ time steps with $2E$ total messages of size $O(|\cup_i S_i - \cap_i S_i|)$ with probability $1-O(t^{-k+2} + nt/q)$. Each non-leaf node in the network requires only $O(|\cup_i S_i - \cap_i S_i|)$ storage. \end{theorem} \begin{proof} Each party constructs its IBLT, and sends it up the communication network to the root. Non-leaf nodes excluding the root combine all the IBLTs from leaves in their subtree, and pass them up to the root. The root gathers all the IBLTs and sends the union of them back down the communication tree. (See Figure~\ref{fig:tree}.) Messages are of size $O(|\cup_i S_i - \cap_i S_i|)$, and each edge of the tree carries a single message in each direction. The IBLTs allow each party to recover all keys with probability $1-O(t^{-k+2} + t/q)$. This follows immediately from Theorem~\ref{thm:mostgen}, as again we are in the setting where $L_1 = \sum_i v_i$ and $L_2 = v_i$. All of the IBLTs will be peeling the same set of keys, so the $O(t^{-k+2})$ term in the failure probability is common to all $n$ parties. The $n$ parties may have different counts associated with different cells, however, from each adding in their own term $(p-n)v_i$, so we take a union bound over the $O(t/q)$ failure probability associated with a false match from the \textsf{keyhashSum} field over the $n$ parties. (We note that if $n = p$, all $n$ parties can use the same IBLT, and this union bound is avoided.) Because of the way IBLTs are combined only $O(|\cup_i S_i - \cap_i S_i|)$ space is required at non-leaf nodes. {\hspace*{\fill}\rule{6pt}{6pt}\smallskip} \end{proof} \subsection{Set Reconciliation via Gossip on General Networks} We now show how gossip spreading techniques (also referred to generally as rumor spreading) allow multi-party set reconciliation over a network in $O(\phi^{-1} \log n)$ rounds of communication, where $\phi$ is the conductance of the network, using our IBLT framework. In gossip spreading, there are generally two different models \cite{shah}. In the single-message model there is one message at a vertex in a graph with $n$ vertices\footnote{For this problem we follow the standard notation for gossip problems and use $n$ for the number of vertices.}, and the goal is for every vertex to obtain that message. In the multi-message model, the standard setting is that for some subset of the vertices, each vertex has a unique message, and all vertices have to obtain all of the messages. With a {\sc PUSH} strategy, in each round every node that has a message contacts a random neighbor, and forwards a single message. The {\sc PULL} strategy is similar, but each node without a message contacts a random neighbor and obtains a message from them. A {\sc PUSH}-{\sc PULL} strategy combines both of these operations in every round. (See, for example, \cite{chierichettialmost,giakkoupis}.) In these models, a vertex can transfer a single message to another vertex in each round. In the multi-message model, one can use network coding by having a vertex send a (usually random) linear combination of the messages it holds at this time, so that messages take (essentially) the same space, but can offer potentially more useful information. (See, for example, \cite{DebMedard,haeupler}.) Our setting does not exactly match any of these situations. For reconciliation, we have multiple vertices each with their own message (the IBLT), and we can combine messages, so we appear to be similar to the multi-message model with network coding. However, in the network coding setting, the eventual goal is to solve a collection of linear equations (corresponding to the combinations of messages received), and in that setting, each new message only provides one more ``degree of freedom'', or one more needed equation, regardless of the component messages it contains. For set reconciliation, we need not solve such a system (although that would be one way to solve the problem); we merely need some appropriate linear combination of all of the IBLTs, as demonstrated in Theorem~\ref{thm:mostgen}. Also, as we are working in the reconciliation setting, we do not require that all vertices obtain the reconciled sets, but only those vertices that begin with a set initially. Because of this, we may more naturally think of the problem as a collection of single-message problems running in parallel, as we now describe. Our approach is as follows. We have $a \leq n$ parties $A_1,\ldots,A_a$ with sets $S_1,\ldots,S_a$ and corresponding IBLTs $I(S_1),\ldots,I(S_a)$. For convenience, without loss of generality we provide the argument where $a = n$, as fewer parties with messages only makes things easier. (Or we may think of parties without a message as having a null message.) Parties can use whatever single-message gossiping algorithm is available. Messages in this setting correspond to random mixtures of IBLTs. That is, here we again think of the IBLT as a vector of entries in $(\mathbb{F}_p)^b$ for a suitably large prime $p$. (Here $p$ will need to be larger than previously to obtain a low failure probability, as we see below.) A message will consist of a linear combination of such vectors $$\sum_{i \in [n]} \alpha_i v_i,$$ where $v_i$ is the IBLT vector for the $i$th party and $\alpha_i$ is a coefficient (modulo $p$). Let the vector of the $j$th party after $\ell$ rounds of communication be given by $\alpha_{ij\ell}$, $i=1,\dots,n$. Our goal is for each party $j$ to obtain a vector $\sum_{i \in [n]} \alpha_{ijL} v_j$ at round $L$ where $\alpha_{ijL} \neq 0$ for all $i$. At that point, as a special case of Theorem~\ref{thm:mostgen}, the $j$th party can reconcile using the combined IBLT by adding $(p-\sum_{i \in [n]} \alpha_{ijL}) v_j$ to this vector, thereby ``canceling out'' any key in the intersection of the sets. To bound the probability that the set $X$ of non-recovered keys is nonempty, we first need to describe how the coefficients come about. The protocol runs as follows. Each vertex holds one linear combination of IBLTs at any time; after round number $\ell$ party $j$ stores $$\sum_{i \in [n]} \alpha_{ij\ell} v_i$$ as well as the coefficient sum $\alpha_{j\ell} = \sum_{i \in [n]} \alpha_{ij\ell}$. To send a message, a party $j$ chooses a random $\kappa_{j\ell} \neq 0$ modulo $p$ and sends $$\kappa_{j\ell} \sum_{i \in [n]} \alpha_{ij\ell} v_i.$$ together with the corresponding coefficient sum $\kappa_{j\ell} \alpha_{j\ell} \bmod p$. Each party receiving a message simply adds it to its current message. The probability of ever ``zeroing out'' a coefficient using this approach is negligible for suitably large $p$. To see this, notice that each coefficient $\alpha_{ij\ell}$ is a (non-zero) multilinear polynomial of degree at most $\ell$ of the random multipliers $\kappa_{j\ell}$ applied to each message. This implies that the probability that $\alpha_{ij\ell}=0$ is $O(\ell/p)$, by the Schwartz-Zippel lemma \cite{Schwartz,Zippel}. Similarly, the probability for each set $T$ that $\sum_{i\in T} \alpha_{ij\ell}$ assumes a given value is $O(\ell/p)$, implying that $X$ is empty with probability $1-O(t\ell/p)$ for each party. Now we examine the protocol from the point of view of a single message. For a single message, the protocol behaves exactly as the single-message protocol; the fact that other IBLTs may be piggy-backing along in a shared message does not make any difference. Hence, we can treat this as multiple single-message problems running in parallel, and apply a union bound on the failure probability. Generally, standard results for single-message problems come with an exponentially decreasing tail bound for the probability of not finishing after a given number of rounds. Assuming this, an additive $O(\log n)$ steps over a standard single-message result are sufficient to guarantee, via union bound, that the $n$ parallel problems all complete. As a specific example, we can consider the best current results on the standard {\sc PUSH}-{\sc PULL} protocol for gossip spreading \cite{giakkoupis}. Let $\phi$ be the conductance of a communications graph $G$ with $n$ vertices, where $a$ of the $n$ vertices wish to reconcile their sets. We can prove the following theorem. \begin{theorem} \label{thm:pushpull} Set reconciliation on a graph of $n$ vertices with $a = O(n)$ parties having sets to reconcile can be accomplished in time $O(\phi^{-1} \log n)$ using the standard randomized {\sc PUSH}-{\sc PULL} protocol with messages of size $$O(|\cup_i S_i - \cap_i S_i|)$$ with success probability $$1-O(t^{-k+2} + nt/q+ntL/p+n^2L/p+n^{-\beta})$$ for any constant $\beta > 0$. \end{theorem} \begin{proof} As mentioned without loss of generality we take the case where $a = n$. We choose a suitable stopping time $L = O(\phi^{-1} \log n)$ based on the choice of the constant $\beta$ that would be suitable for $n$ parallel versions of the single-message {\sc PUSH}-{\sc PULL} gossip protocol to successfully complete with high probability, as guaranteed by Theorem 1.1 of \cite{giakkoupis}. We now apply Theorem~\ref{thm:mostgen}. From our discussion above, we have that after $L$ rounds the $i$th party will store $L_1$ that is a linear combination $\sum_{i \in [n]} \alpha_{ijL} v_i$, as well as $L_2 = v_i$. Here the $\alpha_{ijL}$ are all non-zero with high probability because of the use of random coefficients as discussed above; this probability it at most $O(n^2 L/p)$ by the union bound as there are $n^2$ coefficients and each is 0 with probability at most $O(L/p)$. We further claim the set $X$ is empty for all parties with high probability, also because of the use of random coefficients. To see this, consider any key $x$ in $\cup_i S_i - \cap_i S_i$. For $x$ to not be recoverable by the IBLT by the $i$th party, it would require that $\sum_{j \in T} \alpha_{ijL} = 0 \bmod p$, where $T$ is again the set of parties that have $x$ as a key. This happens with probability at most $O(tL/p)$ as discussed above. Hence, by a union bound over parties, the probability that this happens over all parties is at most $O(ntL/p)$. We therefore obtain full recovery for all parties using randomized {\sc PUSH}-{\sc PULL}, with high probability. Note that again we must take into account that each party has different coefficients in their IBLT, and hence one must apply a union bound to cover the possibility of false positives from the \textsf{keySum} and \textsf{keyhashSum} fields over all IBLTs. However, the recovery process will be the same for all IBLTs, since they all involve the same keys. Our final probability bound includes this accounting. {\hspace*{\fill}\rule{6pt}{6pt}\smallskip} \end{proof} \subsection{Experiments} We briefly describe some experiments designed to test the gossip algorithms approach. We emphasize that the experiments were meant as ``proof-of-concept'', and not an extensive experimental test.\footnote{These experiments were performed by Marco Gentili, who we thank for allowing their use in this paper.} We choose as our test graphs random graphs where each edge is included independently with probability $\frac{2 \ln n}{n}$; this is sufficient to guarantee the graph is connected (with high probability). For our experiments, each each node is a party to the protocol, and each party's set is simply one element, with all sets being distinct. We use IBLTs of $2n$ cells (more than needed) to ensure listing succeeds with high probability. The choice of sets does not significantly impact the failure probability. Every time a party receives a message, it adds the corresponding IBLT multiplied by a random multiplier into its linear combination of IBLTs. We use the {\sc PUSH}-{\sc PULL} protocol as described previously. We also determined with preliminary experiments the number of rounds needed to ensure that all parties would receive the information held from all parties with high probability, and used this many rounds. By doing so, we limit our failures to resulting from the zeroing out of coefficients of the linear combination of IBLTs. This failure probability therefore corresponds to the $O(ntL/p + n^2L/p)$ term from Theorem~\ref{thm:pushpull}. \begin{table}[!htb] \caption{ Success Rate of Listing Keys (Averaged over 1000 trials) with $p=1000000007$}\label{table:p1000000007} \centering \begin{tabular}{|l|l|l|l|l|l|l|} \hline \begin{tabular}[c]{@{}l@{}}\# Parties\end{tabular} & \begin{tabular}[c]{@{}l@{}}\% \\Retrieving \\ All\\ Msgs\end{tabular} & \begin{tabular}[c]{@{}l@{}}\% \\Missing\\ 1 Msg\end{tabular} & \begin{tabular}[c]{@{}l@{}}\% \\Missing\\ \textgreater1 Msg\end{tabular} & \begin{tabular}[c]{@{}l@{}}\# \\Retrieving \\ All Msgs\end{tabular} & \begin{tabular}[c]{@{}l@{}}\# \\Missing\\ 1 Msg\end{tabular} & \begin{tabular}[c]{@{}l@{}}\# \\Missing\\ \textgreater1 Msg\end{tabular} \\ \hline 10 & 100.00\% & 0.00\% & 0.00\% & 10000 & 0 & 0 \\ \hline 20 & 100.00\% & 0.00\% & 0.00\% & 20000 & 0 & 0 \\ \hline 40 & 100.00\% & 0.00\% & 0.00\% & 40000 & 0 & 0 \\ \hline 80 & 100.00\% & 0.00\% & 0.00\% & 80000 & 0 & 0 \\ \hline 160 & 100.00\% & 0.00\% & 0.00\% & 160000 & 0 & 0 \\ \hline 320 & 100.00\% & 0.00\% & 0.00\% & 320000 & 0 & 0 \\ \hline 640 & 100.00\% & 0.00\% & 0.00\% & 640000 & 0 & 0 \\ \hline 1280 & 100.00\% & 0.00\% & 0.00\% & 1279999 & 1 & 0 \\ \hline \end{tabular} \end{table} Table \ref{table:p1000000007} shows the success rate for listing keys using $p = 1000000007$, averaged over 1000 trials. All keys for $n$ up to 640 were reconciled; for $n=1280$, one key from one party is not recovered in one trial. (Various backup measures could be used to easily handle such rare cases.) Note that this value of $p$ would fit into a 32-bit integer and is not unreasonable for calculations. Other experiments with smaller values of $p$ shows that the failures occur at a rate roughly inversely proportional to $p$, as suggested by Theorem~\ref{thm:pushpull}. \section{Conclusion} Previous solutions for set reconciliation, based primarily on Reed-Solomon codes, have not (as far as we are aware) been generalized to the multi-party setting, and it is not immediately clear how to do so. We have shown here that methods based on Invertible Bloom Lookup Tables, which require additional space but only linear time, generalize naturally, and further they do so in a way that allows the application of network-coding based techniques. Hence, by utilizing network coding methods, we can obtain high efficiency in terms of the number of network messages required, as well as small messages because IBLTs and combinations of IBLTs have length proportional to the generalized set difference. While we expect our approach might be improved further, providing better space utilization or smaller probability of error either by theoretical improvements or by careful implementation, we believe this work represents an important step in establishing more practical solutions to the multi-party set reconciliation problem than approaches based on pairwise interactions. There are, of course, quite a number of linear sketches in the literature beyond IBLTs, and combining such sketches is a fairly common technique. Our work emphasizes how IBLTs can naturally lead to reconciliation algorithms that take advantage of methods based on network coding. It would be very interesting if a general statement formalizing this connection more concretely could be developed, or, alternatively, if we can find other cases where the utility of linear sketches can be increased by applying network coding techniques to expand their capabilities or efficiency. We observe that in settings where point-to-point messages can be transmitted more efficiently than by broadcasting, there is potential in some cases to decrease the total size of messages sent and received by each party. For example, this is the case when each set $S_i$ lacks $t$ elements from $\cup_i S_i$, and each of these elements is present in all other sets. Then running a single pairwise set reconciliation protocol with point-to-point messages of size $O(t)$ suffices for each party. However, the IBLT would require each party to send and receive messages of size $O(tn)$. Of course, in this example the parties are making use of the knowledge that all other parties already hold the missing items, but it shows that there are settings where IBLTs will not be optimal. More generally, we leave it as an open problem to explore the possible trade-offs in using or combining various set reconciliation protocols in additional settings. \section*{Acknowledgments} The first author thanks George Varghese for suggesting the problem of multi-party set synchronization. \bibliographystyle{plain}
\section{Introduction} \label{sec:introduction} The idea of simulating a complex, many-body physical system with a simpler model system has a long and rich history. A prominent example is the Bose-Hubbard (BH) model~\cite{Fisher1989}, which governs the dynamics of bosons tunneling between sites of a lattice with energy $t$ and interacting on a given site with energy $U$. Originally proposed to describe the behavior of superfluid $^4$He in porous Vycor glass, it is now applied to a plethora of experimental systems including Josephson junction arrays~\cite{Zant-1992}, ultracold atoms in optical lattices~\cite{Bloch2005}, photonic crystals~\cite{Corrielli2013}, and arrays of coupled cavities~\cite{Hartmann2006,Greentree2006, Angelakis2007,Houck2012}. For repulsive interactions ($U>0$), there is a competition between delocalization due to the tunneling and the tendency to localize due to the energy cost of multiple occupancy of a given site. As a consequence, the model exhibits a quantum (zero-temperature) phase transition~\cite{Fisher1989}. In the strongly interacting (weak tunneling) regime~$t/U\ll 1$, the ground state is predicted to be a Mott insulator (MI)~\cite{Mott1937,Mott1949,Mott1982}, in which each site is occupied by an (identical) integer number of bosons; for $t/U\gg 1$, a superfluid (SF) state results~\cite{Kapitza1938,Allen1938,Landau1941}, in which the bosons become completely delocalized. The transition between these phases was first realized in Bose-Einstein condensates confined in three-dimensional optical lattice potentials~\cite{Greiner2002}, where the ratio $t/U$ was controlled by varying the well depth, and subsequently observed in many other cold-atom optical lattice experiments in one, two and three dimensions~\cite{Stoferle2004,Spielman2007,Gemelke2009,Bakr2010,Haller2010,Trotzky2010}. There has been much recent interest in observing similar quantum phase transitions in cavity quantum electrodynamics~\cite{Hartmann2006,Greentree2006,Angelakis2007}. The principal motivation for employing these environments is the robustness of available technology for producing, manipulating and detecting photons. Unfortunately, photons do not intrinsically interact. Various strategies have been proposed to overcome this limitation, and a vast array of interesting many-body states have been conjectured to appear as a result~\cite{Hartmann2007,Rossini2007,Carusotto2009,Kiffner2010,Halu2013,Hayward2012,Schiro2012}. One model that has attracted particular attention is the Jaynes-Cummings-Hubbard (JCH) model~\cite{Hartmann2006,Hartmann2008a,Makin2008}, comprising a lattice of high-finesse optical cavities each containing one or more two-level atoms, with neighboring cavities coupled via the overlap of their evanescent modes. The (bosonic) photons can then be considered to `tunnel' from cavity to cavity. The atom interacts with the quantized electromagnetic field present within the cavity according to the Jaynes-Cummings model~\cite{Jaynes1963}, and photon interactions are generated via the photon-blockade mechanism~\cite{Birnbaum2005}. Importantly, the lattice polaritons that constitute the spin-photon elementary excitations of the JCH model are also expected to undergo a phase transition from a Mott insulator to a superfluid~\cite{Greentree2006,Hartmann2006,Angelakis2007,Rossini2007,Aichhorn2008,Koch2009,Schmidt2009,Pippan2009,Schmidt2010,Knap2010,Hohenadler2011}. While early proposals for the experimental realization of coupled cavities involved nitrogen vacancies in diamond~\cite{Greentree2006}, self-assembled quantum dots in photonic crystals~\cite{Na2008}, and trapped ions~\cite{Ivanov2009,Mering2009}, more recent proposals favor circuit QED~\cite{Nunnenkamp2011,Wu2011,Houck2012,Schmidt2013}. The close similarity between the properties of the BH and JCH models suggests that the polariton superfluid resulting from arranging cavities along a row should be rather peculiar. In one dimension, the interaction energy of free bosons is strongly enhanced relative to their kinetic energy at low densities~\cite{Bloch2005}, a result that will be discussed in greater detail in Sec.~\ref{subsec:Tonks}. The resulting Tonks-Girardeau gas~\cite{Tonks1936, Girardeau1960} is described as the hard-core limit of the (integrable) Lieb-Liniger model for bosons with delta-function interactions in one dimension~\cite{Lieb1963}. These hard-core bosons can be exactly mapped to non-interacting fermions~\cite{Girardeau1960}; for general densities the state is well-described within the framework of Luttinger liquid theory~\cite{Giamarchi2003}. The Tonks-Girardeau gas was first realized in ultracold atomic Bose gases confined in one-dimensional optical lattices~\cite{Paredes2004, Kinoshita2004}. More recently, the transition between a Luttinger liquid and Mott-insulating state was observed~\cite{Haller2010}. These results suggest that a one-dimensional arrangement of coupled cavities should be sufficient to induce the mobile photons to behave entirely as fermions. In fact, this possibility was noted previously for a dissipative model~\cite{Carusotto2009}. The central goal of the present investigation is to show that the photons are effectively fermionized in the ground-state of the JCH model. A secondary goal is to make a careful, side-by-side comparison of the JCH and BH models in one dimension. In this work, the ground states of the zero-temperature one-dimensional JCH and BH models in the vicinity of the MI phase boundary are obtained numerically using finite-system density matrix renormalization group (DMRG) methods. Several quantities are calculated that provide evidence for the nature of the states, including particle densities, one-particle and two-particle correlation functions, the superfluid fraction, the condensate fraction, and the entanglement entropy. Finite-size scalings are performed in order to infer the value of these quantities in the thermodynamic limit. Two main conclusions can be drawn from our work. First, for small cavity couplings the ground state consists of a polariton MI phase as expected, though with some interesting features not shared by single-component Bose systems. Second, the results clearly reveal the strong fermionization of both the photons and spins in the JCH system throughout the so-called superfluid phase in the low-density limit. Detailed comparisons are made with the ground states of the BH model in the equivalent parameter regimes. The main conclusion that can be drawn is that superfluidity is weakly manifested in this phase, if it exists at all. The manuscript is organized as follows. The JCH and BH models are reviewed in Sec.~\ref{sec:models}, and the DMRG methods used in the characterization of their ground states is described. This section also discusses the properties of the Tonks-Girardeau gas. The numerical results are presented in Sec.~\ref{sec:results}, and the discussion and conclusions are found in Sec.~\ref{sec:conclusions}. \section{Models and Methods} \label{sec:models} In this work we compare the properties of the one-dimensional Jaynes-Cummings-Hubbard (JCH) and Bose-Hubbard (BH) models, using density matrix renormalization group (DMRG) methods. This section briefly provides the background to these models and describes the numerical methods employed in the calculations. \subsection{JCH model} \label{subsec:JCH} The behavior of a single two-level atom, in the pseudospin representation, confined to a single high-finesse cavity is given by the Jaynes-Cummings Hamiltonian~\cite{Jaynes1963}, written within the rotating-wave approximation~\cite{Mandel1995} as \begin{equation} H_{\rm JC} = \omega_c \brackets{a^{\dag}a + \frac{1}{2}} + \frac{1}{2}\omega_a \sigma^z + g \brackets{a^{\dag} \sigma^- + a \sigma^+}. \label{eq:JC} \end{equation} Here, $\omega_c$ is the natural cavity frequency ($\hbar=1$ in this work for convenience), $\omega_a$ is the excitation frequency of the atom, $a$ ($a^{\dag}$) is the photon annihilation (creation) operator, $\sigma^z$ is the spin-$1/2$ representation of the Pauli $z$ operator, $\sigma^{\pm}$ are the spin raising and lowering operators, and $g$ is the strength of the atom-photon coupling, proportional to the magnitude of the inner product between the dipole vector and the local field. In this work, $g$ is assumed to be a real quantity, equivalent to assuming that the dipole and field oscillate in phase. This model describes an isolated system which ignores environmental couplings. The cavity is therefore assumed to have arbitrarily high finesse, and be in the strong-coupling limit. In the rotating-wave approximation the total number of excitations \begin{equation} N_i = a_{i}^{\dag}a_i+\sigma^{+}_i \sigma^{-}_i, \label{eq:JCnumber} \end{equation} is a conserved quantity (note that the spin number operator would normally contribute the term $\sigma^{-}_i \sigma^{+}_i$, but this corresponds to the population of atoms in the ground state; hence, it is not included in the \textit{excitation} number operator). The eigenstates of $H_{\mathrm{JC}}$ are coherent superpositions of photonic and spin excitations with a definite total excitation number, known as polaritons~\cite{Hartmann2006}. Within a particular excitation number block $N$, the eigenstates are given (c.f.\ Appendix A of Ref.~\cite{Koch2009}) by \begin{align} \label{eq:polaritonminus} \ket{N-} = \sin{\theta_N} \ket{Ng} + \cos{\theta_N} \ket{(N-1) e}; \\ \label{eq:polaritonplus} \ket{N+} = \cos{\theta_N} \ket{Ng} - \sin{\theta_N} \ket{(N-1) e}, \end{align} with mixing angle \begin{equation} \theta_N = \frac{1}{2} \arctan{\brackets{\frac{2g\sqrt{N}}{\delta}}}. \label{eq:polaritonmixing} \end{equation} Here $\delta := \omega_c - \omega_a$ is the detuning of the cavity and atomic frequencies. The eigenenergies of Eq.~(\ref{eq:JC}) for $N \geq 1$ are given by \begin{equation} E_{N\pm} = N\omega_c + \frac{\delta}{2} \pm \left[ \left( \frac{\delta}{2} \right)^2 + Ng^2 \right]^{1/2}, \label{eq:eigsJC} \end{equation} while for $N=0$, $E_0=0$. The energy levels are thus arranged in two-dimensional manifolds labeled by the polariton number $N$ (except for the $N=0$ sector, which is one-dimensional), separated by the energy of the single-photon cavity mode, $\omega_c$. The anharmonicity in the eigenenergies~(\ref{eq:eigsJC}) of size $\sqrt{N}$ is the origin of the photon-blockade effect~\cite{Birnbaum2005}, giving rise to effective photon interactions. Consider for simplicity the zero-detuning case $\delta=0$ giving eigenenergies $E_{N\pm}=N\omega_c\pm g\sqrt{N}$. With one photon the cavity has the lower energy eigenvalue $E_{1-}=\omega_c-g$. Na\"\i vely, two independent photons would yield the total energy $2E_{1-}=2\omega_c-2g$, but in fact the two-photon energy eigenvalue is $E_{2-}=2\omega_c-g\sqrt{2}$. The difference between these energies yields an estimate for the effective repulsive photon-photon interaction strength: $E_{2-}-2E_{1-}=(2-\sqrt{2})g$. In the Jaynes-Cummings-Hubbard (JCH) model, the cavity mode leakage is no longer neglected. Instead, one imagines a regular lattice of $L^d$ cavities in $d$ dimensions positioned sufficiently close together that a photon emitted from one cavity can be absorbed into an adjacent cavity with energy (rate) $\kappa$. The JCH model is written as \begin{equation} H_{\mathrm{JCH}} = \displaystyle \sum_i \brackets{H_{\mathrm{JC},i} -\mu N_i} - \kappa \sum_{\langle i,j \rangle} \brackets{a^{\dag}_i a_j + a^{\dag}_j a_i}, \label{eq:JCH1} \end{equation} where $H_{{\rm JC},i}$ and $N_i$ are Eqs.~(\ref{eq:JC}) and (\ref{eq:JCnumber}) respectively with $\{a,a^{\dag},\sigma^z,\sigma^{\pm}\}$ replaced by $\{a_i,a^{\dag}_i,\sigma^z_i,\sigma^{\pm}_i\}$, and $i$ labels the position of a cavity. The $\langle i,j\rangle$ notation indicates that the sum is over nearest neighbors. The chemical potential $\mu$ fixes the mean polariton number, and is employed primarily to connect with the results of the BH model which is generally solved in the grand canonical ensemble. Note that in the JCH model, the atoms are fixed within the cavities, and only the photons are able to `tunnel' from cavity to adjacent cavity. For reasons that will become clearer momentarily, it is convenient to rescale the JCH energy in units of the coupling constant $g$; in the limit of zero detuning $\omega_c=\omega_a$ one can rewrite the Hamiltonian~(\ref{eq:JCH1}) as \begin{eqnarray} \frac{H_{\rm JCH}}{g}&=&-\frac{\kappa}{g}\sum_{\langle i,j\rangle}\left( a_i^{\dag}a_j+a_j^{\dag}a_i\right)\nonumber \\ &+&\sum_i\left[a_i^{\dag}\sigma_i^-+a_i\sigma_i^+ -\left(\frac{\mu-\omega_c}{g}\right)N_i\right], \end{eqnarray} where unimportant additive constant terms are omitted. The first term corresponds to the photon hopping, the second to the local JC term, and the last term can be considered as a rescaled chemical potential for the total polariton density $\tilde{\mu}:=(\mu-\omega_c)/g$. \subsection{BH Model} \label{subsec:BH} In the BH model, bosons tunnel between nearest neighboring sites of a lattice, and experience on-site interactions (which can be either attractive or repulsive in general). The model is described by the Hamiltonian \begin{equation} H_{\mathrm{BH}} =-t\sum_{\langle i,j \rangle} \brackets{b^{\dag}_i b_j + b^{\dag}_j b_i} +\sum_i\left[\frac{U}{2}N_i (N_i - 1) - \mu N_i\right],\ \label{eq:BH} \end{equation} where $b_i$, $b^{\dag}_i$, and $N_i:=b^{\dag}_ib_i$ are respectively the bosonic annihilation, creation, and number operators for site $i$, $t>0$ is the nearest-neighbor tunneling amplitude and $U$ is the on-site interaction energy. The chemical potential fixes the mean boson density on the lattice. In this work only repulsive interactions $U>0$ will be considered. While a direct mapping between the JCH and BH models is not possible because the former has two different kinds of excitations while the latter has only one, the parameters can be chosen in such a way as to simplify comparisons. One can rescale the BH energies in terms of the interaction strength by dividing Eq.~(\ref{eq:BH}) by $U$. In this case the hopping amplitudes are $\kappa/g$ and $t/U$ in the JCH and BH models, respectively. The effective interaction strength between photons is $(2-\sqrt{2})g$, which implies that the two systems should become similar for $t/U\sim\kappa/(2-\sqrt{2})g =\left(1+\frac{1}{\sqrt{2}}\right)(\kappa/g)\approx 1.707(\kappa/g)$. A similar connection can be obtained between the chemical potentials of the two models: $\tilde{\mu}+1\sim(2-\sqrt{2})\mu/U$ or $\mu/U\sim 1.707(\tilde{\mu}+1)$. Note that these scalings are valid only when only a single atom is confined to each cavity. As discussed in the Introduction, in the weak tunneling (strong interactions) limit $t/U\ll 1$ the ground state is an incompressible MI characterized by localized bosons, (constant) integer occupation of a given site, and an energy gap to excitations of order $U$. Deep in this limit, the ground-state wavefunction can be approximated as $|\Psi\rangle\sim\prod_ib_i^{\dag} |\Phi\rangle$, where $|\Phi\rangle$ is the particle vacuum state and the product is over all lattice sites. Because each site is independent of any other, the overlap of the states $b_s|\Psi\rangle$ and $b_r|\Psi\rangle$ is exactly zero unless $r=s$. The one-body boson correlation function \begin{equation} G^{(1)}(r,s)=\langle b_r^{\dag}b_s\rangle \label{eq:G1} \end{equation} and the two-body correlation function \begin{equation} G^{(2)}(r,s)=\langle b_r^{\dag}b_s^{\dag}b_sb_r\rangle \label{eq:G2} \end{equation} will then be zero for all $r\neq s$. In reality, for any finite $t/U$ the gapped ground state will deviate from this simple prediction and the correlation functions should instead decrease exponentially in $|r-s|$ with a characteristic length scale $\xi\sim 1/U$ that scales as the inverse of the gap to excitations~\cite{Hastings2006}. The correlation length diverges as a system becomes critical~\cite{Sachdev2000}. One would therefore expect $\xi$ to increase from $0$ to $\infty$ as the hopping goes from 0 to its critical value at the phase boundary for a fixed chemical potential. The phase boundary in $\mu$-$t$ space, known as the `Mott lobe,' is roughly semi-circular in profile in two and three dimensions~\cite{Fisher1989}. For $\mu\notin\field{Z}$, the system remains in the MI phase with increasing $t$ until some critical value at which point the system undergoes a phase transition to SF; likewise for constant $t$ and increasing $\mu$. The Mott lobe becomes strongly distorted in one dimension~\cite{Kuehner1998}, and the system displays re-entrance: at constant $\mu$, on increasing $t$ the ground state phase changes from MI to SF to MI and back to SF again. In the strong tunneling (weak interactions) limit $t/U\gg 1$, the ground state of the BH model corresponds to an interacting Bose-Einstein condensate. Each boson is highly extended throughout the lattice, and the ground state can be approximated by $|\Psi\rangle\sim\left(\sum_ib_i^{\dag}\right)^{N_B}|\Phi \rangle$, where $N_B$ is the number of bosons. This compressible state is characterized by a gapless linear spectrum and long-range correlation functions~(\ref{eq:G1}) and (\ref{eq:G2}) that are independent of $|r-s|$. In one-dimension, however, true Bose-Einstein condensation is not possible; rather, the ground state corresponds to a quasi-condensate with only algebraic long-finite order and characterized by strong fluctuations~\cite{Popov1987}. Instead one finds $G^{(1)}(r,s)\sim 1/|r-s|^{\alpha}$, where the parameter $\alpha$ characterizes the degree of quasi-condensation ($\alpha\to 0$ for a true condensate). \subsection{Tonks-Girardeau Gas} \label{subsec:Tonks} At very low densities, one-dimensional repulsively interacting bosons form a Tonks-Girardeau gas, and effectively behave as non-interacting fermions~\cite{Girardeau1960}. In the absence of any external potential (other than the ones used for confinement), the ground state properties are governed by the kinetic and interaction potential energies $T$ and $U$. The mean kinetic energy per particle scales as $T/N\sim1/m\ell^2 \sim \langle n\rangle^2/m$, where $\ell$ is the mean interparticle distance which in one dimension scales as the inverse of the mean particle density $\ell\sim 1/\langle n\rangle$. (In the presence of a weak lattice, the bare boson mass $m$ is rescaled to an effective mass $m^*\sim 1/t$). When the interaction potential can be modeled in terms of a pseudopotential (low energy, long-wavelength collisions), one can write the mean interaction potential in one dimension as $U/N\sim\langle n\rangle /m|a_{\rm 1D}|$~\cite{Olshanii1998}, where $a_{\rm 1D}$ is the one-dimensional s-wave scattering length. The Tonks parameter, the ratio of the potential and kinetic energies $\gamma=U/T=2/\langle n\rangle|a_{\rm 1D}|$ is therefore huge at low densities, in marked contrast to the situation in higher dimensions. To minimize the interaction potential, particles prefer to be as far apart from one another as possible, much like fermions. The free fermionic wavefunction can be written in terms of a Slater determinant to guarantee the proper antisymmetrization of the wavefunction. For example, a system of $N$ free fermions on $L$ sites has a wavefunction given in the position representation by \begin{equation} \Psi_F(r_1, \dots, r_L) = \text{det} \left[ \begin{array}{cccc} \phi_1(r_1) & \phi_1(r_2) & \dots & \phi_1(r_L) \\ \phi_2(r_1) & \phi_2(r_2) & \dots & \phi_2(r_L) \\ \vdots & \vdots & \ddots & \vdots \\ \phi_N(r_1) & \phi_N(r_2) & \dots & \phi_N(r_L) \end{array} \right], \end{equation} where the $\set{r_i}$ indicate the positions of the lattice sites and the $\set{\phi_i}$ are single-particle wavefunctions. In the perfectly hard-core limit of the Tonks-Girardeau gas, the ground state of the fermionized bosons is simply \begin{equation} \Psi_B(r_1, \dots, r_L)= \prod_{i<j}^L{\rm sgn}(r_i-r_j)\Psi_F(r_1, \dots, r_L), \label{eq:TonksPsi} \end{equation} where the factor multiplying $\Psi_F$ ensures that all negative signs associated with the interchange of two fermions disappears. Many properties are shared by $\Psi_F$ and $\Psi_B$. For example, the local density profile of both systems in real space is the same, since $|\Psi_B(r_1, \dots, r_N)|^2 = |\Psi_F(r_1, \dots, r_N)|^2$~\cite{Yukalov2005}. Similarly, all density correlation functions are the same~\cite{Cazalilla2011}; for example, for a ring of length $L\to\infty$, the normalized two-body correlation function is \begin{equation} g^{(2)}(r,s)=\frac{\langle b_r^{\dag}b_s^{\dag}b_sb_r\rangle} {\langle b_r^{\dag}b_r\rangle\langle b_s^{\dag}b_s\rangle} = 1-\left[\frac{\sin(\pi n|r-s|)}{\pi n|r-s|}\right]^2, \label{eq:g2Tonks} \end{equation} where $n$ is the mean particle density. The correlation function is zero at $r=s$, reflecting the Pauli exclusion principle; this behavior is referred to as the `exclusion hole'. Away from this point the correlation function grows and displays Friedel oscillations~\cite{Friedel1958} that decay with increasing $|r-s|$. For one-dimensional spinless fermions, the oscillations have wavelength $\lambda_F=1/n=2\pi/k_F$ where $k_F$ is the Fermi wavevector. Thus, the presence of an exclusion hole and Friedel oscillations in the two-body correlation function is a `smoking gun' for the fermionization of bosons in the Tonks-Girardeau gas. For a finite system with $N$ free fermions on $L$ sites with open boundary conditions, such as is considered in this work, a straightforward calculation yields \begin{equation} \label{eq:g2FF} G^{(2)}(r,s) = \brackets{\frac{N}{L+1}}^2\sbrackets{B(r,s) - A(r,s)}, \end{equation} where \begin{align} A(r,s)&=\left[\frac{\cos{\frac{\pi\brackets{N+1}\brackets{r-s}}{2(L+1)}} \sin{\frac{\pi N\brackets{r-s}}{2(L+1)}}} {N\sin{\frac{\pi\brackets{r-s}}{2(L+1)}}}\right. \nonumber \\ &-\left.\frac{\cos{\frac{\pi\brackets{N+1}\brackets{r+s}}{2(L+1)}} \sin{\frac{\pi N\brackets{r+s}}{2(L+1)}}} {N\sin{\frac{\pi\brackets{r+s}}{2(L+1)}}}\right]^2;\nonumber \\ B(r,s)&=\sbrackets{1-\frac{\cos{\frac{\pi\brackets{N+1}r}{(L+1)}}}{N} \frac{\sin{\frac{\pi Nr}{(L+1)}}}{\sin{\frac{\pi r}{(L+1)}}}}\nonumber \\ &\times\sbrackets{1-\frac{\cos{\frac{\pi\brackets{N+1}s}{(L+1)}}}{N} \frac{\sin{\frac{\pi Ns}{(L+1)}}}{\sin{\frac{\pi s}{(L+1)}}}}\label{eq:g2FF2}. \end{align} It is simple to verify that $G^{(2)}(r,r)\to 0$. For large separations between particles $\left|\frac{(r-s)}{(L+1)}\right| \gg 0$, one finds that $A(r,s)$ and $B(r,s)$ oscillate in the vicinity of zero and unity, respectively, so that $G^{(2)}(r,s) \approx \brackets{\frac{N}{L+1}}^2.$ Choosing the location of one particle at the center of the chain $r=\lceil L/2 \rceil$, $G^{(2)}$ far from the center oscillates about a mean value approximately equal to the square of the mean particle density $n$. For $\left|\frac{r-s}{L+1}\right| \gg 0$, the oscillation of $G^{(2)}$ is governed by the last term in the definition of $B(r,s)$ in Eq.~(\ref{eq:g2FF2}). In the thermodynamic limit $N,L\to\infty$ but $n=N/L\to$~const., one obtains $B(\lceil L/2 \rceil,s) \approx 1-\sin(2\pi ns)/2\pi ns$. The Friedel oscillation wavelength is therefore again $\lambda_F=1/n=2\pi/k_F$. The single-body correlation function is not the same for the Tonks-Girardeau and free fermion gases, however: the sign function in Eq.~(\ref{eq:TonksPsi}) does not disappear when inserted into Eq.~(\ref{eq:G1}). The calculation of this quantity is quite involved~\cite{Cazalilla2011}, but the asymptotic behavior $|r-s|\gg 0$ but $|r-s|\ll L$ is found to be \begin{equation} G^{(1)}(r,s)\sim\frac{1}{\sqrt{2n_0L|\sin(\pi|r-s|/L)}}. \label{eq:G1Tonks} \end{equation} For $|r-s|\ll L$ one obtains $G^{(1)}(r,s)\sim 1/\sqrt{|r-s|}$, which indicates that for the Tonks-Girardeau gas the exponent of the power law is $\alpha=1/2$. Another `smoking gun' for the Tonks-Girardeau phase is therefore the power-law behavior of the one-body density matrix with exponent $\alpha=1/2$. The Fourier transform of the one-particle correlation function $G^{(1)}(r)$ is the momentum distribution $n(k)$. For the Tonks-Girardeau gas, the power-law behavior at long distances translates into a power-law divergence of the momentum distribution at long wavelengths, $n(k)\sim 1/|k|^{1/2}$ for $k\to 0$. This highly peaked distribution is reminiscent of the delta-function distribution that one would expect if the bosons formed a Bose-Einstein condensate, except it is now broadened due to the finite-range phase order associated with the quasi-condensation. This distribution is dramatically different from that of a non-interacting Fermi gas, where $n(k)$ is a constant for all $k \leq k_F$ and is zero otherwise ($k_F$ is the Fermi wavevector). The momentum distribution for a Tonks-Girardeau gas in a weak axial trapping potential has been experimentally observed~\cite{Paredes2004,Kinoshita2004}. \subsection{Numerical Methods} \label{subsec:methods} The characteristics of the BH and JCH models were obtained by means of finite-system density matrix renormalization group (DMRG) simulations. We employed the DMRG code from the Algorithms and Libraries for Physics Simulations (ALPS) project~\cite{Albuquerque2007,Bauer2011}. Simulations were carried out for systems of size $L=15,19,23,27,31$ with both open boundary conditions (equivalent to hard-wall boundary conditions) and periodic boundary conditions, and a finite-size scaling analysis was performed for all quantities (unless explicitly noted) in order to infer the results for the thermodynamic limit. We use DMRG as the method because it is suitable for obtaining results that are so precise as to be considered exact~\cite{White1993}, while being able to handle much larger finite-size systems than exact diagonalization~\cite{Schollwock2005,Schollwock2011}. The bulk of the simulations employed open boundary conditions, in order to accelerate convergence. For the BH (JCH) model, a maximum of $N_{\text{max}}=5$ $(6)$ bosons (photons) per site (cavity) were allowed, and we kept $M=80$ (100) states. For the JCH model, this corresponds to a Hilbert dimension $D=12$ for each cavity. For the superfluid fraction, the method chosen necessitated the use of periodic boundary conditions. Usually the number of states kept for these simulations is on the order of the square of the number chosen for open boundary conditions; however, since the method only required ground state energies and not correlation functions, the numerical requirements were not as stringent. The superfluid fraction calculation for the BH (JCH) model was performed using $N_{\text{max}}=7$ $(6)$ and $M=200$ (140). In all cases, we verified that increasing the values of $M$ and $N_{\text{max}}$ did not change the ground state energies or correlation functions. For the BH (JCH) system, these parameters correspond to a maximal Hilbert space dimension of $1.60\times10^5$ $(1.44\times10^6)$ for the simulations with open boundary conditions, and $1.96\times 10^6$ $(2.82 \times 10^6)$ for periodic boundary conditions. We used eight finite-size sweeps for all simulations, and verified that the ground state energy and correlation functions did not change by increasing the number of sweeps. The calculation times for a single run for the simulations for the BH (JCH) models were typically approximately 1 hour (24 hours) when using open boundary conditions, while the runs using periodic boundary conditions required up to approximately 8 (24) hours. In order to keep the calculation time to a minimum, we assumed that the total boson (polariton) number was a conserved quantity in every case. Note that this does not pose a problem even in the superfluid phase because the superfluid density need not correspond to the mean boson (polariton) density. Parameters for the hopping ($t$ or $\kappa$ for the BH or JCH models, respectively) and chemical potential $\mu$ were chosen in order to remain in the vicinity of the MI phase boundary. In 1D, the tip of the Mott $n=1$ lobes in the $(\mu/U,t/U)$ and $(\tilde{\mu},\kappa/g)$ planes for the BH and JCH models are found to be located at approximately $(0.09,0.3)$~\cite{Kuehner1998,Kuehner2000,Ejima2011} and approximately $(-0.95,0.2)$~\cite{Rossini2007,Mering2009}, respectively. This value of $\kappa/g$ is consistent with the rescaling factor of approximately $1.7$ between the BH and JCH models, discussed in Sec.~\ref{subsec:BH}. To span most of the Mott lobe, the range of hopping is therefore chosen to be $t/U\in[0,0.27]$ and $\kappa/g\in[0,0.16]$ for the two models. Likewise, the phase boundaries for $t=0$ and $\kappa=0$ correspond to $\mu/U=n$~\cite{Fisher1989} and $\tilde{\mu}=\sqrt{n}-\sqrt{n+1}$~\cite{Koch2009} for the BH and JCH models in any dimension, respectively. Thus, the transition from the $n=0$ to $n=1$ Mott lobes at zero hopping occurs for $\mu/U=1$ and $\tilde{\mu}=-1$ for the two models; the transition to the $n=2$ Mott lobe occurs for $\mu/U=2$ and $\tilde{\mu}\approx -0.41$. To capture some of the $n=0$ lobe and approximately half of the $n=1$ lobe, we chose chemical potentials in the range $\mu/U\in[-0.17,0.55]$ and $\tilde{\mu}\in[-1.10,-0.68]$. Only the simplest zero-detuning case $\delta=0$ is considered in this work. Previous work has shown that detuning can be a useful parameter, changing the effective strength of interactions and thereby the phase diagrams~\cite{Koch2009,Schmidt2009,Schmidt2010,Schmidt2013}. \section{Results} \label{sec:results} \subsection{Density phase diagrams} \label{subsec:densities} The location of the phase boundary between the gapped MI phase and the SF phase has been previously established numerically in the thermodynamic limit with finite-size DMRG, for both the 1D Bose-Hubbard model~\cite{Kuehner1998,Kuehner2000,Ejima2011} and the 1D JCH model~\cite{Rossini2007}. The results for the mean densities of bosons and polaritons are shown in Fig.~\ref{fig:densities} for the parameter sets discussed in the previous section and the largest number of lattice sites $L=31$ studied. Open boundary conditions are employed in this case; for periodic boundary conditions the density on any site would coincide with the mean density. While such density plots have not to our knowledge been previously shown in the literature, the main purpose of showing these plots here is to orient the reader to the location in phase space for which the simulations have been conducted. The ranges of the normalized hopping parameters $t/U, \kappa/g$ and effective chemical potentials $\mu/U, \tilde{\mu}$ are chosen to be equivalent, based on the scaling assumption given in Sec.~\ref{subsec:BH}. For both models, the goal is to explore the regions in the vicinity of the transition between the MI and SF phases. Unlike most previous studies, this work focuses particularly on the low-density region where fermionization is expected. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.172\textwidth]{JCHdensityL31-crop.pdf} \label{JCH-density-polaritons-31} } \subfigure[][]{ \includegraphics[height=0.172\textwidth]{BHdensityL31_rescaled-crop.pdf} \label{BH-density-31} } \caption{Mean density phase diagrams for (a)~polaritons in the 1D JCH model and (b)~bosons in the 1D BH model, in both cases for a system size $L=31$. Regions of constant mean density correspond to MI states.} \label{fig:densities} \end{figure} The mean densities of either bosons or polaritons can be used to distinguish the two phases in both models, since the mean density is pinned to an integer in the MI regime, but not in the SF regime. Consequently, the locations of the phase boundary for our finite-size systems are quite clearly visible in Fig.~\ref{fig:densities}. The shapes of the phase boundaries closely resemble the thermodynamic limit results of Refs.~\cite{Kuehner1998,Rossini2007}, though their positions are shifted slightly due to the finite-size system. Depicted is the region of the phase diagram where the $n=0$ and $n=1$ Mott lobes meet, as well as the low-density superfluid regions near the boundaries of these lobes. Roughly, the mean density of bosons (polaritons) increases with increasing $t$ ($\kappa$). This work is mainly concerned with the low-density superfluid regions of phase space, corresponding to low hopping and chemical potential. The re-entrant shape of the $n=1$ Mott lobe can be clearly seen as the constant-density region of the BH model in Fig.~\ref{BH-density-31}, but is not as obvious in Fig.~\ref{JCH-density-polaritons-31}. At certain fixed values of $\mu$ within a continuous range, monotonically increasing the hopping parameter $t$ or $\kappa$ from zero causes the system to transition from the MI to the SF, back to the MI and again back to the SF regime~\cite{Rossini2007}. Viewed from within the Mott lobe, the lower part of the phase boundary (the hole boundary) is concave, while the upper part (the particle boundary) is convex. The particle and hole boundaries meet at a sharply cusped tip. The phase transition along the line of constant mean density passing through this point is in the $(d+1)$-dimensional XY universality class, which for $d=1$ is of the Berezinskii-Kosterlitz-Thouless (BKT) type~\cite{Fisher1989,Kosterlitz1973} with Tomonaga-Luttinger parameter $K_b=1/2$~\cite{Kuehner2000,Ejima2011}. The MI-to-SF transition across either the particle or the hole boundary is generic (i.e.\ Gaussian like the condensation transition of an ideal Bose gas) and characterized by $K_b=1$~\cite{Kuehner2000}. This implies that the SF phase near the particle or hole boundaries should be characterized by one-particle correlation functions $G^{(1)}(r)\sim|r|^{-K_b/2}\sim 1/r^{1/2}$, consistent with the Tonks-Girardeau gas scaling. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.175\textwidth]{JCHatomdensityL31-crop} \label{fig:atomdensities-31} } \subfigure[][]{ \includegraphics[height=0.175\textwidth]{JCHphotondensityL31-crop} } \caption{(a) Mean spin and (b) photonic excitation densities in the JCH model with $L=31$. These can vary throughout the MI phase.} \label{fig:atomphotondensities-31} \end{figure} In the JCH model, the mean densities of the spin and photonic species, depicted in Fig.~\ref{fig:atomphotondensities-31}, need not remain proportional to track each other. Consider first the mean spin excitation density, Fig.~\ref{fig:atomdensities-31}. In the atomic limit $\kappa=0$, the cavities are decoupled from each other and thus the overall ground state is the $L$-fold tensor product of the single-cavity lower-lying polariton states \begin{equation} \displaystyle \ket{\psi_{\text{ground}}}=\bigotimes_{i=1}^L \ket{1-}_i. \end{equation} From Eqs.~(\ref{eq:polaritonminus}) and (\ref{eq:polaritonmixing}) in the case of zero-detuning, the single-cavity polariton ground states $\ket{1-}$ are equal-weight superpositions of a photonic and a spin excitation. Consequently, the mean spin excitation density is equal to $\frac{1}{2}$ in this limit. Conversely, in the hopping-dominated limit $\kappa \rightarrow \infty$, the photons and spins decouple. Each atom is then in the ground state so the mean spin excitation density vanishes. At intermediate $\kappa$ between these extremes, the data in Fig.~\ref{fig:atomphotondensities-31} show that across the lower boundary of the Mott lobe (the hole boundary) there is a sharp drop in the mean density of spin excitations, which can be used to distinguish the two phases. By contrast, crossing the upper boundary, the mean spin excitation density varies smoothly from $\frac{1}{2}$ in the atomic limit to lower values, presumably tending towards $0$ in the limit $\kappa \rightarrow \infty$. Next consider the mean photon density. Unlike the spin excitations, the number of photons and hence the mean photon density is unbounded from above in the grand canonical ensemble. In the large-hopping limit at fixed chemical potential, one expects the mean photon density to increase. In fact, for a lattice with coordination number $z_c$, an instability occurs when the hopping roughly satisfies $z_c\kappa/g > -\tilde{\mu}$; for larger values of the hopping, the ground state energy decreases without bound as a function of increasing photon number~\cite{Koch2009}. This regime is not considered in the calculations, as it corresponds to larger densities deep in the SF phase where fermionization is unlikely. The mean photon density within the $n=1$ MI lobe tracks with the mean spin excitation density in such a way that the overall polariton density is pinned to 1 per site, as expected. Crossing the upper (particle) boundary of the Mott lobe, the mean density of photons begins to increase rapidly with increasing hopping. This indicates that in the hopping-dominated limit, the system behaves like a photon superfluid, with the effects of the spins becoming negligible. On the other hand, in the intermediate regions between the $n=0$ and $n=1$ lobes, the mean photon density remains low; the reason is that at low hopping in one dimension, the effective repulsive interactions between photons become strong and thus there is an energy cost associated with adding photons to the system. \subsection{Correlation functions} \label{subsec:correlations} \subsubsection{One-body density matrix} Consider now the single-particle correlation function $G^{(1)}$, defined in Eq.~(\ref{eq:G1}). This has been calculated previously via DMRG for the 1D BH model, to verify the asymptotic predictions of the Luttinger liquid theory~\cite{Pai1996}, and to estimate the location of the critical value of $t/U$ for the BKT transition~\cite{Kuehner1998}. Similar plots of $G^{(1)}$ for varying interaction strengths in the $n=1$ MI lobe of the 1D BH model are shown in Ref.~\onlinecite{Ejima2012}. The normalized version $C_r(s-r):=G^{(1)}(r,s)/\sqrt{N_rN_s}$ was also considered for the 1D BH model with an additional harmonic trapping potential~\cite{Kollath2004} for various different sites $r$, and the coexistence of the two phases was found at certain points in the phase diagram. As discussed in Sec.~\ref{subsec:BH}, in the MI phase one expects this correlation function to decrease exponentially with distance, $G^{(1)}(r,s)\sim\exp(-|r-s|/\xi)$ over a correlation length $\xi$, while in the SF phase it should behave as a power law, $G^{(1)}(r,s)\sim 1/|r-s|^{\alpha}$ where $\alpha$ is some positive constant. The correlation functions are calculated for the JCH model (in which case $b\to a$ and $\sigma$ for photons and spins, respectively) and the BH model for systems of size $L=15$, 19, 23, 27, and 31. Consider for concreteness the shortest length $L=15$, for which the fits are the least reliable; the results are shown in Fig.~\ref{fig:g1repJCH}. This size is chosen convey the worst-quality results for various system sizes. The correlations are measured with respect to the central site, in this case site number $8$. The point where $r=s$ is not plotted or used for the fit, since only the asymptotic form of the correlations for $|r-s| \gg 0$ is of interest. To mitigate boundary effects, the two sites closest to the (open) boundary of the system were also not considered. The correlations for the spins and the photons track each other very closely, so the results for photons are not presented. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.197\textwidth]{g1repJCHatomsMI-crop} \label{fig:JCHG1MI} } \subfigure[][]{ \includegraphics[height=0.197\textwidth]{g1repJCHatomsSF-crop} \label{fig:JCHG1SF} } \caption{Representative plots of the correlation function $G^{(1)}(8,s)$ for the spin excitations in the JCH model with $L=15$. (a)~$n=1$ MI phase at point $\tilde{\mu} = -0.900$, and (b)~SF phase at point $\tilde{\mu} = -0.989$. In both cases the same five values of $\kappa/g$ are chosen with values given in the legends. The correlation functions are consistent with (a)~exponentials and (b)~power-laws.} \label{fig:g1repJCH} \end{figure} The results indicate that the spin excitations and photons behave in close analogy to the bosons of the BH model; that is, the single-particle correlation functions are clearly exponential within the MI lobe and follow power laws in the SF regime. The power law relationship holds up well not only for $L=15$, but also for all larger values of $L$ considered. The exponent associated with the power law, corresponding to the slope of the log-log fit, is increasingly precise for larger values of $L$, with the standard error $\sim 1/\sqrt{L}$. The correlation function decreases most rapidly deep in the MI region but increasingly slowly as the phase boundary is approached. The results shown in Fig.~\ref{fig:JCHG1SF} correspond to the SF phase at constant $\mu$, for values of $\kappa/g$ that range from almost immediately adjacent to the MI lobe ($\kappa/g=0.022$) almost to the edge of the phase diagram ($\kappa/g=0.133$) in Fig.~\ref{fig:atomdensities-31}. Unsurprisingly, the power-law fits are poor near the phase boundary (c.f.\ the points corresponding to $\kappa/g=0.022$) but improve as one moves further away. The correlations exhibited by each type of carrier also track perfectly with each other. This result is in qualitative agreement with Ref.~\cite{Knap2010}, in which the same quantity was calculated using the Variational Cluster Approximation. In fact, this feature persists for all quantities discussed below, unless mentioned explicitly. An intriguing consequence of the identical behavior for the two species of excitations is that even though one normally views the atoms as mediating photonic interactions, one could just as well think of the photons as mediating atomic interactions; though the atoms are each isolated within their own cavities, they nevertheless feel each others' presence. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.216\textwidth]{fssg1photonsJCHMI-crop.pdf} } \subfigure[][]{ \includegraphics[height=0.216\textwidth]{fssg1photonsJCHSF-crop.pdf} \label{fig:scalingSF} } \caption{Representative finite-size scaling analysis of $G^{(1)}$ in the (a) MI and (b) SF regimes of the JCH model. The data points for the MI regime are $(\tilde{\mu},\kappa/g)=(-0.856,0.044)$ (deep within MI lobe), $(-0.856,0.128)$ (near generic transition), and $(-0.900,0.156)$ (near BKT point). The SF regime points are at $(-0.989,0.022)$ (near generic transition), $(-0.989,0.106)$ (deep within SF phase), and $(-0.989,0.156)$ (near BKT point).} \label{fig:g1repfss} \end{figure} The infinite-system values of the correlation length $\xi$ and the power $\alpha$ are estimated using a finite-size scaling analysis, as shown in Fig.~\ref{fig:g1repfss} for a few representative points in phase space. The best exponential or power-law function is fit for $G^{(1)}$ for each value of $L=15$, 19, 23, 27, and 31, such as is shown in Fig.~\ref{fig:g1repJCH} for $L=15$. The values of $\xi$ and $\alpha$ are then plotted as a function of $1/L$, and the data are fit to a line whose intercept is interpreted as the corresponding value in the thermodynamic limit. The data are only weakly dependent on system size deep in the MI phase and near the generic phase boundary, but show a strong dependence near the BKT point. On the SF side the values of $\alpha$ are size-dependent for all three phase space points considered; this likely reflects the fact that along the $\tilde{\mu}=-0.989$ line the Mott boundary remains nearby. The same procedure is carried out for the BH model for comparison (not shown). The phase diagram for the value of $\xi$ in the thermodynamic limit is displayed for the JCH and BH models in Fig.~\ref{fig:g1xiJCH}. Note however that one cannot obtain $\xi$ in this way immediately at the phase boundary. The true correlation length is expected to diverge, and a diverging correlation length cannot be accurately captured by a DMRG procedure with finite truncation $M$~\cite{Schollwock2011}. Furthermore, the precise location of the phase boundary varies depending upon the size of the system, so a particular point in phase space near the boundary displayed for $L=31$ may or may not be in the Mott-insulating lobe, depending upon the system size. In Fig.~\ref{fig:g1xiJCH}, only the phase space points that are unambiguously within the $n=1$ lobe have been included; this accounts for the apparently smaller MI lobes than are depicted in Fig.~\ref{fig:densities}. The numerical results clearly indicate that the correlation length is independent of $\mu$ and is solely a function of $\kappa$ or $t$, smoothly increasing with increasing hopping. Since everywhere within the Mott lobe the state has a well-defined number of excitations $N_{\text{tot}}$, the only effect of varying the chemical potential by an amount $\Delta\mu$ while fixing $\kappa$ or $t$ is to shift the entire spectrum by an amount $N_{\rm tot}\Delta\mu$. The ground state itself at fixed hopping is therefore independent of $\mu$ within the lobe. The correlation function only increases as the BKT point is approached, not near the generic phase boundaries. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.170\textwidth]{JCHg1photonscorrlengthTDL-crop} } \subfigure[][]{ \includegraphics[height=0.170\textwidth]{BHcorrlengthTDL_2_new-crop} } \caption{Phase diagrams for $\xi$ in the thermodynamic limit, assuming $G^{(1)}(r)\sim\exp(-r/\xi)$ in MI regime of the (a) JCH model (photons only) and the (b) BH model. The correlation length $\xi$ approaches the system size at the edge of the Mott lobe near the BKT point.} \label{fig:g1xiJCH} \end{figure} The value of $\alpha$ in the thermodynamic limit is calculated for all points in the superfluid phase for both the JCH and BH models, and the results are shown in Fig.~\ref{fig:g1alphaJCH}. Right at the phase boundary, the exponent approaches values on the order of unity or higher. The same caveats mentioned above for the calculation of $\xi$ apply here as well. In addition, the spatial dependence of $G^{(1)}$ (an example of which is shown in Fig.~\ref{fig:JCHG1SF}) and the finite-size scaling data (an example of which is shown in Fig.~\ref{fig:scalingSF}) are much noisier right near the phase boundary. Everywhere near the phase transition, however, the value of $\alpha$ is close to $0.5$. This value is consistent with what would be expected for a Tonks-Girardeau gas of photons [c.f.\ Eq.~(\ref{eq:G1Tonks})] and matches the Luttinger parameter $K_b=1$, as discussed in Sec.~\ref{subsec:densities}. In fact, $\alpha$ has previously been used to obtain the location of the phase boundary for the BH model, using infinite system DMRG with periodic boundary conditions~\cite{Kuehner1998}. This indicates that the low-density regime outside of the Mott lobes is in fact not strongly superfluid in nature. Rather, the results are consistent with the complete fermionization of the BH bosons and the JCH photons in the equivalent regime. Furthermore, since $G^{(1)}$ for the spin excitations and photons in the JCH model track each other so well (not shown), the spin excitations have also been fermionized, even though the atoms are treated as spins with no particular exchange statistics. The `SF' designation of this phase therefore appears to be a misnomer, but it will be kept for clarity of exposition in what follows. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.172\textwidth]{JCHg1photonsalphaTDL-crop} } \subfigure[][]{ \includegraphics[height=0.172\textwidth]{BHalphaTDL_new-crop} \label{BHalpha} } \caption{Phase diagrams for $\alpha$ in the thermodynamic limit, assuming $G^{(1)}(r)\sim r^{-\alpha}$ in SF regime of the (a) JCH model (photons only) and the (b) BH model. These values of $\alpha$ indicate strong fermionization of the photons in the JCH model.} \label{fig:g1alphaJCH} \end{figure} The value of $\alpha$ decreases for increasing hopping $\kappa$ or $t$, but the trend is slow for the parameter range studied. By the edge of the plots in Fig.~\ref{fig:g1alphaJCH}, the exponent for the SF phase between the $n=0$ and $n=1$ lobes has dropped to almost constant (in terms of $\mu$) values of $0.33$ and $0.29$ for the JCH and BH models, respectively. For larger values of $\mu$ where the density is higher the exponent drops off more rapidly, reaching a range of 0.081-0.254 for the JCH model and 0.226-0.336 for the BH model at the edge of the plots. These values are all quite different from the value $\alpha=0$ that one would expect for an ordinary superfluid, however. \subsubsection{Two-body correlation function} One of the most important signatures of fermionization within the low-density SF phase is found in the two-body correlation function $G^{(2)}(r,s)$ defined in Eq.~(\ref{eq:G2}) and its normalized variant $g^{(2)}(r,s)$ defined in Eq.~(\ref{eq:g2Tonks}), as discussed in Sec.~\ref{subsec:Tonks}. Within the $n=1$ Mott-insulating lobe, one expects $g^{(2)}(r,s)=1-\delta_{r,s}$ irrespective of the model and the excitation, which is exactly what is observed. Of greater interest is the behavior of this correlation function in the SF regime for different mean densities. For a perfect superfluid the two-body correlation function is featureless, $g^{(2)}(r,s)=1$ in the bulk, reflecting the fact that all superfluid carriers occupy the same plane wave state. On the other hand, for a system of free fermions, the two-body correlation function exhibits two important features. The first is the presence of an exclusion hole at $r=s$ reflecting the Pauli exclusion principle. The second is the characteristic Friedel oscillations appearing on either side of the exclusion hole, with a wavelength $\lambda_F$ set by the mean density $n$ or Fermi wavelength $k_F$, as discussed in Sec.~\ref{subsec:Tonks}. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.200\textwidth]{figg2JCH1-crop} } \subfigure[][]{ \includegraphics[height=0.200\textwidth]{figg2JCH2-crop} \label{fig:G2JCH2} }\\ \subfigure[][]{ \includegraphics[height=0.200\textwidth]{figg2BH1-crop} } \subfigure[][]{ \includegraphics[height=0.200\textwidth]{figg2BH2-crop} \label{fig:G2BH2} }\\ \caption{The spatial-dependence of the unnormalized two-body correlation function $G^{(2)}(16,s)$ for $L=31$ is shown for representative points in the SF phase where $\alpha \approx \frac{1}{2}$. Figures (a) and (b) correspond to photons at the points $(\tilde{\mu},\kappa/g)=(-0.989,0.039)$ and $(-0.989,0.106)$ in the JCH model, respectively; `T' and `D' in turn denote `theory' and `data.' Figures (c) and (d) correspond to bosons at the points $(\mu/U,t/U)=(0.019,0.066)$ and $(0.019,0.180)$, respectively.} \label{fig:g2BH} \end{figure} The unnormalized two-body correlation functions $G^{(2)}(r,s)$ for the photons and spin excitations in the JCH model are plotted in Fig.~\ref{fig:g2BH} for representative points in the SF phase where $\alpha\approx\frac{1}{2}$; the BH results are also shown for comparison. (The momentum distribution $n(k)$ in the 1D BH model was previously considered for various interaction strengths within the $n=1$ MI lobe~\cite{Ejima2011}, and for both the MI and SF regimes at unit filling~\cite{Kollath2004}). The data are compared with the analytical expression for free fermions at the same mean density, Eq.~(\ref{eq:g2FF}). The unnormalized two-body correlation function is plotted in order to make apparent the particle densities for the particular points in phase space. The two main signatures of fermionization in the SF phase, the exclusion hole and the Friedel oscillations, are evident in all the plots. At lower densities, the match between the data and the prediction based on non-interacting fermions is excellent. This indicates that the superfluid density at this point in the SF phase is close (if not exactly equal) to zero. One would expect the systems to behave more like a superfluid as the tunneling is increased and the excitation density increases, for points in phase space that are further from the phase boundary. Indeed this is the case; the data depicted in Figs.~\ref{fig:G2JCH2} and \ref{fig:G2BH2} show that the exclusion hole is now slightly filled in, and the Friedel oscillations are increasingly washed out. An interesting feature in the case of the JCH model, not present in the BH model, is that the wavelength of the oscillations for both the photons and the bosons is set by the mean density of polaritons. The amplitude of $G^{(2)}$ for each carrier is determined by the density of just that carrier, however. More precisely, to obtain a fitting curve of the form of Eq.~(\ref{eq:g2FF}) for $G^{(2)}(r,s)$ for the photons (spin excitations), one must use the mean number of photons (spin excitations) as the value of $N$ in the prefactor $\left(N/(L+1)\right)^2$, but the number of polaritons in the oscillatory functions appearing in $A(r,s)$ and $B(r,s)$. Hence, the spin excitations and the photons are individually fermionized, inasmuch as they have inherited this property from the polaritons. \subsection{Other measures of the ground state} \label{subsec:others} \subsubsection{Superfluid fraction} The most compelling evidence for the existence of a SF phase would be the presence of a non-zero superfluid order parameter. An example of such an order parameter is the superfluid fraction or superfluid stiffness $f_s$, the ratio of particles exhibiting superfluid flow to the total number of particles. The factor $f_s$ can be calculated numerically by imposing periodic boundary conditions and applying a phase twist $\Theta \ll \pi$ to the boundary conditions~\cite{Fisher1973,Krauth1991,Singh1994,Roth2003}. In practise, this can be accomplished by means of a Peierls factor applied to the bosonic creation and annihilation operators in the BH and JCH models, which has the effect of modifying the hopping terms via $a_i^{\dagger} a_j \longmapsto a_i^{\dagger} a_j e^{-\mathrm{i} \Theta/L}$~\cite{Roth2003a}. While this induces a velocity $v=(2J)\nabla\Theta$ for each quantum particle ($J$ is the hopping coefficient corresponding to $\kappa$ in the JCH model or $t$ in the BH model) because the current density is $j=nv$, only the particles in the superfluid will respond collectively. As a result, the ground state energy will increase relative to the twist-free case solely due to the kinetic energy of the superfluid particles. From this change of the ground state energy, the superfluid fraction can be determined: \begin{equation} f_s = \frac{L}{2Jn}\left.\frac{\partial^2 E_0(\Theta)}{\partial\Theta^2} \right|_{\Theta=0} \approx\frac{L}{Jn} \frac{E_0(\Theta)-E_0}{\Theta^2}, \label{eq:sffracformula} \end{equation} where $E_0$ is the ground state energy with no phase twist and $E_0(\Theta)$ is ground state energy with overall twist $\Theta\ll\pi$. The latter expression is a finite-difference approximation for the second derivative of the energy with respect to the phase twist, using the central three-point stencil. This calculation of $\rho_s$ has previously been performed at constant density near the BKT transition of the 1D BH model across the tip of the $n=1$ Mott lobe~\cite{Pai1996}; a significant jump in $\rho_s$ across the transition was found in that work. This work instead examines $\rho_s$ at representative points in the low-density SF phase far from the transition, with considerably different results. In the numerical calculations, we considered various points in the SF phase using five different finite-size systems $L=15,19,23,27,31$. Periodic boundary conditions are required, which are computationally more demanding for the DMRG method than are open boundary conditions (see the discussion in Sec.~\ref{subsec:methods}); hence, only a set of representative points in the SF phase were considered. Recall from Sec.~\ref{subsec:methods} that the $n=1$ BKT points in the JCH and BH models are located at approximately $(\tilde{\mu},\kappa/g)=(-0.95,0.2)$~\cite{Rossini2007,Mering2009} and $(\mu/U,t/U)=(0.09,0.3)$~\cite{Kuehner1998,Kuehner2000,Ejima2011}, respectively. We therefore considered these three points in the SF phase of the JCH model: $(-0.944, 0.133)$, $(-0.989, 0,028)$, and $(-1.00, 0.240)$. The first is just left of the BKT point, in the vicinity of the hole boundary, the second is between the $n=0$ and $n=1$ Mott lobes, and the third is to the right of the BKT point. In particular, the first two points correspond to mean polariton densities $n < 1$, while the third has $n >1$. For the BH model we considered the two points: $(0,0.08)$ and $(0,0.5)$; again, the first is left of the BKT point in the vicinity of the hole boundary, while the second point is much to the right, in the deep SF phase well beyond the region depicted in Figs.~\ref{BH-density-31} and \ref{BHalpha}. The first of these points has mean boson density $n < 1$ and the second, $n > 1$. The ground-state energy was obtained for these points in the SF region, for phases $\Theta/\pi\in(0.0, 1.0)$ in increments of $0.2$. Note that the ground state energy is invariant under the transformation $\Theta \longmapsto -\Theta$. For each value of $L$, the second derivative in Eq.~(\ref{eq:sffracformula}) was then calculated using $\Theta=0.2\pi$. The results were then verified by estimating the derivative using a central-difference approximation with five-point ($\Theta/\pi \in \set{0,\pm 0.2,\pm 0.4}$) and seven-point stencils ($\Theta/\pi \in \set{0,\pm 0.2,\pm 0.4,\pm 0.6}$); no discernible difference from the method of Eq.~(\ref{eq:sffracformula}) was observed in the results so obtained. The values of $f_s$ were also obtained by extracting the coefficient for the quadratic term in a polynomial fit of $E_0(\Theta)$. Again, no discernible differences from the central-difference results were found using this approach. Once the value of $f_s$ was obtained for a given system size, a finite-size scaling analysis in $1/L$ was performed to interpolate to the thermodynamic limit. As a final check on the results, the calculation in Eq.~(\ref{eq:sffracformula}) was repeated with smaller values of the phase twist, $\Theta/\pi\in[0.0,0.2]$ in increments of $0.02$. For $\Theta/\pi < 0.06\pi$, the values of $\rho_s(L)$ could not be reliably fitted to determine the value in the thermodynamic limit (the problem of dividing one small number by another). However, for $\Theta/\pi \geq 0.06$, the results were consistent with those obtained with $\Theta/\pi=0.2$. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.236\textwidth]{fssSFfractionJCH-crop} } \subfigure[][]{ \includegraphics[height=0.236\textwidth]{fssSFfractionBH-crop} \label{fsBH} } \caption{Superfluid fraction for the (a) JCH and (b) BH model outside but near the $n=1$ MI lobe. For the JCH model, the symbol $\tilde{\kappa}=\kappa/g$ is used for compactness. For both systems, the curve marked by red squares (color online) is beyond the critical hopping value for the BKT transition for the $n=1$ MI lobe, while the others are not. In both cases, the superfluid fraction $f_s$ tends to 0 in the thermodynamic limit.} \label{fig:SFfraction} \end{figure} \begin{table}[t] \centering \begin{tabular}{|cccc|cccc|} \hline \multicolumn{4}{c}{BH model} & \multicolumn{4}{c}{JCH model} \\ \hline $\mu/U$ & $t/U$ & $f_s$ & $\Delta f_s$ & $\tilde{\mu}$ & $\kappa/g$ & $f_s$ & $\Delta f_s$ \\ \hline 0 & 0.08 & 0.0007 & 0.0008 & -0.944 & 0.133 & -0.0020 & 0.0025 \\ 0 & 0.5 & -0.0001 & 0.0001 & -0.989 & 0.028 & -0.0014 & 0.0004 \\ & & & & -1.00 & 0.240 & 0.0000 & 0.0004 \\ \hline \end{tabular} \caption{Numerical values (with uncertainties) of superfluid fraction $f_s$ in thermodynamic limit, for phase space points indicated in Fig.~\ref{fig:SFfraction}.} \label{tab:fsTDL} \end{table} The finite-size scaling results for the superfluid fraction $f_s$ are shown in Fig.~\ref{fig:SFfraction}, and the resulting values and uncertainties of $f_s$ in the thermodynamic limit are displayed in Table~\ref{tab:fsTDL}. These values are determined by the standard least-squares minimization procedure for linear fitting. In all cases, except for the JCH point marked by red squares in Fig.~\ref{fig:SFfraction}, the obtained thermodynamic limit value of the superfluid density contains 0 within its error interval. For the one exceptional case, the thermodynamic limit $f_s$ is an order of magnitude smaller than the finite-size values (as well as unphysically negative), indicating that the superfluid fraction should vanish in the thermodynamic limit. It is interesting to note that the finite-size scaling plot for the BH point $(\mu/U,t/U)=(0.5,0.5)$ (not shown), which is situated above the particle boundary of the $n=1$ lobe, is identical to the scaling of the $(0.0, 0.5)$ point shown in Fig.~\ref{fsBH}. Similar behavior is found in the JCH model in the SF region to the right of the BKT point. Consider for example the point $(\tilde{\mu},\kappa/g)=(-0.750, 0.240)$ located above the $(-1.000, 0.240)$ point in Table~\ref{tab:fsTDL}. The infinite-system superfluid fraction inferred from finite-size scaling is $f_s=0.0006\pm 0.0033$, consistent with zero. This indicates that the superfluid density is only weakly dependent on the chemical potential for a given hopping strength. The data strongly suggest that the superfluid density is zero throughout the low-density SF region studied. \subsubsection{Condensate fraction} The condensate fraction $f_c$ is defined as the proportion of particles in the lowest-lying single-particle eigenstate of the system. In practice, this can be obtained by the largest eigenvalue of the single-particle density matrix $G^{(1)}(r,r^{\prime})$~\cite{Penrose1956,Penrose1951}. In fact the entire spectrum, known as the entanglement spectrum~\cite{Li2008b}, can be used to identify quantum phases. In the atomic limit, the single-particle ground state for a system of length $L$ is $L$-fold degenerate, since there is no preferred lattice site. In the non-interacting limit, the finite-size systems have a sinusoidal single-particle ground state. At zero temperature, one expects macroscopic occupation of the ground state in the SF regime, tending towards unit occupation for large $J/U$. In the MI regime one expects the particles to be distributed evenly across many nearly-degenerate single particle states, a phenomenon known as fragmentation for large occupation of a single site~\cite{Spekkens1999,Mueller2006}. If the particles have fermionized, however, then no such macroscopic occupation should occur; $f_c$ can therefore be viewed as another signature of fermionization. \begin{figure}[t] \subfigure[]{ \includegraphics[height=0.216\textwidth]{BECfragPhotonsL31-crop} } \subfigure[]{ \includegraphics[height=0.216\textwidth]{BECfractionJCHandBH-crop} } \caption{Maximal eigenvalues of the reduced single-particle density matrix for $L=31$. In (a), the five largest eigenvalues of the photon density matrix are shown for fixed $\tilde{\mu}=-0.678$ as a function of hopping amplitude $\kappa/g$. In (b), the largest eigenvalue of the density matrix (identified with the condensate fraction $f_c$ in the SF phase) for spin excitations and photons in the JCH model, and for bosons in the BH model, are shown as a function of generalized hopping $J/J_0$ at fixed chemical potential $\tilde{\mu}=-0.678$ and $\mu/U=0.55$, respectively. The inset shows the lower density case $\tilde{\mu}=-0.922$ and $\mu/U=0.133$ for comparison.} \label{fig:BECfraction2} \end{figure} The five largest eigenvalues of the single-particle photon density matrix are plotted in Fig.~\ref{fig:BECfraction2}(a) in the JCH model for $\tilde{\mu}=-0.678$ and $L=31$ as a function of hopping strength $\kappa/g$. This corresponds to the constant $\tilde{\mu}$ line near the very top of Fig.~\ref{fig:densities}(a). In the zero-hopping limit $\kappa/g\to 0$, each photon is perfectly localized to each site, and the eigenvalues are precisely $1/L$ (note that the density matrix is normalized to unity rather than the total number of particles). As $\kappa/g$ increases the degeneracy is broken and the largest eigenvalues increase due to the fluctuations of the site occupations, while others decrease to preserve the normalization (not shown). At a critical hopping strength $\kappa/g\approx 0.085$ coinciding with the superfluid transition at this value of $\tilde{\mu}$, one of the eigenvalues increases precipitously relative to the others, signifying the macroscopic occupation of a single mode. This eigenvalue is associated with the (quasi)condensate fraction. For larger hopping strengths the condensate fraction tends towards unity, as expected for a non-interacting Bose gas at zero temperature. Fig.~\ref{fig:BECfraction2}(b) compares the values of the condensate fraction for the BH bosons with the JCH photons and spins along the same line of constant chemical potential considered above, corresponding to $\mu/U=0.55$ in the BH model (recall that $\mu/U\sim 1.707(\tilde{\mu}+1)$ as discussed in Sec.~\ref{subsec:BH}). The value of $f_c$ for each model is plotted for fixed size $L=31$ as a function of the hopping $J$, where $J=\kappa$ ($t$) in the JCH (BH) case, in units of the minimum hopping considered $J_0$ corresponding to $\kappa_0/g=0.00556$ and $t_0/U=1.7\kappa/g=0.00948$. Though the onset of superfluidity occurs for smaller $J/J_0$ in the BH case, consistent with Fig.~\ref{fig:densities}, the condensate fraction does not increase as quickly as that of the photons in the JCH case. This might simply reflect the fact that the mean carrier density is lower for the BH model at the top right point in the phase diagram than for the JCH model, which would discourage condensation. At lower mean particle densities on the SF side $\tilde{\mu}=-0.922$ ($\mu/U=0.133$), shown in the inset of Fig.~\ref{fig:BECfraction2}(b), the $f_c$ for the BH case is larger than that for the photons of the JCH model, but neither reach 50\%. The finite-size results suggest that $f_c$ tracks the mean particle density, as is discussed further below. Interestingly, the spin excitations in the JCH model also show strong evidence of condensation. The value of $f_c$ on the SF side reaches approximately half that for the photons for the largest value of $J/J_0$ considered for this value of $\tilde{\mu}$. For lower values of $\tilde{\mu}$ in the vicinity of the $n=0$ Mott lobe, the ratio of $f_c$ for spin excitations to photons approaches unity (not shown). These results are generally consistent with observations above that indicate that the spin and photon degrees of freedom follow each other closely. The condensation in the spin sector therefore appears to be driven sympathetically by the photons via the polariton excitations. The results are nevertheless somewhat surprising, suggesting that the spin excitations are delocalized. To obtain the condensate fraction in the thermodynamic limit, the values of $f_c$ were obtained throughout the phase diagram for each triple $(\mu, J, L)$, where $L=15,19,23,27,31$ using DMRG subject to open boundary conditions. The finite-size scaling analysis was performed for the SF regime only because the procedure is not robust for the points in the MI regime. Nevertheless, the finite-size results within the MI regime clearly indicate fragmentation, and since the macroscopic degeneracy in the atomic limit is true independent of system size, there is no reason to believe that the results in the thermodynamic limit would be qualitatively different. The condensate fractions for JCH and BH models in the thermodynamic limit throughout the explored SF regime of the phase diagram are shown in Fig.~\ref{fig:BECfraction}(a) and (b), with the MI lobes explicitly zeroed out. Both pictures have the same color scale and are plotted with the axes corresponding to equivalent energy scales. The results are qualitatively similar for both models, but there are some quantitative differences. The value of $f_c$ rises more rapidly with increasing hopping in the JCH model than in the BH model, attaining $f_c=0.800$ at the highest mean densities investigated, as opposed to only $f_c=0.500$ for the BH model. The values of $f_c$ shown in the phase diagrams strongly resemble the mean densities of spins, photons, and bosons, such shown in Fig.~\ref{fig:densities}. Deep in the SF regime the $f_c$ closely follow the mean excitation densities, suggesting that the proclivity toward condensation is governed by the mean density. That said, as the Mott lobe boundary approaches the ratios of $f_c$ to the respective mean densities is found to increase markedly even as $f_c$ approaches zero. \begin{figure}[t] \subfigure[]{ \includegraphics[height=0.176\textwidth]{BECfractionPhotonsTDL2-crop} } \subfigure[]{ \includegraphics[height=0.176\textwidth]{BECfractionBHTDL2-crop} } \caption{Condensate fraction for the (a) JCH model (photons only) and the (b) BH model outside but near $n=1$ MI lobe, in the thermodynamic limit. In both cases, the condensate fraction within the MI lobes is zeroed out.} \label{fig:BECfraction} \end{figure} While there need be no direct relationship between Bose-Einstein condensation and superfluidity, it is nevertheless somewhat surprising that the condensate fraction in the thermodynamic limit would be so large throughout the SF region where the superfluid fraction remains zero. These results are nevertheless consistent with the small values of $\alpha$ in the SF region, shown in Fig.~\ref{fig:g1alphaJCH} (recall that $\alpha\to 0$ for Bose-Einstein condensates), as well as with the filling in of the exclusion hole and the disappearance of the Friedel oscillations found in the spatial-dependence of the two-body correlation function (c.f.\ Fig.~\ref{fig:g2BH}). In addition, the value of $f_c$ is consistently small in the low-density SF regime, as expected for a fermionized gas. \subsubsection{Entanglement properties} \label{subsubsec:entanglement} Much can be learned about the ground states of physical systems by examining the properties of subsystems. A notable example is the entropy of entanglement~\cite{Nielsen2000}. This is obtained by partitioning the system into a block of contiguous lattice sites $b \subset \mathcal{L}$, where $\mathcal{L}$ denotes the full lattice, and its complement $b^{\prime}$. The entropy of entanglement associated with this bipartition is given by \begin{equation} S_{\rho_b} = -\text{Tr}(\rho_{b} \text{ln}\rho_{b}), \end{equation} where $\rho_{b}$ is the reduced density matrix of the state over $b$. Analytical formulas for the entanglement entropy of non-interacting fermions and bosons on a lattice exist for the semi-infinite chain~\cite{Peschel2009} but we are not aware of any for finite-size systems greater than a few sites. The scaling of the entanglement entropy with the size of the subsystem is intimately linked with the utility of DMRG as a simulation method. For a wide array of physical systems, the entanglement entropy obeys an area law~\cite{Eisert2010}, meaning that the entanglement entropy $S_{\rho}$ associated with $\rho_b$ is proportional to the number of sites at the interface between $b$ and $b'$ (henceforth denoted $l$), rather than to its volume or cardinality. In one dimension, the area law corresponds to a value of $S_{\rho}$ that saturates for some finite value of $l$: \begin{equation} S_{\rho}(l) \leq D, \label{eq:arealaw} \end{equation} where $D$ is a constant, independent of $l$ (of course $l=2$ is itself constant). Only states satisfying Eq.~(\ref{eq:arealaw}) can be efficiently simulated by DMRG for large system sizes, because the variational ansatz used by the DMRG algorithm explicitly assumes that the area law is satisfied~\cite{Vidal2003,Eisert2010}. Asymptotic scaling results~\cite{Wolf2006} reveal that free fermions always logarithmically violate the area law in any dimension in both the continuum and on a lattice, whereas free bosons in 1D satisfy the area law away from criticality~\cite{Plenio2005}. The entanglement entropy associated with a finite block can be used to distinguish bosonic and fermionic behavior. The entanglement entropy of the 1D non-interacting Bose gas is a smooth function, whereas that for the non-interacting Fermi gas oscillates with $l$. This is because at zero temperature the bosons condense into the smooth lowest-lying eigenstate of the hopping model, while the Pauli exclusion principle forces fermions into oscillatory excited states. Generally, the entanglement entropy for bosons is larger than for fermions because the number of accessible states $\Omega_b$ is exponentially greater than $\Omega_f$, meaning that when the system is bipartitioned, for each configuration of the left half of the chain there are an exponentially larger number of compatible configurations for the right half in the bosonic case. \begin{figure}[t] \subfigure[][]{ \includegraphics[height=0.200\textwidth]{JCHentropylowdensity} \label{fig:JCHentropylow} } \subfigure[][]{ \includegraphics[height=0.208\textwidth]{JCHentropyhighdensity} \label{fig:JCHentropyhigh} }\\ \subfigure[][]{ \includegraphics[height=0.200\textwidth]{BHentropylowdensity} \label{fig:BHentropylow} } \subfigure[][]{ \includegraphics[height=0.208\textwidth]{BHentropyhighdensity} \label{fig:BHentropyhigh} }\\ \caption{The photon and boson entanglement entropy $S_{\rho}(l)$ for a contiguous block of sites of length $l=3$ to 29 with $L=31$, for the JCH and BH models, respectively. Two mean densities are considered, $n=0.645$ for the (a) JCH and (c) BH models, and $n=1.129$ for the (b) JCH and (d) BH models. The solid blue line is the exact value of $S_{\rho}^b$ for an ideal lattice Bose gas, and the solid red line in (a) and (c) is the corresponding value $S_{\rho}^f$ for the ideal Fermi gas.} \label{fig:entropyprofile} \end{figure} The entanglement entropy $S_{\rho}(l)$ is plotted in Fig.~\ref{fig:entropyprofile} as a function of $l$ for $3\leq l\leq 29$ ($L=31$) along a contour of constant density in the SF regime. The results are displayed for both the JCH and BH models at two different mean densities: $n=0.645$ and $n=1.129$. The corresponding entropy profiles for the free Bose and Fermi gases at the same mean density, calculated numerically, are plotted for comparison. The bosonic entanglement entropy increases approximately linearly with $l$ for $l\ll L$. The fermionic entanglement entropy displays strong oscillations but much weaker $l$-dependence. At the lower mean density considered, the photon entanglement entropy profile is very similar to the ideal Fermi gas for small hopping, again providing strong evidence for fermionization. As the hopping increases, the oscillations become less pronounced and the value of the entropy increases; presumably the entropy profile would converge to that of the ideal Bose gas for very large hopping amplitudes. For the $n>1$ case, which precludes fermionization in this single-band model, the oscillations are almost completely washed out. The magnitude of entanglement entropy remains well below the ideal Bose gas limit, however. While the entanglement entropy can be used to help distinguish quantum phases, it does not directly quantify the possible use of this entanglement for quantum computation. Indeed, it is difficult to conceive of how one could encode quantum algorithms into the (generally delocalized) indistinguishable bosons of the BH model. The JCH model, on the other hand, has distinguishable quantum registers in the spin states of the two-level atoms (i.e.\ qubits) which are each localized to a different optical cavity. If entanglement were generated between cavity atoms by the itinerant photons, then the coupled cavity QED could potentially be a natural environment for quantum computation. It was recently proven that universal quantum computation is possible as long as the entanglement associated with arbitrary bipartitions is non-vanishing~\cite{VanDenNest2013}. One useful measure that is closely related to the entanglement entropy is the localizable entanglement~\cite{Verstraete2004c,Verstraete2004b,Popp2005}. The localizable entanglement $E^{\text{LE}}_{ij}$ between qubits $i$ and $j$ of a multiqubit system is defined as the maximal entanglement that can be concentrated between qubits $i$ and $j$ via local (i.e.\ single-spin) operations and classical communication, and is a non-negative number in the range $\left[0,1\right]$. In a spin network, the localizable entanglement is lower-bounded by the maximal absolute value of the spin-spin correlation function $C^{\alpha,\beta}_{ij} :=\langle\sigma^{\alpha}_i\sigma^{\beta}_j\rangle - \langle\sigma^{\alpha}_i\rangle\langle\sigma^{\beta}_j\rangle$ over all possible axes $\alpha,\beta\in\field{R}^3$~\cite{Popp2005,Amico2008}. Using $\sigma^z_i=\sigma^+_i\sigma^-_i-I/2=n_i^{\rm spin}-I/2$ where $I$ is the $2\times 2$ identity matrix, and if $\alpha=\beta=z$, the localizable entanglement is at least as large as \begin{align*} \left | C^{zz}_{ij} \right | &\equiv \left | \langle N_i N_j\rangle -\langle N_i \rangle \langle N_j \rangle \right |, \end{align*} where here $N_k=\sigma^+_k\sigma^-_k$ is the local spin excitation number operator, satisfying $0\leq\langle N_k\rangle \leq 1$. The localizable entanglement is therefore closely related to the (unnormalized) two-body correlation function, defined in Eq.~(\ref{eq:G2}). For $i\neq j$ (the only case of interest for entanglement), one obtains $C^{zz}_{ij}=G^{(2)}(i,j)-\langle N_i\rangle \langle N_j\rangle\approx G^{(2)}(i,j)-n^2$ (the local excitation number is approximately equal to the mean density -- the total number of excitations over all sites -- if the density is almost constant). The behavior of the spin-spin correlation function in the thermodynamic limit can be estimated under the assumption that the spin excitations have completely fermionized, using Eq.~(\ref{eq:g2Tonks}) for a ring geometry or Eqs.~(\ref{eq:g2FF}-\ref{eq:g2FF2}) for open boundary conditions. Consider the symmetric pair of sites $i_\mp = (L\mp a)/2$, separated by $a$. Using either geometry one obtains $G^{(2)}(i_-,i_+) \rightarrow n^2 \sbrackets{1 - \sin^2 (\pi n a)/(\pi n a)^2}$, assuming short-ranged correlations in the bulk where $a$ remains constant as $L$ increases; for small $n$, the spin-spin correlation function approaches $-n^2$. If the separation instead scales like $a \sim L$, then $G^{(2)}(i_-,i_+) \rightarrow n^2$ and $C^{zz}_{i_-i_+} \rightarrow 0$. The correlation function $C^{zz}_{ij}$ was calculated throughout the phase region investigated. Restricting $i$ and $j$ to lie between $\lceil L/4 \rceil$ and $\lceil 3L/4 \rceil$ to avoid boundary effects, $\left|C^{zz}_{i,j}\right|$ take the largest values when $|i-j|=1$ ($i=j$ is explicitly excluded). The values of $C^{zz}_{i,i\pm 1}$ were found to be close to zero everywhere in the MI phase, consistent with the strong density localization and small density fluctuations which are the hallmark of MI states. Likewise, the values of $C^{zz}_{i,i\pm 1}$ tend to zero for all $i$ at high mean densities and hopping where the BEC fraction is large. This reflects the smooth and relatively constant profile of the two-body correlation function other than the remnant of the exclusion hole at short distances, as shown in Fig.~\ref{fig:g2BH}. At low density and hopping amplitude in the SF regime, the magnitudes of $C^{zz}_{i,i\pm 1}$ generally range between $0.02$ to $0.05$. This is a direct consequence of the large exclusion hole in the spin density-density correlation function in the vicinity of $i\sim j$, inherited from the strongly fermionized polaritons. The spin-spin correlation function reaches a maximum value in the region directly adjacent to the hole boundary of the $n=1$ Mott lobe. At the point $(\tilde{\mu},\kappa/g)=(-0.989, 0.0167)$ where the mean spin excitation density is $n=0.3825$, the numerics yield $\left|C^{zz}_{i,i\pm 1}\right| =0.075$. The fermionized theory predicts the comparable but slightly larger value of $\sin^2(\pi n)/\pi^2\approx 0.088$. The difference between theory and computation is likely partly due to the fact that the spin excitations have not perfectly fermionized. Because we have only considered the $zz$ quadrature of the spin-spin correlation function, the numerical value is a lower-bound to the lower-bound of the localizable entanglement. The actual value of the localizable entanglement in the Tonks-Girardeau regime could well be larger. In any case, the numerical results suggest that there is sufficient entanglement between the atoms in the ground state of coupled cavities to support universal quantum computation, if a suitable strategy to embed this environment in a quantum circuit could be found. \section{Conclusions} \label{sec:conclusions} In this work, we have explored the Mott-Insulator to superfluid transition of the one-dimensional Jaynes-Cummings-Hubbard (JCH) model in the strong-coupling regime with no detuning. The purpose of the work has been two-fold. First and foremost, it has been to study the nature of the ground state in the vicinity of the phase transition, in particular to demonstrate that the photons are in fact strongly fermionized in the low-density `superfluid' phase. Second, it has been to compare and contrast the properties of the ground state to that of the 1D Bose-Hubbard (BH) model, in order to highlight the unique features of the JCH model. The results were obtained by finding the ground state using the finite-system Density Matrix Renormalization Group, computing various static properties, and then performing a finite-size scaling analysis to infer the thermodynamic limit. The main result is that in one dimension, the ground state of the JCH model in the low-density regime outside the Mott-insulating lobes is dramatically different from that of a conventional superfluid. Rather, the system in the region widely characterized as a superfluid (SF) is in fact a Tonks-Girardeau gas of strongly fermionized excitations, much as occurs in the one-dimensional BH model. This is evidenced by the power law of the single-particle density matrix in both the spin and photonic sectors of the model, by the Friedel oscillations and the fermionic exclusion hole in the two-body density matrix for each sector, and by the strongly fermionic profile of the entanglement entropy. Thus coupled cavity QED provides a natural and accessible environment for the realization of strongly correlated photons. The photon fermionization should be readily observable in experiments using standard photon correlation spectroscopy~\cite{Carmichael1996}. We have calculated the superfluid fraction for both species of excitation in the JCH model, both within the low-density SF regime as well as for large tunneling beyond the expected BKT transition point, and found that the superfluid density vanishes in the thermodynamic limit in both cases. At the same time, we have found that the Bose-Einstein condensate fraction for the photons and spin excitations is non-vanishing throughout the SF region. The value of the condensate fraction is low at very low densities, consistent with fermionization, but can reach as high as 80\% at high densities for the photons. This indicates the existence of a (quasi)condensate of photons and even spin excitations, despite the absence of superfluidity. The same static properties for the 1D BH model have been calculated in an equivalent parameter regime in phase space, and the results have been compared in detail to those obtained within the JCH model. Broadly, the behavior of the two models coincide: neither exhibits a true Mott-insulator to conventional superfluid transition. However, various quantitative differences exist. The single-particle correlations in the JCH ground state within the $n=1$ Mott lobe decay more rapidly with increasing hopping that in the BH case, while those in the intermediate region between the $n=0$ and $n=1$ lobes decay more slowly. Consistently with this result, the condensate fraction for the JCH photons rises more rapidly with increasing hopping than for the BH bosons. While a formal mapping between the BH and JCH models does not exist, the results indicate that the well-known manifestations of the BH model will be largely reproduced in physical systems that are well-described by the JCH model. That said, the JCH model has some intriguing features not shared by the BH model, owing to the presence of two species. The spin and photon degrees of freedom are inextricably linked through the fundamental polariton excitations. Thus Bose-Einstein condensation of photons implies that the spin excitations are similarly condensed. The atomic spin states are thereby effectively delocalized across the entire system in spite of the fact that each atom is confined to its respective cavity. This observation opens the intriguing possibility of inducing spin liquid-like states in cavity QED systems. In fact, the possibility of spin dimerization in the JCH model with large positive detuning $\delta\gg 0$ (which induces frustration and is not considered in the present work) has been noted very recently~\cite{Zhu2013}. Likewise, atoms in different cavities are spontaneously entangled via the itinerant photons, as evidenced by the non-zero value of the localizable spin entanglement in the SF regime. It would be intriguing to systematically consider the effect of non-zero detuning on the properties of the ground states, but this is beyond the scope of the present work. The majority of the parameter space explored in this project is in the vicinity of a phase transition, but the convergence of the DMRG algorithm is not guaranteed very close to the phase boundary. It could be fruitful to compare the results with those obtained using a method such as Multiscale Entanglement Renormalization Ansatz~\cite{Vidal2008}, which is specifically tailored to work well with critical systems. In a similar vein, the calculations (in particular the superfluid density) could be repeated with a method that is designed for handling infinite systems directly, such as iDMRG~\cite{McCulloch2008,Crosswhite2008}. This would avoid the need to perform finite-size scaling on small systems to make quantitive statements about the thermodynamic limit. Cavity quantum electrodynamics is a promising candidate for quantum information processing applications, since the local atoms can be used as qubits and the photons can be used to generate entanglement between them. Our results suggest that the localizable entanglement is always finite throughout the SF region. It would be interesting to determine if the ground state for a more complex network of coupled cavities could be a resource for measurement-based quantum computation, where universal quantum algorithms are effected solely via single-qubit measurements conditioned on previous outcomes~\cite{Briegel2009}. Preliminary calculations indicate that the type of correlations between atoms in the current 1D JCH model with zero detuning are probably not suitable for gate teleportation via measurements. This will be explored more fully in future work. \section{Acknowledgements} \label{sec:acknowledgements} The authors are grateful for research funding from the Natural Sciences and Engineering Research Council of Canada, Alberta Innovates -- Technology Futures, and the Canadian Institute for Advanced Research. This research has been enabled by the use of the ALPS software package, particularly the {\tt dmrg} application, as well as by computing resources provided by WestGrid and Compute/Calcul Canada. \bibliographystyle{apsrev} \def{} \def\url#1{}
\section{Introduction} Since its first demonstration over two decades ago, electromagnetically induced transparency (EIT) \cite{Boller1991} has developed into one of the most popular spectroscopic methods in atomic physics \cite{Fleishauer} and has also been applied to molecular \cite{Benabid2005} and solid state physics \cite{Liu2009, Safavi-Naeini2011}. EIT can allow the observation of sub-natural linewidth spectral features, even in room temperature thermal vapours. Although precision spectroscopy is best suited to ultracold gases, for example in optical lattice clocks \cite{Nicholson12}, the apparatus required for such experiments is somewhat bulky and inconvenient when it comes to real world applications. In addition, many atomic species, and most molecules, cannot be laser-cooled. Consequently, devices based on thermal vapours are desirable. In this respect, the commercially available chip scale atom clock SA.45S from Symmetricom \cite{Symmetricom} presents a technological paradigm. The portfolio of fully integrated atomic devices could be expanded to frequency references \cite{Haensch71,Bell,Abel}, Faraday rotators \cite{Siddons09}, slowing and storing light \cite{Phillips01}, optical filters \cite{Menders91}, and many new optical functionalities when combined with micro-optical components \cite{Stern12}. Vapour cell experiments have also recently allowed the demonstration of the cooperative Lamb-shift \cite{Keaveney2012} and extreme dispersion \cite{Keaveney2012a}. A special class within this field is added by the use of highly excited Rydberg states, whose large polarizability and strong interactions allow new features to thermal vapour cell experiments such as e.g. electric field sensors \cite{Daschner2012}, microwave detectors \cite{Sedlacek2012}, electro-optic modulation \cite{Mohapatra2008} and possibly as a single photon source \cite{Dudin2012, Baluktsian2012}. Most of the examples listed above rely on two color excitation schemes, where the manifold of electronic levels allow for various excitation pathways. Although the majority of work on EIT has used $\Lambda$-systems which take advantage of two stable ground states, it is also possible to extend EIT to a ladder system under certain conditions. Since EIT relies on the coherent superposition of two excitation pathways, it is favourable that the upper state lives sufficiently long, as given for high-lying Rydberg states \cite{Mohapatra}, but it works also with short-lived states \cite{Gea-Banacloche1995}. The exaggerated properties of Rydberg states \cite{LoewReview} opens new and exciting opportunities for non-linear quantum optics \cite{PritchardReview} particularly at the single-photon level \cite{Dudin2012,Peyronel,Maxwell} \begin{figure}[b] \includegraphics[width=3in]{figure1.pdf} \caption{(a) Energy level scheme in caesium used experimentally. ${\rm n^{\prime}P, nS}$ are ${\rm 6P, 7S}$ and ${\rm 7P, 32S}$ for Section~\ref{section:Durham} and ~\ref{section:Stuttgart} respectively. ${\rm \Omega_{\rm p,c}}$, ${\rm \Delta_{\rm p,c}}$ and ${\rm \lambda_{\rm p,c}}$ are the Rabi frequencies, detunings and wavelengths of the probe and coupling laser field respectively. ${\rm \Gamma_{n^{\prime}P}}$ and ${\rm \Gamma_{nS}}$ are the decay rates of the excited states. (b) Energy level scheme used in the theoretical model. ${\rm \Gamma}_i $ is the decay rate of the state $| i \rangle $. a, b and c are the branching ratios from the three upper states ${\rm | 3 \rangle}$, ${\rm | 2 \rangle} $ and the loss state ${\rm | L \rangle}$ respectively. } \label{fig:fig1} \end{figure} \begin{figure*}[t] \includegraphics[width=6in]{figure2.pdf} \caption{Wavelength dependence of the EIT. Top: Absorption coefficient per velocity class for (a) $\lambda_{\rm p}=\frac{4}{3} \lambda_{\rm c}$ ($\lambda_{\rm p}>\lambda_{\rm c}$) (b) $\lambda_{\rm p}=\lambda_{\rm c}$ and (c) $\lambda_{\rm p}=\frac{2}{3} \lambda_{\rm c}$ ($\lambda_{\rm p}<\lambda_{\rm c}$). Bottom (d)-(f): Doppler-averaged probe laser transmission signal corresponding to (a)-(c). Simulation parameters: $\Omega_{\rm p}=2\pi\times 0.01$~MHz, $\Omega_{\rm c}=2\pi\times 5$~MHz, $\Gamma_{1}=2\pi\times 5$~MHz, $\Gamma_{2}=2\pi\times 0.1$~MHz, ${\rm T}=20^{\circ}$C and $2$~mm cell length.} \label{fig:fig2} \end{figure*} The main focus of this work is to consider the wavelength ratio of the two excitation lasers used in ladder EIT \cite{Shepherd1996} and the effects of open decay channels in multi-level atoms. To obtain a detailed understanding of all possible dependencies, we perform spectroscopy on two different ladder schemes in caesium, which mainly differ in the lifetime of the upper state. When the upper excited state is short-lived, we find a competition between two-photon absorption and optical pumping which depends on the atom-laser interaction time. When the excited state is long-lived, we find that optical pumping effects are negligible and EIT remains observable as long as the decoherence mechanisms are minimized. In an EIT-ladder scheme, where the probe beam couples the two lower states and the coupling beam the two upper states, we will sometimes encounter a situation where the probe wavelength $\lambda_{\rm p}$ is smaller than the coupling wavelength $\lambda_{\rm c}$ ($\lambda_{\rm p}<\lambda_{\rm c}$) \cite{Carr2011, Carr2012a, Moon12}, what we call an ``inverted'' excitation scheme, since most published work on EIT in ladder systems employs light fields $\lambda_{\rm p}>\lambda_{\rm c}$. Such excitation schemes sometimes present a technological advantage as e.g. in the case of Rydberg states, where the small dipole matrix element can be matched by powerful and convenient infrared lasers. Let us first consider a simple 3-level ladder system, as shown in Figure~\ref{fig:fig1}~(b) with only the levels ${\rm | 1 \rangle}$, ${\rm | 2 \rangle} $ and ${\rm | 3 \rangle}$, ${\rm | 1 \rangle} \rightarrow {\rm | 2 \rangle} $ being the probe transition and ${\rm | 2 \rangle} \rightarrow {\rm | 3 \rangle} $ the coupling transition. A numerical solution of the optical Bloch equations for such a system, including the Doppler broadening, is depicted in Figure~\ref{fig:fig2} for three different wavelength ratios. In the upper panel, the absorption of the probe laser is shown as a function of atomic velocity and coupling laser detuning for the wavelength ratios (a) $\lambda_{\rm p}>\lambda_{\rm c}$ (b) $\lambda_{\rm p}=\lambda_{\rm c}$ and (c) $\lambda_{\rm p}<\lambda_{\rm c}$. In the lower panel, the Doppler averaged probe transmission signals, obtained by summing over the Gaussian-weighted signals for each atomic velocity class, are shown. In the case that $\lambda_{\rm p}<\lambda_{\rm c}$, the EIT feature is significantly smaller. This occurs because the one-photon Doppler shift $\Delta_{\rm 1ph}=-k_{p}v$ has the same sign as the two-photon Doppler shift $\Delta_{\rm 2ph}=-(k_{p}-k_{c})v$. As a result, the transparency window as a function of detuning no longer exists and adding the contribution from each velocity class strongly reduces the probe transmission peak compared to the non-inverted case. For real atoms the transitions in the three-level ladder system are not necessarily closed. Decay to other levels outside the three-level system will reduce the coherence time and by this the visibility of the EIT transparency window, and also introduce optical pumping effects. For instance when the upper transition is no longer closed, atoms can decay to the lower hyperfine level of the ground state via alternative decay channels. This process is known as Double Resonance Optical Pumping (DROP) \cite{Moon2004} and is a time-dependent transfer of atoms between the hyperfine levels of the ground state via a two-photon excitation process. Extensive studies of the interplay between EIT and DROP have been performed in a few non-inverted 3-level ladder system with theoretical models similar to that presented below (\cite{Hayashi10, Moon11, Noh11}). In this paper we demonstrate experimentally and theoretically that in an ``inverted''-wavelength system, the interplay between these coherent and incoherent processes is strongly dependent upon the fine balance between state lifetimes, atom-laser interaction time, defined by the size of the laser beams and the velocities of the atoms, and the laser linewidths. The interplay of all these conditions leads to significant modifications of the optical signal observed. The experimental three-level ladder system in caesium that we use is shown schematically in Figure~\ref{fig:fig1}~(a). The ground state 6S$_{1/2}$ is coupled to the intermediate state n$^{\prime}$P$_{3/2}$ by the probe laser with Rabi frequency $\Omega_{\rm p}$ and wavelength $\lambda_{\rm p}$. This intermediate state is then coupled to the excited state nS$_{1/2}$ by the coupling laser with Rabi frequency $\Omega_{\rm c}$ and wavelength $\lambda_{\rm c}$. \section{Model System} \label{section:Model} All levels and decays observed in the real level scheme (Figure~\ref{fig:fig1}~(a)) can be captured using a simplified model including 5 levels, as shown in Figure~\ref{fig:fig1}~(b). The actual states \mbox{$|6S_{1/2}, F=4 \rangle$}, \mbox{$|n^{\prime}P_{3/2}, F'=3,4,5 \rangle$} and \mbox{$|nS_{1/2}, F''=4 \rangle$} are represented by the states $|1\rangle$, $|2\rangle$ and $|3\rangle$ respectively. The uncoupled lower hyperfine level of the ground state 6S$_{1/2}, F=3$ is included as state $|0\rangle$. Finally, indirect decay from the intermediate and excited states to the ground states is taken into account through the addition of a loss state $|{\rm L}\rangle$. For instance, the decay \mbox{$|7S_{1/2}, F''=4 \rangle $ $\rightarrow$ $| 6P_{3/2}, F'=4 \rangle$ $\rightarrow$ $|6S_{1/2}, F=3 \rangle $} is represented by \mbox{$|3\rangle$ $\rightarrow$ $|L\rangle$ $\rightarrow$ $|0\rangle$} in the model. Using standard semi-classical methods \cite{cohen1992atom} we numerically solve the Liouville von Neumann master equation in steady-state for the density matrix $\rho$, which includes the populations and coherences of the five-level model. Due to thermal motion, the Doppler effect results in an effective detuning $\delta_{\rm Doppler}$ = $-\vec{k}\cdot \vec{v}$, where $k = 2\pi/\lambda$ is the wave vector of the laser field and $v$ the atomic velocity. As we use a co-linear beam configuration, this can be included in the model by averaging over the one-dimensional Gaussian velocity distribution. The finite interaction time of the atoms with the laser beam is included as a transit time decay. This is due to the component of the atomic motion orthogonal to the laser beam. Simply speaking, we re-distribute the atomic population of all states between the hyperfine levels of the ground state with a rate $\Gamma_{\rm tt}$ (\cite{Sagle1996}) with: \begin{equation} \Gamma_{\rm tt} = \frac{1}{a\sqrt{2\log2}}\sqrt{\frac{8k_{\rm B}T}{\pi m}} \end{equation} where $a$ is the $1/e^2$ beam waist, $T$ is the temperature and $m$ is the mass of the atom. Finally, we include laser-induced dephasing in the model. The finite linewidth of the laser can be modelled as a dephasing which acts on the off-diagonal elements of the density matrix $\rho$ \cite{Sultana1994, Gea-Banacloche1995}. \section{Short-lived upper state} \label{section:Durham} \begin{figure}[t] \includegraphics[width=3.3in]{figure3.pdf} \caption{$6S-6P-7P$ experiment: Change in probe transmission as a function of probe beam waist $w_{\rm p}$ for the excitation with $\lambda_{\rm p} = 852$~nm and $\lambda_{\rm c} = 1470$~nm. Parameters: $\Omega_{\rm p}=2 \pi \times 8$~MHz, $\Omega_{\rm c}=2 \pi \times 22$~MHz and $w_{\rm c}=20$~$\mu$m.} \label{fig:fig3} \end{figure} First, we study a three-level system with ``inverted''-wavelengths where the upper level is a short-lived excited state. Experimentally, we realize this situation by tuning the the coupling laser to the $| 6P_{3/2} \rangle \rightarrow | 7S_{1/2} \rangle $ transition. The $| 7S_{1/2} \rangle $ state has a natural linewidth of $\Gamma_{\rm 7S}=2\pi \times 3.3$~MHz comparable to the intermediate state linewidth of $\Gamma_{\rm 6P}=2\pi \times 5.2$~MHz. A circularly polarized 852 nm probe beam, stabilized to the $|6S_{1/2}, F=4 \rangle $ $\rightarrow$ $| 6P_{3/2}, F'=5 \rangle$ transition, passes through a caesium vapour cell. A counter-propagating circularly polarized 1470~nm coupling beam is scanned across the $| 6P_{3/2}, F'=5 \rangle $ $\rightarrow$ $| 7S_{1/2}, F''=4 \rangle $ transition. In order to investigate the atomic dynamics on different timescales, we use tightly focused beams in a 100~$\mu$m long vapour cell. The beam waist of the coupling beam is kept constant at $w_{\rm c}=20$~$\mu$m, and the beam waist of the probe beam $w_{\rm c}$ is varied from 17~$\mu$m to 49~$\mu$m. The beam sizes are approximately constant over the length of the atomic sample as the Rayleigh range is longer than the length of the cell. For 100$^{\circ}$C vapour with most probable velocity $v_{\rm p}=215$~m/s and typical beam size $w_{\rm p}=w_{\rm c}=20$~$\mu$m, the most probable interaction time is $t_{\rm p}\approx90$~ns. This interaction time is comparable to the lifetime of both the intermediate state $\tau_{6P}=30.4$~ns and the excited state $\tau_{\rm 7S}=48.2$~ns. First we consider the effect of applying only the probe laser on resonance $\Delta_{\rm p}/2 \pi=0$~MHz. Atoms with velocities around $v=0$~m/s are excited on the $F=4$ $\rightarrow$ $F'=5$ transition. As this is a closed transition, the atoms cannot enter the lower hyperfine level of the ground state and cycle on this transition until saturation is achieved. Due to the one-photon Doppler shift of the probe laser $\Delta_{\rm 1ph}=-k_{\rm p}v$, non-zero velocity classes around $v=213.85$~m/s and $v=385.10$~m/s are able to excite the $F=4$ $\rightarrow$ $F'=4$ and $F=4$ $\rightarrow$ $F'=3$ transitions respectively. As these transitions are open, atoms can enter the lower hyperfine level and become dark with respect to the probe laser. Now we consider the effect of applying both the probe and coupling laser simultaneously. The change in probe transmission as a function of coupling detuning is investigated for various beam waists and therefore atom-laser interaction times, as shown in Figure~\ref{fig:fig3}. When the coupling laser is on resonance $\Delta_{\rm c}/2 \pi=0$~MHz, a transparency feature is observed. This corresponds to the excitation of atoms with velocities around $v=0$~m/s on the $F=4$ $\rightarrow$ $F'=5$ $\rightarrow$ $F''=4$ two-photon transition. The probe transition is no longer closed and atoms can now decay to the lower hyperfine level of the ground state via hyperfine levels of the intermediate state. This process is known as Double Resonance Optical Pumping (DROP) \cite{Moon2004} and is a time-dependent transfer of atoms to the dark hyperfine state via a two-photon excitation process. For small beam sizes, and therefore short interaction times, the transparency feature at $\Delta_{\rm c}/2\pi=0$~MHz becomes absorptive. When the interaction time is sufficiently short, the atom does not remain in the beam long enough to decay from the excited and intermediate state. As a result, the DROP transparency feature does not occur. Instead, in the very short interaction time domain $w_{\rm p}<30$~$\mu$m, the probe light is absorbed on resonance because atoms are being transferred to the upper state. However, for all interaction times, the feature at $\Delta_{\rm c}/2\pi=105$~MHz remains absorptive. This detuning corresponds to the two-photon detuning $\Delta_{\rm 2ph}=-(k_{\rm p}-k_{\rm c})v$ required to be resonant with the off resonant velocity class excited by the probe laser to the intermediate hyperfine level $F'=4$. When the coupling laser is applied to these open transitions, the decay routes to the lower hyperfine ground state are reduced. This is because atoms can now be excited through $F''=4$ to $F'=5$ which can only decay to the upper hyperfine ground state. As a result, the `openness' of the open transitions is reduced and the probe is absorbed more when the coupling laser is applied. \begin{figure}[t] \includegraphics[width=3.3in]{figure4.pdf} \caption{$6S-6P-7S$ theory: Change in probe transmission as a function of beam waist $w_{\rm 0}$ with $\lambda_{\rm p} = 852$~nm and $\lambda_{\rm c} = 1470$~nm when the coupling laser is on resonance. Inset: Change in probe transmission as a function of coupling laser detuning for two beam waists. Parameters: $\Omega_{\rm p}=2\pi \times8$~MHz, $\Omega_{\rm c}=2 \pi \times 22$~MHz, $a=0.6$ and $c=0.5$.} \label{fig:fig4} \end{figure} This interaction-time dependence of the transmission spectrum is fully reproduced by our model system, where the finite interaction time is included via the transit-time decay rate $\Gamma_{\rm tt}$. Figure~\ref{fig:fig4} shows the calculated probe transmission on resonance for the $F'=5$ intermediate state ($\Delta_{\rm c}/2\pi=0$~MHz), reproducing the change from absorptive for beam sizes $w_{0}<50$~$\mu$m to transmissive for larger beams. The exact position of the frontier between the two regimes in the simulation strongly depends on the Rabi frequencies. We attribute the discrepancy on the position of this frontier ($20$~$\mu$m experimentally versus $50$~$\mu$m theoretically) to the assumption of a constant Rabi frequency in the model, whereas in reality it is Gaussian-shaped. Our model also reproduces the behaviour of the other feature at $\Delta_{\rm c}/2\pi=105$~MHz ($F'=4$ intermediate state). In this system EIT is still present but strongly reduced by the decay from the upper excited state and completely eclipsed by optical pumping effects. We find that there are two distinct time domains defined by the lifetimes of the intermediate and upper state. The response of the atomic system to the probe field is significantly different in these two regimes. For long interaction times optical pumping prevails, whereas for short interaction times, two-photon absorption dominates. \section{Long-lived upper state} \label{section:Stuttgart} \begin{figure}[b] \includegraphics[width=3in]{figure5.pdf} \caption{$6S-7P-32S$ experiment: Change in probe transmission for two different probe Rabi frequencies $\Omega_{\rm p} = 2\pi \times 0.54$~MHz (blue) and $\Omega_{\rm p} = 2\pi \times 8.6$~MHz (green) with $\lambda_{\rm p} = 455$~nm and $\lambda_{\rm c} = 1070$~nm. The coupling Rabi frequency is $\Omega_{\rm c}=2\pi \times 3.0$~MHz. The features at zero detuning corresponds to the intermediate state $|7P_{3/2}, F'=5 \rangle$, discussed here. The other two at $48$~MHz and $86$~MHz correspond to $F'=4$ and $F'=3$ respectively. Note that due to the measurement technique, no absolute transmission scale could be generated. Therefore the amplitude of the two traces are arbitrary and cannot be compared to one another.} \label{fig:fig5} \end{figure} \begin{figure}[t] \includegraphics[width=3in]{figure6.pdf} \caption{$6S-7P-32S$ Theory: Change in probe transmission for the same Rabi frequencies as in Figure~\ref{fig:fig5} for the $|7P_{3/2}, F'=5 \rangle$ intermediate state, with $\lambda_{\rm p} = 455$~nm and $\lambda_{\rm c} = 1070$~nm. The double structure around the resonance is a consequence of the Autler-Townes splitting by the probe laser. Parameters: $\Omega_{\rm p}=2\pi \times 0.54$~MHz and $2\pi \times 8.6$~MHz, $\Omega_{\rm c}=2\pi \times 3.0$~MHz, $w=1$~mm, $T=60^\circ$C, $a=0$, $b=0.55$ and $c=0.5$.} \label{fig:fig6} \end{figure} Next, we discuss the spectroscopy of a three-level ladder system with ``inverted''-wavelengths ($\lambda_p < \lambda_c$) where the upper level is a long-lived Rydberg state. The $|6S_{1/2}, F=4 \rangle \rightarrow |7P_{3/2}, F'=3,4,5 \rangle$ transition represents the probe transition and the $|7P_{3/2}, F'=3,4,5 \rangle \rightarrow |32S_{1/2} \rangle$ transition the coupling transition (see Figure~\ref{fig:fig1}). Such an excitation scheme is advantageous compared to standard non-inverted schemes (used for example in \cite{Mohapatra2008}) since the excitation to the Rydberg state at around $1070$~nm lies in the range of fibre lasers. The much higher output power, compared to diode lasers or Ti:Sa lasers, compensates for the small dipole matrix element and one can achieve coupling strengths comparable to the lower $6S_{1/2} \rightarrow 7P_{3/2}$ transition. Experimentally, both beams are linearly polarized with orthogonal polarizations and are overlapped in the vapour cell in a counter-propagating configuration. The $455$~nm probe beam is stabilized to the $|6S_{1/2}, F=4 \rangle \rightarrow |7P_{3/2}, F'=5 \rangle$ transition and the $1073$~nm coupling laser is scanned over the $|7P_{3/2}, F'=3,4,5 \rangle \rightarrow |32S_{1/2} \rangle$ transitions. As discussed before, the wavelength mismatch of our configuration dramatically reduces the visibility of the EIT peak. To observe a sufficiently large spectroscopy signal lock-in amplification was used. The coupling beam was intensity-modulated using an acousto-optic modulator (AOM) chopped at $20$~kHz. Additionally, the demodulated probe signal was averaged over 32 traces. The cell had a length of $7.5$~cm and was heated to $60^\circ$C. Both beams were collimated with beam waists of $1.5$~mm for the $1073$~nm laser and $1.0$~mm for the $455$~nm laser. Two experimental probe transmission spectra are shown in Figure~\ref{fig:fig5}. Three structures are visible and can be identified with the three dipole-allowed hyperfine transitions $|6S_{1/2}, F=4 \rangle \rightarrow |7P_{3/2}, F'=3,4,5 \rangle$, where the lasers address three velocity classes fulfilling the two-photon resonance condition. A striking feature is the change of transparency to absorption by just changing the intensity of one of the two laser beams. In the following, we will focus on the excitation via the $F'=5$ intermediate state, but our results can be easily extended to the other allowed hyperfine intermediate states. \begin{figure}[b] \includegraphics[width=3in]{figure7.pdf} \caption{$6S-7P-32S$ theory: Foreground: qualitative representation of the experimental change in probe transmission on resonance. Background: theoretical change in probe transmission over the Rabi frequencies $\Omega_{\rm p}$ and $\Omega_{\rm c}$, with $\lambda_{\rm p} = 455$~nm and $\lambda_{\rm c} = 1070$~nm. Parameters: $w=1$~mm, $T=60^\circ$C, $a=0$, $b=0.55$ and $c=0.5$.} \label{fig:fig7} \end{figure} These enhanced transmission (EIT) and enhanced absorption (EA) features are fully reproduced by our 5-level model. Simulated probe transmission traces with the same peak Rabi frequencies as in Figure~\ref{fig:fig5} are shown in Figure~\ref{fig:fig6}. The splitting in the blue curve around the resonance frequency has its origin in an Autler-Townes splitting of the intermediate state induced by the probe beam. The absence of the feature in the experimental spectra might be explained by the fact that the atoms cross the beam, therefore experiencing the Gaussian-distributed Rabi frequencies over the interaction time. The decay rates and branching ratios were matched to the real values, and the transit time effect was included for the corresponding temperature and beam size. A $100$~kHz laser induced dephasing was included, matching our laser linewidths on the timescale of our experiment. We found that including a laser induced dephasing was crucial in order to understand the experimental data. Indeed any decoherence effect reduces the EIT, which is already weak because of the wavelength ratio. We studied the dependence of the transmission signal on resonance (EIT or enhanced absorption) on both probe and coupling Rabi frequencies. The theoretical and experimental results are combined in Figure~\ref{fig:fig7}. The model system reproduces the experimentally observed transition from EIT to EA very well. We attribute the little discrepancy between theory and experiment to the fact that the exact position of this crossover in the simulation was found to be strongly dependent on various parameters, such as the laser linewidths, transit time and branching ratios. The experimental parameters were used as input parameters for the simulations in Figure~\ref{fig:fig7}. Since the upper state is a long-lived Rydberg state with ${\rm 27 \mu s}$ lifetime, optical pumping effects from the upper state like DROP are negligible. The lower transition is however not closed, since the atoms can decay from the $|7P_{3/2} \rangle$ state to the $|7S_{1/2} \rangle$ and $|5D_{3/2, 5/2} \rangle$ states. From our simulations we observe that all of the following conditions have to be met, to reproduce the experimentally observed transition from EIT to EA: (i) the wavelengths are mismatched such that $\lambda_p < \lambda_c$, (ii) a laser induced dephasing is included, (iii) the first transition is open. Otherwise there exists no region of enhanced absorption and only EIT remains. Because of the ``inverted'' wavelength configuration, the EIT is a tiny feature, therefore extremely sensitive to other phenomena that are otherwise negligible. In particular for our system, having an open transition on the first excitation step acts as a decoherence on the system, simultaneously increasing the absorption and ``washing out'' coherent phenomena such as EIT. The finite laser linewidth further reduces the EIT signal, as well as it allows for more absorption around the resonance because of line broadening. \section{Conclusion} To conclude, we have shown that EIT can be observed in ladder systems with ``inverted'' wavelengths. However, it is strongly affected by various incoherent effects due to additional decay channels and optical pumping, and EIT is very likely not to be the dominating effect. With our effective model based on five levels we were able to reproduce the experimentally observed features of two different physical implementations, which was not possible if only four states were included. Therefore we believe that this model will serve as a valuable design tool for all kinds of experiments involving ladder schemes. To date, very few studies have been performed involving ``inverted'' 3-level ladder schemes. Our results show that it is possible to obtain a good spectroscopic signal from these systems. These methods could be used for example as frequency references for further experiments, or as a spectroscopic tool. \section*{Acknowledgments} The authors would like to thank T.~Pfau, H.~K\"{u}bler, S.~Hofferberth, J.~Keaveney, U.~M.~Krohn and I.~G.~Hughes for fruitful discussions, as well as C.~Veit for experimental input. This project was supported by the Carl-Zeiss-Stiftung and the UK Engineering and Physical Sciences Council. RL is indebted to the Baden-W\"{u}rttemberg Stiftung for the financial support of this research by the Eliteprogramme for Postdocs. RR acknowledges financial support from the Landesgraduiertenf\"{o}rderung Baden–W\"{u}rttemberg. KJW acknowledges financial support from Durham University.
\section{Introduction} The quantum enveloping algebra of the loop algebra of $\mathfrak {gl}_n$, or simply quantum affine $\mathfrak {gl}_n$, has two usual definitions, the $R$-matrix one and the Drinfled one, known as Drinfeld's new realisation. Both are presented by generators and relations (see, e.g., \cite[\S2.3]{FM} and the references therein). In \cite[2.3.1,2.5.3]{DDF}, a third presentation is given via the double Ringle--Hall algebra. In this presentation, the Ringel--Hall algebra of the cyclic quiver and its opposite algebra become the $\pm$-part of quantum affine $\mathfrak {gl}_n$. Thus, with this construction, one may consider semisimple or indecomposable generators for quantum affine $\mathfrak {gl}_n$, defined by the semisimple or indecomposable representations of the quiver; see \cite[\S1.4]{DDF}. In particular, one sees easily the fact that the subalgebra generated by simple generators is a {\it proper} subalgebra. This subalgebra is isomorphic to the quantum affine $\mathfrak {sl}_n$. The double Ringel--Hall algebra construction of quantum affine $\mathfrak {gl}_n$ is an affine generalisation of a similar construction for a quantum enveloping algebra ${\mathbf{U}}$ of a finite type quiver via a Ringel--Hall algebra which, as the positive or negative part of ${\mathbf{U}}$, is spanned by the basis of isoclasses of representations of the quiver and whose multiplication is defined by Hall polynomials, see \cite{R90,R932,X97}. However, there is another construction for quantum $\mathfrak {gl}_n$ by Beilinson, Lusztig and MacPherson \cite[5.7]{BLM}, which directly displays a basis for the {\it entire} quantum enveloping algebra ${\mathbf{U}}(\mathfrak{gl}_n)$ and displays the multiplication rules by explicit formulas of basis elements by generators. This construction is geometric in nature and has been {\it partially} generalised to the affine case (more precisely, to affine $\mathfrak {sl}_n$) in \cite{GV,Lu99, VV99}. BLM's geometric approach uses the definition of quantum Schur algebras as the convolution algebras of (partial) flag varieties over finite fields and then, by a process of ``quantumization'', to get a construction over the polynomials ring. Progress on generalising BLM's work to the affine case via an algebraic approach has been made in the works \cite{DF09,DDF,Fu}. In particular, a realisation conjecture \cite[5.5(2)]{DF09} for quantum affine $\frak{gl}_n$ was formulated and was proved in the classical ($v=1$) case in \cite[Ch. 6]{DDF}. We will prove this conjecture in this paper. We will use directly the definition of quantum Schur algebras as endomorphism algebras of certain $\boldsymbol{v}$-permutation modules over the affine Hecke algebra which has a basis indexed by certain double cosets of the affine symmetric group. The double cosets associated with semisimple representations will play a key role in the establishment of multiplication rules of BLM type basis elements by semisimple generators. It should be pointed out that a recent work by Bridgeland constructs quantum enveloping algebras via Hall algebras of complexes and a complete realisation \cite[Th.~4.9]{Bri} is obtained for simply-laced finite types; see \cite{SY} for the general (finite type) case. We obtain here a complete realisation for quantum affine $\mathfrak{gl}_n$. We now describe the main result of the paper. For a positive integer $n$, let $\widehat{\frak{gl}}_n:=M_{\vtg,\cycn}(\mathbb C)$ be the loop algebra of $\mathfrak{gl}_n(\mathbb C)$ consisitng of all matrices $A=(a_{i,j})_{i,j\in\mathbb Z}$ with $a_{i,j}\in\mathbb C$ such that \begin{itemize} \item[(a)]$a_{i,j}=a_{i+n,j+n}$ for $i,j\in\mathbb Z$; \item[(b)] for every $i\in\mathbb Z$, both sets $\{j\in\mathbb Z\mid a_{i,j}\not=0\}$ and $\{j\in\mathbb Z\mid a_{j,i}\not=0\}$ are finite. \end{itemize} A basis for $\widehat{\frak{gl}}_n$ can be described as $\{E^\vartriangle_{i,j}\mid i,j\in\mathbb Z\}$, where the matrix $E^\vartriangle_{i,j}=(e^{i,j}_{k,l})_{k,l\in\mathbb Z}$ is defined by \begin{equation*}e_{k,l}^{i,j}= \begin{cases}1&\text{if $k=i+sn,l=j+sn$ for some $s\in\mathbb Z$,}\\ 0&\text{otherwise}.\end{cases} \end{equation*} Let $\Theta_\vtg(\cycn)=M_{\vtg,\cycn}(\mathbb N)$ be the $\mathbb N$-span of the basis. Then $\Theta_\vtg(\cycn)$ serves as the index set of the PBW basis of the universal enveloping algebra ${\mathcal U}(\widehat{\frak{gl}}_n)$. Let ${\mathbf{U}}(\widehat{\frak{gl}}_n)$ be the quantum enveloping algebra of $\widehat{\frak{gl}}_n$ over $\mathbb Q(\boldsymbol{v})$ and let $$\Theta_\vtg^\pm(\cycn)=\{A\in\Theta_\vtg(\cycn)\mid a_{i,i} =0\text{ for all $i$}\}\text{ and }\mathbb Z_\vtg^{\cycn}=\{({\lambda}_i)_{i\in\mathbb Z}\mid {\lambda}_i\in\mathbb Z,\,{\lambda}_i={\lambda}_{i-n}\ \text{for}\ i\in\mathbb Z\}.$$ Then $\Theta_\vtg^\pm(\cycn)\times\mathbb Z_\vtg^{\cycn}$ serves as an index set of a PBW type basis for ${\mathbf{U}}(\widehat{\frak{gl}}_n)$ (see, e.g., \cite[1.4.6]{DDF}). We will construct a new basis $\{A({\mathbf{j}})\mid A\in \Theta_\vtg^\pm(\cycn),{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$ and prove the following main result. \begin{MThm}\label{MThm} The quantum enveloping algebra ${\mathbf{U}}(\widehat{\frak{gl}}_n)$ is the $\mathbb Q(\boldsymbol{v})$-algebra which is spanned by the basis $\{A({\mathbf{j}})\mid A\in \Theta_\vtg^\pm(\cycn),{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$ and generated by $0({\mathbf{j}})$, $S_\alpha(\mathbf{0})$ and ${}^t\!S_\alpha(\mathbf{0})$ for all ${\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}$ and $\alpha\in\mathbb N^{n}$, where $S_\alpha=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iE^\vartriangle_{i,i+1}$ and ${}^t\!S_\alpha$ is the transpose of $S_\alpha$, and whose multiplication rules are given by the formulas in Proposition \ref{B(bfl,r)A(bfj,r)}(1)--(3). \end{MThm} We organise this paper as follows. We first recall in \S2 some preliminary results on the double Ringel--Hall algebra of a cyclic quiver and affine quantum Schur algebras. In particular, we display a PBW type basis for the former and a basis defined by certain double cosets for the latter. In \S3, we derive in the affine quantum Schur algebra some multiplication formulas (Theorem \ref{[B][A]}) of the basis elements by those associated with semisimple representations of the cyclic quiver. We then prove the main result via Theorem \ref{realization} in \S4. As an application of the work, we obtain certain multiplication formulas in the Ringel--Hall algebra which are not directly seen from the Hall algebra multiplication. In the Appendix, we give a proof for the length formula of the shortest representative of a double coset defined by a matrix. \begin{Not}\label{Notaion} \rm We need the following index sets for bases of the triangular parts of ${\mathbf{U}}(\widehat{\frak{gl}}_n)$. Let $$\Theta_\vtg^+(\cycn):=\{A\in\Theta_\vtg(\cycn)\mid a_{i,j}=0\text{ for }i\geq j\}\text{ and }\Theta_\vtg^-(\cycn)=\{A\in\Theta_\vtg(\cycn)\mid a_{i,j}=0\text{ for }i\leqslant}\def\geq{\geqslant j\}.$$ For $A\in\Theta_\vtg(\cycn)$, we write \begin{equation}\label{A^+,A^-,A^0} A=A^\pm+A^0=A^++A^0+A^- \end{equation} where $A^\pm\in\Theta_\vtg^\pm(\cycn)$, $A^+\in\Theta_\vtg^+(\cycn)$, $A^-\in\Theta_\vtg^-(\cycn)$ and $A^0$ is a diagonal matrix. Further, for $r\geq 0$, let $\mathbb N_\vtg^{\cycn}=\{({\lambda}_i)_{i\in\mathbb Z}\in \mathbb Z_\vtg^{\cycn}\mid {\lambda}_i\ge0\text{ for }i\in\mathbb Z\}$ and let \begin{equation*} \Theta_\vtg(\cycn,r)=\{A\in\Theta_\vtg(\cycn)\mid\sigma}\newcommand{\vsg}{\varsigma(A)=r\} \text{ and } \Lambda_\vtg(\cycn,r)=\{{\lambda}\in\mathbb N_\vtg^{\cycn}\mid\sigma}\newcommand{\vsg}{\varsigma({\lambda})=r\} \end{equation*} where $\sigma}\newcommand{\vsg}{\varsigma(A)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\, j\in\mathbb Z}a_{i,j}$ and $\sigma}\newcommand{\vsg}{\varsigma({\lambda})=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}{\lambda}_i$. Note that $\Theta_\vtg(\cycn,r)$ is the index set of a basis for a certain quotient algebra of ${\mathbf{U}}(\widehat{\frak{gl}}_n)$---the affine quantum Schur algebras. Moreover, we will use the standard notation for Gaussian polynomials. Thus, let ${\mathcal Z}=\mathbb Z[\boldsymbol{v},\boldsymbol{v}^{-1}]$, where $\boldsymbol{v}$ is an indeterminate, and let $\mathbb Q(\boldsymbol{v})$ be the fraction field of ${\mathcal Z}$. For integers $N,t$ with $t\geq 0$ and $\mu\in\mathbb Z_\vtg^{\cycn}$ and ${\lambda}\in\mathbb N_\vtg^{\cycn}$, let \begin{equation*} \left[\!\!\left[{N\atop t}\right]\!\!\right]=\prod\limits_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant t}\frac{\boldsymbol{v}^{2(N-i+1)}-1}{\boldsymbol{v}^{2i}-1} \,\,\text{ and }\,\,\left[\!\!\left[{\mu\atop{\lambda}}\right]\!\!\right]=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\left[\!\!\left[{\mu_i\atop{\lambda}_i}\right]\!\!\right]. \end{equation*} Then $\left[\!\!\left[{N\atop t}\right]\!\!\right]=\frac{\dblr{N}\dblr{N-1}\cdots \dblr{N-t+1}}{\dblr{t}^!}$, where $\dblr{t}^!=\dblr{1}\dblr{2}\cdots \dblr{t}$ with $[\![m]\!]=\frac{\boldsymbol{v}^{2m}-1}{\boldsymbol{v}^2-1}$. We also need the symmetric Gaussian polynomials $\left[{N\atop t}\right]=\boldsymbol{v}^{-t(N-t)}\left[\!\!\left[{N\atop t}\right]\!\!\right].$ For ${\lambda}, {\lambda}^{(1)},\ldots,{\lambda}^{(m)}\in\mathbb N_\vtg^{\cycn}$ with ${\lambda}={\lambda}^{(1)}+\cdots+{\lambda}^{(m)}$, we also need the following notation in \ref{presentation-dbfHa}(2)(e) $$\left[\!\!\left[{{\lambda}\atop{\lambda}^{(1)},\ldots,{\lambda}^{(m)}}\right]\!\!\right]=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\frac{[\![{\lambda}_i]\!]^!}{[\![{\lambda}^{(1)}_i]\!]^!\cdots[\![{\lambda}^{(m)}_i]\!]^!}.$$ \end{Not} \section{Preliminary results} In this section, we briefly discuss the affine symmetric group and its associated Hecke algebra, the affine $\boldsymbol{v}$-Schur algebra, the double Hall algebra interpretation of affine $\mathfrak{gl}_n$ and the connections between them. Let ${\fS_{\vtg,r}}$ be the {\it affine symmetric group} consisting of all permutations $w:\mathbb Z\rightarrow\mathbb Z$ satisfying $w(i+r)=w(i)+r$ for $i\in\mathbb Z$. Let $W_r$ be the subgroup of ${\fS_{\vtg,r}}$, the {\it Weyl group} of affine type $A$, generated by $S=\{s_i\}_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant r}$, where $s_i$ is defined by $s_i(j)=j$ for $j\not\equiv i,i+1\!\!\!\mod\! r$, $s_i(j)=j-1$ for $j\equiv i+1\!\!\!\mod\! r$, and $s_i(j)=j+1$ for $j\equiv i\!\!\!\mod\! r$. Let $\rho$ be the permutation of $\mathbb Z$ sending $j$ to $j+1$ for all $j\in\mathbb Z$. We extend the length function $\ell$ on $W_r$ to ${\fS_{\vtg,r}}$ by setting $\ell(\rho^mw)=\ell(w)$ for all $m\in\mathbb Z,w\in W_r$. The (extended) affine Hecke algebra ${{\mathcal H}_{\!\vartriangle\!}(r)}$ over ${\mathcal Z}$ associated to ${{\frak S}_{{\!\vartriangle\!},r}}$ is the ${\mathcal Z}$-algebra which is spanned by (basis) $\{T_w\}_{w\in{{\frak S}_{{\!\vartriangle\!},r}}}$ and generated by $T_\rho,T_{\rho^{-1}},T_s, s\in S$, and whose multiplication rules are given by the formulas, for all $s\in S$ and $w\in{\frak S}_{{\!\vartriangle\!},r}$, \begin{equation*} \aligned T_sT_w&=\begin{cases} (\boldsymbol{v}^2-1)T_{w}+\boldsymbol{v}^2T_{sw},\quad&\text{ if }\ell(sw)<\ell(w);\\ T_{sw},\quad&\text{ if } \ell(sw)=\ell(w)+1,\end{cases}\\ T_\rho T_w&=T_{\rho w}.\\ \endaligned \end{equation*} Let ${\boldsymbol{\mathcal H}_{\!\vartriangle\!}(r)}={{\mathcal H}_{\!\vartriangle\!}(r)}\otimes_{\mathcal Z}\mathbb Q(v)$. We will discover a similar description for quantum affine $\mathfrak{gl}_n$. For ${\lambda}\in\Lambda_\vtg(\cycn,r)$, let ${\frak S}_{\lambda}:={\frak S}_{({\lambda}_1,\ldots,{\lambda}_n)}$ be the corresponding standard Young subgroup of the symmetric group ${\frak S}_r$. For a finite subset $X\subseteq{{\frak S}_{{\!\vartriangle\!},r}}$, let $$T_X=\sum_{x\in X}T_x\in{{\mathcal H}_{\!\vartriangle\!}(r)}\;\;\text{ and }\;\; x_{\lambda}=T_{{\frak S}_{\lambda}}.$$ The endomorphism algebras over ${\mathcal Z}$ or $\mathbb Q(\boldsymbol{v})$ $${\mathcal S}_{\!\vartriangle\!}(n,r):=\operatorname{End}_{{{\mathcal H}_{\!\vartriangle\!}(r)}}\biggl (\bigoplus_{{\lambda}\in\Lambda_{\!\vartriangle\!}(n,r)}x_{\lambda}{{\mathcal H}_{\!\vartriangle\!}(r)}\biggr)\,\text{ and }\,{\boldsymbol{\mathcal S}}_\vtg(\cycn,r):=\operatorname{End}_{{\boldsymbol{\mathcal H}_{\!\vartriangle\!}(r)}}\biggl (\bigoplus_{{\lambda}\in\Lambda_{\!\vartriangle\!}(n,r)}x_{\lambda}{\boldsymbol{\mathcal H}_{\!\vartriangle\!}(r)}\biggr)$$ are called {\it affine quantum Schur algebras} or, more specifically, {\it affine $\boldsymbol{v}$-Schur algebras} (cf. \cite{GV,Gr99,Lu99}). Note that ${\boldsymbol{\mathcal S}}_\vtg(\cycn,r)\cong{\mathcal S}_{\vtg}(\cycn,r)\otimes_{\mathcal Z}\mathbb Q(v)$. For ${\lambda}\in\Lambda_\vtg(\cycn,r)$, denote the set of shortest representatives of right cosets of ${\frak S}_{\lambda}$ in ${\mathcal S}_{\vtg}(\cycn,r)$ by $$\afmsD_{\lambda}=\{d\mid d\in{{\frak S}_{{\!\vartriangle\!},r}},\ell(wd)=\ell(w)+\ell(d)\text{ for $w\in{\frak S}_{\lambda}$}\}.$$ Note that elements in $\afmsD_{\lambda}$ can be characterised as follows: \begin{equation}\label{minimal coset representative} \aligned d^{-1}\in\afmsD_{\lambda} &\iff d({\lambda}_{0,i-1}+1)<d({\lambda}_{0,i-1}+2)<\cdots<d({\lambda}_{0,i-1}+{\lambda}_i),\,\forall 1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\endaligned \end{equation} where ${\lambda}_{0,i-1}:=\sum_{1\leqslant}\def\geq{\geqslant t\leqslant}\def\geq{\geqslant i-1}{\lambda}_t$. Moreover, $\afmsD_{{\lambda},\mu}:=\afmsD_{{\lambda}}\cap{\afmsD_{\mu}}^{-1}$ is the set of shortest representatives of $({\frak S}_{\lambda},{\frak S}_\mu)$ double cosets. For ${\lambda},\mu\in\Lambda_\vtg(\cycn,r)$ and $d\in\afmsD_{{\lambda},\mu}$, define $\phi_{{\lambda},\mu}^d\in{\mathcal S}_{\!\vartriangle\!}(n,r)$ by \begin{equation*}\label{def of standard basis} \phi_{{\lambda},\mu}^d(x_\nu h)={\delta}_{\mu\nu}\sum_{w\in{\frak S}_{\lambda} d{\frak S}_\mu}T_wh \end{equation*} where $\nu\in\Lambda_\vtg(\cycn,r)$ and $h\in{{\mathcal H}_{\!\vartriangle\!}(r)}$. Then by \cite{Gr99} the set $\{\phi_{{\lambda},\mu}^d\mid {\lambda},\mu\in\Lambda_\vtg(\cycn,r),\, d\in\afmsD_{{\lambda},\mu}\}$ forms a ${\mathcal Z}$-basis for ${\mathcal S}_{\!\vartriangle\!}(n,r)$. For $1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n$, $k\in\mathbb Z$ and ${\lambda}\in\Lambda_\vtg(n,r)$ let ${\lambda}_{k,i-1}:=kr+\sum_{1\leqslant}\def\geq{\geqslant t\leqslant}\def\geq{\geqslant i-1}{\lambda}_t$ and \begin{equation*} R_{i+kn}^{{\lambda}}=\{{\lambda}_{k,i-1}+1,{\lambda}_{k,i-1}+2,\ldots,{\lambda}_{k,i-1}+{\lambda}_i={\lambda}_{k,i}\}, \end{equation*} By \cite[7.4]{VV99} (see also \cite[9.2]{DF09}), there is a bijective map \begin{equation*} {\jmath_{\!\vartriangle\!}}:\{({\lambda}, d,\mu)\mid d\in\afmsD_{{\lambda},\mu},{\lambda},\mu\in\Lambda_\vtg(\cycn,r)\}\longrightarrow\Theta_\vtg(\cycn,r) \end{equation*} sending $({\lambda}, w,\mu)$ to $A=(a_{k,l})$, where $a_{k,l}=|R_k^{\lambda}\cap wR_l^\mu|$ for all $k,l\in\mathbb Z$. For $A\in\Theta_\vtg(\cycn,r)$ let $e_A=\phi_{{\lambda},\mu}^d$ where $A=\jmath_{\!\vartriangle\!}({\lambda},d,\mu)$. Furthermore, let \begin{equation}\label{nbasis} [A]=\boldsymbol{v}^{-d_A}e_{A},\quad\text{ where } \quad d_{A}=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i\geq k,j<l}a_{i,j}a_{k,l}. \end{equation} Later, in \ref{eBeA} and \ref{[B][A]}, we will consider basis elements associated with matrices of the form $M=A+T-\tilde T$ for some $T,\widetilde T\in\Theta_\vtg(\cycn)$. We will automatically set $e_M=0=[M]$ if one of the entry of $M$ is zero. For $A\in\Theta_\vtg^\pm(\cycn)$ and ${\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}$, define elements in ${\boldsymbol{\mathcal S}}_\vtg(\cycn,r)$: \begin{equation}\label{A(dt,la,r),A(dt,r)} \begin{split} A({\mathbf{j}},r)&=\sum_{\mu\in\Lambda_\vtg(n,r-\sigma}\newcommand{\vsg}{\varsigma(A))}v^{\mu\centerdot{\mathbf{j}}} [A+\operatorname{diag}(\mu)];\;\;\;\qquad(\text{cf. \cite{BLM}})\\ \end{split} \end{equation} where $\mu\centerdot{\mathbf{j}}=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\mu_i{\mathbf{j}}_i$. The set $\{A({\mathbf{j}},r)\}_{A\in\Theta_\vtg^\pm(\cycn), {\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}}$ spans ${\boldsymbol{\mathcal S}}_\vtg(\cycn,r)$. \vspace{.3cm} Let $\triangle(n)$ ($n\geq 2$) be the cyclic quiver with vertex set $I=\mathbb Z/n\mathbb Z=\{1,2,\ldots,n\}$ and arrow set $\{i\to i+1\mid i\in I\}$. Let ${\mathbb F}$ be a field. For $i\in I$, let $S_i$ be the irreducible nilpotent representation of $\triangle(n)$ over ${\mathbb F}$ with $(S_i)_i={\mathbb F}$ and $(S_i)_j=0$ for $i\neq j$. For any $A=(a_{i,j})\in\Theta_\vtg^+(\cycn)$, let $$M(A)=M_{\mathbb F}(A)=\bigoplus_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i<j,\,j\in\mathbb Z}a_{i,j}M^{i,j},$$ where $M^{i,j}=M(E^\vartriangle_{i,j})$ is the unique indecomposable nilpotent representation for $\triangle(n)$ of length $j-i$ with top $S_i$. Thus, the set $\{M(A)\}_{A\in\Theta_\vtg^+(\cycn)}$ is a complete set of representatives of isomorphism classes of finite dimensional nilpotent representations of $\triangle(n)$. The Euler form associated with the cyclic quiver $\triangle(n)$ is the bilinear form $\langle-,-\rangle$: $\mathbb Z_\vtg^{\cycn}\times\mathbb Z_\vtg^{\cycn}\rightarrow\mathbb Z$ defined by $\langle{\lambda},\mu\rangle=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}{\lambda}_i\mu_i-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}{\lambda}_i\mu_{i+1}$ for ${\lambda},\mu\in\mathbb Z_\vtg^{\cycn}$. By \cite{Ri93}, for $A,B,C\in\Theta_\vtg^+(\cycn)$, let $\varphi^{C}_{A,B}\in\mathbb Z[\boldsymbol{v}^2]$ be the Hall polynomials such that, for any finite field ${\mathbb F}_q$, $\varphi^{C}_{A,B}|_{\boldsymbol{v}^2=q}$ is equal to the number of submodules $N$ of $M_{{\mathbb F}_q}(C)$ satisfying $N\cong M_{{\mathbb F}_q}(B)$ and $M_{{\mathbb F}_q}(C)/N\cong M_{{\mathbb F}_q}(A)$. By definition, the (generic) twisted {\it Ringel--Hall algebra} $\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)$ of $\triangle(n)$ is the $\mathbb Q(\boldsymbol{v})$-algebra spanned by basis $\{u_A=u_{[M(A)]}\mid A\in\Theta_\vtg^+(\cycn)\}$ whose multiplication is defined by, for all $A,B\in \Theta_\vtg^+(\cycn)$, $$u_{A}u_{B}=\boldsymbol{v}^{\langle {\mathbf{d}} (A),{\mathbf{d}} (B)\rangle}\sum_{C\in\Theta_\vtg^+(\cycn)}\varphi^{C}_{A,B}u_{C},$$ where ${\mathbf{d}}(A)\in\mathbb N I$ is the dimension vector of $M(A)$. By extending ${\boldsymbol{\mathfrak H}_\vtg(n)}$ to Hopf algebras (see \ref{presentation-dbfHa}(2)(b) for multiplication) $$\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\geq0}=\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)\otimes \mathbb Q(\boldsymbol{v})[K_1^{\pm1},\ldots,K_n^{\pm1}]\text{ and } \boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\leq0}= \mathbb Q(\boldsymbol{v})[K_1^{\pm1},\ldots,K_n^{\pm1}]\otimes\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\rm op},$$ we define the double Ringel--Hall algebra ${\boldsymbol{\mathfrak D}_\vtg}(n)$ (cf. \cite{X97} and \cite[(2.1.3.2)]{DDF}) to be a quotient algebra of the free product $\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\geq 0}* \boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\leqslant}\def\geq{\geqslant 0}$ via a certain skew Hopf paring $\psi:\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\geq 0}\times \boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\leqslant}\def\geq{\geqslant 0}\rightarrow \mathbb Q(\boldsymbol{v})$. In particular, there is a triangular decomposition $${\boldsymbol{\mathfrak D}_\vtg}(n)={\boldsymbol{\mathfrak D}^+_\vtg}(n)\otimes{\boldsymbol{\mathfrak D}^0_\vtg}(n)\otimes{\boldsymbol{\mathfrak D}^-_\vtg}(n),$$ where ${\boldsymbol{\mathfrak D}^+_\vtg}(n)=\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)$, ${\boldsymbol{\mathfrak D}^0_\vtg}(n)= \mathbb Q(\boldsymbol{v})[K_1^{\pm1},\ldots,K_n^{\pm1}]$ and ${\boldsymbol{\mathfrak D}^-_\vtg}(n)=\boldsymbol{\mathfrak H}_{\!\vartriangle\!}(n)^{\rm op}$. For $\alpha\in\mathbb N_\vtg^{\cycn}$ let \begin{equation}\label{semisimple} S_\alpha=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iE^\vartriangle_{i,i+1}\in\Theta_\vtg^+(\cycn). \end{equation} Then $M(S_\alpha)=\oplus_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iS_i$ is a semisimple representation of $\triangle(n)$. Let $u_\alpha=u_{S_\alpha}$. For $\alpha,\beta\in\mathbb Z_\vtg^{\cycn}$, define a partial order on $\mathbb Z_\vtg^{\cycn}$ by setting \begin{equation}\label{order1} \alpha\leqslant}\def\ge{\geqslant\beta \iff \alpha_i\leqslant}\def\geq{\geqslant \beta_i\text{ for all }i\in\mathbb Z. \end{equation} We now collects some of the results we need later, see \cite[Th.~2.5.3]{DDF} for part (1) and \cite[2.6.7]{DDF} for part (2)(e). \begin{Thm} \label{presentation-dbfHa} \begin{itemize} \item[(1)] Let ${\mathbf{U}}(\widehat{\frak{gl}}_n)$ be the quantum enveloping algebra of the loop algebra of $\mathfrak{gl}_n$ defined in \cite{Dr88} or \cite[\S2.5]{DDF}. Then there is a Hopf algebra isomorphism ${\boldsymbol{\mathfrak D}_\vtg}(n)\cong{\mathbf{U}}(\widehat{\frak{gl}}_n)$. \item[(2)] The algebra ${\boldsymbol{\mathfrak D}_\vtg}(n)$ is the algebra over $\mathbb Q(\boldsymbol{v})$ which is spanned by basis $$\{u_{A}^+K^{\mathbf{j}} u_{A}^-\mid A\in\Theta_\vtg^+(\cycn),{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}, \text{ where } K{}^{}{\mathbf{j}}=K_1^{j_1}\cdots K_n^{j_n},$$ and generated by $u_\alpha^+$, $K_{i}^{\pm 1}$, $u_\beta^-$ $(\alpha,\beta\in\Theta_\vtg^+(\cycn),\,1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n)$, and whose multiplication is given by the following relations: \begin{itemize} \item[(a)] $K_iK_j=K_jK_i$, $K_iK_i^{-1}=1$; \item[(b)] $K{}^{}{{\mathbf{j}}} u_A^+=\boldsymbol{v}^{\lr{{\mathbf{d}}(A),{\mathbf{j}}}}u_A^+K{}^{}{\mathbf{j}}$, $u_A^-K{}^{}{\mathbf{j}}=\boldsymbol{v}^{\lr{{\mathbf{d}}(A),{\mathbf{j}}}}K{}^{}{\mathbf{j}} u_A^-$; \item[(c)] $u_\alpha^+u_A^+=\sum_{C\in\Theta_\vtg^+(\cycn)}\boldsymbol{v}^{\langle \alpha,{\mathbf{d}}(A)\rangle}\varphi_{S_\alpha,A}^C u_C^+$; \item[(d)] $u_\beta^-u_A^-=\sum_{C\in\Theta_\vtg^+(\cycn)}\boldsymbol{v}^{\langle {\mathbf{d}}(A),\beta\rangle}\varphi_{A,S_\beta}^C u_C^-$; \item[(e)] For ${\lambda},\mu\in\mathbb N_\vtg^{\cycn}$, $u_\mu^-u_{\lambda}^+-u_{\lambda}^+ u_\mu^-=\displaystyle \sum_{\alpha\not=0,\,\alpha\in\mathbb N_\vtg^{\cycn}\atop\alpha\leqslant}\def\geq{\geqslant{\lambda},\,\alpha\leqslant}\def\geq{\geqslant\mu}\sum_{0\leqslant}\def\geq{\geqslant{\gamma}\leqslant}\def\geq{\geqslant\alpha} x_{\alpha,{\gamma}}\widetilde K^{2{\gamma}-\alpha} u_{{\lambda}-\alpha}^+u_{\mu-\alpha}^-,$ where $$\aligned x_{\alpha,{\gamma}}&=\boldsymbol{v}^{\lr{\alpha,{\lambda}-\alpha}+\lr{\mu,2{\gamma}-\alpha}+2\lr{{\gamma},\alpha-{\gamma}-{\lambda}}+2\sigma}\newcommand{\vsg}{\varsigma(\alpha)} \left[\!\!\left[{{\lambda}\atop\alpha-{\gamma}, {\lambda}-\alpha,{\gamma}}\right]\!\!\right]\left[\!\!\left[{\mu\atop\alpha-{\gamma}, \mu-\alpha,{\gamma}}\right]\!\!\right]\frac{{\frak a}_{\alpha-{\gamma}}{\frak a}_{{\lambda}-\alpha} {\frak a}_{\mu-\alpha}}{{\frak a}_{\lambda} {\frak a}_\mu}\\ &\quad\, \times\sum_{m\geq 1,{\gamma}^{(i)}\not=0\,\forall i\atop{\gamma}^{(1)}+\cdots+{\gamma}^{(m)}={\gamma}}(-1)^m\boldsymbol{v}^{2\sum_{i<j}\langle{\gamma}^{(i)},{\gamma}^{(j)}\rangle} {\frak a}_{{\gamma}^{(1)}}\cdots {\frak a}_{{\gamma}^{(m)}}\left[\!\!\left[{\gamma}\atop{\gamma}^{(1)},\ldots,{\gamma}^{(m)}\right]\!\!\right]^2 \endaligned$$ with ${\frak a}_\beta=\displaystyle\prod_{i=1}^n\prod_{s=1}^{\beta_i}(\boldsymbol{v}^{2\beta_i}-\boldsymbol{v}^{2(s-1)})$ as defined in \cite[Lem.~3.9.1]{DDF} and $\widetilde K{}^{}\nu :=(\widetilde K_1)^{\nu_1}\cdots(\widetilde K_n)^{\nu_n}$ with $\widetilde K_i=K_iK_{i+1}^{-1}$ for $\nu\in\mathbb Z_\vtg^{\cycn}$. \end{itemize} \end{itemize} \end{Thm} For $A\in\Theta_\vtg^+(\cycn)$, let $$\widetilde u_A^\pm=\boldsymbol{v}^{\dim \operatorname{End}(M(A))-\dim M(A)}u_A^\pm,$$ and let ${}^t\!A$ be the transpose matrix of $A$. The relationship between ${\boldsymbol{\mathfrak D}_\vtg}(n)$ and ${\boldsymbol{\mathcal S}}_\vtg(\cycn,r)$ can be seen from the following (cf. \cite{GV,Lu99} and \cite[Prop.~7.6]{VV99}). \begin{Thm}[{\cite[3.6.3, 3.8.1]{DDF}}] \label{zr} For $r\geq 1$, the map $\zeta_r:{\boldsymbol{\mathfrak D}_\vtg}(n)\rightarrow {\boldsymbol{\mathcal S}}_\vtg(\cycn,r)$ is a surjective algebra homomorphism such that, for all ${\mathbf{j}}\in \mathbb Z_\vtg^{\cycn}$ and $A\in \Theta_\vtg^+(\cycn)$, $$\zeta_r(K^{\mathbf{j}})=0({\mathbf{j}},r),\;\zeta_r(\widetilde u_A^+)=A(\mathbf{0},r),\;\;\text{and}\;\; \zeta_r(\widetilde u_A^-)=({}^t\!A)(\mathbf{0},r).$$ \end{Thm} \section{Some multiplication formulas in the affine $\boldsymbol{v}$-Schur algebra} We now derive certain useful multiplication formulas in the affine $\boldsymbol{v}$-Schur algebra and, hence, in the quantum affine $\mathfrak{gl}_n$. These formulas will be given in \ref{[B][A]} and \ref{B(bfl,r)A(bfj,r)}. They are the key to the realization of quantum affine $\frak{gl}_n$. We need some preparation before proving \ref{[B][A]} and \ref{B(bfl,r)A(bfj,r)}. The following result is given in \cite[3.2.3]{DDF}. \begin{Lem}\label{double coset} Let ${\lambda},\mu\in\Lambda_\vtg(\cycn,r)$ and $d\in\mathscr D}\newcommand{\afmsD}{{\mathscr D}^{\!\vartriangle\!}_{{\lambda},\mu}^{\!\vartriangle\!}$. Assume $A=\jmath_{\!\vartriangle\!}({\lambda},d,\mu)$. Then $d^{-1}\frak S_{\lambda} d\cap\frak S_\mu=\frak S_\nu$, where $\nu=(\nu^{(1)},\ldots,\nu^{(n)})$ and $\nu^{(i)}=(a_{ki})_{k\in\mathbb Z}=(\ldots,a_{1i},\ldots,a_{ni},\ldots)$. In particular, we have \begin{equation*} x_{\lambda} T_d x_\mu=\displaystyle \prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\dblr{a_{i,j}}^! T_{{\frak S}_{\lambda} d{\frak S}_\mu}. \end{equation*} \end{Lem} Given $A\in\Theta_\vtg(\cycn,r)$ with $A=\jmath_{\!\vartriangle\!}({\lambda},w,\mu)$, let $y_A=w$ be the shortest representative of the double coset ${\frak S}_{\lambda} w{\frak S}_\mu$. \begin{Lem}\label{length of elements in Dla} $(1)$ For ${\lambda}\in\Lambda_\vtg(2,r)$ and $w\in\afmsD_{\lambda}\cap{\frak S}_r$, we have $\ell(w)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant{\lambda}_1}(w^{-1}(i)-i)$. $(2)$ For any $A\in\Theta_\vtg(\cycn,r)$, $\ell(y_A)=\displaystyle\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i<k;j>l}a_{ij}a_{kl}.$ \end{Lem} \begin{proof}Since $w\in{\frak S}_r$, $w^{-1}(1)<\cdots<w^{-1}({\lambda}_1)$ and $w^{-1}({\lambda}_1+1)<\cdots<w^{-1}(r)$, it follows that $$\aligned \ell(w^{-1})&=|\{(i,j)\mid1\leqslant}\def\geq{\geqslant i<j\leqslant}\def\geq{\geqslant r,w^{-1}(i)>w^{-1}(j)\}|\\ &=|\{(i,j)\mid1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant {\lambda}_1,{\lambda}_1+1\leqslant}\def\geq{\geqslant j\leqslant}\def\geq{\geqslant r,w^{-1}(i)>w^{-1}(j)\}|.\endaligned$$ On the other hand, for every $1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant {\lambda}_1$, $w^{-1}(i)-i$ of the numbers $1,2,\ldots, w^{-1}(i)$ must lie in $\{w^{-1}({\lambda}_1+1),\ldots,w^{-1}(r)\}$ which contribute $w^{-1}(i)-i$ inversions. Hence, $\ell(w)=\ell(w^{-1})=(w^{-1}(1)-1)+(w^{-1}(2)-2)+\cdots+(w^{-1}({\lambda}_1)-{\lambda}_1)$, proving part (1). Part (2) is probably known. Since we couldn't find a proof in the literature, a proof is given in the Appendix. \end{proof} \begin{Lem}\label{sum} For $a\geq 0$, $r\geq 1$ and $0\leqslant}\def\geq{\geqslant t\leqslant}\def\geq{\geqslant r$ we have $$\sum_{X\subseteq\{a+1,\cdots,a+r\}\atop |X|=t}v^{2\sum_{x\in X}x}=v^{2at+t(t+1)}\left[\!\!\left[{r\atop t}\right]\!\!\right].$$ \end{Lem} \begin{proof} We proceed by induction on $r$. The case $r=1$ is trivial. Assume now that $r>1$. Then, by induction hypothesis, \begin{equation*} \begin{split} \sum_{X\subseteq\{a+1,\cdots,a+r\}\atop |X|=t}v^{2\sum_{x\in X}x}&= \sum_{X\subseteq\{a+1,\cdots,a+r-1\}\atop |X|=t}v^{2\sum_{x\in X}x}+ \sum_{Y\subseteq\{a+1,\cdots,a+r-1\}\atop |Y|=t-1}v^{2(a+r+\sum_{x\in Y}x)}\\ &=v^{2at+t(t+1)}\left[\!\!\left[{r-1\atop t}\right]\!\!\right]+v^{2(a+r)}v^{2a(t-1)+t(t-1)}\left[\!\!\left[{r-1\atop t-1}\right]\!\!\right]\\ &=v^{2at+t(t+1)}\bigg(\left[\!\!\left[{r-1\atop t}\right]\!\!\right]+v^{2(r-t)}\left[\!\!\left[{r-1\atop t-1}\right]\!\!\right]\bigg)\\ &=v^{2at+t(t+1)}\left[\!\!\left[{r\atop t}\right]\!\!\right], \end{split} \end{equation*} as desired. \end{proof} For $i\in\mathbb Z$ let $\boldsymbol e^\vartriangle_i\in\mathbb N_\vtg^{\cycn}$ be such that \begin{equation*} (\boldsymbol e^\vartriangle_i)_j= \begin{cases} 1&\text{if $j\equiv i \!\!\!\mod\! n$}\\ 0&\text{otherwise}. \end{cases} \end{equation*} \begin{Lem}[{\cite[5.2]{Fu}}]\label{vartheta} Let $\mu\in\Lambda_\vtg(\cycn,r)$, $\beta\in\mathbb N_\vtg^{\cycn}$ and assume $\mu\geq\beta$. $(1)$ If $\alpha=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}(\mu_i-\beta_i)\boldsymbol e^\vartriangle_{i-1}$, ${\delta}=(\alpha_0,\beta_1,\alpha_1,\beta_2,\cdots,\alpha_{n-1},\beta_n)$ and $$\mathpzc Y=\{(Y_0,Y_1,\cdots,Y_{n-1})\mid Y_i\subseteq R_{i+1}^\mu,\,|Y_i|=\alpha_i,\ for\ 0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\},$$ then there is a bijective map $$g:\afmsD_{\delta}\cap{\frak S}_\mu\rightarrow\mathpzc Y$$ defined by sending $w$ to $(w^{-1}X_0,w^{-1}X_1,\cdots,w^{-1}X_{n-1})$ where $X_i= \{\mu_{0,i}+1,\mu_{0,i}+2,\cdots,\mu_{0,i}+\alpha_i\},$ with $\mu_{0,i}=\sum_{1\leqslant}\def\geq{\geqslant s\leqslant}\def\geq{\geqslant i}\mu_s$ and $\mu_{0,0}=0$. $(2)$ If ${\gamma}=\mu-\beta$, $\theta} \newcommand{\vth}{\vartheta=(\beta_1,{\gamma}_1,\beta_2,{\gamma}_2,\cdots,\beta_n,{\gamma}_n)$ and $$\mathpzc Y'=\{(Y_1',Y_2',\cdots,Y_{n}')\mid Y_i'\subseteq R_{i}^{\mu},\,|Y_i'|={\gamma}_i,\ for\ 1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\},$$ then there is a bijective map $$g':\afmsD_{\theta} \newcommand{\vth}{\vartheta}\cap{\frak S}_{\mu}\rightarrow\mathpzc Y'$$ defined by sending $w$ to $(w^{-1}X_1',w^{-1}X_2',\cdots,w^{-1}X_{n}')$ where $X_i'= \{\mu_{0,i-1}+\beta_i+1,\mu_{0,i-1}+\beta_i+2,\cdots,\mu_{0,i}\}.$ \end{Lem} The injection can be seen easily by noting that $$(\alpha_0+\beta_1,\alpha_1+\beta_2,\cdots,\alpha_{n-1}+\beta_n)=(\mu_1,\mu_2,\cdots,\mu_n)=(\beta_1+{\gamma}_1,\beta_2+{\gamma}_2,\cdots,\beta_n+{\gamma}_n)$$ and $X_i$ (reps., $X_{i+1}'$) consists of the first $\alpha_i$ (reps., the last ${\gamma}_{i+1}$) numbers in $R_{i+1}^\mu$ for all $0\leqslant}\def\geq{\geqslant i<n$, while the subjection is to define $w=g^{-1}(Y_0,Y_1,\ldots,Y_{n-1})$ by $w^{-1}(\mu_{0,i}+s)=k_{i,s}$ for all $0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1$ and $1\leqslant}\def\geq{\geqslant s\leqslant}\def\geq{\geqslant \mu_{i+1}$, where $Y_i=\{k_{i,1},\ldots,k_{i,\alpha_i}\}$, $R_{i+1}^\mu\backslash Y_i=\{k_{i,\alpha_i+1},k_{i,\alpha_i+2},\ldots,k_{i,\mu_{i+1}}\}$, and both are strictly increasing. For $A\inM_{\vtg,\cycn}(\mathbb Z)$ with $\sigma}\newcommand{\vsg}{\varsigma(A)=r$, we denote $e_A=[A]=0\in{\mathcal S}_{\vtg}(\cycn,r)$ if $a_{i,j}<0$ for some $i,j\in\mathbb Z$. There is a natural map \begin{equation}\label{ti} \widetilde\ :\Theta_\vtg(\cycn)\rightarrow\Theta_\vtg(\cycn)\;\;A=(a_{i,j})\longmapsto\widetilde A=(\widetilde a_{i,j}), \end{equation} where $\widetilde a_{i,j}=a_{i-1,j}$ for all $i,j\in\mathbb Z$. We are now ready to establish multiplication formulas of an arbitrary basis elements $e_A$ by certain basis elements $e_B$ in the affine Schur algebra ${\mathcal S}_{\vtg}(\cycn,r)$ over ${\mathcal Z}$, where $B^+$ or ${}^t(B^-)$ defines a semisimple representation of the cyclic quiver. The significance of these formulas is the generalisation of \cite[3.5]{Lu99} (cf. \cite[3.1]{BLM}) from real roots to all roots including all imaginary roots. \begin{Prop}\label{eBeA} Let $A\in\Theta_\vtg(\cycn,r)$ and $\mu=\text{\rm ro}(A)$. Assume $\beta\in\mathbb N_\vtg^{\cycn}$ and $\beta\leqslant}\def\geq{\geqslant\mu$. Let $\alpha=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}(\mu_i-\beta_i)\boldsymbol e^\vartriangle_{i-1}$ and ${\gamma}=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}(\mu_i-\beta_i)\boldsymbol e^\vartriangle_{i}=\mu-\beta$, $B=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iE^\vartriangle_{i,i+1}+\operatorname{diag}(\beta)$, and $C=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}{\gamma}_iE^\vartriangle_{i+1,i}+\operatorname{diag}(\beta)$. Then the following identities hold in ${\mathcal S}_{\vtg}(\cycn,r)$. \begin{itemize} \item[(1)] $e_Be_A=\sum\limits_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j>l} (a_{i,j}-t_{i-1,j})t_{i,l}}\prod\limits_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right] e_{A+T-\widetilde T};$ \item[(2)] $e_Ce_A =\sum\limits_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)={\gamma}} v^{2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j<l}(a_{i,j}-t_{i,j})t_{i-1,l}}\prod\limits_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\left[\!\!\left[{a_{i,j}-t_{i,j}+t_{i-1,j}\atop t_{i-1,j}}\right]\!\!\right] e_{A-T+\widetilde T}.$ \end{itemize} \end{Prop} \begin{proof} We only prove (1). The proof for (2) is entirely similar. Let ${\lambda}=\text{\rm ro}(B)$ and $\nu=\text{\rm co}(A)$. Assume $d_1\in\mathscr D}\newcommand{\afmsD}{{\mathscr D}^{\!\vartriangle\!}^{\!\vartriangle\!}_{{\lambda},\mu}$ and $d_2\in\mathscr D}\newcommand{\afmsD}{{\mathscr D}^{\!\vartriangle\!}^{\!\vartriangle\!}_{\mu,\nu}$ defined by $\jmath_{\!\vartriangle\!}({\lambda}, d_1,\mu)=B$ and $\jmath_{\!\vartriangle\!}(\mu, d_2,\nu)=A$. Then ${\lambda}_i=\alpha_i+\beta_i$ and $\mu_i=\alpha_{i-1}+\beta_i$ for all $1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n$. From \ref{double coset} we see that \begin{equation*} \begin{split} e_Be_A(x_\nu)&=T_{\frak S_{\lambda} d_1\frak S_\mu}\cdot T_{d_2}\cdot T_{\afmsD_\omega\cap\frak S_\nu}\\ &=\frac{1}{\sum_{w\in\frak S_\mu}v^{2\ell(w)}}T_{\frak S_{\lambda} d_1\frak S_\mu}\cdot T_{\frak S_\mu d_2\frak S_\nu}\\ &=\frac{1}{\sum_{w\in\frak S_\mu}v^{2\ell(w)}}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\frac{1}{\dblr{a_{i,j}}^!}T_{\frak S_{\lambda} d_1\frak S_\mu}\cdot T_{\frak S_\mu}\cdot T_{d_2}\cdot T_{\frak S_\nu}\\ &=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\frac{1}{\dblr{a_{i,j}}^!}T_{\frak S_{\lambda} d_1\frak S_\mu}\cdot T_{d_2}\cdot T_{\frak S_\nu}\\ &=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\frac{1}{\dblr{a_{i,j}}^!}T_{\frak S_{\lambda}}\cdot T_{d_1}\cdot T_{\afmsD_{\delta}\cap\frak S_\mu}\cdot T_{d_2}\cdot T_{\frak S_\nu} \end{split} \end{equation*} where $\frak S_\omega=d_2^{-1}\frak S_\mu d_2\cap\frak S_\nu$, $\frak S_{\delta}=d_1^{-1}\frak S_{\lambda} d_1\cap{\frak S}_\mu$ with ${\delta}=(\alpha_0,\beta_1,\alpha_1,\beta_2,\cdots,\alpha_{n-1},\beta_n)$. By \eqref{minimal coset representative}, we have $d_1=\rho^{-\alpha_0}$ (so $\ell(d_1)=0$). This together with the fact that $d_2\in\afmsD_\mu$ implies that $\ell(d_1wd_2)=\ell(d_1)+\ell(w)+\ell(d_2)=\ell(w)+\ell(d_2)$ for $w\in\afmsD_{\delta}\cap\frak S_\mu$. Thus, we have \begin{equation}\label{eq1 for fundentemental formulas} e_Be_A(x_\nu)=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\frac{1}{\dblr{a_{i,j}}^!}\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}}T_{\frak S_{\lambda}}T_{d_1wd_2}T_{\frak S_\nu} \end{equation} For $w\in\afmsD_{\delta}\cap\frak S_\mu$ let $C^{(w)}=(c_{i,j}^{(w)})\in\Theta_\vtg(\cycn,r)$, where $c_{i,j}^{(w)}=|R_i^{\lambda}\cap d_1wd_2R_j^\nu|$, and let $T^{(w)}=(t_{i,j}^{(w)})\in\Theta_\vtg(\cycn)$, where $t_{i,j}^{(w)}=|w^{-1}X_i\cap d_2 R_j^\nu|$ with $X_i= \{\mu_{0,i}+1,\mu_{0,i}+2,\cdots,\mu_{0,i}+\alpha_i\}$ for $1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n$ and $j\in\mathbb Z$. Then $\text{\rm ro}(T^{(w)})=\alpha$ and $\text{\rm co}(T^{(w)})\leqslant}\def\geq{\geqslant \nu$. Since $d_1^{-1}R_i^{\lambda}=\alpha_0+R_i^{\lambda}=(R_i^\mu\backslash X_{i-1}\cup X_i)$, we see that $c_{i,j}^{(w)}=|R_i^{\lambda}\cap d_1wd_2R_j^\nu|=|w^{-1}d_1^{-1}R_i^{\lambda}\cap d_2 R_j^\nu|=a_{i,j}-t_{i-1,j}^{(w)}+t_{i,j}^{(w)}$ (see the proof of \cite[5.3]{Fu}). In other words, for all $w\in\afmsD_{\delta}\cap{\frak S}_\mu$, \begin{equation}\label{eq2 for fundentemental formulas} C^{(w)}=A+T^{(w)}-\widetilde T^{(w)}. \end{equation} In particular, $y_{C^{(w)}}\in {\frak S}_{\lambda} d_1wd_2{\frak S}_\nu\cap \afmsD_{{\lambda}\nu}$. Putting ${\frak S}_{\alpha_w}=y_{C^{(w)}}^{-1}{\frak S}_{\lambda} y_{C^{(w)}}\cap{\frak S}_\nu$, we have by \ref{double coset} \[ \begin{split} \sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}}T_{\frak S_{\lambda}}T_{d_1wd_2}T_{\frak S_\nu} &=\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta},\,d_1wd_2=w'y_{C^{(w)}}w''\atop w'\in{\frak S}_{\lambda},\,w''\in{\frak S}_\nu\cap\afmsD_{\alpha_w}}T_{\frak S_{\lambda}} T_{w'}T_{y_{C^{(w)}}}T_{w''}T_{\frak S_\nu}\\ &=\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta},\,d_1wd_2=w'y_{C^{(w)}}w''\atop w'\in{\frak S}_{\lambda},\,w''\in{\frak S}_\nu\cap\afmsD_{\alpha_w}}v^{2(\ell(w')+\ell(w''))}T_{\frak S_{\lambda}} T_{y_{C^{(w)}}} T_{\frak S_\nu}\\ &=\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}}v^{2(\ell(w)+\ell(d_2)-\ell(y_{C^{(w)}}))}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\in\mathbb Z}\left[\!\!\left[ c_{i,j}^{(w)}\right]\!\!\right]^! e_{C^{(w)}}(x_\nu). \end{split} \] Now by \eqref{eq1 for fundentemental formulas} and \eqref{eq2 for fundentemental formulas} and noting $\text{\rm ro}(T^{(w)})=\alpha$ for $w\in\afmsD_{\delta}\cap{\frak S}_\mu$, we have \begin{equation}\label{eq3 for fundentemental formulas} \begin{split} e_Be_A&=\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}}v^{2(\ell(w)+\ell(d_2)-\ell(y_{C^{(w)}}))} \prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\in\mathbb Z}\frac{\left[\!\!\left[ c_{i,j}^{(w)}\right]\!\!\right]^!}{\dblr{a_{i,j}}^!} e_{C^{(w)}} \\ &=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z} \frac{\dblr{a_{i,j}-t_{i-1,j}+t_{i,j}}^!}{\dblr{a_{i,j}}^!}v^{2(\ell(d_2)-\ell(y_{A+T-\widetilde T}))}\bigg(\sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}\atop T^{(w)}=T}v^{2\ell(w)}\bigg)e_{A+T-\widetilde T}. \end{split} \end{equation} Given $T\in\Theta_\vtg(\cycn)$ with $\text{\rm ro}(T)=\alpha$ let $$\mathpzc Z(T)=\{Z=(Z_{i,j})_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1,\,j\in\mathbb Z}\mid |Z_{i,j}|=t_{i,j},\,Z_{i,j}\subseteq R_{i+1}^\mu\cap d_2R_j^\nu,\ \text{for}\ 0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1,\,j\in\mathbb Z\}.$$ If $T=T^{(w)}$ then the bijective map $g$ in \ref{vartheta} induces a bijective map $$h_T: \{w\in\afmsD_{\delta}\cap{\frak S}_\mu\mid T^{(w)}=T\}\rightarrow\mathpzc Z(T)$$ defined by sending $w$ to $(w^{-1}(X_i)\cap d_2R_j^\nu)_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1,\,j\in\mathbb Z}$. Since for $0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1$ and $j\in\mathbb Z$ $$ R_{i+1}^\mu\cap d_2R_j^\nu=\bigg\{\mu_{0,i}+\sum_{s\leqslant}\def\geq{\geqslant j-1}a_{i+1,s}+1,\mu_{0,i}+\sum_{s\leqslant}\def\geq{\geqslant j-1}a_{i+1,s}+2,\cdots,\mu_{0,i}+\sum_{s\leqslant}\def\geq{\geqslant j}a_{i+1,s}\bigg\},$$ it follows from \ref{length of elements in Dla} and the definition of $g^{-1}$ that, for $Z=(Z_{i,j})_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1,\,j\in\mathbb Z}\in\mathpzc Z(T)$, \begin{equation*} \begin{split} \ell(h_T^{-1}(Z)) &=\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1}\bigg(\sum_{k\in Z_{i,j}\atop j\in\mathbb Z}k -\sum_{1\leqslant}\def\geq{\geqslant j\leqslant}\def\geq{\geqslant\alpha_i}(\mu_{0,i}+j)\bigg)\\ &=\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j\in\mathbb Z,\,k\in Z_{i,j}}k-\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1}\bigg(\alpha_i\mu_{0,i}+\frac{\alpha_i(\alpha_i+1)}{2}\bigg).\\ \end{split} \end{equation*} This implies that \begin{equation*} \begin{split} \sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}\atop T^{(w)}=T}v^{2\ell(w)}&= \sum_{Z\in\mathpzc Z(T)}v^{2\ell(\kappa_T^{-1}(Z))}\\ &=v^{-2\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1}(\alpha_i\mu_{0,i}+\alpha_i(\alpha_i+1)/2)}\prod_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j\in\mathbb Z}\bigg(\sum_{Z_{i,j}\subseteq R_{i+1}^\mu\cap d_2R_j^\nu\atop |Z_{i,j}|=t_{i,j}}v^{2\sum_{k\in Z_{i,j}}k}\bigg). \end{split} \end{equation*} Consequently, by \ref{sum}, we have \begin{equation}\label{eq4 for fundentemental formulas} \begin{split} \sum_{w\in{\frak S}_\mu\cap\afmsD_{\delta}\atop T^{(w)}=T}v^{2\ell(w)} &=v^{2a_T}\prod_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j\in\mathbb Z}\left[\!\!\left[{a_{i+1,j}\atop t_{i,j}}\right]\!\!\right]=v^{2a_T}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\left[\!\!\left[{a_{i,j}\atop t_{i-1,j}}\right]\!\!\right] \end{split} \end{equation} where $$a_T=\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j\in\mathbb Z}\bigg(t_{i,j}\big(\mu_{0,i} +\sum_{s\leqslant}\def\geq{\geqslant j-1}a_{i+1,s}\big)+\frac{t_{i,j}(t_{i,j}+1)}{2}\bigg)-\sum _{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1}\bigg(\alpha_i\mu_{0,i}+\frac{\alpha_i(\alpha_i+1)}{2}\bigg).$$ Since $\text{\rm ro}(T)=\alpha$ we have $\alpha_i=\sum_{j\in\mathbb Z}t_{i,j}$ and $\alpha_i^2=\sum_{j\in\mathbb Z}t_{i,j}^2+2\sum_{j<l}t_{i,j}t_{i,l}$ for $0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1$. This implies that $$a_T=\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j,s\in\mathbb Z,\, s<j}a_{i+1,s}t_{i,j}-\sum_{0\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n-1\atop j,l\in\mathbb Z,\,j<l}t_{i,j}t_{i,l}.$$ Since $d_2=y_A$ is the shortest representative in the double coset associated with $A$, by \ref{length of elements in Dla}(2), $$\ell(d_2)-\ell(y_{A+T-\widetilde T})= \sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}a_{i,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}a_{i+1,l}t_{i,j}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l} t_{i,j}t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}t_{i-1,j}t_{i,l}.$$ It follows that $$a_T+\ell(d_2)-\ell(y_{A+T-\widetilde T})=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}a_{i,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}t_{i-1,j}t_{i,l}.$$ Consequently, by \eqref{eq3 for fundentemental formulas}, \eqref{eq4 for fundentemental formulas} and noting $\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\in\mathbb Z}\dblr{ t_{i,j}}^!=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\in\mathbb Z}\dblr{t_{i-1,j}}^!$ we have \begin{equation*} \begin{split} e_Be_A &=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{2(a_T+\ell(d_2)-\ell(y_{A+T-\widetilde T}))}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z} \frac{\dblr{a_{i,j}-t_{i-1,j}+t_{i,j}}^!}{\dblr{t_{i-1,j}}^!\cdot \dblr{a_{i,j}-t_{i-1,j}}^!}e_{A+T-\widetilde T}\\ &=\sum\limits_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j>l} (a_{i,j}-t_{i-1,j})t_{i,l}}\prod\limits_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right] e_{A+T-\widetilde T}, \end{split} \end{equation*} proving (1). \end{proof} Let $\bar\ :{\mathcal Z}\rightarrow{\mathcal Z}$ be the ring homomorphism defined by $\bar v=v^{-1}$. We now use \ref{eBeA} to derive the corresponding formulas for the normalised basis $\{[A]\}_{A\in\Theta_\vtg(\cycn,r)}$ defined in \eqref{nbasis}. \begin{Prop} \label{[B][A]} Let $A\in\Theta_\vtg(\cycn,r)$ and $\alpha,{\gamma}\in\mathbb N_\vtg^{\cycn}$. $(1)$ For $B\in\Theta_\vtg(\cycn,r)$, if $B-\sum\limits_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iE^\vartriangle_{i,i+1}$ is a diagonal matrix and $\text{\rm co}(B)=\text{\rm ro}(A)$, then $$[B][A]=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{\beta(T,A)}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}[A+T-\widetilde T],$$ where $\beta(T,A)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\geq l}(a_{i,j}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j>l}(a_{i+1,j}-t_{i,j})t_{i,l}$. $(2)$ For $C\in\Theta_\vtg(\cycn,r)$, if $C-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}{\gamma}_iE^\vartriangle_{i+1,i}$ is a diagonal matrix and $\text{\rm co}(C)=\text{\rm ro}(A)$, then $$[C][A]=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)={\gamma}}v^{\beta'(T,A)}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\overline{\left[\!\!\left[{a_{i,j}-t_{i,j}+t_{i-1,j}\atop t_{i-1,j}}\right]\!\!\right]}[A-T+\widetilde T],$$ where $\beta'(T,A)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,l\geq j}(a_{i,j}-t_{i,j})t_{i-1,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,l>j}(a_{i,j}-t_{i,j})t_{i,l}$. \end{Prop} \begin{proof} We only prove (1). The proof for (2) is entirely similar. Note that we have $$\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}=v^{2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\, j\in\mathbb Z}(a_{i,j}-t_{{i-1,j}})t_{i,j}}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}$$ Thus by \ref{eBeA}(1) we have $$[B][A]=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{\beta(T,A)}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}\overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}[A+T-\widetilde T],$$ where \begin{equation*} \begin{split} \beta(T,A)&=2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j>l} (a_{i,j}-t_{i-1,j})t_{i,l}+d_{A+T-\widetilde T}-d_A-d_B+2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z}(a_{i,j}-t_{{i-1,j}})t_{i,j}\\ &=2\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\geq l} (a_{i,j}-t_{i-1,j})t_{i,l}+d_{A+T-\widetilde T}-d_A-d_B. \end{split} \end{equation*} Fix $T\in\Theta_\vtg(\cycn)$ satisfying $\text{\rm ro}(T)=\alpha$. Then by definition we have $d_B=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}b_{i,i}\alpha_i$ and \begin{equation*} \begin{split} &\qquad d_{A+T-\widetilde T}-d_A\\ &= \sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i\geq k,\,j<l}a_{i,j}(t_{k,l}-t_{k-1,l}) +\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i\geq k,\,j<l}a_{k,l}(t_{i,j}-t_{i-1,j})+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i\geq k,\,j<l}(t_{i,j}-t_{i-1,j})(t_{k,l}-t_{k-1,l})\\ &=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j<l}a_{i,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j<l}a_{i+1,l}t_{i,j}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j<l}(t_{i,j}-t_{i-1,j})t_{i,l}\\ &=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j<l}(a_{i,j}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}(a_{i+1,j}-t_{i,j})t_{i,l}. \end{split} \end{equation*} Furthermore, since $\text{\rm ro}(T)=\alpha$ and $\text{\rm co}(B)=\text{\rm ro}(A)$ we have $b_{i,i}=\sum_{j\in\mathbb Z}(a_{i,j}-t_{i-1,j})$ and $\alpha_i=\sum_{l\in\mathbb Z}t_{i,l}$ for each $i$, and hence $$d_{A+T-\widetilde T}-d_A-d_B=-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l}(a_{i,j}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}(a_{i+1,j}-t_{i,j})t_{i,l}.$$ Consequently, $\beta(T,A)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j\geq l}(a_{i,j}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,j>l}(a_{i+1,j}-t_{i,j})t_{i,l}$. The proof is completed. \end{proof} \section{Proof of the main theorem} We now construct explicitly a subalgebra of the algebra ${\boldsymbol{\mathcal S}}_\vtg(n):=\prod_{r\geq 1}{\boldsymbol{\mathcal S}}_\vtg(\cycn,r)$ and prove that this subalgebra is isomorphic to ${\mathbf{U}}(\widehat{\frak{gl}}_n)$. Recall the elements $A({\mathbf{j}},r)$ defined in \eqref{A(dt,la,r),A(dt,r)} and let \begin{equation}\label{basis-fB} A({\mathbf{j}})=(A({\mathbf{j}},r))_{r\geq 0}\in{\boldsymbol{\mathcal S}}_\vtg(n)\quad\text{ and }\quad{\frak B}=\{A({\mathbf{j}})\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}. \end{equation} Then ${\frak B}$ is linearly independent by \cite[Prop.4.1(2)]{DF09}. Let $\boldsymbol{{\mathcal V}}_\vtg(n)$ be the $\mathbb Q(v)$-subspace of ${\boldsymbol{\mathcal S}}_\vtg(n)$ spanned by ${\frak B}$. We will prove that $\boldsymbol{{\mathcal V}}_\vtg(n)$ is a subalgebra of ${\boldsymbol{\mathcal S}}_\vtg(n)$ isomorphic to ${\mathbf{U}}(\widehat{\frak{gl}}_n)$ or ${\boldsymbol{\mathfrak D}_\vtg}(n)$ by \ref{presentation-dbfHa}(a). For this purpose, we need a larger spanning set containing ${\frak B}$: $$\widetilde{\frak B}=\{A({\mathbf{j}}, {\lambda})\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn},\,{\lambda}\in\mathbb N_\vtg^{\cycn}\}$$ where $A({\mathbf{j}},{\lambda})=(A({\mathbf{j}},{\lambda},r))_{r\geq 0}$ with $A({\mathbf{j}},{\lambda},r)$ defined by \begin{equation} A({\mathbf{j}},{\lambda},r)=\sum_{\mu\in\Lambda_\vtg(n,r-\sigma}\newcommand{\vsg}{\varsigma(A))}v^{\mu\centerdot{\mathbf{j}}} \left[{\mu\atop{\lambda}}\right][A+\operatorname{diag}(\mu)]\;\quad(\text{cf. \cite[\S2]{Fu1}}) \end{equation} Note that, for $\sigma({\lambda})\leqslant}\def\geq{\geqslant r$, $0({\mathbf{j}},{\lambda}, r)=\sum_{\mu\in\Lambda_\vtg(n,r),{\lambda}\leqslant}\def\geq{\geqslant\mu}v^{\mu\centerdot{\mathbf{j}}}\left[{\mu\atop{\lambda}}\right][\operatorname{diag}(\mu)]$ \begin{Lem}\label{spanning set of afbfVn} The space $\boldsymbol{{\mathcal V}}_\vtg(n)$ is spanned by the set $\widetilde {\frak B}$. In other words, every $A({\mathbf{j}}, {\lambda})\in\boldsymbol{{\mathcal V}}_\vtg(n)$. \end{Lem} \begin{proof} Let $\boldsymbol{{\mathcal V}}_\vtg^0(n)$ be the $\mathbb Q(v)$-subalgebra of ${\boldsymbol{\mathcal S}}_\vtg(n)$ generated by $0(\pm\boldsymbol e^\vartriangle_i)$ for $1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n$. Then the set $\{0({\mathbf{j}})\mid{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$ forms a $\mathbb Q(v)$-basis for $\boldsymbol{{\mathcal V}}_\vtg^0(n)$. Since $$0({\mathbf{j}},{\lambda},r)=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\bigg(0(\boldsymbol e^\vartriangle_i,r)^{j_i} \prod_{1\leqslant}\def\geq{\geqslant s\leqslant}\def\geq{\geqslant{\lambda}_i}\frac{0(\boldsymbol e^\vartriangle_i,r)v^{-s+1}-0(-\boldsymbol e^\vartriangle_i,r)v^{s-1}}{v^s-v^{-s}}\bigg),\text{ where }\sigma({\lambda})\leqslant}\def\geq{\geqslant r,$$ we have \begin{equation}\label{0span} 0({\mathbf{j}},{\lambda})=\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\bigg(0(\boldsymbol e^\vartriangle_i)^{j_i} \prod_{1\leqslant}\def\geq{\geqslant s\leqslant}\def\geq{\geqslant{\lambda}_i}\frac{0(\boldsymbol e^\vartriangle_i)v^{-s+1}-0(-\boldsymbol e^\vartriangle_i)v^{s-1}}{v^s-v^{-s}}\bigg)\in\boldsymbol{{\mathcal V}}_\vtg^0(n). \end{equation} On the other hand, by the proof of \cite[3.4]{Fu1}), we have $$0({\mathbf{j}},{\lambda})A(\mathbf{0})=v^{\text{\rm ro}(A)\centerdot({\mathbf{j}}+{\lambda})}A({\mathbf{j}},{\lambda})+ \sum_{\mu\in\mathbb N^{n},\,\mathbf{0}<\mu\leqslant}\def\geq{\geqslant{\lambda}} v^{\text{\rm ro}(A)\centerdot({\mathbf{j}}+{\lambda}-\mu)}\left[{\text{\rm ro}(A)\atop\mu}\right] A({\mathbf{j}}-\mu,{\lambda}-\mu).$$ By induction, we see that $A({\mathbf{j}},{\lambda})\in\operatorname{span}\{0({\mathbf{j}},{\lambda})A(\mathbf{0})\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn},\,{\lambda}\in\mathbb N_\vtg^{\cycn}\}$. Thus, by \eqref{0span}, $A({\mathbf{j}},{\lambda})\in\operatorname{span}\{0({\mathbf{j}})A(\mathbf{0})\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$. This span equals $\boldsymbol{{\mathcal V}}_\vtg(n)$ by \cite[(4.2.1)]{DF09} (i.e., \ref{B(bfl,r)A(bfj,r)}(1) below). \end{proof} For $T=(t_{i,j})\in\Theta_\vtg(\cycn)$ let ${\delta}_T$ be the diagonal of $T$, i.e., $${\delta}_T=(t_{i,i})_{i\in\mathbb Z}\in\mathbb N_\vtg^{\cycn}.$$ We now use \ref{[B][A]} to derive multiplication formulas of an arbitrary basis element by a ``semisimple generators'', which is the key to solving the realisation problem. Recall the notation in \eqref{A^+,A^-,A^0}. \begin{Prop}\label{B(bfl,r)A(bfj,r)} Let ${\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}$, $A\in\Theta_\vtg^\pm(\cycn)$, $\alpha\in\mathbb N_\vtg^{\cycn}$, and $S_\alpha=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_iE^\vartriangle_{i,i+1}$. The following identities holds in $\boldsymbol{{\mathcal V}}_\vtg(n)$: \begin{itemize} \item[(1)] $0({\mathbf{j}}')A({\mathbf{j}})=\boldsymbol{v}^{{\mathbf{j}}'\centerdot\text{\rm ro}(A)}A({\mathbf{j}}'+{\mathbf{j}})$ and $A({\mathbf{j}})0({\mathbf{j}}')=\boldsymbol{v}^{{\mathbf{j}}'\centerdot\text{\rm co}(A)}A({\mathbf{j}}'+{\mathbf{j}})$ {\rm(\cite[(4.2.1)]{DF09})}. \item[(2)] $\displaystyle S_\alpha(\mathbf{0})A({\mathbf{j}})= \sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{f_{A,T}}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i} \overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}(A+T^\pm-\widetilde T^\pm)({\mathbf{j}}_T,{\delta}_T)$,\\ where ${\mathbf{j}}_T={\mathbf{j}}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}(\sum_{j<i}(t_{i,j}-t_{i-1,j}))\boldsymbol e^\vartriangle_i$ and\vspace{-1ex} \begin{equation*} \begin{split} f_{A,T}&=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i} a_{i,j}t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i+1}a_{i+1,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i}t_{i-1,j}t_{i,l}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i,\,j\not=i+1}t_{i,j}t_{i,l}\\ &\qquad+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j<i+1}t_{i,j}t_{i+1,i+1}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}j_i(t_{i-1,i}-t_{i,i}); \end{split} \end{equation*} \item[(3)] $\displaystyle {}^t\!S_\alpha(\mathbf{0})A({\mathbf{j}})= \sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{f_{A,T}'}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i} \overline{\left[\!\!\left[{a_{i,j}-t_{i,j}+t_{i-1,j}\atop t_{i-1,j}}\right]\!\!\right]}(A-T^\pm+\widetilde T^\pm)({\mathbf{j}}'_T,{\delta}_{\widetilde T})$,\\ where ${\mathbf{j}}'_T={\mathbf{j}}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}(\sum_{j>i}(t_{i-1,j}-t_{i,j}))\boldsymbol e^\vartriangle_i$ and \begin{equation*} \begin{split} f_{A,T}'&=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop l\geq j,\,j\not=i} a_{i,j}t_{i-1,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop l>j,\,j\not=i}a_{i,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,l\not=i}t_{i-1,j}t_{i,l}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,l\not=i,\,l\not=i+1}t_{i,j}t_{i,l}\\ &\qquad+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i<j}t_{i,j}t_{i-1,i}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}j_i(t_{i,i}-t_{i-1,i}). \end{split} \end{equation*} \end{itemize} In particular, $\boldsymbol{{\mathcal V}}_\vtg(n)$ is closed under the multiplication by the ``generators'' $0({\mathbf{j}}), S_\alpha(\mathbf{0}), {}^t\!S_\alpha(\mathbf{0})$ for all ${\mathbf{j}}\in\mathbb N$ and $\alpha\in\mathbb N^{n}$. \end{Prop} \begin{proof} We only prove (2). If $r<\sigma(A)$, the $r$-th components of both sides are 0. Assume now $r\geq\sigma(A)$. By \ref{[B][A]} and noting the fact that, for $X,Y\in\Theta_\vtg(\cycn,r)$, $[X][Y]\neq0\implies \text{\rm co}(X)=\text{\rm ro}(Y)$ the $r$-th component of $S_\alpha(\mathbf{0})A({\mathbf{j}})$ becomes \begin{equation*} \begin{split} S_\alpha(\mathbf{0},r)A({\mathbf{j}},r)&=\sum_{{\gamma}\in\Lambda_\vtg(n,r-\sigma}\newcommand{\vsg}{\varsigma(A))}v^{{\gamma}\centerdot{\mathbf{j}}} \bigg[S_\alpha+\operatorname{diag}\bigg({\gamma}+\text{\rm ro}(A)-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}\alpha_i\boldsymbol e^\vartriangle_{i+1}\bigg)\bigg][A+\operatorname{diag}({\gamma})]\\ &=\sum_{T\in\Theta_\vtg(\cycn)\atop\text{\rm ro}(T)=\alpha} \prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i}\overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}x_T \end{split} \end{equation*} where $$x_T= \sum_{{\gamma}\in\Lambda_\vtg(n,r-\sigma}\newcommand{\vsg}{\varsigma(A))} v^{{\gamma}\centerdot{\mathbf{j}}+\beta(T,A+\operatorname{diag}({\gamma}))}\overline{\left[\!\!\left[{\gamma}+{\delta}_T-{\delta}_{\widetilde T}\atop{\delta}_T)\right]\!\!\right]} [A+\operatorname{diag}({\gamma})+T-\widetilde T]. $$ Let $A+\operatorname{diag}({\gamma})=(a_{i,j}^{\gamma})$. Then $a_{i,j}^{\gamma}=a_{i,j}$ for $i\not=j$ and $a_{i,i}^{\gamma}={\gamma}_i$. Let $\nu={\gamma}+{\delta}_T-{\delta}_{\widetilde T}$. Then \begin{equation*} \begin{split} \beta(T,A+\operatorname{diag}({\gamma}))&=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l}(a_{i,j}^{\gamma}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l}(a_{i+1,j}^{\gamma}-t_{i,j})t_{i,l}\\ &=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i}(a_{i,j}-t_{i-1,j})t_{i,l}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i\geq l}(\nu_i-t_{i,i})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i+1}(a_{i+1,j}-t_{i,j})t_{i,l}\\ &\qquad -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i+1>l}(\nu_{i+1}-t_{i+1,i+1})t_{i,l}\\ &=\beta_{A,T}+\beta_{\nu,T}, \end{split} \end{equation*} where $\beta_{\nu,T}=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,i\geq l}\nu_it_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n,\,i+1>l}\nu_{i+1}t_{i,l}$ and \begin{equation*} \begin{split} \beta_{A,T}&=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i}(a_{i,j}-t_{i-1,j})t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i+1}a_{i+1,j}t_{i,l}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i,i+1}t_{i,j}t_{i,l}\\ &\qquad -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n}t_{i,i}^2+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i+1>l}t_{i+1,i+1}t_{i,l}. \end{split} \end{equation*} Clearly, we have $\overline{\big[\!\!\big[{\nu\atop{\delta}_T}\big]\!\!\big]}=v^{{\delta}_T\centerdot({\delta}_T-\nu)}\big[{\nu\atop {\delta}_T}\big]$, $\beta_{A,T}+{\delta}_T\centerdot{\delta}_T+{\mathbf{j}}\centerdot({\delta}_{\widetilde T}-{\delta}_T)=f_{A,T}$ and $\beta_{\nu,T}+\nu\centerdot({\mathbf{j}}-{\delta}_T)=\nu\centerdot{\mathbf{j}}_T$. This implies that \begin{equation*} \begin{split} x_T&=v^{\beta_{A,T}+{\delta}_T\centerdot{\delta}_T+{\mathbf{j}}\centerdot({\delta}_{\widetilde T}-{\delta}_T)} \sum_{\nu\in\Lambda_\vtg(n,r-\sigma}\newcommand{\vsg}{\varsigma(A+T^\pm-\widetilde T^\pm))}v^{\beta_{\nu,T}+\nu\centerdot({\mathbf{j}}-{\delta}_T)}\left[{\nu\atop{\delta}_T}\right]\\ &\qquad\times [A+T^\pm-\widetilde T^\pm+\operatorname{diag}(\nu)]\\ &=v^{f_{A,T}}(A+T^\pm-\widetilde T^\pm)({\mathbf{j}}_T,{\delta}_T,r), \end{split} \end{equation*} proving (2). \end{proof} For $A,B\in\Theta_\vtg^\pm(\cycn)$, define the ordering on $\Theta_\vtg(\cycn)$ by setting \begin{equation}\label{order2} A\preceq B\iff \sum\limits_{s\leqslant}\def\geq{\geqslant i,t\geq j}a_{s,t}\leqslant}\def\geq{\geqslant \sum\limits_{s\leqslant}\def\geq{\geqslant i,t\geq j}b_{s,t},\,\forall i<j,\text{ and } \sum\limits_{s\geq i,t\leqslant}\def\geq{\geqslant j}a_{s,t}\leqslant}\def\geq{\geqslant \sum\limits_{s\geq i,t\leqslant}\def\geq{\geqslant j}a_{s,t},\,\forall i>j. \end{equation} \begin{Prop}\label{triangular formula in A(bfj)} With the notation in \eqref{A^+,A^-,A^0} we have, for any $A\in\Theta_\vtg^\pm(\cycn)$ and ${\mathbf{j}}\in\mathbb N_\vtg^{\cycn}$, \[ A^+(\mathbf{0})0({\mathbf{j}})A^-(\mathbf{0})=v^{{\mathbf{j}}\centerdot(\text{\rm co}(A^+)+\text{\rm ro}(A^-))}A({\mathbf{j}})+ \sum_{B\in\Theta_\vtg^\pm(\cycn) \atop B\prec A,\,{\mathbf{j}}'\in\mathbb N_\vtg^{\cycn}}f_{A,{\mathbf{j}}}^{B,{\mathbf{j}}'}B({\mathbf{j}}'), \] where $f_{A,{\mathbf{j}}}^{B,{\mathbf{j}}'}\in\mathbb Q(v)$. \end{Prop} \begin{proof} Let ${\boldsymbol{\mathfrak D}^+_\vtg}(n)$ be the subspace of ${\boldsymbol{\mathfrak D}_\vtg}(n)$ spanned by the elements $u_A^+$ for $A\in\Theta_\vtg^+(\cycn)$. According to \cite[6.2]{DDX}, the algebra ${\boldsymbol{\mathfrak D}^+_\vtg}(n)$ is generated by the elements $\widetilde u_{S_\alpha}^+$ for $\alpha\in\mathbb N_\vtg^{\cycn}$, where $S_\alpha$ is defined as in \eqref{semisimple}. This together with \ref{zr} implies that $A^+(\mathbf{0})$ can be written as a linear combination of monomials in $S_\alpha(\mathbf{0})$. Thus, by \ref{B(bfl,r)A(bfj,r)} and \ref{spanning set of afbfVn}, we conclude that there exist $f_{A,{\mathbf{j}}}^{B,{\mathbf{j}}'}\in\mathbb Q(v)$ (independent of $r$) such that \begin{equation}\label{eq1 triangular formula in A(bfj)} A^+(\mathbf{0})0({\mathbf{j}})A^-(\mathbf{0})=\sum_{B\in\Theta_\vtg^\pm(\cycn) \atop {\mathbf{j}}'\in\mathbb N_\vtg^{\cycn}}f_{A,{\mathbf{j}}}^{B,{\mathbf{j}}'}B({\mathbf{j}}'). \end{equation} On the other hand, by the triangular relation given in \cite[3.7.3]{DDF}, we have $$ A^+(\mathbf{0},r)0({\mathbf{j}},r)A^-(\mathbf{0},r)=v^{{\mathbf{j}}\centerdot(\text{\rm co}(A^+)+\text{\rm ro}(A^-))}A({\mathbf{j}},r)+f $$ where $f$ is a $\mathbb Q(v)$-combination of $[B]$ with $B\in\Theta_\vtg(\cycn,r)$ and $B\prec A$. Combining this with \eqref{eq1 triangular formula in A(bfj)} proves the assertion. \end{proof} The maps $\zeta_r$ given in \ref{zr} induce an algebra homomorphism $$\zeta=\prod_{r\geq 1}\zeta_r:{\boldsymbol{\mathfrak D}_\vtg}(n)\longrightarrow{\boldsymbol{\mathcal S}}_\vtg(n)=\prod_{r\geq 1}{\boldsymbol{\mathcal S}}_\vtg(\cycn,r).$$ We now prove the conjecture formulated in \cite[5.5(2)]{DF09}. \begin{Thm}\label{realization} The $\mathbb Q(v)$-space $\boldsymbol{{\mathcal V}}_\vtg(n)$ is a subalgebra of ${\boldsymbol{\mathcal S}}_\vtg(n)$ with $\mathbb Q(v)$-basis ${\mathfrak B}$. Moreover, the map $\zeta$ is injective and induces a $\mathbb Q(v)$-algebra isomorphism ${\boldsymbol{\mathfrak D}_\vtg}(n)\overset\zeta\cong\boldsymbol{{\mathcal V}}_\vtg(n)$. \end{Thm} \begin{proof} According to \cite[4.1]{DF09}, the set $\frak B$ forms a $\mathbb Q(v)$-basis for $\boldsymbol{{\mathcal V}}_\vtg(n)$. This together with \ref{triangular formula in A(bfj)} implies that the set $\{A^+(\mathbf{0})0({\mathbf{j}})A^-(\mathbf{0})\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$ forms another basis for $\boldsymbol{{\mathcal V}}_\vtg(n)$. Note that, by \ref{presentation-dbfHa}(2), the set $\{\widetilde u_{A^+}^+K^{\mathbf{j}}\widetilde u_{{}^t\!(A^-)}^-\mid A\in\Theta_\vtg^\pm(\cycn),\,{\mathbf{j}}\in\mathbb Z_\vtg^{\cycn}\}$ forms a $\mathbb Q(v)$-basis for ${\boldsymbol{\mathfrak D}_\vtg}(n)$. Furthermore, by \ref{zr}, $$\zeta(\widetilde u_{A^+}^+K^{\mathbf{j}}\widetilde u_{{}^t\!(A^-)}^-)=A^+(\mathbf{0})0({\mathbf{j}})A^-(\mathbf{0}).$$ Thus, $\zeta$ takes a basis for ${\boldsymbol{\mathfrak D}_\vtg}(n)$ onto the basis for $\boldsymbol{{\mathcal V}}_\vtg(n)$. It follows that $\zeta$ is injective and $\zeta({\boldsymbol{\mathfrak D}_\vtg}(n))=\boldsymbol{{\mathcal V}}_\vtg(n)$. The proof is completed. \end{proof} Now, the Main Theorem \ref{MThm} follows immediately. We end the paper with an application to the (untwisted) Ringel--Hall algebra of a cyclic quiver. Let $\boldsymbol{{\mathcal V}}_\vtg^+(n)$ be the subalgebra of $\boldsymbol{{\mathcal V}}_\vtg(n)$ spanned by $A(\mathbf{0})$ for all $A\in\Theta_\vtg^+(\cycn)$. Then, by \ref{zr}, the map sending $\widetilde u_A$ to $A(\mathbf{0})$ is an algebra isomorphism from ${\boldsymbol{\mathfrak D}^+_\vtg}(n)={\boldsymbol{\mathfrak H}_\vtg(n)}$ to $\boldsymbol{{\mathcal V}}_\vtg^+(n)$. In particular, the formula \ref{B(bfl,r)A(bfj,r)}(2) gives the following multiplication formula in the Ringel--Hall algebra ${{\mathfrak H}_\vtg(n)}$ over ${\mathcal Z}$: $$\widetilde u_\alpha\widetilde u_A= \sum_{T\in\Theta_\vtg^+(\cycn)\atop\text{\rm ro}(T)=\alpha}v^{f_{A,T}}\prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i} \overline{\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}\widetilde u_{A+T-\widetilde T^+}$$ where $$f_{A,T}=\displaystyle\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i} a_{i,j}t_{i,l}-\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i+1}a_{i+1,j}t_{i,l} -\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\geq l,\,j\not=i}t_{i-1,j}t_{i,l}+\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j>l,\,j\not=i,\,j\not=i+1}t_{i,j}t_{i,l}.$$ Untwisting the multiplication for ${{\mathfrak H}_\vtg(n)}$ yields the following. \begin{Thm} The Ringel--Hall algebra ${{\mathfrak H}_\vtg(n)}^\diamond$ is the algebra over the polynomial ring $\mathbb Z[{\boldsymbol{q}}]$ (${\boldsymbol{q}}=\boldsymbol{v}^2$) which is spanned by the basis $\{u_A\mid A\in\Theta_\vtg^+(\cycn)\}$ and generated by $\{u_\alpha\mid\alpha\in\mathbb N_\vtg^{\cycn}\}$ and whose (untwisted) multiplication is given by the formulas: for any $A\in\Theta_\vtg^+(\cycn)$ and $\alpha\in\mathbb N_\vtg^{\cycn}$, $$ u_\alpha \diamond u_A \sum_{T\in\Theta_\vtg^+(\cycn)\atop\text{\rm ro}(T)=\alpha} {\boldsymbol{q}}^{\sum_{1\leqslant}\def\geq{\geqslant i \leqslant}\def\geq{\geqslant n,\,l< j}(a_{i,j}t_{i,l}-t_{i,j}t_{i+1,l})} \prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i} {\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}u_{A+T-\widetilde T^+}$$ \end{Thm} In other words, if the Hall polynomial $\varphi_{S_\alpha,A}^B$ is nonzero, then there exists $T=(t_{i,j})\in\Theta_\vtg^+(\cycn)$ with $\text{\rm ro}(T)=\alpha$ such that $B=A+T-\widetilde T^+$ and $$\varphi_{S_\alpha,A}^{A+T-\widetilde T^+}={\boldsymbol{q}}^{\sum_{1\leqslant}\def\geq{\geqslant i \leqslant}\def\geq{\geqslant n,\,l< j}(a_{i,j}t_{i,l}-t_{i,j}t_{i+1,l})} \prod_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop j\in\mathbb Z,\,j\not=i} {\left[\!\!\left[{a_{i,j}+t_{i,j}-t_{i-1,j}\atop t_{i,j}}\right]\!\!\right]}.$$ Note that, when $\alpha$ defines a simple module or $A$ defines another semisimple module, the formula coincides with the formulas given in \cite[Th.~5.4.1]{DDF} (built on \cite[Th.~4.2]{DF09}) and \cite[Cor.~1.5]{DDM}. \def{\text{Inv}}{{\text{Inv}}} \section{Appendix --- Proof of Lemma \ref{length of elements in Dla}(2)} For $w\in{{\frak S}_{{\!\vartriangle\!},r}}$ and $t\in\mathbb Z$, let \begin{equation}\label{Inv(w)} \aligned \text{Inv}(w,t)&=\{(i,j)\in\mathbb Z^2\mid1+t\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant r+t,\ i<j,\ w(i)>w(j)\}\\ {\text{Inv}}(w)&={\text{Inv}}(w,0).\endaligned \end{equation} Then the number of inversions $\ell'(w):=|\text{Inv}(w,t)|$ is clearly independent of $t$. \begin{Prop}\label{inversion} If $w=ys$ with $y,w\in{{\frak S}_{{\!\vartriangle\!},r}}$ and $s\in S$ satisfies $\ell(w)=\ell(y)+1$, then $\ell'(w)=\ell'(y)+1$. \end{Prop} \begin{proof}Suppose $s=s_{i_0}$ for some $1\leqslant}\def\geq{\geqslant i_0\leqslant}\def\geq{\geqslant r$. By the hypothesis and \cite[4.2.3]{Shi86}, we have $y(i_0)<y(i_0+1)$. Fix $t\in\{0,1\}$ with $i_0, i_0+1\in[1+t,r+t]$. Let ${\mathcal W}=\text{Inv}(w,t)\text{ and }{\mathcal Y}=\text{Inv}(y,t).$ We want to prove that $|{\mathcal W}|=|{\mathcal Y}|+1$. For $j\in\mathbb Z$ and $i\in[i_0+1,i_0+r]$, let $$c(i_0,i,j):=|\{i_0,i_0+1\}\cap\{i,\bar j\}|,$$ where $\bar j$ denote the unique integer in $[1+t,r+t]$ such that $j\equiv\bar j\mod r$. For each $x\in\{0,1,2\}$, let $${\mathcal W}_x=\{(i,j)\in{\mathcal W}\mid c(i_0,i,j)=x\}\text{ and } {\mathcal Y}_x=\{(i,j)\in{\mathcal Y}\mid c(i_0,i,j)=x\}.$$ Then we have disjoint unions ${\mathcal W}={\mathcal W}_0\cup{\mathcal W}_1\cup{\mathcal W}_2$ and ${\mathcal Y}={\mathcal Y}_0\cup{\mathcal Y}_1\cup{\mathcal Y}_2$. For $j\in\mathbb Z$ and $i\in[i_0+1,i_0+r]$, if $c(i_0,i,j)=0$, then $w(i)=y(i)$ and $w(j)=y(j)$. This implies ${\mathcal W}_0={\mathcal Y}_0$. Hence, $|{\mathcal W}_0|=|{\mathcal Y}_0|$. Since $y(i_0)<y(i_0+1)$, it follows that $${\mathcal W}_2=\{(i_0,i_0+1+kr)\mid k\in\mathbb Z_{\geq0}, y(i_0+1)>y(i_0)+kr\},$$ while $${\mathcal Y}_2=\{(i_0+1,i_0+kr)\mid k\in\mathbb Z_{>0}, y(i_0+1)>y(i_0)+kr\}.$$ Hence, $|{\mathcal W}_2|=|{\mathcal Y}_2|+1$. It remains to prove that $|{\mathcal W}_1|=|{\mathcal Y}_1|$. In this case, we have $${\mathcal W}_1={\mathcal W}_{1,(i_0,\bullet)}\cup {\mathcal W}_{1,(i_0+1,\bullet)}\cup{\mathcal W}_{1,(\bullet,i_0)}\cup{\mathcal W}_{1,(\bullet,i_0+1)},$$ where ${\mathcal W}_{1,(i_0,\bullet)}=\{(i,j)\in{\mathcal W}_1\mid i=i_0\}$, etc. Define ${\mathcal Y}_{1,(\bullet,\bullet)}$ similarly to get a similar partition for ${\mathcal Y}_1$. Then the condition $y(i_0)<y(i_0+1)$ implies the following $$\aligned {\mathcal Y}_{1,(i_0,\bullet)}\subseteq {\mathcal W}_{1,(i_0,\bullet)},\qquad{\mathcal Y}_{1,(i_0+1,\bullet)}\supseteq {\mathcal W}_{1,(i_0+1,\bullet)},\\ {\mathcal Y}_{1,(\bullet,i_0)}\supseteq {\mathcal W}_{1,(\bullet,i_0)},\qquad{\mathcal Y}_{1,(\bullet,i_0+1)}\subseteq {\mathcal W}_{1,(\bullet,i_0+1)}. \endaligned $$ But, since $$\aligned {\mathcal W}_{1,(i_0,\bullet)}\backslash {\mathcal Y}_{1,(i_0,\bullet)}&=\{(i_0,j)\mid i_0<j,j\in\mathbb Z,\bar j\not\in\{i_0,i_0+1\},y(i_0)<y(j)<y(i_0+1)\}\text{ and}\\ {\mathcal Y}_{1,(i_0+1,\bullet)}\backslash {\mathcal W}_{1,(i_0+1,\bullet)}&=\{(i_0+1,j)\mid i_0+1<j,j\in\mathbb Z,\bar j\not\in\{i_0,i_0+1\},y(i_0)<y(j)<y(i_0+1)\},\\ \endaligned $$ it follows that $|{\mathcal W}_{1,(i_0,\bullet)}|+|{\mathcal W}_{1,(i_0+1,\bullet)}|=|{\mathcal Y}_{1,(i_0,\bullet)}|+|{\mathcal Y}_{1,(i_0+1,\bullet)}|$. Similarly, one proves that $|{\mathcal W}_{1,(\bullet,i_0)}|+|{\mathcal W}_{1,(\bullet,i_0+1)}|=|{\mathcal Y}_{1,(\bullet,i_0+1)}|+|{\mathcal Y}_{1,(\bullet,i_0)}|$. This completes the proof. \end{proof} The following result given in \cite[(3.2.1.1)]{DDF} without proof follows immediately. \begin{Coro}\label{combinatorial description of length} For $w\in{{\frak S}_{{\!\vartriangle\!},r}}$ we have $\ell(w)=\ell'(w)$. \end{Coro} We now generalise the construction for the shortest representatives of double cosets of the symmetric group \cite[\S3]{Du} to the affine case. For $A\in\Theta_\vtg(\cycn,r)$ with ${\lambda}=\text{\rm ro}(A)$ and $\mu=\text{\rm co}(A)$, define a pseudo matrix $A^-$ as follows: the entry $a_{i,j}$ is replaced by the sequence $$\underline{c}_{i,j}=\underline{c}_{i,j}(A)=\bigg({\lambda}_{k_0,i_0-1}+\sum_{t\leqslant}\def\geq{\geqslant j-1}a_{i,t}+1,\cdots,{\lambda}_{k_0,i_0-1}+\sum_{t\leqslant}\def\geq{\geqslant j-1}a_{i,t}+(a_{i,j}-1) ,{\lambda}_{k_0,i_0-1}+\sum_{t\leqslant}\def\geq{\geqslant j}a_{i,t}\bigg)$$ where $i=i_0+k_0n,j\in\mathbb Z$ with $1\leqslant}\def\geq{\geqslant i_0\leqslant}\def\geq{\geqslant n$ and $k_0\in\mathbb Z$. We define $\widetilde y_{A}\in{{\frak S}_{{\!\vartriangle\!},r}}$ by $$\widetilde y_{A}(i+kr)=a_i+kr,\text{ for all }1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant r,k\in\mathbb Z,$$ where $(a_1,a_2,\cdots,a_r)$ is the sequence obtained by reading the numbers in column 1 inside the subsequences from left to right and from top to bottom, and then in column 2, etc., and then in column $n$. In other words, it is the sequence obtained by ignoring 0's from $((\underline{c}_{k,1})_{k\in\mathbb Z}, (\underline{c}_{k,2})_{k\in\mathbb Z},\cdots,(\underline{c}_{k,n})_{k\in\mathbb Z})$ with $(\underline{c}_{k,i})_{k\in\mathbb Z}=(\cdots,\underline{c}_{1,i},\underline{c}_{2,i},\cdots,\underline{c}_{n,i},\cdots)$. We are ready to prove Lemma \ref{length of elements in Dla}(2). \begin{Prop} Let $A\in\Theta_\vtg(\cycn,r)$. Then $\widetilde y_A$ is the shortest representative of the double coset defined by $A$, i.e., $y_A=\widetilde y_A$, and $$\ell(y_A)=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i<k;j>l}a_{ij}a_{kl}.$$ \end{Prop} \begin{proof} Let ${\lambda}=ro(A)$ and $\mu=co(A)$. For $w\in{\frak S}_{\lambda} y_A{\frak S}_\mu$, let $${\mathcal N}=\bigcup_{1\leqslant}\def\geq{\geqslant l\leqslant}\def\geq{\geqslant n\atop i<k;j>l}(R_i^{\lambda}\cap w(R_j^\mu))\times(R_k^{\lambda}\cap w(R_l^\mu))\text{ and }N=|{\mathcal N}|=\sum_{1\leqslant}\def\geq{\geqslant i\leqslant}\def\geq{\geqslant n\atop i<k;j>l}a_{ij}a_{kl}.$$ Then there is a injective map $\varphi_w$ defined as follows: \begin{equation*} \varphi_{w}:{\mathcal N}\longrightarrow{\text{Inv}}(w),\;\;(c,d)\longmapsto (w^{-1}(d),w^{-1}(c)). \end{equation*} Hence by \ref{combinatorial description of length} we have $N\leqslant}\def\geq{\geqslant \ell(w)$. In particular, we have $\ell(y_A)\geq N$. For $i,j\in\mathbb Z$ let $C_{ij}(A)$ denote the set of members of the sequence $\underline{c}_{ij}(A)$. By definition we have, for $i\in\mathbb Z$ and $1\leqslant}\def\geq{\geqslant j\leqslant}\def\geq{\geqslant n$, $$R_i^{\lambda}=\bigcup_{l\in\mathbb Z}C_{il}(A),\quad\text{and}\quad \widetilde y{_A}(R_j^\mu)=\bigcup_{k\in\mathbb Z}C_{kj}(A).$$ It is easy to see that $C_{ij}(A)+tr=C_{i+tn,j+tn}(A)$ for $i,j,t\in\mathbb Z$. Hence, for $j=j_0+tn$ with $1\leqslant}\def\geq{\geqslant j_0\leqslant}\def\geq{\geqslant n$ and $t\in\mathbb Z$, $$\widetilde y{_A}(R_j^\mu)=tr+ w{_A}(R_{j_0}^\mu)=tr+\bigcup_{k\in\mathbb Z}C_{kj_0}(A)=\bigcup_{k\in\mathbb Z}C_{k+tn,j_0+tn}(A)=\bigcup_{k\in\mathbb Z}C_{k,j}(A).$$ Thus, we have $C_{ij}(A)=R_i^{\lambda}\cap \widetilde y{_A}(R_j^\mu)$ for $i,j\in\mathbb Z$ and so $a_{ij}=|R_i^{\lambda}\cap \widetilde y{_A}(R_j^\mu)|$ for $i,j\in\mathbb Z$. This implies that $\widetilde y_A\in{\frak S}_{\lambda} y_A{\frak S}_\mu$ and, hence, $\ell(\widetilde y_A)\geq\ell(y_A)\geq N$. Observe that $\widetilde y{_A}(C_{ji}({}^t\!A))=C_{ij}(A)$ for $i,j\in\mathbb Z$, where ${}^t\!A$ is the transpose matrix of $A$. We now prove that $\ell(\widetilde y_A)=N$ by showing that $\varphi_{\widetilde y{_A}}$ is surjective. Let $(a,b)\in{\text{Inv}}(\widetilde y{_A})$. Since $\mathbb Z=\bigcup_{s,t\in\mathbb Z}C_{s,t}(A)$, there exist $i,j,k,l\in\mathbb Z$ such that $\widetilde y{_A}(a)\in C_{kl}(A)$ and $\widetilde y{_A}(b)\in C_{ij}(A)$. Since $1\leqslant}\def\geq{\geqslant a\leqslant}\def\geq{\geqslant r$ we have $1\leqslant}\def\geq{\geqslant l\leqslant}\def\geq{\geqslant n$. Since $\widetilde y{_A}(a)>\widetilde y{_A}(b)$ we have either $i<k$ or $i=k$ and $j<l$. On the other hand, the conditions $a\in \widetilde y{_A}^{-1}(C_{kl}(A))=C_{lk}({}^t\!A)$, $b\in \widetilde y{_A}^{-1}(C_{ij}(A))=C_{ji}({}^t\!A)$ and $a<b$ imply either $l<j$ or $l=j$ and $k<i$. Hence, we must have $i<k$ and $l<j$. Therefore, the map $\varphi_{\widetilde y{_A}}$ is bijective, proving $\ell(\widetilde y{_A})=N$. \end{proof}
\section{Introduction} Complex networks are used to represent and analyze a wide range of systems~\cite{Dorogovtsev2002,Boccaletti2006,Newman2010}. Models of complex networks usually aim for simplicity and attempt to keep the number of parameters as low as possible. However, real data is more complex than any simple model which makes it difficult to draw clear links between data and models. To capture the increasingly available massive real data~\cite{Gonzales2007}, we need high-dimensional models where the number of parameters grows with the number of nodes. An example of such a model is the latent space model~\cite{Hoff02} where nodes are assigned independent and identically distributed vectors and the probability of a link connecting two nodes depends only on the distance of their vectors. While there are plenty of simple (and not so simple) network models, little is known as to which of them are really supported by data. While calibration of complex network models often uses standard statistical techniques, their validation is typically based on comparing their aggregate features (such as the degree distribution or clustering coefficient---see~\cite{Costa2007,Kolaczsyk2009} for detailed accounts on network measurements) with what is seen in real networks (see~\cite{Papa2012,Li2013} for recent examples of this approach). The focus on aggregate quantities naturally reduces the discriminative power of model validation which is often further harmed by the use of inappropriate statistical methods~\cite{Stumpf12}. As a result, we still lack knowledge of what is to date the best model explaining the growth of the scientific citation network, for example. We argue that network models need to be evaluated by robust statistical methods~\cite{Bos07,Freedman}, especially by those that are suited to high-dimensional models~\cite{Buehlmann2011}. This is exemplified in~\cite{Leskovec08} where various low-dimensional microscopic mechanisms for evolution of social networks are compared on the basis of their likelihood of generating the observed data. Prohibitive computational complexity of maximum likelihood estimation is often quoted as a reason for its limited use in the study of real world complex networks~\cite{Leskovec10}. However, as we shall see here, even small subsets of data allow to discriminate between models and point clearly to those that are actually supported by the data. This, together with the ever-increasing computational power at our disposal, opens the door to the likelihood analysis of complex network models. We analyze here a recent network growth model~\cite{Medo2011} which naturally leans itself to high-dimensional analysis. This model generalizes the classical preferential attachment (PA; often referred to as the Barabási-Albert model in the complex networks literature) \cite[Sections~7,~8]{Albert2002} by introducing node relevance which decays in time and co-determines (together with node degree) the rate at which nodes acquire new links. If either the initial relevance values or the functional form of the relevance decay are heterogeneous among the nodes, this model is able to produce various realistic degree distributions. By contrast to~\cite{Eom2011} which modifies preferential attachment by introducing an additive heterogeneous term, in~\cite{Medo2011} relevance combines with degree in a multiplicative way which means that once it reaches zero, the degree growth stops. This makes the model an apt candidate for modeling information networks where information items naturally lose their pertinence with time and the growth of their degree eventually stops. (See~\cite{Holme2012} for a review of work on temporal networks.) This model has been recently used to quantify and predict citation patterns of scientific papers~\cite{Wang2013}. Before methods for high-dimensional parameter estimation are applied to real data, we calibrate and evaluate them on artificial data where one has full control over global network parameters (size, average degree, etc.) and true node parameter values are known. For simplicity, we limit our attention to the case where the functional form of relevance decay is the same for all nodes and only the initial relevance values differ. We present here various estimation methods and evaluate their performance. Plain maximum likelihood~\cite[Chapter~7]{Freedman} produces unsatisfactory results, especially in the case of sparse networks which are commonly seen in practice. We enhance the method by introducing an additional term which suppresses undesired correlation between node age and estimates of initial relevance. We then introduce a mean-field approach which allows us to reduce high-dimensional estimation to a low-dimensional one. Calibration and evaluation of these parameter-estimation methods is done on artificial data. Real data is then used to employ the established framework and compare the statistical evidence for several low- and high-dimensional network models on the given data. Analysis of small subsets of input data is shown to efficiently discriminative among the available models. Since this work focuses on model evaluation, estimated parameter values are thus of secondary importance to us. Necessary conditions for obtaining precise estimates and the potential risk of large errors~\cite{Owhadi2013} are therefore left for future research (see Sec.~\ref{sec:conclusions}). \section{Model} \label{sec:model} The original model of preferential attachment with relevance decay (PA-RD) has been formulated for an undirected network where the initial node degree is non-zero because of links created by the node on its arrival~\cite{Medo2011}. To allow zero-degree nodes to collect links, some additive attractiveness or random node selection need to be introduced. When these two mechanisms are combined with PA-RD, the probability that a new link created at time $t$ attaches to node $i$ can be written as \begin{equation} \label{PARD} P(i, t) = \lambda\,\frac{R_i(t)\big(k_i(t)+A\big)}{\sum_{j=1}^{n(t)} R_j(t)\big(k_j(t)+A\big)} + \frac{1-\lambda}{n(t)}. \end{equation} Here $k_i(t)$ and $R_i(t)$ are degree and relevance of node $i$ at time $t$, respectively, $n(t)$ is the number of nodes present at time $t$, and $A$ is the additive attractiveness term. Finally, $\lambda$ is the probability that the node is chosen by the PA-RD mechanism; the node is chosen at random with the complementary probability $1-\lambda$. When $A=0$ and $\lambda=1$, a node of zero degree will never attract new links. \eref{PARD} can be used to model a monopartite network where nodes link to each other as well as a bipartite network where one set of nodes is unimportant and we can thus speak of outside links attaching to nodes. For example, one can use the model to describe the dynamics of item popularity in a user-item bipartite network representing an e-commerce system~\cite{Ming-Sheng10}. There are now two points to make. Firstly, the model is invariant with respect to the rescaling of all relevance values, $R_i(t)\to \xi R_i(t)$. This may lead to poor convergence of numerical optimization schemes because $R_i(t)$ values can drift in accord without affecting the likelihood value. The convergence problems can be avoided by imposing an arbitrary normalization constraint on the relevance values as we do below. Secondly, $A$ and $\lambda$ act in the same direction: they introduce randomness in preferential attachment-driven network growth (in particular, as $A\to\infty$ and/or $\lambda\to0$, preferential attachment loses all influence). One can therefore expect that $A$ and $\lambda$ are difficult to be simultaneously inferred from the data. This is especially true for the original preferential attachment without decaying relevance. If node relevance decays to zero, node attraction due to $A$ eventually vanishes while the random-attachment part proportional to $\lambda$ remains---it is therefore possible, at least in principle, to distinguish between the two effects. To better focus on the high-dimensional likelihood maximization of node parameters, we assume $\lambda=1$ in all our simulations. The PA-RD model has been solved in~\cite{Medo2011} for a case where $\lambda=1$, $A=0$, and the initial degree of all nodes equal to one. It was further assumed that $T_i:=\int_0^{\infty} R_i(t)\,\mathrm{d} t$ is finite for all nodes and the distribution of $T$ values among the nodes, $\varrho(T)$, decays exponentially or faster. The probability normalization term $\sum_j R_j(t)(k_j(t)+A)$ then eventually fluctuates around a stationary value $\Omega^*$ and the expected final degree of node $i$ can be written as $\avg{k_i^F}=\exp(T_i/\Omega^*)$. It has been shown that the network's degree distribution, shaped mainly by $\varrho(T)$, can take on various forms including exponential, log-normal, and power-law. \subsection{Description of artificial data} \label{sec:artif_data} We begin by describing bipartite network data with temporal information. We consider a simplified bipartite case where links arrive from outside and thus only their target nodes matter---see Fig.~\ref{fig:network}a for illustration. Links are numbered with $l=1,\dots,E$ and the times at which they are introduced are $t_1\leq t_2\leq\cdots\leq t_E$. Nodes are numbered with $i=1,\dots,N$ and the times at which they are introduced are $\tau_1\leq\tau_2\leq\cdots\leq\tau_N$. At time $t$, there are $n(t)$ target nodes in the network. Degree of node $i$ at time $t_l$ when link $l$ is added is $k_i(t_l)$ and the target node of link $l$ is $n_l$. The average node degree is $z:=E/N$ (the factor of two is missing here because we consider a bipartite network where $E$ edges point to $N$ nodes of interest). We use the PA-RD model to create artificial networks with well-defined properties. There are initially $n_I$ nodes with zero degree. After every $\Delta T$ time steps, a new node of zero degree is introduced in the network. In each time step, one new link is created and chooses its target node according to \eref{PARD}. The network growth stops once there are $E_F = \lceil zn_I / (1-z/\Delta T)\rceil$ links and $n_F = n_I + \lfloor E_F/\Delta T\rfloor$ nodes in the network. At that point, the average node degree is approximately $z$. It must hold that $z<\Delta T$; in the opposite case, the average degree $z$ cannot be achieved because new nodes dilute the network too fast. Each node has the relevance decay function $R_i(t) = I_i \exp[-(t-\tau_i)/\varTheta]$ where $\varTheta$ is the decay time scale and $I_i$ is the initial relevance of node $i$. Initial relevance values are drawn from the exponential distribution $f(I)=\mathrm{e}^{-I}$. When the decay parameter $\varTheta$ is sufficiently high, this setting produces broad degree distributions~\cite{Medo2011} which are similar to distributions often seen in real information networks~\cite[Chapter~4]{Newman2010}. We use $n_I=10$, $\Delta T = 16$, $\varTheta=50$, $A=1$, and $\lambda=0$ for all artificial networks studied here; their sample degree distributions are shown in Fig.~\ref{fig:network}b. \begin{figure} \centering \vspace*{4pt} \includegraphics[scale = 0.34]{net-illustration} \caption{(a) Illustration of a bipartite network where only links' target nodes are of interest. (b) Sample degree distributions of networks produced according to Sec.~\ref{sec:artif_data}.} \label{fig:network} \end{figure} \section{Parameter estimation methods} \subsection{Maximum likelihood estimation} \label{sec:MLE} We first use the standard maximum likelihood estimation (MLE) to estimate parameters of the PA-RD model~\cite{Bos07}. A generic form of log-likelihood of realization $\mathcal{D}$ for a network growth model $\mathcal{M}$ has the form \begin{equation} \ln\mathcal{L}(\mathcal{D}\vert\mathcal{M}) = \sum_{l=1}^E \ln P(n_l, t_l\vert\mathcal{M}). \end{equation} where $P(n_l, t_l\vert\mathcal{M})$ is the probability of link $l$ arriving at node $n_l$ at time $t_l$ under model $\mathcal{M}$. It is convenient to transform this quantity into log-likelihood per link by dividing it with the number of links, $\ln\mathcal{L}(\mathcal{D}\vert\mathcal{M})/E$. For model $\mathcal{M}$ represented by its attachment probability $P(n_l, t_l\vert\mathcal{M})$ and a vector of model parameters $\vek{p}$, log-likelihood can be maximized with respect to these parameters and yields their estimates $\tilde{\vek{p}}$. Given a network realization obtained with Eq.~(\ref{PARD}), there are several parameters to estimate: initial relevance values of all nodes, additional attractiveness term $A$, and parameters of the relevance decay function. (Note that we make the estimation task easier by assuming that the functional form of relevance decay is known.) Greedy (uphill) maximization of log-likelihood is made possible by the profile of the likelihood function which does not feature multiple local maxima in the space of initial relevance values (see Sec.~\ref{sec:maxima} for an explanation). Starting from a random initial guess, we sequentially update all model parameters by quadratic extrapolation and repeat this process until the difference between new and old estimates is less than some sufficiently small threshold (we use $10^{-3}$ here). Due to the scale-invariance of relevance values, they can be normalized after each iteration so that their average is one, which improves convergence. While each evaluation of log-likelihood is time consuming and this straightforward approach is thus computationally expensive, it is often, as we shall show, viable. \subsection{Mean-field approximation to MLE (MF-MLE)} \label{sec:MF-MLE} As mentioned in Sec.~\ref{sec:model}, when the number of nodes is large and their relevance decays to zero, fluctuations of the denominator in \eref{PARD} become small and one can therefore replace it with a constant term $\Omega^*$. This mean-field approximation decouples the dynamics of nodes which then compete for new links with the external field $\Omega^*$ instead of competing with the other nodes present in the system. Eq.~(\ref{PARD}) then simplifies to \begin{equation} P(i, t) = \frac{R_i(t\vert\vek{\eta}_i)\big(k_i(t)+A\big)}{\Omega^*} \end{equation} where $\vek{\eta}$ is a vector of parameters of node $i$ and we again assume $\lambda=1$. In our case, the initial relevance value $I_i$ is the only node-specific parameter and thus $\vek{\eta}_i = (I_i)$. Since $\Omega^*$ is the same for all nodes, we can subsume it in $I_i$ due to the aforementioned scale invariance. The likelihood function for node $i$ is then constructed by evaluating all links created after this node has been introduced in the network. For link $l$, we assess whether the link points to node $i$ (then $\delta_{i, n_l} = 1$) or not (then $\delta_{i, n_l} = 0$). We get \begin{gather} \label{LL-mean_field} \ln\mathcal{L}_i(\mathcal{D}\vert \vek{\eta}_i) = \notag\\ = \sum_{\genfrac{}{}{0pt}{}{l=1}{t_l\geq\tau_i}}^E \ln\big[P(i, t_l)\delta_{i,n_l} + (1-P(i, t_l))(1-\delta_{i,n_l})\big] \end{gather} where we ignore links that are older than node $i$. This function can be maximized with respect to $\vek{\eta}_i$ for any given $A$. Global model parameters such as, in our case, $A$ and the time scale of relevance decay $\varTheta$ can be estimated by minimizing $\sum_i\ln\mathcal{L}_i(\mathcal{D}\vert\tilde{\vek{\eta}}_i)$ with respect to them (estimates $\tilde{\vek{\eta}}_i$ then need to be updated to reflect new values of the global parameters). MF-MLE makes it easy to change the functional form of relevance decay $R(t)$ for any individual node and thus classify their behavior (see~\cite[Chapter~8]{Han11} for more information on classification problems). While we do not pursue this direction here, it is of particular significance to the analysis of real data where various behavioral classes of nodes are likely to coexist. Also, the vector of node parameters can be easily extended by, for example, making the decay time $\varTheta$ node-dependent, while still maintaining the low-dimensional nature of the resulting likelihood optimization. \section{Estimation evaluation} To evaluate various estimation methods, we assess the maximal likelihood that they are able to achieve. Parameter estimation is simplified by assuming that the functional form of relevance decay is known and only model parameters $A, \varTheta, \{I_i\}_i$ are to be estimated. Since the true parameter values are available to us, we also measure Pearson's correlation between true values $I$ and their estimates $\tilde I$, $r(I, \tilde I)$ (the higher the value, the better the estimates). In evaluating this correlation, nodes with final degree four and less are excluded because their estimates are too noisy due to the lack of data. The advantage of using Pearson's correlation to measure the accuracy of estimates lies in its invariance with respect to rescaling of $\tilde I_i$ which fits well with the scale-invariance of the PA-RD model itself. The accuracy of estimates of $A$ and $\varTheta$ is measured as well. \begin{figure} \centering \includegraphics[scale = 0.34]{R-LL}\\[4pt] \includegraphics[scale = 0.34]{R-r} \caption{Results of a constrained MLE procedure where $A$ and $\varTheta$ are fixed at various values and log-likelihood is thus maximized only with respect to the initial relevance of each node. These results were obtained for one realization of the artificial network model with $z=12$ which corresponds to $n_F=30$ and $E_F=480$.} \label{fig:profile} \end{figure} Simulations reveal that MLE sometimes converges to estimates which are far from the true parameter values. To explain the reason for this behavior, Fig.~\ref{fig:profile} shows the results of constrained likelihood maximization where we artificially fix $A$ and $\varTheta$ at various values, many of which are far from the true values $A=1$ and $\varTheta=50$. The corresponding maximal log-likelihood values exhibit a shallow maximum in $A$ with the optimal value $2.7$ lying significantly above the true value $1$. Worse, the maximum in $\varTheta$ is non-existent: as $\varTheta$ increases, log-likelihood increases too and saturates at a value which is maintained also in the limit $\varTheta\to\infty$ (\emph{i.e.}, no relevance decay). Resulting $\tilde\varTheta$ thus depends on the initial values of model parameters and the procedure in which they are iteratively improved in the search for maximal likelihood. While Fig.~\ref{fig:profile} shows results for one network realization, the same behavior can be seen for all realizations of the input artificial network. Inspection of the initial relevance values estimated for large $\varTheta$ makes it clear that the lack of relevance decay is then compensated by later nodes being assigned higher initial relevance than earlier ones. As a result, MLE estimates then do not reflect the true initial relevance values but rather the order in which nodes are introduced in the network. This is demonstrated by the second panel of Fig.~\ref{fig:profile} where $r(I, \tilde I)$ reaches maximum for $\varTheta$ close to the true value of $50$ and then quickly drops to negative values for larger values of $\varTheta$. The negative correlation values are observed here because in this particular network realization, node arrival times are negatively correlated with their initial relevance values. The overall maximum of $r(I,\tilde I)$ lies at $A=0.94$ and $\varTheta=51$. \begin{figure} \centering \includegraphics[scale = 0.4]{r_vs_omega} \caption{Impact of the log-likelihood penalization term given in \eref{LLpenalized} as a function of $\omega$ for various values of $z$. Mean values and their standard errors were obtained by estimation in 1000 independent network realizations (the same applies to Tab.~\ref{tab:evaluation} and Fig.~\ref{fig:comparison}). Estimates of $\varTheta$ are large for $z=10$ and thus missing in the inset.} \label{fig:omega} \end{figure} The problem of excessive estimated decay time $\varTheta$ can be solved by introducing an additional term in log-likelihood with the aim to penalize solutions with high $\varTheta$. This is similar to regularization schemes such as LASSO~\cite{Tibshirani1996} which are often used to constraint solutions in high-dimensional optimization problems~\cite{Buehlmann2011}. We choose here to maximize \begin{equation} \label{LLpenalized} \frac1E\ln\mathcal{L}_i(\mathcal{D}\vert I_i, A, \varTheta)- \omega r(\tau,\tilde I) g[r(\tau,\tilde I)] \end{equation} where $g(x) = x$ for $x>0$ and $0$ otherwise; the additional term penalizes positive correlation between nodes' arrival times and their estimated initial relevances. As shown in Fig.~\ref{fig:omega}, MLE estimates with the correlation term $r(\tau, \tilde I)$ are superior to the original ones over a broad range of $\omega$. The difference is particularly large for sparse networks where standard MLE strongly overestimates $\varTheta$. Note that unlike for large parts of Fig.~\ref{fig:profile}b, the average correlation value $r(I,\tilde I)$ in Fig.~\ref{fig:omega} is positive even when $\tilde\varTheta$ is large. This is because while Fig.~\ref{fig:profile} presents outcome for a single network realization, Fig.~\ref{fig:omega} averages over many of them and the resulting average correlation is positive. \begin{table} \centering \begin{ruledtabular} \begin{tabular}{rrrrr} method & $\ln\mathcal{L}/E$ & $\tilde A$ & $\tilde\varTheta$ & $r(I, \tilde I)$\\ \hline MLE, $\omega=0$ & $-1.350(5)$ & $2.86(2)$ & $54.4(1)$ & $0.767(5)$\\ MLE, $\omega=1$ & $-1.350(5)$ & $2.85(2)$ & $51.8(1)$ & $0.835(3)$\\ MF-MLE & $-1.381(5)$ & $6.09(9)$ & $65.2(2)$ & $0.722(2)$\\ \end{tabular} \end{ruledtabular} \caption{Estimates obtained with respective methods for $z=14$ (which corresponds to $n_F=80$, $E_F=1120$). Numbers in brackets report uncertainty of the last digit given by the standard error of the mean.} \label{tab:evaluation} \end{table} We proceed now to a direct comparison of the discussed estimation methods. As can be seen in Tab.~\ref{tab:evaluation}, all methods produce log-likelihood of a comparable magnitude but their parameter estimates (only $\tilde A$ and $\tilde\varTheta$ are shown here) differ. Notably, while $\tilde\varTheta$ is close to the true value, the error of $\tilde A$ is substantial. One can say that MLE tends to overestimate the tendency to random connections (as $A$ grows, the influence of preferential attachment vanishes) in this artificial system. We found no signs of this error vanishing with the data size which can be probably attributed to the network growth which constantly injects new nodes with zero degree in the system. The highest correlation between $I$ and $\tilde I$ is achieved with penalized MLE estimation when $\omega=1$. Performance of the methods is further illustrated in Fig.~\ref{fig:comparison} as a function of $z$. As expected, $r(I, \tilde I)$ increases with $z$ for both exact MLE methods and penalized estimations always outperform the unpenalized ones. The behavior is different for results obtained with the mean-field MLE whose quality slowly deteriorates as $z$ increases. Obviously, further improvements are necessary to make this otherwise promising method applicable in practice. \begin{figure} \centering \includegraphics[scale = 0.4]{comparison-DT=16} \caption{Performance of estimation methods vs mean degree of artificial networks $z$.} \label{fig:comparison} \end{figure} \begin{table*} \centering \begin{ruledtabular} \begin{tabular}{rrrrrrrrrr} model $M$ & $P_i(t)$ & $k_M$ & $\ln\mathcal{L}/E$ & $AICc(M)$ & $w_M$ & $\tilde A$ & $\tilde R_{\infty}$ & $\tilde\varTheta$ & $\tilde\beta$\\ \hline RAND & $1$ & $0$ & $-5.805$ & $285,364$ & $0$ & --- & --- & --- & ---\\ PA & $k_i(t)+A$ & $1$ & $-5.767$ & $283,519$ & $0$ & $137$ & --- & --- & ---\\ PA-H & $k_i(t)+A_i$ & $N$ & $-4.641$ & $229,931$ & $0$ & --- & --- & --- & ---\\ PA-HD & $k_i(t)+A_i(t)$ & $N+2$ & $-4.111$ & $203,872$ & $0$ & --- & --- & $1.9$ & $0.42$\\ PA-R & $(k_i(t)+A)R_i$ & $N+1$ & $-4.641$ & $229,905$ & $0$ & $1108$ & --- & --- & ---\\ PA-RD & $(k_i(t)+A)R_i(t)$ & $N+4$ & $-4.043$ & $200,536$ & $1$ & $60$ & $0.0088$ & $3.2$ & $0.64$\\ \end{tabular} \end{ruledtabular} \caption{Maximum likelihood estimates and model selection results for the Econophysics Forum data. Model weights $w_M$ defined by Eq.~(\ref{w_M}) show that there is overwhelming evidence in favor of the model with decaying relevance (PA-RD).} \label{tab:results} \end{table*} \section{Analysis of real data} \label{sec:real_data} To illustrate the potential of high-dimensional statistical analysis of network data, we finally apply it to real data and compare the level of support which it gives to various network models. The analyzed data originates from the interdisciplinary physics community web site \emph{Econophysics Forum} (see \url{www.unifr.ch/econophysics}) which is run by the research group of Yi-Cheng Zhang at the University of Fribourg since 1998. We parsed server web log files collected from 6th July 2010 until 31st March 2013 (a time span of 1000 days). Activities of web bots and other automated access were removed from the data. While web logs contain all user actions on the web site, we kept only entries corresponding to downloads of papers posted on the Econophysics Forum. The corresponding user-paper bipartite network consists of 844 paper-nodes and their 24,581 links~\cite{data}. As expected, the degree distribution of paper-nodes is broad (the maximal degree of 741 is much greater than the average degree of 29), making this data a good candidate for being explained by preferential attachment or related models. We use this data to evaluate two low-dimensional and four high-dimensional models. The low-dimensional models are: random attachment to an existing node (RAND) and the standard preferential attachment (PA). The high-dimensional models are: preferential attachment with heterogeneous (node-dependent) additive term (PA-H), preferential attachment with heterogeneous and decaying additive term (PA-HD) which has been introduced in~\cite{Eom2011}, preferential attachment with constant relevance (PA-R; such constant relevance is usually referred as fitness in past works~\cite{Bianconi01}), and finally preferential attachment with relevance decay (PA-RD). The functional form of the probability of a new link attaching to an existing node at time $t$, $P_i(t)$, is shown in Tab.~\ref{tab:results} for each model. The form of $A_i(t)$ suggested for PA-HD in~\cite{Eom2011} is generalized to $A_i(t) = I_i(t) \exp[-((t-\tau_i)/\varTheta)^{\beta}]$ which in our case performs better than the original form without $\beta$. Note that~\cite{Huberman13} reports a similar behavior in the popularity growth of stories in \url{digg.com}. For simplicity, we assume a similar form of $R_i(t)$ in PA-RD, $R_i(t) = I_i(t) \exp[-((t-\tau_i)/\varTheta)^{\beta}] + R_{\infty}$, which in fact roughly corresponds to the empirical relevance decay results presented in~\cite{Medo2011}. A non-vanishing absolute term $R_{\infty}$ is needed here to allow for links occasionally attaching to old nodes. The log-normal decay form reported in \cite{Radicchi08,Wang2013} does not yield better fit in our case, perhaps as a result of immediate response of the Econophysics Forum users which makes the increasing relevance phase provided by log-normal curves unfitting. For PA-RD, we report results obtained with the penalization term ($\omega=1$) which, however, differ little from the results obtained with $\omega=0$. To maximize the likelihood functions we use the iterative extrapolating approach described in Sec.~\ref{sec:MLE}. This procedure is run ten times with independent random initial configurations; the best result obtained with each method is reported in Table~\ref{tab:results}. In addition, the table shows also the number of model parameters $k_M$ and the corrected Akaike information criterion \begin{equation} AICc(M) = -2\ln\big(\!\max\mathcal{L}(\mathcal{D}\vert M)\big)+\frac{2k_ME}{E-k_M-1} \end{equation} where the maximum is taken over the whole parameter space of model $M$. $AICc(M)$ measures how well model $M$ fits the data and corrects for a finite sample size~\cite{Burnham2002}. It can be used to construct model weights~\cite{Claeksens2008} in the form \begin{equation} \label{w_M} w_M\sim\exp\big[(\min_{M'} AICc(M')-AICc(M))/2\big] \end{equation} where the proportionality factor is obtained by requiring the sum of all model weights to equal one. Finally, we report the values of global model parameters that maximize data likelihood for each model. Our comparison of models contains several notable outcomes. Firstly, both low-dimensional models are clearly insufficient to explain the data. In fact, preferential attachment yields only marginally better fit than random attachment. Secondly, high-dimensional models without time decay perform significantly worse than their counterparts with time decay. This is not surprising because we fit the models to an information network where, as argued in~\cite{Medo2011}, aging of nodes is of prime importance. Thirdly, while the log-likelihood values obtained with PA-HD and PA-RD are both substantially better than those obtained for other models, the difference between them is big enough for the Akaike information criterion to assign an overwhelming weight to PA-RD. (The resulting weight of PA-HD, which has been truncated to zero in Table~\ref{tab:results}, is around $10^{-724}$.) For PA-RD, the effective lifetime corresponding to the obtained relevance decay parameters is \begin{equation} \avg{t} := \frac{\int_0^{\infty} t R(t)\,\mathrm{d} t}{\int_0^{\infty} R(t)\,\mathrm{d} t} = \frac{\tilde\varTheta\,\Gamma(2/\tilde\beta)}{\tilde\beta\,\Gamma(1+1/\tilde\beta)}\approx 8\,\mathrm{days} \end{equation} where we neglect $R_{\infty}$ which is small, yet it formally causes the above-written expression to diverge. This lifetime well agrees with the fact that papers typically spend one week on the front page of the Econophysics Forum. The value of the additive term $\tilde A \approx 60$ is relatively high in comparison with the average node degree of $29$ which suggests that in the studied dataset, the influence of preferential attachment (\emph{i.e.}, attachment probability proportional to node degree) is relatively weak. An alternative explanation is that our assumed relevance decay function $R(t)$ disagrees with the data and thus an increased proportion of ``random'' connections is necessary to model the data. A more detailed analysis is necessary to establish what is the real reason behind this apparent randomness. \begin{figure} \centering \includegraphics[scale = 0.4]{weights_vs_size} \caption{Model weights as a function of subset length for PA, PA-HD, and PA-RD. Lines show the mean weights and shaded areas mark their standard deviation based on 1000 random subsets drawn for each subset length.} \label{fig:weights} \end{figure} Since likelihood computation is costly and during its maximization in numerous variables it needs to be carried out many times, obtaining the results presented in Table~\ref{tab:results} on a standard desktop computer takes several hours. It is thus natural to ask whether significant evidence in favor of one of the models cannot be obtained by analyzing subsets of the data which would save considerable computational time. To this end, we evaluated weights of three representative models (PA, PA-HD, and PA-RD) on data subsets corresponding to time spans (which we refer to as subset lengths, $L$) ranging from 4 to 100 days; the starting. We generated many subsets for each $L$ by choosing their starting day at random. Results shown in Fig.~\ref{fig:weights} demonstrate that while particularly short subsets favor the low-dimensional PA model, the situation quickly changes and this model is virtually eliminated as soon as $L\gtrsim30$. Two high-dimensional models, which enjoy comparable support until $L\lesssim30$, are clearly distinguished at $L=60$ and above. Meanwhile, evaluation of multiple small-scale subsets is fast: the computational time required for one likelihood maximization of PA-RD drops from 10 minutes for the whole 1000-day data spanning to 2 seconds for a 100-day subset. We can conclude that this approach allows us to efficiently discriminate between models even when no particularly efficient approach to likelihood maximization is available. \section{Discussion} \label{sec:conclusions} We studied the use of maximum likelihood estimation in analysis of high-dimensional models of growing networks. Artificially created networks with preferential attachment and decaying relevance~\cite{Medo2011} were used to show that a near-flat likelihood landscape makes the standard likelihood maximization rather unreliable and sensitive to the initial choice of model parameters. Introducing a penalization term effectively modifies the landscape and helps to avoid ``wrong'' solutions. The resulting MLE-based scheme outperforms the standard likelihood maximization for a wide range of model networks. On the other hand, both original and modified MLE overestimate the additive parameter $A$ which is crucial in the early stage of a node's degree growth. How to improve on that remains an open question. We then tested the previously developed methods on real data where both preferential attachment and relevance decay are expected to play a role. In this part, the focus is on comparing various competing network models that may be used to explain the data. We show that the data shows overwhelming evidence in favor of one of the models and that sufficiently strong evidence can be achieved by studying small subsets of the data. Model evaluation by such subset sampling is of particular importance to large-scale datasets where straightforward likelihood maximization is prohibitively time-consuming. Up to now, models of complex networks have been appraised mostly by comparing aggregate characteristics of the produced networks (degree distribution or clustering coefficient, for example) with features seen in real data. The caveat of this approach is that many network characteristics are computed on static network snapshots and are thus of little use for the measurement of growing networks. Empirical node relevance~\cite{Medo2011} is designed especially for growing networks but more metrics, targeted at specific situations and questions, are needed. Despite potential improvements in this direction, to gain \emph{real} evaluative and discriminative power over network models, robust statistical methods such as maximum likelihood estimation need to be relied on. We have made a step in this direction which, hopefully, will contribute to consolidating and further developing the field of network models. Open issues include estimates of parameter uncertainty in the case of real data by bootstrap methods~\cite{Davison97,Shalizi10}, identification of situations where maximum likelihood estimates converge to true parameter values (including model misspecification as in~\cite{Owhadi2013} which is of particular importance to parameter estimates in complex systems), and improvements of the mean-field likelihood estimation which was introduced in Sec.~\ref{sec:MF-MLE}. It needs to be stressed that the potential impact of parameter estimation far exceeds the academic problem of model validation: Model parameters, once known, can be directly useful in practice. In the case of preferential attachment with relevance decay, for example, the overall rate of relevance/interest decay is closely connected to the most successful strategy in the competition for attention~\cite{Huberman13}. On the other hand, the initial, current, or total relevance values of individual items can be used to detect which items deserve to be examined more closely. \begin{acknowledgments} This work was supported by the EU FET-Open Grant No.~231200 (project QLectives) and by the Swiss National Science Foundation Grant No.~200020-143272. \end{acknowledgments}
\section{Introduction} In applications like wireless communications, audio signal processing, radar and microphone array processing, adaptive beamforming has been intensively researched and developed in the past years. However, under certain circumstances, adaptive beamformers suffer a performance degradation due to several reasons which include short data records, the presence of the desired signal in the training data, or imprecise knowledge of the steering vector of the desired signal. In order to improve the performance of adaptive beamformers in the presence of steering vector mismatches, RAB techniques have been developed. Different from the standard designs \cite{r1}, the design principles of RAB MVDR beamformers \cite{r6} include: the generalized sidelobe canceller, diagonal loading \cite{r4,r5}, subspace projection \cite{jio_el,jio_lcmv_esp,jidf_radar,jio_radar}, worst-case optimization \cite{r3,r14} and steering vector estimation with presumed prior knowledge \cite{r7,r8,r15,r16}. However, RAB designs based on these principles have some drawbacks such as their ad hoc nature, high probability of subspace swap at low SNR and high computational cost \cite{r7}. Some recent design approaches have considered combining different design principles together to improve RAB performances. The algorithm which jointly estimates the mismatched steering vector using Sequential Quadratic Program (SQP) \cite{r8} and the interference-plus-noise covariance (INC) matrix using a shrinkage method \cite{r10} has been reported recently. Later, another similar approach which jointly estimates the steering vector using SQP and the INC matrix using a covariance reconstruction method \cite{r11}, presents outstanding performance compared to other RAB techniques. However, the cost of the algorithm in \cite{r11} is high due to the required matrix reconstruction process. In this paper, we propose an RAB algorithm with low complexity, which requires very little in terms of prior information, and has a superior performance to previously reported RAB algorithms. The proposed technique estimates the steering vector using a Low-Complexity Shrinkage-Based Mismatch Estimation (LOCSME) algorithm. LOCSME estimates the covariance matrix of the input data and the INC matrix using the Oracle Approximating Shrinkage (OAS) method. The only prior knowledge that LOCSME requires is the angular sector in which the desired signal steering vector lies. Given the sector, the subspace projection matrix of this sector can be computed in very simple steps \cite{r7,r8,r9,r10,r11}. In the first step, an extension of the OAS method \cite{r12} is employed to perform shrinkage estimation for both the cross-correlation vector between the received data and the beamformer output and the received data covariance matrix. LOCSME is then used to estimate the mismatched steering vector and does not involve any optimization program, which results in a lower computational complexity. In a further step, we estimate the desired signal power using the desired signal steering vector and the received data. As the last step, a strategy which subtracts the covariance matrix of the desired signal from the data covariance matrix estimated by OAS is proposed to obtain the INC matrix. The advantage of this approach is that it circumvents the use of direction finding techniques for the interferers, which are required to obtain the INC matrix. This paper is structured as follows. The system model is described in Section II. In Section III, the proposed LOCSME algorithm is presented. Section IV shows and discusses the simulation results. Finally, Section V gives the conclusion. \section{System Model} Consider a linear antenna array of $M$ sensors and $K$ narrowband signals received at $i$th snapshot as expressed by \begin{equation} {\bf x}(i)={\bf A}({\boldsymbol \theta}){\bf s}(i)+{\bf n}(i), \end{equation} where ${\bf s}(i) \in {\mathbb C}^{K \times 1}$ presents the uncorrelated source signals, ${\boldsymbol\theta}=[{\theta}_1,\dotsb,{\theta}_K]^T \in {\mathbb R}^K$ is the vector containing the directions of arrivals (DoAs), ${\bf A}({\boldsymbol \theta})=[{\bf a}({\theta}_1 )+{\bf e}, \dotsb, {\bf a}({\theta}_K)] \in {\mathbb C}^{M \times K}$ is the matrix which contains the steering vector for each DoA and ${\bf e}$ is the mismatch of the steering vector of the desired signal, ${\bf n}(i) \in {\mathbb C}^{M \times 1}$ is assumed to be complex Gaussian noise with zero mean and variance ${\sigma}^2_n$. The beamformer output is given by \begin{equation} y(i)={\bf w}^H{\bf x}(i), \end{equation} where ${\bf w}=[w_1,\dotsb,w_M]^T \in {\mathbb C}^{M\times1}$ is the beamformer weight vector, where $({\cdot})^H$ denotes the Hermitian Transpose. The optimum beamformer can be computed by maximizing the signal-to-interference-plus-noise ratio (SINR) given by \begin{equation} SINR=\frac{{\sigma}^2_1{\lvert{\bf w}^H{\bf a}\rvert}^2}{{\bf w}^H{\bf R}_{i+n}{\bf w}}. \end{equation} Assume that the steering vector ${\bf a}$ is known precisely (${\bf a}={\bf a}({\theta}_1 )$), where ${\sigma}^2_1$ is the desired signal power and ${\bf R}_{i+n}$ is the INC matrix, then problem (3) can be transformed into an optimization problem \begin{equation} \begin{aligned} & \underset{\bf w} {\text{minimize}} && {\bf w}^H{\bf R}_{i+n}{\bf w} \\ & \text{subject to} && {\bf w}^H{\bf a}=1, \end{aligned} \end{equation} which is known as the MVDR beamformer or Capon beamformer \cite{r1}. The optimum weight vector is given by ${\bf w}_{opt}= \frac{{{\bf R}^{-1}_{i+n}}{\bf a}}{{\bf a}^H{{\bf R}^{-1}_{i+n}}{\bf a}}.$ Since ${\bf R}_{i+n}$ is usually unknown in practice, it is estimated by the sample covariance matrix (SCM) of the received data as \begin{equation} \hat{\bf R}(i)=\frac{1}{i}\sum\limits_{k=1}^i{\bf x}(k){{\bf x}^H}(k), \end{equation} which will result in the Sample Matrix Inversion (SMI) beamformer ${\bf w}_{SMI}=\frac{\hat{\bf R}^{-1}{\bf a}}{{\bf a}^H\hat{\bf R}^{-1}{\bf a}}$. However, the SMI beamformer requires a large number of snapshots and is sensitive to steering vector mismatches \cite{r10,r11}. \section{Proposed LOCSME Algorithm} In this section, the proposed LOCSME algorithm is introduced. The idea of LOCSME is to estimate the steering vector and the INC matrix separately as in previous approaches. The estimation of the steering vector is described as the projection onto a predefined subspace matrix of an iteratively shrinkage-estimated cross-correlation vector between the beamformer output and the array observation. The INC matrix is obtained by subtracting the desired signal covariance matrix from the data covariance matrix estimated by the OAS method. \subsection{{Steering Vector Estimation using LOCSME}} The cross-correlation between the array observation vector and the beamformer output can be expressed as \begin{equation} {\bf d}=E\lbrace{\bf x}y^*\rbrace. \end{equation} We assume that ${\lvert{{\bf a}_m^H{\bf w}}\rvert}\ll{\lvert{{\bf a}_1^H{\bf w}}\rvert}$ for $m=2,\dotsb,K$, all signal sources and the noise have zero mean, and the desired signal and every interferer are independent from each other. By substituting (1) and (2) into (6), we suppose the interferers are sufficiently canceled such that they fall much below the noise floor and the desired signal power is not affected by the interference so that ${\bf d}$ can be rewritten as \begin{equation} {\bf d}=E\lbrace{{{{\sigma}_1}^2{\bf a}_1^H{\bf w}{\bf a}_1}+{\bf n}{\bf n}^H{\bf w}}\rbrace. \end{equation} In order to eliminate the unwanted part of ${\bf d}$ and obtain an estimate of the steering vector ${\bf a}_1$, ${\bf d}$ can be projected onto a subspace \cite{r9} that collects information about the desired signal. Here the prior knowledge amounts to providing an angular sector range in which the desired signal is located, say $[{\theta}_1-{\theta}_e,{\theta}_1+{\theta}_e]$. The subspace projection matrix ${\bf P}$ is given by \begin{equation} {\bf P}=[{\bf c}_1,{\bf c}_2,\dotsb,{\bf c}_p][{\bf c}_1,{\bf c}_2,\dotsb,{\bf c}_p]^H, \end{equation} where ${\bf c}_1,\dotsb,{\bf c}_p$ are the $p$ principal eigenvectors vectors of the matrix ${\bf C}$, which is defined by \cite{r8} \begin{equation} {\bf C}= \int \limits_{{\theta}_1-{\theta}_e}^{{\theta}_1+{\theta}_e}{\bf a}({\theta}){\bf a}^H({\theta})d{\theta}. \end{equation} At this point, LOCSME will use the OAS method to compute the correlation vector ${\bf d}$ iteratively. The aim is to devise a method that estimates ${\bf d}$ more accurately with the help of the shrinkage technique. An accurate estimate of ${\bf d}$ can help to obtain a better estimate of the steering vector. Let us define \begin{equation} \hat{\bf F}=\hat{\nu} {\bf I}, \end{equation} where $\hat{\nu} = {{\rm tr}(\hat{\bf S})}/M$ and $\hat{\bf S} = {\rm diag}({\bf x}y^*)$. Then, a reasonable tradeoff between covariance reduction and bias increase can be achieved by shrinkage of $\hat{\bf S}$ towards $\hat{\bf F}$ \cite{r12} and subsequently using it in a vector shrinkage form, which results in \begin{equation} \hat{\bf d} = \hat{\rho}{\rm diag}(\hat{\bf F}) + (1-\hat{\rho}){\rm diag}(\hat{\bf S}), \end{equation} which is parameterized by the shrinkage coefficient $\hat{\rho}$. If we define $\hat{\bf D}={\rm diag}(\hat{\bf d})$ then the goal is to find the optimal value of $\hat{\rho}$ that minimizes the mean square error (MSE) of $E[{\lVert{\hat{\bf D}(i)-\hat{\bf F}(i-1)}\rVert}^2]$ in the $i$th snapshot, which leads to \begin{equation} \hat{\bf d}(i) = \hat{\rho}(i){\rm diag}(\hat{\bf F}(i)) + (1-\hat{\rho}(i)){\rm diag}(\hat{\bf S}(i)), \end{equation} \begin{equation} \hat{\rho}(i+1) = \frac{(1-\frac{2}{M}){\rm tr}(\hat{\bf D}(i)\hat{\bf S}^*(i)) + {\rm tr}(\hat{\bf D}(i)){\rm tr}(\hat{\bf D}^*(i))}{(i+1-\frac{2}{M}){\rm tr}(\hat{\bf D}(i)\hat{\bf S}^*(i))+(1-\frac{i}{M}){\rm tr}(\hat{\bf D}(i)){\rm tr}(\hat{\bf D}^*(i))}, \end{equation} where the derivation is shown in the Appendix and $\hat{\bf S}(i)$ is the sample correlation vector (SCV) given by \begin{equation} \hat{\bf S}(i)={\rm diag}\Big(\frac{1}{i}\sum\limits_{k=1}^i{\bf x}(k)y^*(k)\Big). \end{equation} As long as the initial value of $\hat{\rho}(0)$ is between $0$ and $1$, the iterative process in (12) and (13) is guaranteed to converge \cite{r12}. Once the correlation vector $\hat{\bf d}$ is obtained by the above OAS method, the steering vector is estimated by \begin{equation} {\hat{\bf a}_1}(i)=\frac{{\bf P}\hat{\bf d}(i)}{{\lVert{{\bf P}\hat{\bf d}(i)}\rVert}_2}, \end{equation} where $\hat{\bf a}_1(i)$ gives the final estimate of the steering vector. \subsection{{Interference-Plus-Noise Covariance Matrix Estimation}} In order to estimate the INC matrix, the data covariance matrix (which contains the desired signal) is required. The SCM in (5) is necessary as a preliminary approximation. In the next step, similar to using OAS to estimate the cross-correlation vector $\hat{\bf d}$, the SCM is also processed with the OAS method as a further shrinkage estimation step. Let us define the following quantity \begin{equation} \hat{\bf F}_0={\hat{\nu}_0}{\bf I}, \end{equation} where $\hat{\nu}_0={{\rm tr}(\hat{\bf R})}/M$. Then, we use the shrinkage form again \begin{equation} \tilde{\bf R}={\hat{\rho}_0}{\hat{\bf F}_0}+(1-{\hat{\rho}_0})\hat{\bf R}. \end{equation} By minimizing the MSE described by $E[{\lVert{{\tilde{\bf R}(i)}-{\hat{\bf F}_0}(i-1)}\rVert}^2]$, we obtain the following recursion \begin{equation} \tilde{\bf R}(i)={\hat{\rho}_0}(i){\hat{\bf F}_0}(i)+(1-{\hat{\rho}_0}(i))\hat{\bf R}(i), \end{equation} \begin{equation} {\hat{\rho}_0}(i+1)=\frac{(1-{\frac{2}{M}}){\rm tr}({\tilde{\bf R}(i)}{\hat{\bf R}(i)})+{\rm tr}^2(\tilde{\bf R}(i))}{(i+1-{\frac{2}{M}}){\rm tr}({\tilde{\bf R}(i)}{\hat{\bf R}(i)})+(1-{\frac{i}{M}}){\rm tr}^2(\tilde{\bf R}(i))}. \end{equation} Provided that $0<{\hat{\rho}_0}(0)<1$, the iterative process in (18) and (19) is guaranteed to converge \cite{r12}. In order to eliminate the unwanted information of the desired signal in the covariance matrix and obtain the INC matrix, the desired signal power ${\sigma}^2_1$ must be estimated. Let us rewrite the received data as \begin{equation} {\bf x}=\sum\limits_{k=1}^K{\bf a}_k{s_k}+{\bf n}. \end{equation} Pre-multiplying the above equation by ${\bf a}_1^H$, we have \begin{equation} {\bf a}_1^H{\bf x}={\bf a}_1^H{\bf a}_1{s_1}+{\bf a}_1^H(\sum\limits_{k=2}^K{\bf a}_k{s_k}+{\bf n}). \end{equation} Assuming that ${\bf a}_1$ is uncorrelated with the interferers, we obtain \begin{equation} {\bf a}_1^H{\bf x}={\bf a}_1^H{\bf a}_1{s_1}+{\bf a}_1^H{\bf n}. \end{equation} Taking the expectation of $E[|{\bf a}_1^H{\bf x}|^2]$, we obtain \begin{equation} |{\bf a}_1^H{\bf x}|^2=E[({\bf a}_1^H{\bf a}_1{s_1}+{\bf a}_1^H{\bf n})^*({\bf a}_1^H{\bf a}_1{s_1}+{\bf a}_1^H{\bf n})]. \end{equation} If the noise is statistically independent of the desired signal, we have \begin{equation} |{\bf a}_1^H{\bf x}|^2=|{\bf a}_1^H{\bf a}_1|^2|s_1|^2+{\bf a}_1^H{\bf n}{\bf n}^H{\bf a}_1, \end{equation} where $|s_1|^2$ is the desired signal power which can be replaced by its estimate $\hat{\sigma}^2_1$, ${\bf n}{\bf n}^H$ represents the noise covariance matrix ${\bf R}_n$ which can be replaced by ${\sigma}^2_n{\bf I}_M$. Replacing ${\bf a}_1$ by its estimate ${\hat{\bf a}_1}(i)$ the desired signal power estimate is given by \begin{equation} \hat{\sigma}^2_1(i)=\frac{|{{\hat{\bf a}_1}^H}(i){\bf x}(i)|^2-{{\hat{\bf a}_1}^H}(i){\hat{\bf a}_1}(i){\sigma}^2_n}{|{{\hat{\bf a}_1}^H}(i){\hat{\bf a}_1}(i)|^2}. \end{equation} As the last step, the desired signal covariance matrix is subtracted and the INC matrix is given by \begin{equation} {\tilde{\bf R}_{i+n}}(i)=\tilde{\bf R}(i)-\hat{\sigma}^2_1(i){\hat{\bf a}_1}(i){{\hat{\bf a}_1}^H}(i). \end{equation} The advantage of this step compared to SMI and existing methods is that it does not require direction finding and is suitable for real-time applications. With the estimates for the steering vector and the INC matrix, the beamformer is computed by \begin{equation} \hat{\bf w}(i)=\frac{{\tilde{\bf R}^{-1}_{i+n}}(i)\hat{\bf a}_1(i)}{{\hat{\bf a}_1}^H(i){\tilde{\bf R}^{-1}_{i+n}}(i)\hat{\bf a}_1(i)}. \end{equation} Table I summarizes LOCSME in steps. From a complexity point of view, the main computational cost is due to the following steps: SCM of the observation data, OAS estimation for SCM, norm computations of the covariance matrix and the INC matrix. Each of these steps has a complexity of ${\mathcal{O}}(M^3)$. Additionally, compared to the previous RAB algorithms in \cite{r7}, \cite{r8}, \cite{r10} and \cite{r11} which have complexity equal or higher than ${\mathcal{O}}(M^{3.5})$, LOCSME has a lower cost (${\mathcal{O}}(M^3)$). \begin{table} \small \begin{center} \caption{Proposed LOCSME Algorithm} \vspace{-0.5em} \begin{tabular}{|l|} \hline Initialize: \\ ${\bf C}=\int\limits_{{\theta}_1-{\theta}_e}^{{\theta}_1+{\theta}_e}{\bf a}({\theta}){\bf a}^H({\theta})d{\theta}$ \\ $[{\bf c}_1,\dotsb,{\bf c}_p]$: p princical eigenvectors of ${\bf C}$ \\ Subspace projection ${\bf P}=[{\bf c}_1,\dotsb,{\bf c}_p][{\bf c}_1,\dotsb,{\bf c}_p]^H$ \\ $\hat{\bf R}(0)={\bf 0}$; $\hat{\bf S}(0)={\bf 0}$; ${\bf w}(0)={\bf 1}$; \\ $\hat{\rho}(1)=\rho(0)=\hat{\rho}_0(1)=\rho_0(0)=1$; \\ For each snapshot index $i=1,2,\dotsb$: \\ $\hat{\bf R}(i)=\frac{1}{i}\sum\limits_{k=1}^i{\bf x}(k){{\bf x}^H}(k)$ \\ $\hat{\bf S}(i)= {\rm diag}(\frac{1}{i}\sum\limits_{k=1}^i{\bf x}(k)y^*(k))$ \\ $\hat{\nu}(i) = {{\rm tr}(\hat{\bf S}(i))}/M$ \\ $\hat{\bf F}(i)=\hat{\nu}(i){\bf I}$ \\ $\hat{\bf d}(i) = \hat{\rho}(i) {\rm diag}(\hat{\bf F}(i)) + (1-\hat{\rho}(i)){\rm diag}(\hat{\bf S}(i))$ \\ $\hat{\bf D}(i)= {\rm diag}(\hat{\bf d}(i))$ \\ $\hat{\rho}(i+1) = \frac{(1-\frac{2}{M}){\rm tr}(\hat{\bf D}(i)\hat{\bf S}^*(i)) + {\rm tr}(\hat{\bf D}(i)){\rm tr}(\hat{\bf D}^*(i))}{(i+1-\frac{2}{M}){\rm tr}(\hat{\bf D}(i)\hat{\bf S}^*(i))+(1-\frac{i}{M}){\rm tr}(\hat{\bf D}(i)){\rm tr}(\hat{\bf D}^*(i))}$ \\ ${\hat{\bf a}_1}(i)=\frac{{\bf P}\hat{\bf d}(i)}{{\lVert{{\bf P}\hat{\bf d}(i)}\rVert}_2}$ \\ $\hat{\nu}_0(i)={{\rm tr}(\hat{\bf R}(i))}/M$ \\ $\hat{\bf F}_0(i)={\hat{\nu}_0}(i){\bf I}$ \\ $\tilde{\bf R}(i)={\hat{\rho}_0}(i){\hat{\bf F}_0}(i)+(1-{\hat{\rho}_0}(i))\hat{\bf R}(i)$ \\ ${\hat{\rho}_0}(i+1)=\frac{(1-{\frac{2}{M}}){\rm tr}({\tilde{\bf R}(i)}{\hat{\bf R}(i)})+{\rm tr}^2(\tilde{\bf R}(i))}{(i+1-{\frac{2}{M}}){\rm tr}({\tilde{\bf R}(i)}{\hat{\bf R}(i)})+(1-{\frac{i}{M}}){\rm tr}^2(\tilde{\bf R}(i))}$ \\ $\hat{\sigma}^2_1(i)=\frac{|{{\hat{\bf a}_1}^H}(i){\bf x}(i)|^2-{{\hat{\bf a}_1}^H}(i){\hat{\bf a}_1}(i){\sigma}^2_n}{|{{\hat{\bf a}_1}^H}(i){\hat{\bf a}_1}(i)|^2}$ \\ $\tilde{\bf R}(i)=\tilde{\bf R}(i)+{{\lVert{\tilde{\bf R}(i)}\rVert}_2} {\bf I}$ \\ ${\tilde{\bf R}_{i+n}}(i)=\tilde{\bf R}(i)-\hat{\sigma}^2_1(i){\hat{\bf a}_1}(i){{\hat{\bf a}_1}^H}(i)$ \\ ${{\tilde{\bf R}}_{i+n}}(i)={{\tilde{\bf R}}_{i+n}}(i) {\frac{2{\sigma}^2_n}{{\lVert{\tilde{\bf R}_{i+n}(i)}\rVert}_2}}$ \\ $\hat{\bf w}(i)=\frac{{\tilde{\bf R}^{-1}_{i+n}}(i)\hat{\bf a}_1(i)}{{\hat{\bf a}_1}^H(i){\tilde{\bf R}^{-1}_{i+n}}(i)\hat{\bf a}_1(i)}$ \\ \hline \end{tabular} \end{center} \end{table}\vspace{-1em} \section{Simulations} In our simulations, a uniform linear array (ULA) of $M=12$ omnidirectional sensors with a spacing of half wavelength is considered. Three source signals include the desired signal which is presumed to arrive at ${\theta}_1=10^\circ$ and two interferers which are impinging on the antenna array from directions ${\theta}_2=50^\circ$ and ${\theta}_3=90^\circ$. The signal-to-interference ratio (SIR) is fixed at 20dB. Only one iteration is performed per snapshot and we employ $i=50$ snapshots and 100 repetitions to obtain each point of the curves. The beamformer computed with LOCSME is compared to existing beamformers in terms of the output SINR. For the beamformers of \cite{r7}, \cite{r8}, \cite{r10}, \cite{r11} and the beamformer with LOCSME, the angular sector is chosen as $[{\theta}_1-5^\circ,{\theta}_1+5^\circ]$ and $p=8$ principal eigenvectors are used. The number of eigenvectors of the subspace projection matrix $p$ is selected manually with the help of simulations. For the beamformers of \cite{r7}, \cite{r8}, \cite{r10} and \cite{r11} which also require an optimization technique, the CVX software is used. The SINR performance versus snapshots and SNR of the algorithms is shown in Figs. 1 and 2 and the number of snapshots is $50$ for the SINR versus SNR plots. The average execution time of the algorithms in \cite{r7}, \cite{r8}, \cite{r10} and \cite{r11} is around $0.3$ sec/snapshot, while LOCSME only requires $0.021$ sec/snapshot. \subsection{Mismatch due to Coherent Local Scattering} In this case, the steering vector of the desired signal is affected by a local scattering effect and modeled as \begin{equation} {\bf a}={\bf p}+\sum\limits_{k=1}^4{e^{j{\varphi}_k}}{\bf b}({\theta}_k), \end{equation} where ${\bf p}$ corresponds to the direct path while ${\bf b}({\theta}_k)(k=1,2,3,4)$ corresponds to the scattered paths. The angles ${\theta}_k(k=1, 2, 3, 4)$ are randomly and independently drawn in each simulation run from a uniform generator with mean $10^\circ$ and standard deviation $2^\circ$. The angles ${\varphi}_k(k=1, 2, 3, 4)$ are independently and uniformly taken from the interval $[0,2\pi]$ in each simulation run. Notice that ${\theta}_k$ and ${\varphi}_k$ change from trials while remaining constant over snapshots \cite{r3}. Figs. 1 (a) and 2 (a) illustrate the SINR performance versus snapshots and SNR under the coherent scattering case. LOCSME outperforms the other algorithms and is close to the optimum SINR. \subsection{ Mismatch due to Incoherent Local Scattering} In the incoherent local scattering case, the desired signal has a time-varying signature and the steering vector is modeled by \begin{equation} {\bf a}(i)=s_0(i){\bf p}+\sum\limits_{k=1}^4{s_k(i)}{\bf b}({\theta}_k), \end{equation} where $s_k(i)(k=0, 1, 2, 3, 4)$ are i.i.d zero mean complex Gaussian random variables independently drawn from a random generator. The angles ${\theta}_k(k=0, 1, 2, 3, 4)$ are drawn independently in each simulation run from a uniform generator with mean $10^\circ$ and standard deviation $2^\circ$. This time, $s_k(i)$ changes both from run to run and from snapshot to snapshot. Figs. 1 (b) and 2 (b) depict the SINR performance versus snapshots and SNR. Compared to the coherent scattering results, all the algorithms have a performance degradation due to the effect of incoherent local scattering. However, LOCSME is able to outperform the remaining robust beamformers over a wide range of input SNR. The reason for the improved performance of LOCSME is the combined use of accurate estimates of the INC matrix and of the steering vector mismatch. Further testing with a larger number of antenna array elements indicates that the performance of all algorithms degrades (e.g. LOCSME has around 2dB degradation when $M=60$). In addition, inappropriate choice for the angular sector in which the desired signal is assumed to be located will lead to obvious performance degradation. \vspace{-1.0em} \begin{figure}[!htb] \begin{center} \def\epsfsize#1#2{0.875\columnwidth} \epsfbox{fig1.eps} \vspace*{-1.2em} \caption{SINR versus snapshots.} \label{1} \end{center} \end{figure} \vspace{-1.0em} \begin{figure}[!htb] \begin{center} \def\epsfsize#1#2{0.875\columnwidth} \epsfbox{fig2.eps} \vspace*{-1.2em} \caption{SINR versus SNR.} \label{2} \end{center} \end{figure} \vspace{1.4em} \section{Conclusion} We have proposed LOCSME that only requires prior knowledge of the angular sector of the desired signal and is less costly than existing methods. Simulation results have shown that LOCSME outperforms prior art in both coherent local scattering and incoherent local scattering cases. \begin{appendix} \textit{Derivation of $\hat{\rho}(i)$}: Equation (12) can be rewritten in an alternative way in matrix version as $\hat{\bf D}(i)=\hat{\rho}(i)\hat{\bf F}(i)+(1-\hat{\rho}(i))\hat{\bf S}(i)$. By using (10), then the shrinkage intensity $\hat{\rho}(i)$ can be computed from the following optimization problem \vspace{-0.5em} \begin{equation} \begin{aligned} & \underset{\hat{\rho}(i), \hat{\nu}(i)} {\text{min}} && E[{\lVert{\hat{\bf D}(i)-\hat{\bf F}(i-1)}\rVert}^2] \\ & \text{subject to} && \hat{\bf D}(i)=\hat{\rho}(i)\hat{\nu}(i){\bf I}+(1-\hat{\rho}(i))\hat{\bf S}(i). \end{aligned} \end{equation} Since $E[\hat{\bf S}(i)]=E[\hat{\bf F}(i-1)]$, the objective function in (30) can be rewritten as $\hat{\rho}^2(i){\lVert{\hat{\bf F}(i-1)-\hat{\nu}(i){\bf I}}\rVert}^2+(1-\hat{\rho}(i))^2E[{\lVert{\hat{\bf S}(i)-\hat{\bf F}(i-1)}\rVert}^2]$ \cite{r13}. The optimal value of $\hat{\nu}(i)$ is obtained as the solution to a problem that does not depend on $\hat{\rho}(i)$ as given by $\underset{\hat{\nu}(i)}{\text{min}}{\lVert{\hat{\bf F}(i-1)-\hat{\nu}(i){\bf I}}\rVert}^2$, which can be solved by computing the partial derivative of the argument with respect to $\hat{\nu}(i)$ and equating the terms to zero. By substituting the optimal value of $\hat{\nu}(i)$ into (30), computing the partial derivative of the argument with respect to $\hat{\rho}(i)$, equating the terms to zero and solving for $\hat{\rho}(i)$, we obtain \vspace{-0.5em} \begin{equation} \hat{\rho}(i)=\frac{E[{\lVert{\hat{\bf S}(i)-\hat{\bf F}(i-1)}\rVert}^2]}{{\lVert{\hat{\bf F}(i-1)-\mu(i){\bf I}}\rVert}^2+E[{\lVert{\hat{\bf S}(i)-\hat{\bf F}(i-1)}\rVert}^2]}. \end{equation} By further Gaussian assumptions as in \cite{r12}, replacing $\hat{\bf F}(i-1)$ by its estimate $\hat{\bf D}(i)$ and the data sample number $n$ by the snapshot index $i$, equation (13) can be obtained. \end{appendix} \newpage
\section{Introduction} It has been suggested quite recently that the galactic halo possesses some of the characteristics needed to support traversable wormholes \cite{RKRI}. For such wormholes the detection by means of gravitational lensing becomes a distinct possibility, especially when viewed from a strong-field perspective. \section{Background} It has been known for a long time that the rotation curves of neutral hydrogen clouds in the outer regions of the galactic halo cannot be explained in terms of ordinary luminous matter. The usual explanation is that galaxies and even clusters of galaxies are pervaded by dark matter, which does not emit electromagnetic waves nor interact with normal matter. To pursue the study of wormholes in this context, we will rely on the Navarro-Frenk-White density profile \cite{NFW96}: \begin{equation}\label{E:rho} \rho(r)=\frac{\rho_s}{\frac{r}{r_s}\left( 1+\frac{r}{r_s} \right)^2}, \end{equation} where $r_s$ is the characteristic scale radius and $\rho_s$ the corresponding density, to be described below. Eq. (\ref{E:rho}) resulted from $N$-body simulations in the search for the structure of dark halos in the standard CDM cosmology. One of the most important tools for the possible detection of exotic objects such as wormholes is gravitational lensing. Since the exotic matter inside a wormhole antigravitates, Cramer et al. \cite{Cramer95} and Safanova \cite{STR02} examined the lensing effects of negative masses on light rays from point sources. Torres et al. \cite{TRA98} suggested that wormholes can be probed using light curves of gamma-ray bursts. A method for calculating the gravitational microlensing effect of the Ellis wormhole is derived in Ref. \cite{tA10} and continued in Ref. \cite{TKAA10}. More recently, gravitational lensing has been studied from a strong-field perspective, often using an analytical method pioneered by Bozza \cite{vB02} to calculate the deflection angle. This method was used in Refs. \cite{TL05, TL12} to study two special models by Lemos et al. \cite{jL03}. Gravitational lensing by a $C$-field wormhole is discussed in Ref. \cite{RKC07}. \section{Wormhole structure} As noted in the Introduction, the existence of dark matter can be deduced from the observed flat rotation curves of neutral hydrogen clouds in the outer regions of the halo. The neutral hydrogen clouds must therefore be treated as test particles moving in circular orbits. Accordingly, the spacetime in the galactic halo is characterized by the line element \cite{RKRI} \begin{equation}\label{E:line1} ds^2=-B_0r^ldt^2+e^{2g(r)}dr^2 +r^2(d\theta^2+\sin^2\theta\,d\phi^2), \end{equation} using units in which $c=G=1.$ Here $l = 2(v^\phi)^2$, where $v^{\phi}$ is the rotational velocity and $B_0$ is an integration constant. (According to Ref. \cite{kN09}, $l\approx 0.000001$.) In a wormhole setting a more convenient form is \cite {MT88} \begin{equation}\label{E:line2} ds^2=-e^{2\Phi (r)}dt^2+\frac{dr^2} {1-\frac{b(r)}{r}} +r^2(d\theta^2+\sin^2\theta\,d\phi^2), \end{equation} where $e^{2\Phi (r)}=B_0r^{l}$. So by letting $B_0=1/b_0^{l}$, the spacetime metric becomes \begin{equation}\label{E:line3} ds^2=-\left(\frac{r}{b_0}\right)^ldt^2+\frac{dr^2} {1-\frac{b(r)}{r}} +r^2(d\theta^2+\sin^2\theta\,d\phi^2). \end{equation} From the Einstein field equation \begin{equation}\label{E:Einstein} \frac{b'(r)}{r^2}=8\pi\rho(r), \end{equation} it is readily shown that \begin{equation}\label{E:shape} b(r)= 8 \pi \rho_s r_s^3 \left[\ln \left( 1+\frac{r}{r_s} \right) + \frac{1}{ 1+\frac{r}{r_s} }+C \right], \end{equation} where $C$ is an integration constant. We now recall that if Eq. (\ref{E:line2}) represents a wormhole, then\\ (1) The \emph{redshift function}, $\Phi (r)$, must remain finite to prevent an event horizon. According to Eq. (\ref{E:line3}), this condition is met. It also follows that the wormhole spacetime is not asymptotically flat.\\ (2) The \emph{shape function}, $b(r)$, must obey the following conditions at the throat $r = r_{th}$ : $b(r_{th}) = r_{th}$ and $b^\prime(r_{th}) < 1$, the so-called flare-out condition. Moreover, Eq. (\ref{E:Einstein}) implies that $b'(r)$ must be positive.\\ (3) For $b=b(r)$ we also have $b(r)<r$ for $r >r_{th}$.\\ Finally, the expressions for the radial pressure, $p_r$, and the lateral pressure, $p_t$, are given in Ref. \cite{RKRI}. It is also shown that $\rho+p_r<0$ near the throat. Before analyzing the shape function, we return to Ref. \cite{NFW96} to note some of the proposed forms of $\rho(r)$: \begin{equation*} \rho(r)\propto \frac{1}{r\left(1+\frac{r}{r_s}\right)^3} \quad \text{and} \quad \rho(r)\propto \frac{1}{r\left(1+\frac{r}{r_s}\right)^2}. \end{equation*} These forms suggest a more general starting point: \begin{equation}\label{E:start} \rho(r)= \frac{K}{r\left(1+\frac{r}{r_s}\right)^n}, \quad n>1, \end{equation} for some $K>0$. This form yields \begin{equation} b(r)=\frac{8\pi Kr_0^2}{(n-1)(n-2)}\frac{(n-1) \frac{r}{r_s}+1}{\left(1+\frac{r}{r_s}\right)^{n-1}}+C \end{equation} and \begin{equation} b'(r)=\frac{8\pi Kr_s}{n-2}\frac{1-\frac{(n-1)r/r_s+1} {1+r/r_s}} {\left(1+\frac{r}{r_s}\right)^{n-1}},\quad n\neq 2. \end{equation} We now see that if $n>2$, the second fraction of this product is negative, but if $n<2$, the first fraction is negative. Either way, $b'(r)<0$ for all $n\neq 2$, and we do not obtain a wormhole. So the only case allowed is $n=2$, which corresponds to Eq. (\ref{E:rho}). Here $b'(r)>0$: \begin{equation}\label{E:der} b'(r)=8\pi Kr_s\frac{\frac{r}{r_s}} {\left(1+\frac{r}{r_s}\right)^2}. \end{equation} We also have $b'(r)<1$ for all $r$, provided that $K<8\pi r_s$. This condition is easily met for all the cases discussed in TABLE 1 of Ref. \cite{NFW96}: $r_s$ is defined in terms of the ``virial" radius, which ranges from 177 kpc for a dwarf galaxy to 3740 kpc for a rich galaxy cluster. [See Ref. \cite{NFW96} for details.] That $b'(r)<1$ near the throat can also be seen from Fig. 1. Returning now to the shape function, Eq. (\ref{E:shape}), one way to determine the radius of the throat is to define the function $B(r)=b(r)-r$ and locate the root $r=r_{th}$ (if it exists) of $B(r)=b(r)-r=0$. In other words, we are treating $B(r)$ as if it were a function in rectangular coordinates. This approach is possible because of the spherical symmetry: we can move radially outward in any direction, thereby forming the $r$-axis. These facts will be exploited further in Sec. \ref{S:location}. Next, we observe that since $B'(r)= b'(r)-1$, the condition $b'(r)<1$ near the throat and $b'(r)>0$ imply that \begin{equation}\label{E:Bprime} -1<B'(r)<0. \end{equation} So $b(r)-r$ is strictly decreasing. If $b(r)-r$ does not intersect the $r$-axis, it must have a horizontal asymptote, so that $\text{lim}_{r\rightarrow \infty}(b'(r)-1)=0$. But this is impossible since $\text{lim}_{r\rightarrow\infty}b'(r)=0$ by Eq. (\ref{E:der}). So for some $r=r_{th}$, $b(r_{th})=r_{th}$, which is the throat of the wormhole. We will see in the next section that this conclusion can be reached more easily from the graph of $b=b(r)$, thereby yielding an alternative to the method in Ref. \cite{RKRI}. \section{Further discussion of $\rho(r)$} \label{S:rho} Since the discussion of our wormhole structure depends on Eq. (\ref{E:rho}), we need to take a closer look at the parameters used, as well as the coordinate system. In particular, a more complete description of the density is \begin{equation}\label{E:rho2} \rho(r)=\frac{\delta_c\rho_{crit}}{\frac{r}{r_s} \left(1+\frac{r}{r_s}\right)^2}. \end{equation} Here $\rho_{crit}$ is the critical density that is given by $\rho_{crit}=3H^2/8\pi G$, where $H$ is the current value of the Hubble constant. The other parameter is \begin{equation*} \delta_c=\frac{200}{3}\frac{c^3}{\text{ln}\,(1+c) -\frac{c}{1+c}} \end{equation*} with the incorporated concentration parameter \[ c=\frac{r_{200}}{r_s}, \] which is defined in terms of two additional parameters, $r_{200}$, the ``virial" radius and $r_s$, a characteristic scale radius. [See Ref. \cite{NFW96} for details.] As already noted, TABLE 1 in Ref. \cite{NFW96} lists the values of these parameters for different systems ranging from dwarf galaxies to rich galactic clusters. Being primarily interested in our own galaxy, we will use the values in Line 5 of TABLE 1, i.e., $r_s/r_{200} =0.060$, $r_{200}=348\,\text{kpc}$, resulting in $r_s=20.88$ kpc. For the time being we need to assume that the center of our wormhole is located at the origin $O$. This is described in Ref. \cite{NFW96} as the center of the halo, which, in turn, is the center of mass of certain ``clumps" \cite{NFW96}. Other possible locations of these wormholes are discussed in Sec. \ref{S:location}. Returning once again to the shape function, Eq. (\ref{E:shape}), observe that, qualitatively, $b(r)$ has the form \begin{equation}\label{E:form1} b(r)=A\left[\text{ln}\,\left(1+\frac{r}{B}\right) +\frac{1}{1+\frac{r}{B}}+C\right]. \end{equation} Fig. 1 shows that a throat of radius $r=r_{th}$ will \begin{figure}[htbp] \includegraphics[scale=.6]{throatold.eps} \caption{Qualitative features of the shape function showing the existence and position of the throat.} \label{fig:shape2} \end{figure} always exist. However, based on our model, there is no way to determine its size from purely geometric considerations. In other words, additional information is needed and this comes from the gravitational lensing discussed in the next section. \section{Gravitational lensing} As noted in the Introduction, Bozza \cite{vB02} provided an analytic method for calculating the deflection angle for any spherically symmetric spacetime in the strong field limit. To apply these methods to wormholes, we will follow the procedure in Refs. \cite{TL05, TL12}. As in most discussions on gravitational lensing, the line element is taken to be \begin{equation} ds^2=-A(x)dt^2+B(x)dx^2 +C(x)(d\theta^2+\sin^2\theta\,d\phi^2), \end{equation} where $x$ is the radial distance defined in terms of the Schwarzschild radius $x=r/2M$. Then \begin{equation}\label{E:closest} x_0=\frac{r_0}{2M} \end{equation} denotes the closest approach of the light ray. Using the lens equation in Ref. \cite{VE00}, it is shown in Refs. \cite{TL05, TL12} that the deflection angle $\alpha(x_0)$ consists of the sum of two terms: \begin{equation}\label{E:angle1} \alpha(x_0)=\alpha_e+I(x_0). \end{equation} Here \begin{equation}\label{E:angle2} \alpha_e=-2\, \text{ln}\left(\frac{2a}{3}-1\right) -0.8056 \end{equation} is due to the external Schwarzschild metric outside the wormhole's mouth $r=a$; $I(x_0)$ is the contribution from the internal metric, given by \begin{equation}\label{E:deflection1} I(x_0)=2\int^{\infty}_{x_0}\frac{\sqrt{B(x)}\,dx} {\sqrt{C(x)}\sqrt{\frac{C(x)A(x_0)}{C(x_0)A(x)}-1}}. \end{equation} From our line element (\ref{E:line3}) and the shape function (\ref{E:shape}), we obtain \begin{equation}\label{E:deflection2} I(x_0)=\\ \int^a_{x_0}Q(x)\,dx, \end{equation} where \begin{multline} Q(x)=\\ \frac{2}{\sqrt{x^2\left\{1- \frac{k}{2M}\frac{1}{x}\left[\text{ln}\, \left(1+\frac{x}{x_s}\right) +\frac{1}{1+\frac{x}{x_s}}+C\right]\right\}} \sqrt{\frac{x^{2-l}}{x_0^{2-l}}-1}}. \end{multline} Here $k=8\pi\rho_sr_s^3$ from Eq. (\ref{E:shape}). Observe that $\rho_sr_s^2$ is dimensionless, so that $\rho_sr_s^2=\rho_sx_s^2$. It follows that $k$ has units of length; hence, in Schwarzschild units, \begin{equation}\label{E:k} \frac{k}{2M}=8 \pi \rho_s x_s^2 \frac{r_s}{2M}=8 \pi \rho_s x_s^3. \end{equation} To see where this integral diverges, we make the change of variable $y=x/x_0$: \begin{equation}\label{E:deflection3} I(x_0)=\int_1^{a/x_0}R(y)\,dy, \end{equation} where \begin{multline} R(y)=\\ \frac{2}{\sqrt{(y^{4-l}-y^2) \left\{1-\frac{k}{2Mx_0}\frac{1}{y}\left[\text{ln} \left(1+\frac{yx_0}{x_s}\right)+ \frac{1}{1+\frac{yx_0}{x_s}}+C\right]\right\}}}. \end{multline} The radicand $F(y)$ in the denominator can be expanded in a Taylor series around $y=1$. Letting \begin{equation}\label{E:form2} g(y)=1-\frac{k/2M}{x_0}\frac{1}{y}\left[\text{ln} \left(1+\frac{yx_0}{x_s}\right)+ \frac{1}{1+\frac{yx_0}{x_s}}+C\right], \end{equation} we obtain \begin{multline} F(y)=(2-l)g(1)(y-1)\\+ \left[\frac{1}{2}(5-l)(2-l)g(1)+(2-l)g'(1)\right] (y-1)^2\\+ \text{higher powers}. \end{multline} If $g(1)\neq 0$, the integral converges due to the leading term $(y-1)^{1/2}$ resulting from the integration. If $g(1)=0$, then the second term leads to $\text{ln}\,(y-1)$, which causes the integral to diverge. (As an illustration, the integral \begin{equation*} \int_1^2\frac{dx}{\sqrt{3(x-1)+2(x-1)^2+(x-1)^3+ 5(x-1)^4}} \end{equation*} converges, but the integral \begin{equation*} \int_1^2\frac{dx}{\sqrt{2(x-1)^2+(x-1)^3+ 5(x-1)^4}} \end{equation*} does not.) So we need to examine $g(y)$ more closely. Observe that from Fig. 1, for any $y_m>0$, there exists a constant $C_1$ such that \begin{equation}\label{E:form3} \text{ln}\left(1+\frac{y_mx_0}{x_s}\right)+ \frac{1}{1+\frac{y_mx_0}{x_s}}+C_1=y_m. \end{equation} [The form of the function remains that of Eq. (\ref{E:form1}).] Hence from Eq. (\ref{E:form2}), \[ 1-\frac{k/2M}{x_0}\frac{1}{y_m}(y_m)=0 \] and $x_0=k/2M$. So for some $C_2$, i.e., translating in the vertical direction, \begin{multline}\label{E:form4} g(1)=\\1-\frac{k/2M}{x_0\times 1}\left[\text{ln} \left(1+\frac{1\times x_0}{x_s}\right)+ \frac{1}{1+\frac{1\times x_0}{x_s}}+C_2\right] =0, \end{multline} and again $x_0=k/2M$. In this manner we have obtained the value of the closest approach in Schwarzschild units. By reinterpreting Eq. (\ref{E:form4}), i.e., \begin{multline}\label{E:form5} g(1)=\\1-\frac{1}{2Mx_0}\left\{k\left[\text{ln} \left(1+\frac{x_0}{x_s}\right)+ \frac{1}{1+\frac{x_0}{x_s}}+C_2\right]\right\}=\\ 1-\frac{1}{r_0}\left\{k\left[\text{ln} \left(1+\frac{r_0}{r_s}\right)+ \frac{1}{1+\frac{r_0}{r_s}}+C_2\right]\right\}=0, \end{multline} we see that \[ 1-\frac{b(r_0)}{r_0}=0 \quad \text{and} \quad b(r_0)=r_0. \] So $r_0$, the closest approach, coincides with the throat, i.e., $r_0=r_{th}$. According to Refs. \cite{TL05, TL12}, this is also the radius of the photon sphere. It remains to determine the value of $k$. From Eq. (\ref{E:k}), \[ \frac{k}{2M}=8\pi\rho_sx_s^2\frac{r_s}{2M}=x_0 =\frac{r_0}{2M}, \] and we may revert to $k=8\pi\rho_sr_s^3=r_0=r_{th}$. We now find that \[ r_{th}=3.7466\times 10^{15}\, \text{m}\approx 0.40\, \text{ly}. \] \section{The location of the wormhole}\label{S:location} We saw earlier that the throat is the $r$-intercept of $B(r)=b(r)-r$ since, due to the spherical symmetry, $B(r)$ is the same function along any outward ray emanating from the origin. In similar manner, if a shape function is viewed as a translation to or from the origin along this ray, this would correspond to a horizontal translation in Fig. 1. Upon closer examination, however, such a translation would affect the radius of the throat. But recalling that $b(r)$ has the form in Eq.(\ref{E:form1}), namely, \begin{equation*} b(r)=A\left[\text{ln}\,\left(1+\frac{r}{B}\right) +\frac{1}{1+\frac{r}{B}}+C\right], \end{equation*} a larger $B$ will cause the curve to ``flatten," as shown in Fig. 2. Since $r_s\approx 20.88$ kpc is \begin{figure}[htbp] \includegraphics[scale=.6]{throatnew.eps} \caption{A horizontal translation of $b(r)$ has a negligible effect on the throat size.} \label{fig:shape2} \end{figure} indeed very large compared to $r$ not too far from the origin $O$, the effect of the translation on the throat size is small or even negligible. Suppose that the center of the wormhole is at $O'$, the new origin (Fig. 3). Then the throat radius relative to $O'$ \begin{figure}[htbp] \includegraphics[scale=.6]{throatoprime.eps} \caption{A wormhole centered at $O$ cannot be distinguished from a wormhole centered at $O'$.} \label{fig:shape2} \end{figure} is approximately the same as in Fig. 1. So mathematically speaking, $b(r)$ has the same properties regardless of the location along the ray. In other words, mathematically, and hence physically, we cannot distinguish between a wormhole centered at $O$ from a wormhole centered at $O'$. In particular, the radius of the throat is approximately 0.40 ly for all wormholes in our model, provided, of course, that they exist. Since the radius of the throat corresponds to the radius of the photon sphere, such a sphere would be detectable, thereby providing observational evidence for the existence of the wormhole. To obtain further evidence, it would be highly desirable to obtain at least a rough estimate of the radius $r=a$ of the mouth of the wormhole. Here we return to the origin $O$. As noted earlier, the wormhole spacetime is not asymptotically flat, raising the question of a possible cutoff at $r=a$ and the junction to an external Schwarzschild spacetime. Here we are going to assume that even though we are still in the halo region, the exterior solution is close enough to Schwarzschild to permit a rough estimate of $r=a$. To that end we recall that the mass of the wormhole is $M= \frac{1}{2}b(a)$, where $a<r_s$. Using the redshift function in line element (\ref{E:line3}), $a$ is determined from the condition $(a/b_0)^l =1-2M/a$, i.e., \begin{equation}\label{E:junction} \left(\frac{a}{b_0}\right)^{l}=1- \frac{k\left[\text{ln}\left(1+\frac{a}{r_s} \right)+\frac{1}{1+\frac{a}{r_s}}+C\right] }{a}, \end{equation} where $b_0$ is a constant of integration. Before continuing, we need to estimate the value of $C$, given the earlier parameters. Since $k= r_{th}$ and $b(r_{th})=r_{th}$, we have from the shape function \[ b(r)=k\left[\text{ln}\left(1+\frac{r}{r_s} \right) +\frac{1}{1+\frac{r}{r_s}}+C\right] \] that \[ \text{ln}\left(1+\frac{r_{th}}{r_s}\right) +\frac{1}{1+\frac{r_{th}}{r_s}}+C=1, \] which leads to $C\approx -1.7\times 10^{-11}$. It follows that $C$ can be safely neglected in Eq. (\ref{E:junction}). A return to the parameters discussed in Sec. \ref{S:rho} now offers some posibilities: suppose we let $b_0=r_{200}=348\,\text{kpc}$, the virial radius, and then use $r_s=20.88\,\text{kpc}=6.44\times 10^{20}\, \text{m}$, $l=0.000001$, and $k= 3.7466\times 10^{15}\,\text{m}$. These values suggest that we should expect $r=a$ to be in excess of 1 $r_s$. Evidence for the existence of a mouth can therefore be sought in this range. It has been suggested that for wormholes in our halo, the Large Magellanic Cloud can be used for applying the gravitational lensing technique, possibly even using past data. For the location of the mouth, we need to recall that according to Refs. \cite{TL05, TL12}, photons with closest approach greater than the wormhole's mouth have a Schwarzschild lensing effect. For both throat and mouth, these effects can in principle be detected. \\ \\ \emph{Remark:} While it is known that the NFW model predicts velocities in the central parts that are too low \cite{gG14}, these discrepancies vanish in the more remote regions of the halo \cite{cT06, aM12}. \section{Summary} Given the Navarro-Frenk-White density profile of halos, it was shown in Ref. \cite{RKRI} (and confirmed in this paper) that these halos posses some of the characteristics that could give rise to traversable wormholes. Given that $n=2$ is the only allowed value in Eq. (\ref{E:start}), we also obtained a partial converse: the existence of traversable wormholes helps determine the basic form of the Navarro-Frenk-White density profile. The shape function, obtained from this profile, meets the flare-out condition at the throat. Such a throat always exists, but its location cannot be determined from the geometry. However, using a method for calculating the deflection angle pioneered by Bozza \cite{vB02}, it is shown that the deflection angle diverges at the throat. The resulting photon sphere has a radius of about 0.40 ly regardless of the location of the wormhole. Detection may be possible using past data. Since the dark matter in the halo region does not interact with light, a suitable vehicle would be ordinary light from the Large Magellanic Cloud.
\section{Introduction} We study a general class of complex non-linear segmentation energies with high-order regional terms. Such energies are often desirable in the computer vision tasks of image segmentation, co-segmentation and stereo \cite{kkz:iccv03,freedman:pami04,rother:cvpr06,Ismail:LevelSetsWithArea08,ismail:cvpr10,KC11:iccv,linesearchcuts:12,FTR:cvpr13} and are particularly useful when there is a prior knowledge about the appearance or the shape of an object being segmented. We focus on energies of the following form: \begin{equation} \label{eq:ERL} \min_{S\in \Omega}E(S) = R(S) + \lambda L(S), \end{equation} where $S$ is a binary segmentation, $L(S)$ is a standard length-based smoothness term, and $R(S)$ is a (generally) non-linear {\em regional functional} discussed below. \begin{figure}[t] \begin{center} \includegraphics[width = 0.4\textwidth]{yuri_lena.pdf} \bf\caption{\rm% Gradient descent ({\em level sets}) and {\em trust region}. \label{fig:yuri_lena}} \end{center} \end{figure} Let $I:\Omega\to\mathbb{R}^m$ be an image defined in $\Omega \subset \mathbb{R}^n$. The most basic type of regional terms used in segmentation is a linear functional $U(S)$, which can be represented via an arbitrary scalar function $f:\Omega\to\mathbb{R}$ \begin{align}\label{eq:LinearRegionalFunctional} U(S) = \int_S f\,\text{d}x = \int_\Omega f\cdot 1_S\,\text{d}x =: \vecprod{f}{S}. \end{align} Usually, $f$ corresponds to an {\em appearance model} based on image intensities/colors $I$, e.g.,~$f(x)$ could be a log-likelihood ratio for intensity $I(x)$ given particular object and background intensity distributions. The integral in \eqref{eq:LinearRegionalFunctional} can be seen as a dot product between scalar function $f$ and $1_S$, which denotes the characteristic function of set $S$. We use notation $\vecprod{f}{S}$ to refer to such linear functionals. More general \emph{non-linear regional functional} $R(S)$ can be described as follows. Assume $k$ scalar functions $f_1,\ldots,f_k:\Omega\to\mathbb{R}$, each defining a linear functional $\vecprod{f_i}{S}$ of type~\eqref{eq:LinearRegionalFunctional}, and one differentiable non-linear function $F(v_1,\ldots,v_k)\colon\mathbb{R}^k\to\mathbb{R}$ that combines them, \begin{align}\label{eq:NonlinearRegionalFunctional} R(S) = F\left(\vecprod{f_1}{S},\ldots,\vecprod{f_k}{S}\right). \end{align} Such general regional terms could enforce non-linear constraints on volume or higher-order shape moments of segment $S$. They could also penalize $L_2$ or other distance metrics between the distribution (or non-normalized bin counts) of intensities/colors inside segment $S$ and some given target. For example, $S$ could be softly constrained to specific volume $V_0$ via quadratic functional $$R(S) = (\vecprod{1}{S}-V_0)^2$$ using $f_1(x)=1$ and $F(v)=(v-V_0)^2$, while the Kullback-Leibler (KL) divergence between the intensity distribution in $S$ and a fixed target distribution $q=(q_1,\ldots,q_k)$ could be written as $$R(S) = \sum_i^k \frac{\vecprod{f_i}{S}}{\vecprod{1}{S}}\log\left(\frac{\vecprod{f_i}{S}}{\vecprod{1}{S}\cdot q_i}\right).$$ Here scalar function $f_i(x)$ is an indicator for pixels with intensity $i$ (in bin $i$). More examples of non-linear regional terms $R(S)$ are discussed in Section~\ref{sec:experiments}\@. In general, optimization of non-linear regional terms is NP-hard and cannot be addressed by standard global optimization methods. Some earlier papers developed specialized techniques for particular forms of non-linear regional functionals. For example, the algorithm in \cite{ismail:cvpr10} was developed for minimizing the Bhattacharyya distance between distributions and a dual-decomposition approach in \cite{woodford:iccv09} applies to convex piece-wise linear regional functionals \eqref{eq:NonlinearRegionalFunctional}. These methods are outside the scope of our paper since we focus on more general techniques. Combinatorial techniques \cite{kkz:iccv03,rother:cvpr06} apply to general non-linear regional functionals, and they can make large moves by globally minimizing some approximating energy. However, as pointed out in \cite{linesearchcuts:12}, these methods are known to converge to solutions that are not even a local minimum of energy \eqref{eq:ERL}. Level sets are a well established in the litterature as a gradient descent framework that can address arbitrary differentiable functionals \cite{LevelSetsBook:Ismail11}, and therefore, are widely used for high-order terms \cite{Adam2009,Foulonneau2009,BenAyed2009,Michailovich2007,freedman:pami04}. This paper compares two known optimization methods applicable to energy \eqref{eq:ERL} with any high-order regional term \eqref{eq:NonlinearRegionalFunctional}: {\em trust region} and {\em level sets}. To address non-linear term $R(S)$, both methods use its first-order functional derivative \begin{align}\label{eq:dR} \frac{\partial R}{\partial S} = \sum_{i=1}^k \frac{\partial F}{\partial v_i} \left(\vecprod{f_1}{S},\ldots,\vecprod{f_k}{S}\right) \cdot f_i \end{align} either when computing gradient flow for \eqref{eq:ERL} in level sets or when approximating energy \eqref{eq:ERL} in trust region. However, despite using the same first derivative $\frac{\partial R}{\partial S}$, the level sets and trust region are fairly different approaches to optimize~\eqref{eq:ERL}. The structure of the paper is as follows. Sections~\ref{sec:ls}-\ref{sec:tr} review the two general approaches that we compare: gradient descent implemented with level sets and trust region implemented with graph cuts. While the general trust region framework \cite{FTR:cvpr13} can be based on a number of underlying global optimization techniques, we specifically choose a graph cut implementation (versus continuous convex relaxation approach), since it is more appropriate for our CPU-based evaluations. Our experimental results and comparisons are reported in Section~\ref{sec:experiments} and Section~\ref{sec:conclusion} presents the conclusions. \section{Overview of Algorithms}\label{sec:overview} Below we provide general background information on level sets and trust region methods, see Sections \ref{sec:ls}--\ref{sec:tr}. High-level conceptual comparison of the two frameworks is provided in Section \ref{sec:comp}. \subsection{Standard Level Sets}\label{sec:ls} In the level set framework, minimization of energy $E$ is carried out by computing a partial differential equation (PDE), which governs the evolution of the boundary of $S$. To this end, we derive the Euler-Lagrange equation by embedding segment $S$ in a one-parameter family $S(t)$, $t \in \mathbb{R}^{+}$, and solving a PDE of the general form: \begin{equation} \label{eq:pde} \frac{\partial S}{\partial t} = -\frac{\partial E(S)}{\partial S} = - \frac{\partial R(S)}{\partial S} - \frac{\partial L(S)}{\partial S} \end{equation} where $t$ is an artificial time step parameterizing the descent direction. The basic idea is to describe segment $S$ implicitly via an embedding function $\phi: \Omega \rightarrow \mathbb{R}$: \begin{eqnarray} \label{level_set_region_membership} S &=& \{ x \in \Omega | \phi(x)\leq 0 \} \nonumber \\ \Omega \setminus S &=& \{ x \in \Omega | \phi(x) > 0 \}, \end{eqnarray} and evolve $\phi$ instead of $S$. With the above representation, the terms that appear in energy (\ref{eq:ERL}) can be expressed as functions of $\phi$ as follows \cite{Chan01ip,LevelSetsBook:Ismail11}: \begin{eqnarray}\label{eq:LSlength} L(S) &=& \int_{\Omega} \|\nabla H(\phi)\|dx = \int_{\Omega}\delta (\phi)\|\nabla \phi \|dx \nonumber \\ \vecprod{f_i}{S} &=& \int_\Omega H(\phi) f_i dx. \end{eqnarray} Here, $\delta$ and $H$ denote the Dirac function and Heaviside function, respectively. Therefore, the evolution equation in (\ref{eq:pde}) can be computed directly by applying the Euler-Lagrange descent equation with respect to $\phi$. This gives the following gradient flow: \begin{equation} \label{curve-flow-withoutSDF} \frac{\partial \phi}{\partial t} = \left[-\frac{\partial R(S)}{\partial S}+\lambda\kappa\right]\delta(\phi) \end{equation} with $\kappa:=\Div\left(\frac{\nabla\phi}{\norm{\nabla\phi}}\right)$ denoting the curvature of $\phi$'s level lines. The first term in (\ref{curve-flow-withoutSDF}) is a regional flow minimizing $R$, and the second is a standard curvature flow minimizing the length of the segment's boundary. In standard level set implementations, it is numerically mandatory to keep the evolving $\phi$ close to a distance function \cite{Li2005a,Osher2002}. This can be done by re-initialization procedures \cite{Osher2002}, which were intensively used in classical level set methods \cite{Caselles97ijcv}. Such procedures, however, rely on several {\em ad hoc} choices and may result in undesirable side effects \cite{Li2005a}. In our implementation, we use an efficient and well-known alternative \cite{Li2005a}, which adds an internal energy term that penalizes the deviation of $\phi$ from a distance function: \begin{equation} \label{distance-function-penalty} \frac{\mu}{2} \int_{\Omega} \left (1 - \|\nabla \phi\| \right)^2 dx . \end{equation} In comparison to re-initialization procedures, the implementation in \cite{Li2005a} allows larger time steps (and therefore faster curve evolution). Furthermore, it can be implemented via simple finite difference schemes, unlike traditional level set implementations which require complex upwind schemes \cite{Sethian1999}. With the distance-function penalty, the gradient flow in (\ref{curve-flow-withoutSDF}) becomes: \begin{equation} \label{curve-flow-withSDF} \frac{\partial \phi}{\partial t} = \mu \left [ \Delta\phi - \kappa \right ] + \left[-\frac{\partial R(S)}{\partial S}+\lambda \kappa\right]\delta(\phi) \end{equation} For all the experiments in this paper, we implemented the flow in (\ref{curve-flow-withSDF}) using the numerical prescriptions in \cite{Li2005a}. For each point $p$ of the discrete grid, we update the level set function as \begin{equation} \label{Discrete-Level-Set-Updates} \phi^{j+1}(p) = \phi^{j}(p) + \Delta t\cdot A(\phi^{j}(p)), \end{equation} where $\Delta t$ is the discrete time step and $j$ is the iteration number. $A(\phi^{j}(p))$ is a numerical approximation of the right-hand side of (\ref{curve-flow-withSDF}), where the spatial derivatives of $\phi$ are approximated with central differences and the temporal derivative with forward differences. The Dirac function is approximated by $\delta_{\epsilon}(t) = \frac{1}{2\epsilon} [1+\cos(\frac{\pi t}{\epsilon})]$ for $|t| \leq \epsilon$ and $0$ elsewhere. We use $\epsilon = 1.5$ and $\mu=0.05$. In the context of level set and PDE methods, it is known that the choice of time steps should follow strict numerical conditions to ensure stability of front propagation, e.g., the standard Courant-Friedrichs-Lewy (CFL) conditions \cite{Estellers2012}. These conditions require that $\Delta t$ should be smaller than a certain value $\tau$ that depends on the choice of discretization. The level set literature generally uses fixed time steps. For instance, classical upwind schemes \cite{Sethian1999} generally require a small $\Delta t$ for stability, whereas the scheme in \cite{Li2005a} allows relatively larger time steps. The optimum time step is not known a priori and finding a good $\Delta t < \tau$ via an adaptive search such as back-tracking \cite{Boyd2004} seems attractive. However, to apply a back-tracking scheme, we would have to evaluate the energy at each step. In the case of level sets, this requires a discrete approximation of the original continuous energy. We observed in our experiments that the gradient of such discrete approximation of the energy does not coincide with the gradient obtained in the numerical updates in~ \eqref{Discrete-Level-Set-Updates}. Therefore, with a back-tracking scheme, level sets get stuck very quickly in a local minimum of the discrete approximation of the energy (See the adaptive level set example in Fig. \ref{fig:volume}). We believe that this is the main reason why, to the best of our knowledge, back-tracking approaches are generally avoided in the level-set literature. Therefore, in the following level-set experiments, we use a standard scheme based on fixed time step $\Delta t$ during each curve evolution, and report the performance at convergence for several values $\Delta t \in \{1 \ldots 10^3\}$. \subsection{Trust Region Framework}\label{sec:tr} Trust region methods are a class of iterative optimization algorithms. In each iteration, an approximate model of the optimization problem is constructed near the current solution. The model is only ``trusted'' within a small region around the current solution called ``trust region'', since in general, approximations fit the original non-linear function only locally. The approximate model is then globally optimized within the trust region to obtain a candidate iterate solution. This step is often called {\em trust region sub-problem}. The size of the trust region is adjusted in each iteration based on the quality of the current approximation. Variants of trust region approach differ in the kind of approximate model used, optimizer for the trust region sub-problem step and a protocol to adjust the next trust region size. For a detailed review of trust region methods see \cite{TRreview:Yuan}. Below we outline a general version of a trust region algorithm in the context of image segmentation. The goal is to minimize $E(S)$ in Eq.~\eqref{eq:ERL}. Given solution $S_j$ and distance $d_j$, the energy $E$ is approximated using \begin{equation}\label{eq:TR_subproblem} \widetilde{E}(S) = U_0(S) + L(S), \end{equation} where $U_0(S)$ is the first order Taylor approximation of the non-linear term $R(S)$ near $S_j$. The trust region sub-problem is then solved by minimizing $\widetilde{E}$ within the region given by $d_j$. Namely, \begin{equation} \label{eq:constrained} S^*= \underset{||S-S_j||<d}{\operatorname{argmin}} \tilde{E}(S). \end{equation} Once a candidate solution $S^*$ is obtained, the quality of the approximation is measured using the ratio between the actual and predicted reduction in energy. The trust region is then adjusted accordingly. For the purpose of our CPU-based evaluations we specifically selected the Fast Trust Region (FTR) implementation \cite{FTR:cvpr13} which includes the following components for the trust region framework. The non-linear term $R(S)$ is approximated by the first order Taylor approximation $U_0(S)$ in \eqref{eq:TR_subproblem} using first-order functional derivative \eqref{eq:dR}. The trust region sub-problem in \eqref{eq:constrained} is formulated as unconstrained Lagrangian optimization, which is globally optimized using one graph-cut (we use a floating point precision in the standard code for graph-cuts \cite{BK:PAMI04}). Note that, in this case, the length term $L(S)$ is approximated using Cauchy-Crofton formula as in \cite{GeoCuts:ICCV03}. More details about FTR can be found in \cite{FTR:cvpr13}. \subsection{Conceptual Comparison}\label{sec:comp} \begin{figure*}[t] \begin{center} \includegraphics[width = 1\textwidth]{volume.pdf} \bf\caption{\rm% Volume constraint with boundary length regularization. We set the weights to $\lambda_{Length}=1$, $\lambda_{Volume}=10^{-4}$. \label{fig:volume}} \end{center} \end{figure*} Some high-level conceptual differences between the {\em level sets} and {\em trust region} optimization frameworks are summarized in Figure \ref{fig:yuri_lena}. Standard level sets methods use fixed $\Delta t$ to make steps $-\Delta t \cdot \frac{\partial E}{\partial S}$ in the gradient descent direction. Trust region algorithm \cite{FTR:cvpr13} moves to solution $S^*$ minimizing approximating functional $\tilde E(S)$ within a circle of given size $d$. The trust region size is adaptively changed from iteration to iteration based on the observed approximation quality. The blue line illustrates the spectrum of trust region moves for all values of $d$. Solution $\tilde{S}$ is the global minimum of approximation $\tilde{E}(S)$. For example, if $\tilde{E}(S)$ is a 2nd-order Taylor approximation of $E(S)$ at point $S_0$ then $\tilde{S}$ would correspond to a Newton's step. \section{Experimental Comparison}\label{sec:experiments} In this section, we compare trust region and level sets frameworks in terms of practical efficiency, robustness and optimality. We selected several examples of segmentation energies with non-linear regional constraints. These include: 1) quadratic volume constraint, 2) shape prior in the form of $L_2$ distance between the target and the observed shape moments and 3) appearance prior in the form of either $L_2$ distance, Kullback-Leibler divergence or Bhattacharyya distance between the target and the observed color distributions. In all the experiments below, we optimize energy of a general form $E(S) = R(S) + L(S)$. To compare optimization quality, we will plot energy values for the results of both level sets and trust region. Note that the Fast Trust Region (FTR) implementation uses a discrete formulation based on graph-cuts and level sets are a continuous framework. Thus, the direct comparison of their corresponding energy values should be done carefully. While numerical evaluation of the regional term $R(S)$ is equivalent in both methods, they use completely different numerical approaches to measuring length $L(S)$. In particular, level sets use the approximation of length given in \eqref{eq:LSlength}, while the graph cut variant of trust region relies on integral geometry and Cauchy-Crofton formula popularized by \cite{GeoCuts:ICCV03}. \begin{figure*}[t] \begin{center} \includegraphics[width = 0.8\textwidth]{liver.pdf} \bf\caption{\rm% Shape prior constraint with length regularization and log-likelihood models. Target shape moments and appearance models are computed from the provided ellipse. We used 100 intensity bins, moments up to order $l=2$, $\lambda_{Length}=10$, $\lambda_{Shape}=0.01$ and $\lambda_{App}=1$. The continuous energy is plotted starting from $4^{th}$ iteration to reduce the range of the y-axis. \label{fig:liver}} \end{center} \end{figure*} Since the energies are not comparable by its actual number, we study instead the robustness of each method independently and compare the resulting segmentation with one another. Note that level sets have much smaller oscillation for small time steps, which supports the theory of the CFL-conditions. In each application below we examine the robustness of both trust region and level sets methods by varying the running parameters. In the trust region method we vary the multiplier $\alpha$ used to change the size of the trust region from one iteration to another. For the rest of the parameters we follow the recommendations of~\cite{FTR:cvpr13}. In our implementation of level sets we vary the time-step size $\Delta t$, but keep the parameters $\epsilon=1.5$ and $\mu=0.05$ fixed for all the experiments. The top-left plots in figures \ref{fig:volume}-\ref{fig:mushroom} report energy $E(S)$ as a function of the CPU time. At the end of each iteration, both level sets and trust region require energy updates, which could be computationally expensive. For example, appearance-based regional functionals require re-evaluation of color histograms/distributions at each iteration. This is a time consuming step. Therefore, for completeness of our comparison, we report in the top-middle plots of each figure energy $E(S)$ versus the number of energy evaluations (number of updates) required during the optimization. \subsection{Volume Constraint} First, we perform image segmentation with a volume constraint with respect to a target volume $V_0$, namely, $$R(S) = (\vecprod{1}{S}-V_0)^2.$$ We choose to optimize this energy on a synthetic image without appearance term since the solution to this problem is known to be a circle. Figure \ref{fig:volume} shows that both FTR and level sets converge to good solutions (nearly circle), with FTR being 25 times faster, requiring 150 times less energy updates and exhibiting more robustness to the parameters. \begin{figure*}[t] \begin{center} \includegraphics[width = 0.8\textwidth]{soldier_length.pdf} \bf\caption{\rm% $L_2$ norm between the observed and target color bin counts with length regularization. We used 100 bins per channel, $\lambda_{App} =1$ and $\lambda_{Length} =1$. \label{fig:soldier_L2L}} \end{center} \end{figure*} \subsection{Shape Prior with Geometric Shape Moments} Next, we perform image segmentation with a shape prior constraint in the form of $L_2$ distance between the geometric shape moments of the segment and a target. Our energy is defined as $E(S) = \lambda_{Shape}R(S) + \lambda_{Length}L(S) + \lambda_{App}D(S)$, where $D(S)$ is a standard log-likelihood unary term based on intensity histograms. In this case, $R(S)$ is given by $$ R(S) =\sum_{p+q\leq l}(\vecprod{x^py^q}{S}-m_{pq})^2, $$ with $m_{pq}$ denoting the target geometric moment of order $l=p+q$. Figure \ref{fig:liver} shows an example of liver segmentation with the above shape prior constraint. The target shape moments as well as the foreground and background appearance models are computed from the user provided input ellipse as in \cite{KC11:iccv,FTR:cvpr13}. We used moments of up to order $l=2$ (including the center of mass and shape covariance but excluding the volume). Both trust region and level sets obtain visually pleasing solutions. The trust region method is two orders of magnitude faster and requires two orders of magnitude less energy updates (top-left and top-middle plots). Since the level sets method was forced to stop after 10000 iterations, we show the last solution available for each value of parameter $\Delta t$ . The actual convergence for this method would have taken more iterations. In this example, the oscillations of the energy are especially pronounced (top-right plot). \subsection{Appearance Prior} \exclude{ \begin{figure*}[t] \begin{center} \includegraphics[width = 0.8\textwidth]{imgs/soldier.pdf} \end{center} \caption{$L_2$ norm between the observed and target color bin counts, no length regularization ($\lambda_{Length}=0$). In this case, the discrete and continuous energies are equal. We used 100 bins per channel.\label{fig:soldier_L2}} \end{figure*} } In the experiments below, we apply both methods to optimize segmentation energies where the goal is to match a given target appearance distribution using either the $L_2$ distance between the observed and target color bin counts, or the Kullback-Leibler divergence and Bhattacharyya distance between the observed and target color distributions. Here, our energy is defined as $E(S) = \lambda_{App}R(S) + \lambda_{Length}L(S)$. We assume $f_i$ is an indicator function of pixels belonging to bin $i$ and $q_i$ is the target count (or probability) for bin $i$. The target appearance distributions for the object and the background were obtained from the ground truth segments. We used 100 bins per color channel. The images in the experiments below are taken from \cite{RKB:SIGGRAPH04}. \paragraph{$L_2$ distance constraint on bin counts:} Figure \ref{fig:soldier_L2L} shows results of segmentation with $L_2$ distance constraint between the observed and target bin counts regularized by length. The regional term in this case is $$R(S) = \sqrt{\sum_{i=1}^{k} (\vecprod{f_i}{S}-q_i)^2}.$$ Since the level sets method was forced to stop after 15000 iterations for values of $\Delta t=1,5$, we show the last solution available. Full convergence would have taken more iterations. For higher values of $\Delta t$, we show results at convergence. We observe two orders of magnitude difference between the trust region and level sets method in terms of the speed and the number of energy updates required. \paragraph{Kullback-Leibler divergence:} Figure \ref{fig:llama} shows results of segmentation with KL divergence constraint between the observed and target color distributions. The regional term in this case is given by $$R(S)= \sum_{i=1}^k\frac{\vecprod{f_i}{S}}{\vecprod{1}{S}}\log\left(\frac{\vecprod{f_i}{S}}{\vecprod{1}{S}q_i}\right).$$ The level sets method converged for times steps $\Delta t=50$ and $1000$ , but was forced to stop for other values of the parameter after $10000$ iterations. We show the last solution available. Full convergence would have required more iterations. The trust region method obtains solutions that are closer to the ground truth, runs two order of magnitude faster and requires less energy updates. \begin{figure*}[t] \begin{center} \includegraphics[width = 0.8\textwidth]{llama.pdf} \bf\caption{\rm% KL divergence between the observed and the target color distribution. We used 100 bins per channel, $\lambda_{App}=100$ and $\lambda_{Length}=0.01$. Continuous energy is plotted starting from forth iteration to reduce the range of the y-axis. \label{fig:llama}} \end{center} \end{figure*} \paragraph{Bhattacharyya divergence:} Figure \ref{fig:mushroom} shows results of segmentation with Bhattacharyya distance constraint between the observed and target color distributions. The regional term in this case is given by $$R(S)= -\log\left(\sum_{i=1}^k \sqrt{\frac{\vecprod{f_i}{S}}{\vecprod{1}{S}}q_i}\right).$$ Also for this image, since the level sets method had not (yet) converged after 10000 iterations for any set of the parameters, we show the last solution available. Further increasing parameter $\Delta t$ would increase the oscillations of the energy (see top-right plot). \section{Conclusions}\label{sec:conclusion} For relatively simple functionals \eqref{eq:NonlinearRegionalFunctional}, combining a few linear terms ($k$ is small), such as constraints on volume and low-order shape moments, the quality of the results obtained by both methods is comparable (visually and energy-wise). However, we observe that the number of energy updates required for level sets is two orders of magnitude larger than for trust region. This behavior is consistent with the corresponding CPU running time plots. The segmentation results on shape moments were fairly robust with respect to parameters (time step $\Delta t$ and multiplier $\alpha$) for both methods. The level sets results for volume constraints varied with the choice of $\Delta t$. In general, larger steps caused significant oscillations of the energy in level sets thereby affecting the quality of the result at convergence. When optimizing appearance-based regional functionals with large number of histogram bins (corresponding to large $k$), level sets proved to be extremely slow. Convergence would require more than $10^4$ iterations (longer than 1 hour on our machine). In some cases, the corresponding results were far from optimal both visually and energy-wise. This is in contrast to the results by trust region approach, which consistently converged to plausible solutions with low energy in less than a minute or $100$ iterations. \exclude{ \FS{ It is worth noting that the slow convergence of the level set method is not only due to the curvature flow (see the example without length in Fig.~\ref{fig:soldier_L2}), but also due to the distance function penalty in \eqref{distance-function-penalty}. In our experiment we observed that increasing the time step in level sets does not necessarily correspond to larger moves within each iteration. }{} } We believe that our results will be useful for many practitioners in computer vision and medical imaging when selecting an optimization technique. \begin{figure*}[t] \begin{center} \includegraphics[width = 0.8\textwidth]{mushroom.pdf} \bf\caption{\rm % Bhattacharyya distance between the observed and the target color distribution. We used 100 bins per channel, $\lambda_{App}=1000$ and $\lambda_{Length}=0.01$. \label{fig:mushroom}} \end{center} \end{figure*} {\small \bibliographystyle{ieee}
\section{Introduction} Waveguides with non-zero curvature are basic constituents of matter-wave circuits in atom chip technology \cite{atomchip,chips}, as well as its ion \cite{ionchip}, molecular \cite{molecularchip}, and electron \cite{electronchip} counterparts. Their relevance is further enhanced by the development of flexible techniques to create optical waveguides for ultracold gases. In this context, waveguide trapping potentials can be engineered by a variety of methods including the time-averaging painted potential technique \cite{painters}, the use of an intensity mask \cite{Anderson07,Heinzen09}, and holographic methods, in particular, digital holography \cite{GH12}. Circular ring traps have attracted a considerable amount of attention \cite{ring1,ring2,ring3,ring4,ring5,ring6,painters}, most recently to study Josephson junction dynamics \cite{Wright2013, Ryu13}. Other curved waveguides have also been engineered, such as a stadium-shaped potential trap \cite{stadium04,Heller06}. The propagation of matter-waves in bent waveguides generally differs from that in straight waveguides due to the appearance of a purely attractive local quantum potential of geometrical origin \cite{SRS77,daCosta81,daCosta82}. Under tight-transverse confinment, the magnitude of this curvature-induced potential (CIP) is proportional to the square of the curvature of the waveguide, and affects both the single-particle and many-body physics of the confined matter-waves \cite{EB89,GJ92,CB96,Clark98,EV99,LP01,Schwartz06}. As a result, the scattering properties of a curved tight waveguide are modified, e.g., by the appearance of bound states \cite{EB89,GJ92}. Advances in the design of bent waveguides, in which curvature-induced effects are tailored and suppressed, are required for the miniaturization of matter-wave circuits. It is to this problem that we turn our attention. \section{Results} In this manuscript we design bent waveguides for matter-wave circuits free from spurious quantum mechanical effects associated with CIPs. Three novel ideas are presented: (i) Exploiting the interplay of geometry and supersymmetry in quantum mechanics, we relate pairs of waveguides whose CIPs are isospectral and share the same scattering properties. (ii) We then identify waveguides which are reflectionless for coherent matter-waves at all energies. (iii) Furthermore, we show that by tailoring the depth of the waveguide trap, it is possible to cancel the CIP, rendering the dynamics of the guided matter-waves equivalent to that in straight waveguides. Let us consider the dynamics of matter-waves confined in a tight waveguide whose axis follows the curve $\gamma$, parametrized as a function of the arc length $q_1$ by the vector ${\bf r}={\bf r}(q_1)$, with tangent ${\bf t}(q_1)$. We start by recalling the fundamental theorem of curves which asserts that a curve is completely determined, up to its position in space, by its curvature $\kappa$ and torsion $\tau$ \cite{Struik88}. Indeed, the expressions $\kappa=\kappa(q_1)$ and $\tau=\tau(q_1)$ constitute the natural intrinsic equations of a curve. A parametrization of the curve can be obtained by integration of the Frenet-Serret equations \begin{eqnarray} d_{q_1} \begin{pmatrix} \hat{\rm\bf t} \\ \hat{\rm\bf n} \\ \hat{\rm\bf b} \\ \end{pmatrix} = \begin{pmatrix} 0 & \kappa & 0\\ -\kappa & 0 & \tau \\ 0 & -\tau & 0 \\ \end{pmatrix} \begin{pmatrix} \hat{\rm\bf t} \\ \hat{\rm\bf n} \\ \hat{\rm\bf b} \\ \end{pmatrix}, \end{eqnarray} where $\hat{\rm\bf n}=(d\hat{\rm\bf t}/dq_1)/\kappa$ (provided $d\hat{\rm\bf t}/dq_1\neq 0$) and $\hat{\rm\bf b}=\hat{\rm\bf t}\times\hat{\rm\bf n}$ are the principal normal and binormal unit vectors, and the curvature and torsion at the point of arc length $q_1$ are defined as $\kappa(q_1)=|{\rm d}{\bf t}(q_1)/dq_1|$ and $\tau(q_1)=-\hat{\rm\bf n}\cdot d\hat{\rm\bf b}/dq_1$, respectively. Let $(q_2,q_3)$ be the transverse local coordinates and consider a transverse confining potential ${\rm U}_{\lambda}(q_2,q_3)$ such that in the limit of tight confinement $\lambda\rightarrow\infty$ the particle is bounded to $\gamma$. Under dimensional reduction, the purely-attractive CIP emerges \cite{daCosta81,daCosta82,EB89,GJ92,CB96} \begin{eqnarray} \label{daCostapot} {\rm V}(q_1)=-\frac{\hbar^2}{8m}\kappa(q_1)^2. \end{eqnarray} This result is independent of ${\rm U}_{\lambda}$ and holds in particular under an isotropic transverse harmonic confinement ${\rm U}_{\omega_{\perp}}=m\omega_{\perp}^2(q_2^2+q_3^2)/2$ with ground state width $\sigma_0=[\hbar/(m\omega_{\perp})]^{1/2}$ \cite{LP01,Schwartz06}. The conditions for the dimensional reduction to be valid explicitly read \begin{eqnarray} \label{cc} \kappa \sigma_0\ll 1,\qquad |\kappa'|\sigma_0\ll|\kappa|,\qquad |\kappa''|\sigma_0\ll\kappa^2, \end{eqnarray} where primes denote derivatives with respect to $q_1$. We next pose the problem of identifying pairs of waveguides with the same scattering properties, and engineering a waveguide which minimizes the effect of the CIP. Generally, direct integration of the Frenet-Serret equations is not possible. However, the quantum mechanical behavior of matter waves bounded to isometric curves with different torsion but the same curvature remains the same \cite{daCosta81,daCosta82} because ${\rm V}$ is independent of $\tau$. In addition, matter-wave circuits in atom chips and optical realizations of waveguides are often associated with curves $\gamma$ on a plane, for which $\tau=0$. As a result, we focus on planar curves, given by the parametrization ${\bf r}(q_1)=(x(q_1),y(q_1))$. We shall return to the case of $\tau\neq0$ whenever the curve $\gamma$ exhibits multiple points in which the waveguide self-intersects. The CIP depends only on $\kappa^2$ and remains invariant under the mapping $\kappa\rightarrow {\rm sgn}(g(q_1))\kappa$ with an arbitrary real function $g(x)$, a symmetry which we shall exploit to engineer the guiding potential. Provided $\tau=0$, and $\kappa(q_1)\neq 0$ for all $q_1$ it is always possible to integrate the Frenet-Serret equations, and to find the natural representation of the curve in terms of the arc length \begin{eqnarray} \label{IFSE} \nonumber x(q_1)&=&x_0+\int_{q_1^0}^{q_1}\cos\left(\int_{q_1^0}^{\bar{s}}\kappa(s)ds\right)d\bar{s},\\ \\ \nonumber y(q_1)&=&y_0+\int_{q_1^0}^{q_1}\sin\left(\int_{q_1^0}^{\bar{s}}\kappa(s)ds\right)d\bar{s}. \end{eqnarray} In what follows we shall take $(x_0,y_0)=(0,0)$ without loss of generality, and extend Eq. (\ref{IFSE}) to $q_1<0$ to sample the full range of ${\rm V}$. \subsection{Supersymmetric partner waveguides} In the Witten model of supersymmetric quantum mechanics (SUSY QM) \cite{Witten81,CKS95}, a pair of SUSY partner Hamiltonians is considered with the factorization, \begin{eqnarray} H_- :={\cal A}^{\dag}{\cal A}\geq 0,\qquad H_+:={\cal A}\A^{\dag}\geq 0, \end{eqnarray} where the annihilation and creation operators are defined by ${\cal A}:=\frac{ip}{\sqrt{2m}}+\Phi(q_1)$ and ${\cal A}^{\dag}:=-\frac{ip}{\sqrt{2m}}+\Phi(q_1)$, in terms of the superpotential $\Phi(q_1)$. The SUSY partner Hamiltonians can be explicitly written as $H_{\pm}=-\frac{\hbar^2}{2m}\partial_{q_1}^2+V_\pm$, where $V_\pm(q_1):=\Phi^2(q_1)\pm\frac{\hbar}{\sqrt{2m}}\Phi'(q_1)$ are partner potentials. We consider the case in which the SUSY partner potentials are both induced by curvature, i.e. $V_\pm={\rm V}_\pm=-\frac{\hbar^2}{8m}\kappa_\pm^2$. It follows that SUSY partner Hamiltonians are associated with curves whose curvatures $\kappa_{\pm}$ are related by \begin{eqnarray} \label{SUSYkappa} \kappa_{+}^2(q_1)&=&\kappa_{-}^2(q_1)-\frac{8\sqrt{2m}}{\hbar}\Phi'(q_1),\\ &=&\kappa_{-}^2(q_1)-\frac{8m}{\hbar^2}[{\cal A},{\cal A}^{\dag}], \end{eqnarray} where we have used the relation between the derivative of the superpotential and the commutator of the creation and annihilation operators in the second line \cite{DKS88}. The relation between $\kappa_\pm$, can be further developed by making reference to the ground state $\psi_0$ of $H_-$ satisfying $H_-\psi_0=0$, in terms of which the superpotential reads $\Phi(q_1) =-\frac{\hbar}{\sqrt{2m}}\frac{\partial_{q_1}\psi_0}{\psi_0}$. Using this expression in (\ref{SUSYkappa}), it follows that \begin{eqnarray} \label{SUSYkappa2} \kappa_{+}^2(q_1)=\kappa_{-}^2(q_1)+8\bigg[\frac{\partial_{q_1}^2\psi_0}{\psi_0}-\left(\frac{\partial_{q_1}\psi_0}{\psi_0}\right)^2\bigg]. \end{eqnarray} According to the fundamental theorem of curves, the shape and length of a planar curve is completely determined by its (single-valued and continuous) curvature. It follows that the SUSY partner potentials ${\rm V}_\pm$ are associated with the family of curves $\{\gamma_\pm\}$ (with curvatures $\kappa_\pm$ satisfying $\kappa_\pm^2=-\frac{2m}{\hbar}\Phi^2\mp\frac{\sqrt{2m}}{\hbar}\Phi'$), that we shall refer to as SUSY partner curves. This set can be extended to include curves associated with a family of shape-invariant potentials \cite{Gendenshtein83}, as discussed in Methods, or using higher-order SUSY QM \cite{CKS95}. \begin{figure} \centering{\includegraphics[width=0.7\linewidth]{SUSYcurvesFig1.pdf}} \caption{ {\bf Supersymmetric reflectionless waveguides.} (a) Waveguide whose CIP is the reflectionless P\"oschl-Teller potential with $\nu=1$. Its SUSY partner curve is the straight line $\gamma_-$ with zero curvature. (b) The multiple point can be removed to obtain a simple waveguide using the curvature mapping $\kappa(q_1)\rightarrow {\rm sgn}(q_1)\kappa(q_1)$ under which the CIP remains invariant. (c) For higher curvature values, as in the waveguide associated with the $n=2$ Sukumar reflectionless potential shown here, integration of the Frenet-Serret equations leads to curves with several multiple points. (d) Such CIPs can be engineered in a non-planar waveguide, with non-zero torsion $\tau$. } \label{ReflectionlessCurves} \end{figure} \begin{figure*} \centering \centering{\includegraphics[width=0.95\linewidth]{SUSYcurvesFig2.pdf}} \caption{ {\bf Scattering dynamics in bent waveguides.} Sequence of snapshots of the time-evolution of the density profile of a wavepacket along a planar bent waveguide with the curvature (\ref{MPTkappa}) and $\nu=1/2$ (left), and $\nu=1$ (right), as in Fig.\,\ref{ReflectionlessCurves}(b). Generally, the wavepacket is split by the CIP into a transmitted and a reflected component. Despite the high degree of bending shown in the inset, whenever $\nu$ is an integer, the waveguide becomes reflectionless and exhibits unit transmission probability for all energies of the impinging matter-wave beam. The color coding varies from white to red as the probability density increases.The dimensions of each waveguide image are $908\sigma_0 \times 47\sigma_0$ and the time interval between successive images is $1920/\omega_{\perp}$. The initial wavepacket has FWHM = 235$\sigma_0$ and momentum $(1/32)m\omega_{\perp}\sigma_0$, and $\alpha = 1/8$. } \label{ReflectionlessScattering} \end{figure*} Why are waveguides along these curves of interest? The main physical feature of SUSY partner curves is that they exhibit the same scattering properties, a distinguishing feature directly inherited from ${\rm V}_\pm$ \cite{CKS95,propagators}. Let $\gamma_\pm$ be open waveguides with finite curvature as $q_1\rightarrow\pm\infty$, so that ${\rm V}_{\pm}(q_1\rightarrow\pm\infty)\rightarrow\Phi(q_1\rightarrow\pm\infty)^2=:\Phi_{\pm}^2$, and consider the scattering states of momentum $k$ and energy $E=\hbar^2k^2/(2m)$, with reflection and transmission amplitudes $R_{\pm}(k)$ and $T_{\pm}(k)$, respectively. It follows that $R_-(k)=\frac{\Phi_-+i\hbar k/\sqrt{2m}}{\Phi_- -i\hbar k/\sqrt{2m}}R_+(k)$ and $T_-(k)=\frac{\Phi_+-i\hbar k'/\sqrt{2m}}{\Phi_- -i\hbar k/\sqrt{2m}}T_+(k)$ where $k=[2m(E-\Phi_-^2)]^{1/2}/\hbar$ and $k'=[2m(E-\Phi_+^2)]^{1/2}/\hbar$, that is, the reflection as well as the transmission probabilities are the same for SUSY partner curves. Further, the Hamiltonians $H_\pm$ associated with SUSY partner curves are isospectral, except for the lowest energy level of $H_-$ with zero-energy, which is absent in the spectrum of $H_+$. \subsection{Design of reflectionless curves} CIPs are of attractive character and as a result can lead to quantum reflection \cite{PSU58a,PSU58b,Henkel96,FT04}. The dynamics of a guided matter-wave on a bent waveguide is generally affected by the curvature. We next illustrate the power of the SUSY partner waveguides in designing reflectionless curves. An obvious instance where the CIP vanishes is that of an infinite straight waveguide, with $\kappa_-=0$ and superpotential $\Phi=A\tanh\alpha q_1$ with $A>0$. This configuration is of relevance to guided atom lasers \cite{GAL1,GAL2}, and we wish to mimic it in bent waveguides. SUSY QM allows us to find SUSY partners which are reflectionless. In this case, $V_+(q_1)$ is given by the modified P\"oschl-Teller potential $V_{\rm curv}(q_1)=-\frac{\hbar^2}{2m}\frac{\nu(\nu+1)}{\cosh^2(\alpha q_1)}$ \cite{CKS95}, so that the curvature of the SUSY $\gamma_+$ curve reads \begin{eqnarray} \label{MPTkappa} \kappa_+(q_1)=2\alpha\sqrt{\nu(\nu+1)}{\rm sech}\alpha q_1, \end{eqnarray} where $\nu$ is a positive integer. Provided that the dimensional reduction is valid, the transmission probability for a waveguide with curvature (\ref{MPTkappa}) and arbitrary $\nu$, is given by $|T_+|^2(k)=\frac{\mu^2}{1+\mu^2}$ with $\mu=\frac{{\rm sinh}(\pi k/\alpha)}{\sin \pi \nu}$ and $k=\sqrt{2mE/\hbar}$. Such a waveguide becomes reflectionless for integer values of $\nu$. Different reflectionless waveguides are plotted in Fig. \ref{ReflectionlessCurves}, where it is shown that the number of multiple points increases with the magnitude of the curvature. A simple waveguide without junctions can then be engineered by exploiting the invariance of the CIP with respect to changes in the sign of the curvature, or by considering a nonzero torsion $\tau\neq0$, whose realization might be achieved using an extended version of the painted potential technique \cite{painters}. We emphasize that there are infinitely many instances of reflectionless waveguides. Let us illustrate this by adapting the algorithm developed by Sukumar \cite{Sukumar86} to CIPs. One can construct reflectionless waveguides supporting $n$ bound states. Let us fix the bound state energies to be $E_n=-\frac{\hbar^2}{2m}\eta_n^2$ with $\eta_n^2>\eta_{n-1}^2>\dots\eta_1^2$. The symmetric reflectioneless curves are described by the equation \begin{eqnarray} \kappa_n^2(q_1)=8\partial_{q_1}^2{\rm ln} \det D_n, \end{eqnarray} where $[D_n]_{ij}=\frac{1}{2}\eta_j^{i-1}[\exp(\eta_jq_1)+(-1)^{i+j}\exp(-\eta_jq_1)]$. For a single bound state, one obtains $\kappa^2=8\eta_1^2{\rm sech}\eta_1q_1$, closely related to the SUSY curves associated with (\ref{MPTkappa}). Figure \ref{ReflectionlessCurves} (lower panels) shows the reflectionless curves corresponding to a Sukumar potential supporting two bound states with $\eta_1=1$ and $\eta_2=3/2$, where multiple points in \ref{ReflectionlessCurves}(c) are avoided by a non-zero torsion ($\tau=20q_0^{-1}$) in \ref{ReflectionlessCurves}(d). The axis of the associated waveguide follows the curve $(x(s),y(s),\tau s)$ with (squared) curvature $\kappa^2(s)=-8m{\rm V}(s)/\hbar^2$, torsion $\tau$ and arc length $q_1=\sqrt{1+\tau^2}s$. At variance with (\ref{MPTkappa}), the relative angle between the asymptotes can be tuned by adjusting the value of $\eta_2$ relative to $\eta_1$, which will allow for the engineering of reflectionless bends through a range of desired angles. Further examples of reflectionless waveguides can be found by using the infinite family of reflectionless potentials discussed by Shabat \cite{Shabat92} and Spiridonov \cite{Spiridonov92}. The reflectionless character of the SUSY waveguides becomes apparent in the dynamics of guided matter waves. Figure \ref{ReflectionlessScattering} shows an elongated Gaussian beam being guided in a bent waveguide with curvature given by (\ref{MPTkappa}). $\gamma_+$ is asymptotically flat for $q_1\rightarrow\pm\infty$. For a general non-integer value of $\nu$, the traveling beam is substantially reflected off the bent region. For integer $\nu$ there exists a delocalized critical bound-state with zero energy and the waveguide becomes reflectionless for all scattering energies. However, the degree of bending increases with $\nu$. As a result, reflectionless waveguides provide a remarkable counterexample to the common expectation that the reflection probability increases with the degree of bending of the waveguide. In addition, the numerical simulations correspond to the propagation in a waveguide with finite transverse width, for which the explicit form of the curvature induced potential \cite{SRS77,EB89,Schwartz06} is more complex than that in Eq. (\ref{daCostapot}) used to design the reflectionless SUSY waveguide, and where excitations of the transverse waveguide modes are possible. The fact that despite the finite transverse width the waveguide remains reflectionless illustrates the robustness of its design against imperfections. We also note that the reflectivity of the P\"oschl-Teller potential changes only gradually as $\nu$ departs from an integer value. \begin{figure} [t] \centering{\includegraphics[width=0.7\linewidth]{SUSYcurvesFig3.pdf}} \caption{ {\bf Canceling out the curvature-induced potential.} Elliptical waveguide potentials of increasing eccentricity (top) and corresponding ground state densities (middle). Bottom row shows ground state densities when the CIP is compensated by modulating the depth of the trap. The dimensionless density profile $n(q_1)\sigma_0$ is scaled up by a factor $10^3$, the perimeter of the ellipse is $L=150\sigma_0$ and the plotted area is $80\sigma_0\times 50\sigma_0$. } \label{noGIP} \end{figure} \subsection{Canceling out the geometry-induced potential} A variety of experimental techniques to design matter-wave circuits, such as the painted potential technique, offers an alternative way to control the design of $\gamma$: the modulation of the potential depth of the waveguide. Consider two arbitrary isometric waveguides $\gamma$ and $\widetilde{\gamma}$ (both either open or closed, and without multiple points), with CIPs ${\rm V}_{\gamma}(q_1)$ and ${\rm V}_{\widetilde{\gamma}}(q_1)$, respectively. Under the consistency conditions (\ref{cc}), it is then possible to make $\widetilde{\gamma}$ isospectral to $\gamma$ by modulating the depth of the waveguide potential, i.e., by creating a potential barrier of the form $U(q_1)=-[{\rm V}_{\widetilde{\gamma}}(q_1)-{\rm V}_{\gamma}(q_1)]$. A natural case is that in which ${\rm V}_{\gamma}(q_1)$ either vanishes or is an irrelevant constant energy shift. $U(q_1)$ is then the potential required to flatten out the depth of the global potential of $\widetilde{\gamma}$. In addition, the acceleration of the guided matter-waves towards the region of high-curvature is prevented. To explore in detail this possibility, we consider an elliptical trap \cite{SRS77,SK01,Schwartz06}, associated with the curve ${\bf r}(u)=(a\cos u,b\sin u)$, with $a\geq b>0$, and circumference $L$. The CIP in an elliptical trap reads \begin{eqnarray} {\rm V}(u)=-\frac{\hbar^2}{8m}\frac{a^2b^2}{(b^2\cos^2u+a^2\sin^2u)^3}. \end{eqnarray} The eccentricity of an ellipse is defined by $\epsilon = [1-\left(b/a\right)^2]^{\frac 1 2}\in [0,1]$ and can be used to quantify the deformation from a circle (for which $a=b$, $\epsilon=0$). For a ring of radius $a=b$ ($\gamma$, with $\epsilon=0$), the curvature is $\kappa(q_1)=1/a$ and the CIP becomes constant, and the ground state density profile is uniform along the arc length $q_1$. For $\widetilde{\gamma}$ with $\epsilon>0$, the CIP comes into play and creates two attractive double wells, centered around the points with higher curvature $q_1=\{0,L/2\}$ ($b<a$) and with the minimum value $-\frac{\hbar^2}{8m}\frac{a^2}{b^4}$. The extent to which geometry-induced effects can be cancelled out by painting a barrier $U(q_1)=-{\rm V}_{\widetilde{\gamma}}(q_1)$ is illustrated in Fig. \ref{noGIP}. Such cancellation is effective as long as the consistency conditions for the dimensional reduction hold, which ceases to be the case as $\epsilon$ is increased while the transverse width $\sigma_0$ remains fixed. The ground state density profile is a fairly robust quantity, but we note that this compensation is efficient as well for dynamical processes involving all spectral properties of the waveguide. Consider the time evolution of the density profile of an initially localized wavepacket released in the elliptical trap, displayed in Fig. \ref{carpet}. \begin{figure} [t] \centering{ \includegraphics[angle=0,width=1.1\linewidth]{SUSYcurvesFig4.pdf} } \caption{ {\bf Curvature-induced suppression of temporal Talbot oscillations.} Time evolution of the density profile $n(q_1,t)=\int dq_{\perp}n(q_1,q_\perp,t)$ of an initially tightly-localized wavepacket released in a two-dimensional elliptical waveguide. (a) For a ring trap ($\epsilon=0$) $n(q_1,t)$ exhibits Talbot oscillations as a result of the quadratic dispersion relation (left). Two Talbot oscillations are displayed. (b) Whenever $\epsilon>0$, the CIP lifts the degeneracies in the spectrum and suppresses Talbot oscillations ($\epsilon=0.9$). (c) The CIP can be cancelled out by modulating the depth of the trap ($\epsilon=0.9$). $L=150\sigma_0$ in all cases and hence the revival time $\tau_R = m L^{2}/(\pi \hbar)$ is constant for different values of $\epsilon$. } \label{carpet} \end{figure} For a ring trap, where $\epsilon=0$, the evolution of the density profile $n(q_1,t)$ weaves a highly structured interference pattern with ``scars'' in the plane $(q_1,t)$, known as a ``quantum carpet'' \cite{Berry96}. Such quantum carpets exhibit a temporal analogue of the Talbot effect: in wave optics, the near-field diffraction pattern of a wave incident upon a periodic grating is characterized by a spatial periodicity \cite{Talbot1836,BMS01}. The quantum dynamics of an initially localised wavepacket which is released in a two-dimensional ring trap exhibits a periodic revival of the initial state, with period $\tau_R=mL^2/(\pi\hbar)$ (the temporal analogue of optical Talbot oscillations). The reconstruction of the density profile at $t=0$ can be traced back to the quadratic dispersion relation of the trap and the degeneracies it entails \cite{Schleich1,Schleich2}. This phenomenon has been experimentally observed in a variety of systems \cite{Chapman95,Mark11}. In a two-dimensional elliptical waveguide, the spectrum is modified and the dispersion relation ceases to be quadratic. For $\epsilon>0$, the geometry-induced potential lifts the degeneracy in the spectrum, leading to the suppression of Talbot oscillations. Nonetheless, the dynamics corresponding to a ring trap can be effectively recovered in an ellipcal trap with $\epsilon>0$ after compensating the depth of the waveguide potential. The reapperance of Talbot oscillations in compensated elliptical waveguides signals the isospectral properties with respect to the ring trap, illustrating the suppression of curvature-induced effects. \section{Discussion} The dynamics of matter waves in bent waveguides is severely distorted by the appearance of an attractive curvature-induced quantum potential. As matter wave circuits shrink in size and atomic velocities must be reduced to maintain single mode propagation, curvature-induced potentials impose practical limitations on minimum velocities, and methods to reduce their effects are needed. Using methods of supersymmetric quantum mechanics, we have introduced a framework to design sets of bent waveguides which share the same scattering properties. As a relevant example, an infinite family of reflectionless waveguides with a controllable number of bound states has been presented. As a complementary approach, we have discussed the possibility of tailoring curvature-induced effects by controlling the depth of the waveguide trapping potential. Our discussion has been focused on the effects of curvature on guided matter-waves which are experimentally realizable by a variety of techniques including atom chip technology and the painted potential technique based on a time-averaged optical dipole potential \cite{painters}. Our results are however directly applicable to other systems such as optical waveguides and photonic lattices \cite{GJ92}, in which curvature-induced potentials \cite{CIP10,Schleich13}, reflectionless potentials \cite{RP10}, and concepts of supersymmetric quantum mechanics \cite{SOW,SUSY14} have already been implemented in the laboratory, and that provide a natural alternative platform to experimentally explore the interplay between geometry and supersymmetry in quantum mechanics. \section{Methods} Let us consider the case in which the superpotential depends on a collective set of parameters $a_0$, $\Phi=\Phi(q_1;a_0)$. The partner potentials ${\rm V}_\pm$ are shape-invariant if they are related by ${\rm V}_+(q_1;a_0)={\rm V}_-(q_1;a_1)+R(a_1)$ where the residual term $R(a_1)$ is independent of $q_1$ and $a_1=f(a_0)$ is a new set of parameters obatined form $a_0$ by the action of the function $f(\cdot)$. By iteration, one can construct the series of Hamiltonians $\{H_k|k=0,1,\dots\}$ with $H_0=H_-$ and $H_1=H_+$, such that $H_s=H_0+\sum_{k=1}^sR(a_k)$, with $a_k=f^k(a_0)$, i.e., obtained by the $f$ function iterated $k$ times. It follows that the squared curvatures of the SUSY partner curves $\{\gamma_{s-1},\gamma_{s}\}$ are related by a constant shift \begin{eqnarray} \label{SIkappa} \kappa_{s}^2(q_1)&=&\kappa_{s-1}^2(q_1)-\frac{8m}{\hbar^2}R(a_s) =\kappa_{0}^2(q_1)-\frac{8m}{\hbar^2}\sum_{k=1}^sR(a_k).\nonumber\\ \end{eqnarray}
\section*{MAGNA: Maximizing Accuracy in Global Network Alignment} \textbf{Vikram Saraph and Tijana Milenkovi\'{c}$^*$}\\ \textbf{Department of Computer Science and Engineering}\\ \textbf{University of Notre Dame, Notre Dame, IN 46556, USA}\\ $^*$Corresponding Author (E-mail: <EMAIL>) \section*{Summary} Biological network alignment aims to identify similar regions between networks of different species. Existing methods compute node ``similarities'' to rapidly identify from possible alignments the ``high-scoring'' alignments with respect to the overall node similarity. However, the accuracy of the alignments is then evaluated with some other measure that is different than the node similarity used to construct the alignments. Typically, one measures the amount of conserved edges. Thus, the existing methods align similar \emph{nodes} between networks \emph{hoping} to conserve many \emph{edges} (\emph{after} the alignment is constructed!). Instead, we introduce MAGNA to directly ``optimize'' edge conservation \emph{while} the alignment is constructed. MAGNA uses a genetic algorithm and our novel function for ``crossover'' of two ``parent'' alignments into a superior ``child'' alignment to simulate a ``population'' of alignments that ``evolves'' over time; the ``fittest'' alignments survive and proceed to the next ``generation'', until the alignment accuracy cannot be optimized further. While we optimize our new and superior measure of the amount of conserved edges, MAGNA can optimize \emph{any} alignment accuracy measure. In systematic evaluations against existing state-of-the-art methods (IsoRank, MI-GRAAL, and GHOST), MAGNA improves alignment accuracy of \emph{all} methods. \bigskip \medskip \twocolumn \section{Introduction} \subsection{Motivation and background}\label{sect:motivation} Genomic sequence alignment has led to breakthroughs in our understanding of how cells work. It identifies regions of similarity between sequences of individual genes that are a likely consequence of evolutionary relationships between the sequences. However, genes, i.e., their protein products, do not act alone but instead interact with each other to carry out cellular processes. And this is exactly what \emph{protein-protein interaction (PPI) networks} model. (While we focus on PPI networks, our ideas are applicable to \emph{any} network type.) Then, network alignment (NA) can be used to find regions of similarities between PPI networks of different species that are a likely consequence of evolutionary relationships between the networks. Unlike sequence alignment that ignores genes' interconnectivities, NA allows for studying complex cellular events that are a consequence of the \emph{collective} behavior of the genes' protein products. As such, NA is promising to \emph{further} our biological understanding \cite{Sharan2006}. As recent biotechnological advances continue to yield large amounts of PPI data \cite{BIOGRID}, alignment of PPI networks of different species continues to gain importance \cite{Sharan2006}. This is because NA could guide the transfer of biological knowledge across species between conserved (aligned) network regions \cite{Sharan2006}. This is important, since many proteins remain functionally uncharacterized even for well studied species \cite{Sharan2007}. Traditionally, the across-species transfer of biological knowledge has relied on sequence alignment. However, since PPI networks and sequences can capture complementary functional slices of the cell, implying that PPI networks can uncover function that cannot be uncovered from sequences by current methods, restricting alignment to sequences may limit the knowledge transfer \cite{Memisevic10b}. Unfortunately, the mathematics of complexity theory dictates that exact network (or graph) comparison is computationally intractable. The underlying problem is that of subgraph isomorphism, which asks whether one graph (the source) appears as an exact subgraph of another graph (the target). Answering this is NP-complete \cite{Cook1971}. Furthermore, simply answering the subgraph isomorphism problem is not enough when comparing PPI networks, since one PPI network is rarely an exact subnetwork of another due to biological variation \cite{GRAAL}. It is much more desirable to answer how similar two networks are and in what regions they share similarity. NA can be used for this purpose. NA is a less restrictive problem than that of subgraph isomorphism, as it seeks to ``fit'' the source into the target in the ``best possible way'' even if the source is not an exact subgraph of the target. An alignment is a mapping between nodes of the source and nodes of the target that is expected to conserve as much structure (or topology) as possible between the two networks. (Note that methods exist that can align more than two networks \cite{IsoRankN,Flannick2008}, but we focus on \emph{pairwise} NA.) Since NA is computationally hard, heuristic methods must be sought. NA can be \emph{local} (LNA) or \emph{global} (GNA). Initial solutions for NA have aimed to match local network regions \cite{PathBlast,Sharan2005,Flannick2006,Mawish,Berg04,Liang2006a,Berg2006}. That is, in LNA, subnetworks, rather than the entire networks, are aligned. However, aligned regions can overlap, leading to ``ambiguous'' many-to-many node mappings. Thus, GNA solutions have been proposed \cite{Singh2007,Flannick2008,Singh2008,GraphM,IsoRankN,GRAAL,HGRAAL,MIGRAAL,GHOST,Narayanan2011,NATALIE,NETAL}. In contrast to LNA, GNA compares entire networks, typically by aligning every node in the source to exactly one unique node in the target. We focus on \emph{GNA}, but our ideas are also applicable to LNA. Traditionally, GNA has relied on biological information \emph{external} to network topology, e.g., sequence similarity \cite{Sharan2006}. To extract the most from each source of biological information, it would be good to know how much of new biological knowledge can be uncovered \emph{solely} from network topology \emph{before} integrating it with other sources of biological information \cite{GRAAL,HGRAAL,MIGRAAL,GHOST,NETAL}. Only after methods for topological GNA are developed that result in alignments of good topological \emph{and} biological quality, it would be beneficial to integrate them with other biological data sources to further improve the quality. Thus, we focus on \emph{topological} GNA, but additional biological data can easily be added. Existing GNA methods, of which the more prominent ones (and which we consider in our study) are outlined below, typically use a two-step approach: (1) score the ``similarity'' of pairs of nodes from different networks, and (2) feed these scores into an alignment strategy to identify ``high-scoring'' alignments from all possible alignments. \emph{IsoRank} \cite{Singh2007} scores nodes from two networks by a PageRank-based spectral graph theoretic principle: two nodes are a good match if their neighbors are good matches. After these topological scores are computed, biological scores can be added to get final node scores. An alignment is then constructed by greedily matching the high-scoring node pairs. IsoRank has evolved into \emph{IsoRankN} to allow for \emph{multiple} GNA \cite{IsoRankN} and \emph{many-to-many} node mapping, but this is out of the scope of our study. The \emph{GRAAL} family of algorithms \cite{GRAAL,HGRAAL,CGRAAL,MIGRAAL}, developed in parallel with the IsoRank family, use graphlet (or small induced subgraph) counts to compute mathematically rigorous topological node similarity scores \cite{Milenkovic2008,MMGP_Roy_Soc_09,Solava2012}. Intuitively, two nodes are a good match if their \emph{extended} network neighborhoods are ``topologically similar'' with respect to the graphlet counts. Also, \emph{MI-GRAAL} \cite{MIGRAAL}, the latest of the family members, can automatically add other (biological) node similarity scores into final scores. It is the alignment strategies of the GRAAL family members that are different. MI-GRAAL combines alignment strategies of the other members, thus outperforming each of them \cite{MIGRAAL}. More recent \emph{GHOST} \cite{GHOST} uses ``spectral signatures'' to score node pairs topologically while also allowing for inclusion of biological node scores. Similar to MI-GRAAL, GHOST's alignment strategy is also seed-and-extend, except that GHOST solves a quadratic assignment problem, while MI-GRAAL solves a linear assignment problem. \subsection{Our contribution} Recall that the existing GNA methods construct alignments by scoring all node pairs with respect to the nodes' similarities and by rapidly identifying ``high-scoring'' alignments from all possible alignments. Here, ``high-scoring'' alignments are typically those that ``maximize'' (greedily or optimally) the node similarity score totaled over all mapped nodes \cite{Singh2007,GRAAL,HGRAAL,MIGRAAL,GHOST}. However, the accuracy (or quality) of the alignments is then evaluated with respect to some other measure of an inexact fit of two networks, which is different than the node scoring function that is used to construct the alignments in the first place. Typically, one measures the amount of conserved edges (see below) \cite{MIGRAAL,GHOST}. (One also evaluates the alignments biologically with respect to functional knowledge.) Thus, the existing methods align ``similar'' \emph{nodes} between networks with the goal (or \emph{hope}!) of conserving as many \emph{edges} as possible under the alignment (\emph{after} the alignment is constructed!). Instead, we introduce \emph{MAGNA}, a new framework for directly ``\textbf{m}aximizing'' (or ``optimizing'') \textbf{a}ccuracy in \textbf{GNA} with respect to the amount of conserved edges \emph{while} the alignment is being constructed, which is our first contribution. Optimizing the amount of conserved edges would require finding a global optimum over the search space consisting of all possible node mappings. Due to the large size of the space, exhaustive search is computationally intractable. But, approximate techniques exist with solutions very close to optimal, such as genetic algorithms \cite{Cross2000}. Hence, we adapt the idea of genetic algorithms to the problem of GNA to develop MAGNA as a conceptually novel GNA framework. This is our second contribution, since to our knowledge, genetic algorithms have not been used for PPI GNA thus far. MAGNA simulates a ``population'' of alignments that ``evolves'' over time (the initial population can consist of random alignments or of alignments produced by the existing methods). Then, the ``fittest'' candidates (those that conserve the most edges) survive and proceed to the next generation. This is repeated until the algorithm converges, i.e., until the amount of conserved edges cannot be optimized further. Much of what defines any genetic algorithm is the crossover function, which ``combines'' two candidates (i.e., alignments) into a new one. And since genetic algorithms have not been used for GNA thus far, we had to devise a novel (and to our knowledge, \emph{the first ever}) function for crossover of two parent alignments into a child alignment that reflects (ideally the best of) each parent. The alignment crossover function is the third and a major contribution of our study, because it allows MAGNA not only to combine alignments produced by \emph{any} existing method to \emph{improve} them but also to produce \emph{its own new} superior alignments. It is not obvious how to measure the quality of an alignment \cite{HGRAAL}, i.e., which measure to optimize as the ``fitness'' function within the genetic algorithm. Clearly, a good alignment should maximize the amount of conserved edges. Different measures have been proposed to quantify this, all of which are heuristics and thus correctly reflect the actual alignment quality in some cases but fail to do so in other cases. Thus, as our fourth contribution, we introduce a new and superior alignment quality measure that takes the best from each existing measure. While we optimize with MAGNA this new measure as well as the existing measures of the amount of conserved edges, importantly, MAGNA can optimize \emph{any} measure of alignment quality, topological \emph{or} biological, which is another of our contributions. We evaluate MAGNA against IsoRank, MI-GRAAL, and GHOST by aligning a high-confidence yeast PPI network with its noisy counterparts, where the true node mapping is known. This popular evaluation test \cite{GRAAL,MIGRAAL,GHOST} allows for a systematic method comparison. MAGNA improves alignment quality of \emph{all} of the existing methods. \section{Methods} \subsection{Alignment crossover function}\label{sect:crossover} In this section, we provide the mathematical rigor necessary to define our novel ``crossover function'', which is at the heart of MAGNA. (The description of MAGNA is given in Section \ref{sect:MAGNA}.) The crossover function should take two ``parent'' alignments and produce a ``child'' alignment that is intended to reflect (ideally the best of) both parents. Let $G_1(V_1, E_1)$ and $G_2(V_2, E_2)$ be two networks with $V_i$ and $E_i$ as the sets of nodes and edges, respectively. Let $m = |V_1|$ and $n = |V_2|$. Without loss of generality, suppose $|V_1| \le |V_2|$. An \emph{alignment} of $G_1$ to $G_2$ is a total injective function $f : V_1 \rightarrow V_2$. That is, every element of $V_1$ is matched uniquely with an element of $V_2$. If $|V_1| = |V_2|$, when $f$ is an injective function, then in fact $f$ is a bijection. Let $V_1 = \{x_1, \ldots, x_m\}$ and $V_2 = \{y_1, \ldots, y_n\}$. Let $[n] = \{1, \ldots, n\}$ be the set of natural numbers from 1 to $n$. A \emph{permutation} is a bijection $\sigma : [n] \rightarrow [n]$. Then, with the assumption that $m = n$, and given this fixed number labeling of nodes as above, we can represent any alignment $f$ with a corresponding permutation $\sigma$ that maps node labels to node labels. Even though it is rare that $|V_1| = |V_2|$, we can easily force this condition to be true by adding dummy, zero-degree nodes $z_i$ to $V_1$, as follows: $\bar{V_1} = V_1 \cup \{z_{m+1}, \ldots, z_n\}$. Thus, from now on, we will simply assume that $|V_1| = |V_2|$, without explicitly referring to $\bar{V_1}$. Therefore, any alignment can be represented as a permutation, and for the remainder of this section, we use ``permutation'' and ``alignment'' interchangeably. This representation is critical to our crossover function. Let $S_n$ denote the set of all permutations. Notice that $|S_n| = n!$, which is large. In theory, to find an alignment of maximum quality (with respect to a given criterion), we could ``simply'' enumerate all permutations and evaluate the quality of each one. However, this is impractical due to the large size of $S_n$, so we require a clever search heuristic. We design such a heuristic as follows. First, we create a graph with $S_n$ as the set of nodes in which two permutations (alignments) are connected by an edge if the alignments are ``adjacent'' (see below). Second, since intuitively the alignment quality is continuous in alignment ``adjacency'' (in the sense that two ``adjacent'' alignments should be of similar quality, or in other words, a small perturbation of an alignment should not greatly affect its quality), we exploit the topology of this graph to define a function for crossover of two alignments. Namely, we define the child alignment as the alignment that is ``in the middle'' between two given parent alignments in this graph. Formal details are as follows. Given two permutations $\sigma$ and $\tau$, we define what it means for $\sigma$ and $\tau$ to be adjacent. A \emph{transposition} of a permutation is a new permutation that fixes every element of the original permutation, except two elements, which are swapped. Then, two permutations are \emph{adjacent} if they differ by a transposition; that is, $\sigma$ and $\tau$ are adjacent if there is a transposition $\rho$ such that $\sigma = \rho \circ \tau$. We create graph $\Gamma_n$ with the set of nodes $S_n$ and the set of edges $E_n$, where an edge between $\sigma$ and $\tau$ is in $E_n$ if and only if $\sigma$ and $\tau$ are adjacent. Then, we define $\sigma \otimes \tau$, the \emph{crossover} of \emph{any} two permutations $\sigma$ and $\tau$ from $S_n$, as a permutation which is the midpoint on a shortest path from $\sigma$ to $\tau$ in $\Gamma_n$. This definition captures what we desire from a crossover function. More precisely, it can be shown that for randomly selected permutations $\sigma$ and $\tau$, $|\sigma \cap (\sigma \otimes \tau)| / n \rightarrow 1/2$ and $|\tau \cap (\sigma \otimes \tau)| / n \rightarrow 1/2$ as $n \rightarrow \infty$. That is, $\sigma \otimes \tau$ is expected to share approximately half of its aligned pairs with $\sigma$, and likewise with $\tau$. A proof of the above statement relies on the fact that the expected number of cycles in a permutation is $\Theta(\log(n))$. We leave out further discussion on this, as it would require more basics of abstract algebra, which is beyond the scope of this paper; see \cite{Knuth1997,Dummit2004} for details. \subsection{MAGNA: genetic algorithm-based GNA framework}\label{sect:MAGNA} A genetic algorithm mimics the evolutionary process, guided by the ``survival of the fittest'' principle \cite{Back1996}. It begins with an initial ``population'' of a given number of ``members''. Members of a population ``crossover'' with one another to produce new members. The ``child'' resulting from a crossover should resemble both of its ``parents''. Crossing over different pairs of members at a given generation yields new members, which comprise the new ``generation'' of members. The probability of a member being given a chance to crossover with another member is determined by its ``fitness,'' so that fitter members are more likely to crossover. To prevent the size of the population to grow without bound, the size is kept constant across all generations, with only the fittest members surviving from one generation to the next. To ensure that the maximum fitness of the population is nondecreasing, with each generation, a designated ``elite'' class of the fittest members is automatically passed to the next generation. As the algorithm progresses, newer generations are produced, with fitness (hopefully) increasing, until a stopping criterion is reached. To specify a genetic algorithm, we need to specify all of the above parameters. In MAGNA, members of a population are alignments. We use different types of initial populations: 1) all random alignments, 2) random alignments mixed with an IsoRank's alignment, 3) random alignments mixed with a MI-GRAAL's alignment, and 4) random alignments mixed with a GHOST's alignment. Since we focus on \emph{topological} network alignments (Section \ref{sect:motivation}), we produce all alignments by using only topological information in the existing methods' node scoring function. For each type of initial population, we test populations of different sizes: 200, 500, 1,000, 2,500, 5,000, 10,000, and 15,000. (It is because the population sizes are large that we cannot form an initial population consisting only of alignments produced by an existing method, due to large computational complexity of the existing methods. Instead, we use in the initial population an existing alignment and fill the remaining part of the population with random alignments.) The mathematical machinery from Section \ref{sect:crossover} gives us a suitable crossover function for producing a child alignment that resembles both of its parent alignments. Our fitness function is the measure of alignment quality we choose to optimize; in our case, it is edge correctness (EC), induced conserved structure (ICS), or symmetric substructure score (S$^3$) (Section \ref{sect:S3}), but it can be \emph{any} measure. In every generation, we keep the best half of the population from the previous generation, and we fill the remaining half of the population with alignments produced by crossovers. We select pairs of alignments to be crossed as follows. At a given generation of population size $p$, we have ${p \choose 2}$ crossover possibilities. This is too large a number to consider all of them. Thus, to select crossover pairs, we use \emph{roulette wheel selection}, which is a commonly adopted selection strategy for genetic algorithms \cite{Back1996}. Roulette wheel selection chooses members with probability in linear proportion to the members' fitness. We let MAGNA run for many generations. We vary the number of generations from 0 to 2,000 in increments of 200. The fittest alignment from the last generation is reported as the \emph{final} alignment. We describe MAGNA in the pseudocode (Supplementary Algorithm S1). We provide MAGNA's implementation upon request. Our implementation of the alignment crossover function takes $O(|V|)$ time. MAGNA's bottleneck, though, tends to be the computation of alignment quality $F$. If the measure of $F$ is EC, ICS, or S$^3$, then for a given alignment, it takes $O(|E|\log(|E|))$ time to compute $F$. Finally, sorting each generation of size $p$ takes $O(p\log(p))$ time, though this is typically negligible compared to the computation of $F$. If MAGNA is run for $N$ generations, this brings the overall time complexity of MAGNA to $O(N(p|V| + p|E| \log(|E|) + p \log(p)))$. Note that most of MAGNA is embarrassingly parallelizable, which can lead to a very high degree of speedup. \subsection{New alignment quality measure}\label{sect:S3} To motivate our new measure of alignment quality, the \emph{symmetric substructure score} (S$^3$), we first present drawbacks of existing \emph{edge correctness} (EC) and \emph{induced conserved structure} (ICS) measures. Let $G_1(V_1, E_1)$ and $G_2(V_2, E_2)$ be two networks, and let $f : V_1 \rightarrow V_2$ be an alignment between them. If $X \subseteq V_2$, let $G_2[X]$ be the induced subnetwork of $G_2$ with node set $X$. Also, if $H$ is a subnetwork of $G_2$, let $E(H)$ be its edge set. Let $f(E_1) = \{(f(u), f(v)) \in E_2 : (u, v) \in E_1\}$, and let $f(V_1) = \{ f(v) \in V_2 : v \in V_1\}$. EC of $f$ is the ratio of the number of edges conserved by $f$ to the number of edges in the source network: $\text{EC}(f) = \dfrac{|f(E_1)|}{|E_1|}$ \cite{GRAAL}. Because EC is defined with respect to the source but not the target network, it fails to penalize alignments mapping sparser network regions to denser ones (Fig. \ref{fig:S3}). \begin{figure}[H] \centering \includegraphics[scale=0.6]{Figure1} \caption{Illustration of our new S$^3$ measure and its difference with EC and ICS. The illustrated alignment between nodes in network $G$ and nodes in network $H$ has an EC of $4/5=0.8$, an ICS of $4/5=0.8$, but an S$^3$ of $4/6=0.67$. EC rewards for aligning four edges in $G$ to four edges in $H$ and penalizes for misaligning an edge in $G$ to a non-edge in $H$, but it fails to penalize for misaligning a non-edge in $G$ to an edge in $H$. Similarly, ICS rewards for aligning four edges in $G$ to four edges in $H$ and penalizes for misaligning an edge in $H$ (between the aligned nodes) to a non-edge in $G$, but it fails to penalize for misaligning an edge in $G$ to a non-edge in $H$. Like EC and ICS, S$^3$ also rewards for aligning four edges in $G$ to four edges in $H$, but unlike EC or ICS, S$^3$ penalizes for misaligning \emph{both} an edge in $G$ to a non-edge in $H$ and a non-edge in $G$ to an edge in $H$.}\label{fig:S3} \end{figure} ICS of $f$ is the ratio of the number of edges conserved by $f$ to the number of edges in the subnetwork of $G_2$ induced on the nodes in $G_2$ that are aligned to the nodes in $G_1$: $\text{ICS}(f) = \dfrac{|f(E_1)|}{|E(G_2[f(V_1)])|}$ \cite{GHOST}. Because ICS is defined with respect to the target but not the source network, it fails to penalize alignments mapping denser network regions to sparser ones (Fig. \ref{fig:S3}). Therefore, we define S$^3$ with respect to \emph{both} the source network and the target network: $\text{S}^3(f) = \dfrac{|f(E_1)|}{|E_1| + |E(G_2[f(V_1)])| - |f(E_1)|}$. The difference between EC, ICS, and S$^3$ is the denominator. Intuitively, if $G_1$ and $G_2[f(V_1)]$ are overlaid into a composite graph, then the denominator of S$^3$ is the number of unique edges in this composite graph. Thus, S$^3$ of an alignment is 100\% if and only if $f$ is a \emph{perfect} embedding. As such, S$^3$ penalizes \emph{both} alignments that map denser network regions to sparser ones and alignments that map sparser network regions to denser ones (Fig. \ref{fig:S3}). \section{Results and discussion}\label{sect:results_discussion} \subsection{Validating MAGNA on networks with \emph{known} node mapping}\label{sect:hc} \subsubsection{Data description} We aim to validate MAGNA by analyzing the largest connected component of the high-confidence yeast \emph{S. cerevisiae} PPI network \cite{Collins07} with 1,004 proteins and 8,323 PPIs. We align this network with the same network augmented with lower-confidence PPIs from the same study \cite{Collins07}. We analyze different noise levels, by adding 0\%, 5\%, 10\%, 15\%, 20\%, and 25\% of lower-confidence PPIs; we add higher-scoring lower-confidence PPIs first. Since the networks being aligned are defined on the same set of nodes and differ only in the number of edges, we know the correct node mapping. An additional advantage of aligning these networks is that the original is an exact subgraph of each noisy network. \subsubsection{MAGNA parameters} MAGNA requires several parameters: the type of initial population, population size, maximum number of generations (i.e., iterations of the genetic algorithm), and optimization function (i.e., alignment quality measure) (Sections \ref{sect:MAGNA} and \ref{sect:S3}). We evaluate MAGNA comprehensively and systematically, by varying values of each parameter. We use four different population types: random, IsoRank, MI-GRAAL, and GHOST. The random population aims to produce a high-quality alignment from scratch (by relying only on our new alignment crossover function), while the other three population types try to improve upon the existing methods. We test seven population sizes from 200 to 15,000. We vary the maximum number of generations up to 2000, in increments of 200. We optimize three alignment quality measures: EC, ICS, and S$^3$. See Sections \ref{sect:MAGNA} and \ref{sect:S3} for details. Each combination of initial population type, population size, maximum number of generations, and optimization function results in one final (best) alignment. This comprehensive testing has resulted in the total of 5,544 \emph{final} alignments. \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of the initial population type.} Since we aim to compare MAGNA against IsoRank, MI-GRAAL, and GHOST (and also random alignments), we continue by considering all four initial population types and we discuss their effect on the alignment quality in more detail below. \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of population size.} We find that, in general, larger population size is always preferred, independent of the initial population type, maximum number of generations, and optimization measure (Supplementary Section S1 and Supplementary Fig. S1). Henceforth, we continue with the largest population size of 15,000. \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of the maximum number of generations.} We find that, in general, the larger the population size, the larger number of generations is preferred, which is $\sim$2,000 for random initial population, independent on the optimization measure, and $\sim$400-1,200 for IsoRank, MI-GRAAL, or GHOST initial population, depending on the optimization measure (Supplementary Section S1 and Supplementary Fig. S1). In general, GHOST initial population ``converges'' faster than MI-GRAAL and IsoRank populations. Because of the design of MAGNA, the alignment quality never drops from one generation to the following one. Thus, even with IsoRank, MI-GRAAL, and GHOST populations, the results are never worse at the 2,000$^{th}$ generation compared to the 400-1,200$^{th}$ generation. Thus, henceforth, we continue with the maximum number of generations of 2,000, since this helps for at least one population type without harming others. However, it is encouraging that some methods can converge very fast, indicating that MAGNA can produce high-quality alignments in short time. \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of the optimization measure.} Since we aim to compare our new S$^3$ measure with existing EC and ICS measures, we continue by considering all three and we discuss their effect on the alignment quality in more detail below. \subsubsection{MAGNA evaluation and comparison with existing methods} For each of the six noise levels, four initial population types (each of size 15,000), and three optimization measures, we obtain with MAGNA one final alignment, i.e., the best alignment (with respect to the given optimization measure) at the 2,000$^{th}$ generation. In addition, we study the original alignments produced by the existing methods. Then, we compare these original alignments to those produced by MAGNA to see whether MAGNA improves the alignment quality of the existing alignments. Note that independent on which of the three alignment quality measures (EC, ICS, or S$^3$) we optimize, the question remains on how to best evaluate the correctness of the resulting final alignment. Certainly, we could use \emph{any} of the three alignment quality measures for this purpose. However, since the true node mapping is known when aligning the high-confidence yeast PPI network to its noisy counterparts, we can actually evaluate each method more fairly by counting the number of correctly aligned node pairs (or ``node correctness''). \begin{figure}[H] \centering \includegraphics[scale=0.6]{Figure2} \caption{ Correctness of alignments produced on noisy yeast networks (for noise levels in 0\%-25\% range), with respect to the number of correctly aligned node pairs. Panel \textbf{(a)} shows results for alignments produced by four existing algorithms (Random, IsoRank, MI-GRAAL, and GHOST) as well as by running MAGNA on populations containing the alignments produced by the existing algorithms. We use four different populations, corresponding to the four existing algorithms. For each population, we show results for an original alignment produced by the existing algorithm (O), as well as for MAGNA's alignments produced when optimizing each of the following: EC, ICS, and $S^3$. All results are for population size of 15,000 and for 2,000 generations. Correctness of the alignments with respect to additional criteria, including EC, ICS, and S$^3$, are shown in Supplementary Fig. S2 and S3. Panel \textbf{(b)} shows, for each noise level, comparison of results from panel (a) between MAGNA's best alignment (over all initial population types and optimization measures) and the original alignments of the existing methods. (In most cases, original random alignments have scores close to 0 and are thus not visible.) }\label{fig:hc} \end{figure} When we do this, we find that MAGNA improves \emph{all} of the original alignments (i.e., all of the existing methods), across all levels of noise, and for each of the three optimization measures (Fig. \ref{fig:hc}). If we compute the ``improvement'' of MAGNA over an existing method as the ratio of MAGNA's node correctness to the existing method's node correctness, then MAGNA's improvement is up to 2,588\% upon IsoRank, up to 256\% upon MI-GRAAL, and up to 118\% upon GHOST, depending on the noise level and optimization measure. In general, the higher the noise level, the larger our improvements upon the existing methods (Fig. \ref{fig:hc}). \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of the initial population type.} GHOST's \emph{original} alignments are overall slightly superior or comparable to MI-GRAAL's original alignments, depending on the noise level and the optimization measure, both are superior to IsoRank's original alignments, and all three are superior to random original alignments (Fig. \ref{fig:hc}). These results are consistent to those in the existing literature \cite{GRAAL,MIGRAAL,GHOST}. Thus, one might expect that \emph{MAGNA's improved} alignments of GHOST would be of better quality than MAGNA's improved alignments of MI-GRAAL, that both would be of higher quality than MAGNA's improved alignments of IsoRank, and that all three would be of higher quality than MAGNA's improved alignments of random alignments. However, interestingly, we find that this is not always the case (Fig. \ref{fig:hc}). Actually, in many cases, there are surprising effects of the choice of the initial population type. For example, our improved alignments of MI-GRAAL are sometimes better than our improved alignments of GHOST. Or, even more interestingly, for larger noise levels, it is the \emph{random} population that results in the best alignments; that is, we improve more when starting from completely random alignments than we do when starting from the original alignments of IsoRank, MI-GRAAL, or GHOST (Fig. \ref{fig:hc}). More precisely, when we measure, for each of the six noise levels, which initial population type results in the final alignment with the highest node correctness score over all four population types, we find that GHOST's initial population is the best for three out of six noise levels (5\%, 10\%, and 15\%), random initial population is the best for two noise levels (20\% and 25\%), and MI-GRAAL's initial population is the best for the remaining noise level (0\%). The above results suggest that MAGNA is not only capable of improving alignments generated by the existing methods, but it is also capable of generating from completely random alignments its own new alignments that are superior, especially for the higher noise levels. Interesting implications are as follows. First, because current PPI networks are likely even noisier than those used in this section \cite{Mering02,Venkatesan2009}, our results suggest that one might be able to improve upon the current best PPI network alignments (of different species, when the actual node mapping is unknown) simply by using MAGNA on completely random alignments of the PPI networks. Second, recall that random initial population converges the slowest of all populations, if at all. And recall that we stop MAGNA after 2,000 iterations, as all initial population types but random one converge even before that. Because of this, and because random population is superior for larger noise levels, it is possible that for such noise levels, the alignment quality could be improved even further by running MAGNA longer, as dictated by the available computing resources. \vspace{0.1cm} \hspace{-0.35cm}\textbf{The effect of the optimization measure.} No single optimization measure (out of EC, ICS, and S$^3$) is always superior with respect to the node correctness as the alignment quality measure; the results depend on the choice of MAGNA's parameters. Over all noise levels, random initial population prefers (in the sense that it results in the highest node correctness for) EC and S$^3$ equally, IsoRank initial population prefers ICS, MI-GRAAL's initial population prefers S$^3$, and GHOST initial population prefers EC (Fig. \ref{fig:hc}). Hence, S$^3$, as well as EC, seem to be preferred overall in this context. Over all population types, four of the six noise levels (5\%, 10\%, 15\%, and 25\%) prefer EC, one noise level (0\%) prefers S$^3$, and the remaining noise level (20\%) prefers ICS. Hence, EC seems to be preferred overall in this context. We even further study the effect of the three optimization measures by computing Pearson correlation between the ``node correctness'' on one hand and EC, ICS, or S$^3$ on the other, across all alignments from Fig. \ref{fig:hc}. A higher and more statistically significant correlation would indicate that the given optimization measure is capable of uncovering more correct alignments. The node correctness correlates the best and the most significantly with our new S$^3$ measure, suggesting its superiority over the existing measures (Table \ref{tab:correlation}). \vspace{0.2cm} \begin{table}[htbp] \begin{center} \begin{tabular}{ccc} \hline measure & correlation & $p$-value\\ \hline EC & 0.7538 & $3.9 \times 10^{-19}$ \\ ICS & 0.8339 & $2.7 \times 10^{-26}$ \\ S$^3$ & 0.8980 & $1.4 \times 10^{-35}$ \\ \hline \end{tabular} \end{center} \vspace{-0.2cm} \caption{Correlation between the node correctness and each of EC, ICS, and S$^3$.}\label{tab:correlation} \end{table} \section{Concluding remarks} We present a conceptually novel framework for ``optimizing'' pairwise global network alignment with respect to any alignment quality measure, which outperforms the existing state-of-the-art methods. Given the tremendous amounts of biological network data that are being produced, network alignment will only continue to gain importance, as it can be used to transfer biological knowledge from well characterized species to poorly characterized ones between aligned network regions. Also, analogous to sequence alignment, network alignment can be used to infer species' phylogeny based on similarities of their biological networks. Thus, it could lead to new discoveries about the principles of life, evolution, disease, and therapeutics. \section*{Acknowledgements} We thank Dr. H. Bunke for suggestions regarding the parameters of the genetic algorithm and Drs. R. Patro and C. Kingsford for their assistance with running GHOST. This work was funded by the NSF CCF-1319469 and EAGER CCF-1243295 grants.
\section{Introduction} \noindent The recently completed calculation of the next-to-next-to-leading order (NNLO) corrections to the total hadronic top-quark pair-production cross section \cite{Baernreuther:2012ws, Czakon:2012zr, Czakon:2012pz, Czakon:2013goa} has been preceded by the derivation of the velocity enhanced terms in a threshold expansion in Ref.~\cite{Beneke:2009ye}. The latter publication exploited the advances in soft-gluon resummation \cite{Czakon:2009zw, Beneke:2009rj} and dealt with the incorporation of potential effects, which were proven to factorize in Ref.~\cite{Beneke:2010da}. After factorization of the Born cross section, the threshold expansion is given for the two leading channels, quark annihilation and gluon fusion, in terms of inverse powers of the top-quark velocity, $\beta$, and its logarithms. Unfortunately, the expansion formulae are missing $\beta$-independent terms, which are more difficult to derive. These terms are of some phenomenological interest, since they propagate through resummation to higher orders \cite{Cacciari:2011hy, Beneke:2011mq, Beneke:2012wb}. In principle, they can be obtained by expanding the fits to the numerical results provided in Refs.~\cite{Baernreuther:2012ws, Czakon:2013goa}. Nevertheless, due to the inherent lack of numerical precision in the strict threshold region, this approach leads to large uncertainties. In particular, Ref.~\cite{Czakon:2013goa} quotes a 50 \% uncertainty on the constant in the gluon fusion case. It is thus interesting, whether these constant terms can be obtained by other methods. Reviewing the same problem at the next-to-leading order (NLO), we notice that the $\beta$-indepen\-dent terms have only been obtained after an exact analytic calculation of the total cross sections \cite{Czakon:2008ii}, and their projections onto the singlet and octet color configurations of the top-quark pair \cite{Czakon:2008cx}. Currently, it is hardly conceivable to perform similar analytic calculations at NNLO. However, the threshold behavior of cross sections for heavy-flavor production is much better understood and it seems that the necessary information can be inferred from soft-gluon factorization. In this case, the effect of the radiation is contained in color configuration dependent soft functions, which are convoluted with hard functions representing the purely virtual contributions. As long as one is only interested in NNLO expansions of cross sections including $\beta$-independent terms, an additional factorization of the potential effects is not necessary. The hard and soft functions are needed for color singlet and octet configurations of the final state. Unfortunately, apart from the color singlet soft function \cite{Belitsky:1998tc}, they are unknown beyond NLO. At NLO, the soft function for the production of a massive color octet state at threshold has been evaluated in Refs.~\cite{Czakon:2009zw, Beneke:2009rj, Idilbi:2009cc}. Interestingly, Ref.~\cite{Idilbi:2009cc} demonstrates its application to the production of a fundamental scalar. The purpose of this work is to evaluate the color octet soft function at NNLO. The hard functions will be presented together with the complete virtual corrections in a subsequent publication. Beyond total cross sections, there are related developments aiming at the derivation of expansions and resummations of differential distributions in different kinematical regimes of top-quark pair-production. Most recent results are to be found in Refs.~\cite{Ferroglia:2012ku,Ferroglia:2012uy, Ferroglia:2013zwa, Ferroglia:2013awa, Zhu:2012ts, Li:2013mia}. They are extensions of previous analyses from Refs.~\cite{Ahrens:2011mw, Kidonakis:2010dk}. The paper is organized as follows. In the next section, we discuss the factorization of cross sections in the threshold limit with emphasis on hard functions. Subsequently, we define the soft function and provide the details of our calculation, the results for the bare soft function in $d$-dimensions, as well as the renormalized expression to be used in applications. Conclusions and outlook close the main text, which is supplemented with two appendices, one on Wilson lines and the other containing the anomalous dimensions, which are needed for renormalization. \section{Cross section factorization and hard functions} \noindent Consider the total hadronic cross section for the production of a heavy quark-anti-quark pair accompanied by any number of gluons and massless quarks. We will denote with $Q$ the invariant mass of the final state at threshold, i.e.\ $Q = 2m$, where $m$ is the heavy-quark mass. Notice that our considerations can also be applied to the production of an elementary state, e.g.\ a color octet scalar. In this case, there are no potential exchange effects in the final state and the discussion is slightly simplified. The cross section can be written as \begin{equation} \sigma_{h_1 h_2} = \sum_{ab} \hat\sigma^0_{ab} \otimes \phi^0_{a/h_1} \otimes \phi^0_{b/h_2} \; , \end{equation} where $\hat\sigma^0_{ab}$ is the partonic cross section for the initial state partons $a$ and $b$, while $\phi^0_{a/h_1}$ and $\phi^0_{b/h_2}$ are the parton distribution functions (PDFs) for the partons $a$ and $b$ inside the hadrons $h_1$ and $h_2$. The superscript 0 underlines that the quantities are not collinearly renormalized, while the symbol $\otimes$ denotes convolution in the momentum of the partons. There are only two possible channels in the Born approximation to the partonic cross section: quark-anti-quark annihilation and gluon-gluon fusion. We now assume that we are only interested in the production close to threshold, where the total energy of any additional radiation is strongly restricted from above, and the final state heavy quarks are non-relativistic. This condition can be enforced at the hadronic level by the available collider energy. Nevertheless, we will not discuss its phenomenological relevance in realistic situations. Let us first ignore any potential (e.g.\ Coulomb) interactions between the heavy quarks, which are also enhanced at threshold. The partonic matrix elements factorize in this soft limit. Any radiation can be approximated by emissions from eikonal lines, which can be described as Wilson lines at the operator level (see \ref{sec:wilson} for the definition and properties of Wilson lines, and section \ref{sec:soft} for their relation to eikonal lines). In the soft approximation, the matrix elements of the basic hard $2 \to 2$ production process without radiation are taken at threshold (potential effects are ignored at this point), while the partonic cross section for this process is only affected by radiation through the phase space volume. Let us denote the four-momentum of the radiation by $P^\mu = (\omega, \vec p)$, while the initial parton momenta in the center-of-mass system by $p_1$ and $p_2$. The phase space volume for the two particle state depends on \begin{equation} (p_1+p_2-P)^2 = \hat s-2\sqrt{\hat s} \, \omega + (\omega^2-\vec p^2) \approx \hat s - 2 \sqrt{\hat s} \, \omega \; , \end{equation} where $\hat s = (p_1+p_2)^2$, and the last approximation amounts to only keeping the leading behavior in the $\omega \rightarrow 0$ limit. If we now parameterize $\omega$ as \begin{equation}\label{eq:z1} \omega = \frac{Q}{2}(1-z) \; , \end{equation} then the volume of the phase space will be given by \begin{equation}\label{eq:z2} \hat s-\sqrt{\hat s} Q(1-z) \approx \hat s z \; . \end{equation} The factorization formula is \begin{equation}\label{eq:fac1} \hat \sigma^0 = \sum_{\alpha\beta} H^0_{\alpha\beta} \otimes S^0_{\alpha\beta} \; , \end{equation} where we have suppressed the parton indices, and the sum runs over color structures $\alpha$ and $\beta$. $H^0_{\alpha\beta}$ are the hard functions, i.e.\ cross sections for the $2 \to 2$ process, while $S^0_{\alpha\beta}$ are the soft functions containing the effect of the soft radiation from the Wilson lines. The convolution is performed in the $z \in [0,1]$ variable defined by Eqs.~(\ref{eq:z1},\ref{eq:z2}). The origin of the color structure indices is best understood by inspecting a schematic representation of the factorization given in Fig.~\ref{fig:fac}. The vertical dashed line represents the unitarity cut. On the left hand side, we have the matrix element, while on the right hand side its complex conjugate. The factorization occurs at the matrix element level for each color structure represented by the $\otimes$ symbol. The sum over color configurations is thus coherent. \begin{figure}[h] \begin{center} \includegraphics[width=90mm,angle=0]{factorization.eps} \caption{\sf Schematic representation of soft factorization. $H$ stands for the hard function, whereas $S$ for the soft function. The double lines denote Wilson lines, whereas the $\otimes$ symbols stand for the insertion of the color structure of the hard matrix element. The sum over different color structures is suppressed.\label{fig:fac}} \end{center} \end{figure} The choice of a basis for the color structures in the factorization formula is crucial. Indeed, if the hard matrix elements are decomposed into singlet and octet configurations of the final state, then the out-going Wilson lines of the soft functions at threshold, i.e.\ having the same velocity, can be combined into one as shown in \ref{sec:wilson}. The summation in Eq.~(\ref{eq:fac1}) becomes diagonal \cite{Beneke:2009rj}, since the color configurations of the singlet and octet are orthogonal (for initial state gluons, there are two octet configurations, symmetric and anti-symmetric - they are also orthogonal by Bose symmetry). The final factorization formula is now \begin{equation}\label{eq:fac2} \hat \sigma^0 = \sum_{\alpha} H^0_{\alpha} \otimes S^0_{\alpha} \; , \end{equation} where the single color index in the hard and soft functions specifies diagonal elements of both, and runs over singlet and octet configurations. At this point, we have to take into account the effect of potential interactions between the non-relativistic final state quarks. Fortunately, it turns out that we do not have to make substantial changes in our exposition \cite{Beneke:2010da}. As long as the hard function contains $s$-wave effects only and is decomposed into irreducible color representations of the final state, non-relativistic effects factorize from the soft effects. This factorization implies that the heavy-quark velocity must be set to zero in the soft component, as we already assumed. On the other hand, the hard amplitudes are to be expanded in the velocity, rather than just evaluated at threshold. In principle, we could now further factorize the potential effects in the hard functions as explained in Ref.~\cite{Beneke:2010da}, but we shall not do that, since we are not interested in their resummation, but rather in the fixed order expansion of the cross sections at threshold. This is achieved by formula Eq.~(\ref{eq:fac2}) after inclusion of the renormalization of the PDFs in the soft limit in order to yield finite results. A last subtlety concerns the restriction to $s$-wave contributions in the hard functions. Indeed, it is to be expected that higher partial waves will occur in the velocity independent terms of their threshold expansions beyond NLO. Nevertheless, they are not enhanced by soft radiation at NNLO and, consequently, do not spoil the expansion generated with Eq.~(\ref{eq:fac2}). Although the partonic cross section on the left hand side of Eq.~(\ref{eq:fac2}) suffers from initial state collinear divergences only, factorization introduces additional divergences into the hard and soft functions on the right hand side. We now consider the renormalization of these infrared divergences of the hard functions. Since the latter are given by total cross section contributions due to color projected purely virtual amplitudes, the information we need is contained in the singularities of the virtual amplitudes themselves. It turns out that the complete divergence structure of a UV renormalized amplitude $|M(\epsilon,\{\underline{p}\}, \{\underline{m}\})\rangle$ is encoded in the following equation \begin{equation} {\bf Z}_M^{-1}(\epsilon,\{\underline{p}\},\{\underline{m}\},\mu)\,\,|M(\epsilon,\{\underline{p}\},\{\underline{m}\})\rangle\,=\,\text{finite} \; , \end{equation} where the $\overline{\rm MS}$ renormalization constant ${\bf Z}_M$ is a matrix in color space and has a non-trivial dependence on the kinematics $\{\underline{p}\}=\{p_1,...,p_n\}$, and by the same on the masses $\{\underline{m}\}=\{m_1,...,m_n\}$ of the $n$ external partons. It can be derived from the differential equation \begin{equation} \frac{d}{d\ln\mu}\,{\bf Z}_M(\epsilon,\{\underline{p}\},\{\underline{m}\},\mu)\,=\,-{\bf \Gamma}_M(\{\underline{p}\},\{\underline{m}\},\mu) \, {\bf Z}_M(\epsilon,\{\underline{p}\},\{\underline{m}\},\mu) \; , \end{equation} where the color space matrix anomalous dimension is given by \cite{Ferroglia:2009ii} \begin{equation}\label{eq:Gamma} \begin{split} {\bf \Gamma}_M(\{\underline{p}\},\{\underline{m}\},\mu)\,&=\,\sum\limits_{(i,j)}\frac{{\bf T}_i\cdot{\bf T}_j}{2}\,\gamma_{\text{cusp}}(\alpha_s)\,\ln\frac{\mu^2}{-s_{ij}}\,+\,\sum\limits_i \gamma^i(\alpha_s)\\ &-\,\sum\limits_{(I,J)}\frac{{\bf T}_I\cdot{\bf T}_J}{2}\,\gamma_{\text{cusp}}(\beta_{IJ},\alpha_s)\,+\,\sum\limits_I \gamma^I(\alpha_s)\,+\,\sum\limits_{I,j}{\bf T}_I\cdot{\bf T}_j\,\gamma_{\text{cusp}}(\alpha_s)\,\ln\frac{m_I\,\mu}{-s_{Ij}}\\ &+\,\sum\limits_{(I,J,K)}i\,f^{abc}\,{\bf T}_I^a\,{\bf T}_J^b\,{\bf T}_K^c\,F_1(\beta_{IJ},\beta_{JK},\beta_{KI})\\ &+\,\sum\limits_{(I,J)}\sum\limits_k\,i\,f^{abc}\,{\bf T}_I^a\,{\bf T}_J^b\,{\bf T}_k^c\,f_2\left(\beta_{IJ},\ln\frac{-\sigma_{Jk}\,v_J\cdot p_k}{-\sigma_{Ik}\,v_I\cdot p_k}\right)\,+\,\mathcal{O}(\alpha_s^3) \; . \end{split} \end{equation} Its structure for massless partons has been determined already in Ref.~\cite{Aybat:2006mz}. The structure of the massive case has been studied in Refs.~\cite{Mitov:2009sv, Becher:2009kw}, with explicit expressions for the second line given in Refs.~\cite{Becher:2009kw, Czakon:2009zw}. Finally, the third and fourth lines have been determined in Ref.~\cite{Ferroglia:2009ii} (see also \cite{Mitov:2010xw}). The summations in Eq.~(\ref{eq:Gamma}) run over massless (indices $i, j, k$) and massive (indices $I, J, K$) partons, with the notation $(i,j,...)$ denoting unordered tuples of different indices. The color operators ${\bf T}^a_i$ act on the color indices of the respective partons. If the particle is a gluon carrying a color index $c$, we have $({\bf T}^a)_{bc}=-i\, f^{abc}$, assuming the result has been projected on color index $b$. Similarly, for an outgoing quark (or incoming anti-quark) the generator is $({\bf T}^a)_{bc} = T^a_{bc}$, whereas for an incoming quark (or outgoing anti-quark) the generator is $({\bf T}^a)_{bc} = -T^a_{cb}$. The kinematic dependence is contained in $s_{ij} = 2\sigma_{ij} p_i \cdot p_j + i0^+$, where the sign factor $\sigma_{ij} = +1$ if the momenta $p_i$ and $p_j$ are both incoming or outgoing, and $\sigma_{ij} = -1$ otherwise. For massive partons there is $p_I^2 = m_I^2$, $v_I = p_I/m_I$, and $\cosh \beta_{IJ} = -s_{IJ}/2m_I m_J$. The cusp anomalous dimensions, $\gamma_{\rm cusp}$, for the massless and massive cases, and the functions $F_1,f_2$ can be found in Ref.~\cite{Ferroglia:2009ii} and references therein. It is interesting to note that the triple color correlations given in the third and fourth lines of Eq.~(\ref{eq:Gamma}) cannot contribute to the divergences of spin and color summed amplitudes at NNLO, as long as the Born amplitudes do not contain complex couplings or masses. This implies in particular that they will not contribute to top-quark pair-production amplitudes, which was noticed for the quark annihilation channel in Ref.~\cite{Czakon:2009zw} and for both channels in Ref.~\cite{Ferroglia:2009ii}. In the general case, the argument is as follows. First, notice that one can decompose the Born amplitude treated as a vector in color and spin space in terms of color structures \begin{equation} | M^{(0)} \rangle = \sum_\alpha | M^{(0)}_\alpha \rangle \otimes | c_\alpha \rangle \; , \end{equation} where the vectors $| c_\alpha \rangle$ are made of $T^a_{bc}$ and $i f^{abc}$ only. The amplitudes $|M^{(0)}_\alpha \rangle$ are stripped of all color factors generated from QCD vertices. Now, for $i,j,k$ all different (the indices make no distinction this time between massive and massless partons), there is \begin{equation}\label{eq:antisymmetry} \langle c_\alpha | \, if^{abc} {\bf T}^a_i {\bf T}^b_j {\bf T}^c_k \, | c_\beta \rangle^* = -\langle c_\beta | \, if^{abc} {\bf T}^a_i {\bf T}^b_j {\bf T}^c_k \, | c_\alpha \rangle \; , \end{equation} simply because the ${\bf T}^a_i$ operators are hermitian and commute with each other as long as the parton indices are different, and because the structure constants are real. On the other hand, both sides of Eq.~(\ref{eq:antisymmetry}) are real, since they can be evaluated with the Cvitanovi\'c algorithm \cite{Cvitanovic:1976am} containing only real expressions. The color matrix elements of the triple color correlator are thus anti-symmetric in the color indices and we have \begin{eqnarray}\label{eq:matrix} \langle M^{(0)} | \, if^{abc} {\bf T}^a_i {\bf T}^b_j {\bf T}^c_k \, | M^{(0)} \rangle &=& \sum_{\alpha\beta} \langle M^{(0)}_\alpha | M^{(0)}_\beta \rangle \langle c_\alpha | \, if^{abc} {\bf T}^a_i {\bf T}^b_j {\bf T}^c_k \, | c_\beta \rangle \\ \nonumber &=& \frac{1}{2} \sum_{\alpha\beta} \left(\langle M^{(0)}_\alpha | M^{(0)}_\beta \rangle - \langle M^{(0)}_\beta | M^{(0)}_\alpha \rangle \right) \langle c_\alpha | \, if^{abc} {\bf T}^a_i {\bf T}^b_j {\bf T}^c_k \, | c_\beta \rangle \; . \end{eqnarray} Since $\langle M^{(0)}_\alpha | M^{(0)}_\beta \rangle^* = \langle M^{(0)}_\beta | M^{(0)}_\alpha \rangle$, the right hand side of Eq.~(\ref{eq:matrix}) vanishes if $\langle M^{(0)}_\alpha | M^{(0)}_\beta \rangle$ is real. Due to elementary spin summation rules, this is the case if there are no complex parameters in the Lagrangian (the case of complex parameters is generally of interest, since we might want to describe unstable particles). The argument presented above is, of course, also valid for diagonal matrix elements between color projected Born amplitudes, which shows that the NNLO renormalization constants for our hard functions can be determined from the dipole correlations given in the first two lines of Eq.~(\ref{eq:Gamma}). Since the initial partons are either in the fundamental, or the adjoint representation, while the final state may be in a singlet or octet configuration, we have to consider a set of color vectors $| c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \rangle$, where $\bm{R} \in \{ \bm{3}, \bm{8} \}$ denotes the representation of the initial partons, while $\bm{R'} \in \{ \bm{1}, \bm{8}, \bm{8_A}, \bm{8_S} \}$ that of the final state. The subscripts $\bm{S}$ and $\bm{A}$ in the latter case stand for symmetric and anti-symmetric octets in the case of an $\bm{8}\otimes\bm{8}$ initial configuration. The bare hard functions for heavy flavor production are given by \begin{equation} H^0_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} = {\cal N} \int d \mbox{PS}_2 \sum_{\alpha\beta} \frac{\langle \alpha | c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \rangle \langle c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} | \beta \rangle}{\langle c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} | c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \rangle} \langle M_\alpha | M_\beta \rangle \; , \end{equation} where the right hand side is expanded in the heavy-quark velocity, $\beta$, up to and including terms of order $\beta$. ${\cal N}$ denotes the product of the flux, and color and spin average factors, while $| M_\alpha \rangle$ are the purely virtual UV renormalized amplitudes for heavy flavor production, where the initial state is specified by the color representation $\bm{R}$. The expansion in $\beta$ in the definition of the hard functions would directly correspond to the soft approximation of taking the matrix element at threshold and keeping the exact phase space, if there were no potential effects. Due to the latter, the result contains inverse powers and logarithms of $\beta$. The renormalization of the hard functions is now achieved with \begin{equation} H^0_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} = Z^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}_H(\mu/Q) \, H_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \; , \end{equation} where the renormalization constant $Z_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}$ satisfies the equation \begin{equation} \frac{d}{d \ln \mu} \, Z_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}(\mu/Q) = -\Gamma_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}(\mu/Q) \, Z_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}(\mu/Q) \; , \end{equation} with \begin{equation} \Gamma_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}(\mu/Q) = \lim_{\beta \to 0} 2\,\Re\left( \langle c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} |\,{\bf \Gamma}_M\left(\beta,\cos\theta,\frac{\mu}{Q} \right) | c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \rangle\Big/ \langle c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} | c^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}} \rangle \right) \; . \end{equation} Contrary to the hard function itself, the hard anomalous dimensions are finite in the $\beta \to 0$ limit, and do not depend on the scattering angle $\theta$. They can be found in \ref{sec:AnomalousDimensions}. We mentioned before that the same formalism can be applied to the production of a fundamental object, e.g.\ a color octet scalar. Clearly, the constant $Z_H^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}$ will also be used for the renormalization of the soft function, which does not depend on the nature of the final state. This means that the latter does not play any role in the divergences of the amplitude at threshold. Indeed, this is true for the real part of the anomalous dimension ${\bf \Gamma}_M$ (the imaginary part contains Coulomb phases). The final state anomalous dimension coefficients in Eq.~(\ref{eq:Gamma}) for massive states have a purely soft origin. \section{Soft function} \label{sec:soft} \noindent We define the bare soft function for color octet production at rest as follows \begin{eqnarray}\label{eq:SoftFunction} S_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}^0(\omega) &=& \frac{Q}{2} \sum_{X} \delta(\omega - E_X) \\ \nonumber &\times& \sum_{abc} \left| \sum_{a'b'c'} C^{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}_{a'b'c'} \langle X | T \left[ \Phi^{(\bm 8)}_{v,aa'}(+\infty,0) \Phi^{(\bm R)}_{n,b'b}(0,-\infty) \Phi^{(\bm{\bar R})}_{\bar n,c'c}(0,-\infty) \right] | 0 \rangle\right|^2 \; , \end{eqnarray} where the summation in the first line is taken over all possible states $X$ with energy $E_X$, and in the second line over color indices. The Wilson line operators $\Phi^{(\bm R)}_{\beta,cd}(b,a)$ discussed in \ref{sec:wilson} have directions $n^\mu = (1,\vec{\bm 0}^{(d-2)},1)$, $\bar{n}^\mu = (1,\vec{\bm 0}^{(d-2)},-1)$ for the two incoming light-cone states, and $v^\mu = (1,\vec{\bm 0}^{(d-1)})$ for the outgoing octet at rest. Notice that time ordering is in fact only necessary for the product of the two incoming lines. The notation $\bm{R} \otimes \bm{\bar R} | {\bm{R'}}$ specifies the representation of the initial states as $\bm{R}$ and $\bm{\bar R}$, and the restriction to the hard scattering color structure corresponding to the irreducible component $\bm{R'}$ of the tensor product $\bm{R} \otimes \bm{\bar R}$. We consider three cases \begin{equation}\label{eq:ColorStructures} C^{\bm{3} \otimes \bm{\bar{3}} | {\bm{8}}}_{abc} = \sqrt{\frac{2}{N_c^2-1}} T^a_{cb} \; , \;\; C^{\bm{8} \otimes \bm{8} | {\bm{8_A}}}_{abc} = \sqrt{\frac{1}{N_c(N_c^2-1)}} i f^{abc} \; , \;\; C^{\bm{8} \otimes \bm{8} | {\bm{8_S}}}_{abc} = \sqrt{\frac{N_c}{(N_c^2-1)(N_c^2-4)}} d^{abc} \; , \end{equation} where $N_c = 3$, and the symmetric tensor $d^{abc}$ is defined through $\mbox{Tr} \; T^a T^b T^c = 1/4(d^{abc}+i f^{abc})$. The invariance of the color structures together with the gauge transformation properties of the Wilson lines, Eq.~(\ref{eq:covariance}), assure the gauge invariance of the soft function. Up to NNLO, the singularities in $\epsilon$ of $S_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}^0$ are known to only depend on the Casimir invariants of the representations of the Wilson lines \cite{Czakon:2009zw, Beneke:2009rj}. We have checked explicitly that this is valid for the exact $\epsilon$ dependence as well. Since the out-going massive Wilson line is always in the octet representation, the result depends on $\bm{R} \otimes \bm{\bar R} | {\bm{R'}}$ only through $C_R$. Therefore, we will drop the subscript in $S_{\bm{R} \otimes \bm{\bar R} | {\bm{R'}}}^0$ in all the subsequent formulae. The prefactor $Q/2$ in Eq.~(\ref{eq:SoftFunction}) and the normalization of the color structures in Eq.~(\ref{eq:ColorStructures}) guarantee that the soft function reduces to $\delta(1-z)$ at leading order. Dimensional analysis allows to write the following perturbative expansion \begin{equation}\label{eq:S0} S^{0}(z) = \delta(1-z)+\frac{1}{1-z}\,\sum\limits_{n=1}^\infty \left(\frac{Z_{\alpha_s}\,\alpha_s}{\pi}\right)^{n}\,\left(\frac{\mu}{Q\,(1-z)}\right)^{2 n\epsilon}\,s^{(n)} \; , \end{equation} where the coefficients $s^{(n)}$ depend on $\epsilon$, color (through $C_A$ and $C_R$) and the number of light quark flavors $n_f$ (as usual, through $T_F n_f$). We now apply a Mellin transform \begin{equation} \begin{split} S^{0}(N)&= 1+\,\sum\limits_{n=1}^\infty \left(\frac{Z_{\alpha_s}\,\alpha_s}{\pi}\right)^{n}\,\left(\frac{\mu}{Q}\right)^{2 n\epsilon}\,\left[\int\limits_0^1 z^{\tilde{N}-1}\,(1-z)^{-1-2 n \epsilon}\,dz\right]\,s^{(n)}\\ &= 1+\,\sum\limits_{n=1}^\infty \left(\frac{Z_{\alpha_s}\,\alpha_s}{\pi}\right)^{n}\,\left(\frac{\mu}{Q}\right)^{2 n\epsilon}\,\frac{\Gamma(\tilde{N})\,\Gamma(-2 n \epsilon)}{\Gamma(\tilde{N}-2 n \epsilon)}\,s^{(n)}\\ &= 1+\,\sum\limits_{n=1}^\infty \left(\frac{Z_{\alpha_s}\,\alpha_s}{\pi}\right)^{n} \,e^{2n\epsilon(L-\gamma_E)}\,\Gamma(-2 n \epsilon)\,s^{(n)}\,+\,\mathcal{O}\left(\frac{1}{N}\right) \; , \end{split} \end{equation} where $\tilde{N} = N e^{-\gamma_{{E}}}$ and $L = \ln(\mu N/Q)$. We only keep the leading behavior in $1/N$, since the limit $N \rightarrow \infty$ in Mellin space corresponds to the soft limit $z \rightarrow 1$ . The last line of the above equation demonstrates that the soft function depends on $L$ rather than separately on $\mu/Q$ and $N$. The finiteness of the partonic cross section implies that the renormalized soft function is given by \begin{equation}\label{eq:SR} S(L) = Z_H(\mu/Q) Z_\phi^2(N) S^{0}(L) \; , \end{equation} where we have suppressed the dependence on the initial state parton in both $Z_H$ and $Z_\phi$, the latter being the renormalization constant of the parton distribution function. $Z_\phi$ satisfies the equation \begin{equation} \frac{dZ_\phi(N)}{d\ln\mu^2} = -P(N)Z_\phi(N) \; , \end{equation} where $P(N)$ is the soft expansion of the Altarelli--Parisi splitting kernel in Mellin space, which can be found in \ref{sec:AnomalousDimensions}. In consequence, the soft function obeys the Renormalization Group Equation (RGE) \begin{equation}\label{eq:RGE} \frac{d S(L)}{dL} = -\left(\Gamma_H(\mu/Q) + 4 P \left(N\right)\right) S(L) \; , \end{equation} which shows that the logarithmic dependence on $\mu/Q$ in $\Gamma_H$ and on $N$ in $P$ must combine into a dependence on their product as given in $L$. This is a demonstration of the well known fact that the singular part of the soft limit of the splitting kernels is given by the same soft anomalous dimension, which governs the soft-collinear singularities of the virtual amplitudes. The RGE Eq.~(\ref{eq:RGE}) can be used to resum large logarithms of $N$. For this purpose, it is sufficient to evolve from the scale $\mu = \mu_0/N$, where $L = \ln(\mu_0/Q)$ is small, to the actual scale $\mu = \mu_0$. We are, however, concerned with the fixed order perturbative expansion, and will now present the calculation and results at NNLO. \subsection{Calculation up to ${\cal O}(\alpha_s^2)$} The ${\cal O}(\alpha_s)$ contribution to the soft function for color octet production at threshold has been obtained in \cite{Idilbi:2009cc, Czakon:2009zw}, and for general representations of the three Wilson lines in \cite{Beneke:2009rj}. With the conventions of Eq.~(\ref{eq:S0}), the result reads \begin{equation} s^{(1)} = -e^{\gamma_{{E}}\epsilon}\frac{\Gamma(1-\epsilon)}{\Gamma(1-2\epsilon)}\left(C_A\,\frac{1}{1-2 \epsilon}+\,C_R\,\frac{2}{\epsilon}\right) \; . \end{equation} The ${\cal O}(\alpha_s^2)$ contribution is conveniently evaluated in the momentum representation, where Wilson lines become eikonal lines. Each emission of a gluon with momentum $q_1$ from a hard parton with momentum $p$ contributes a factor \begin{equation} i g_s^0 \frac{i p^\mu}{\pm p \cdot q + i\epsilon} \, {\bf T}^{(\bm R) \, a} \; , \end{equation} where $q$ is the sum of $q_1$ and the momenta of the gluons emitted before for an in-going line (minus sign in the denominator), or after for an out-going line (plus sign in the denominator). The phase space integrations are performed with \begin{equation} \begin{split} d\mbox{PS}_{{1}} &= \frac{Q}{2} \int \frac{d\Omega_{{d-1}}\,dE}{(2\pi)^{d-1}}\,\frac{E^{d-3}}{2}\,\delta(\omega-E) \; , \\ d\mbox{PS}_{{2}} &= \frac{Q}{2} \int \frac{d\Omega_{{d-1}}^{(1)}\,dE_{{1}}}{(2\pi)^{d-1}}\frac{d\Omega_{{d-1}}^{(2)}\,dE_{{2}}}{(2\pi)^{d-1}}\,\frac{E_{{1}}^{d-3}}{2}\,\frac{E_{{2}}^{d-3}}{2}\,\delta(\omega-E_{{1}}-E_{{2}}) \; , \end{split} \end{equation} for the real-virtual (one-loop corrections to single gluon emission), and double-real (double-gluon or quark-anti-quark emission) cases respectively. For a gluon pair in the final state, an additional factor of $1/2$ has to be included. Our result for the bare ${\cal O}(\alpha_s^2)$ contribution in $d$-dimensions is presented in the form of four contributions \begin{equation} s^{(2)} = s_{{\Square}}^{(2)} + s_{{\bigtriangleup}}^{(2)} + s_{{\Circle}}^{(2)} + s_{{\CIRCLE}}^{(2)} \; . \end{equation} The first three of them correspond to double-real radiation, whereas the last one to the real-virtual corrections. The two cases are discussed separately. \subsection{Double-real corrections} \begin{figure}[h] \begin{center} \begin{tabular}{|m{1cm} m{3cm} m{3cm} m{3cm}|} \hline &&&\\ &\includegraphics[width=29mm,angle=0]{Eikonal1.eps} &\includegraphics[width=29mm,angle=0]{Eikonal2.eps}&\includegraphics[width=29mm,angle=0]{Eikonal3.eps}\\ (A)&&&\\ &\includegraphics[width=29mm,angle=0]{Eikonal4.eps} &\includegraphics[width=29mm,angle=0]{Eikonal5.eps}&\includegraphics[width=29mm,angle=0]{Eikonal6.eps}\\ \hline &&&\\ (B)&\includegraphics[width=29mm,angle=0]{Eikonal10.eps} &\includegraphics[width=29mm,angle=0]{Eikonal11.eps}&\includegraphics[width=29mm,angle=0]{Eikonal7.eps}\\ \hline &&&\\ (C)&\includegraphics[width=29mm,angle=0]{Eikonal12.eps} &\includegraphics[width=29mm,angle=0]{Eikonal9.eps}&\includegraphics[width=29mm,angle=0]{Eikonal8.eps}\\ \hline \end{tabular} \end{center} \caption{\sf Complete set of double-real emission graphs contributing to the soft function. The graphs are divided into: (A) two emissions from eikonal lines, (B) gluon splitting after emission from an eikonal line, (C) massless quark-pair emission. \label{fig:EikonalRR}} \end{figure} \noindent The complete set of double-real emission graphs is depicted in Fig.~\ref{fig:EikonalRR}. We have divided the diagrams into three categories: (A) two emissions from eikonal lines, (B) gluon splitting after emission from an eikonal line, (C) massless quark-pair emission. The division into (A) and (B) is not gauge invariant. In order to uniquely define the contributions of the interferences between the (A) and (B) categories, we mention that we work in the Feynman gauge and that we take the gluon polarization sums to be \begin{equation} \sum_\lambda \epsilon_\mu(q,\lambda) \epsilon^*_\nu(q,\lambda) \rightarrow -g_{\mu\nu} \; . \end{equation} In consequence, we also have to take ghost pairs into account, when evaluating the square of the sum of the diagrams from (B). In our calculation, we have not made use of non-abelian exponentiation \cite{Mitov:2010rp, Gardi:2013ita} (see, however, next subsection for an application in the real-virtual case). The color factors are obtained for $SU(N_c)$ and subsequently translated into Casimir operators. The occurring phase space integrals can be evaluated analytically by the following method. After canceling numerators with denominators where possible, partial fractioning is used to obtain denominators with the smallest number of different scalar products from the set $n_i \cdot q_j$, $n_i \cdot (q_1 + q_2)$, $q_1 \cdot q_2$, where $q_1$ and $q_2$ are the momenta of the emitted gluons. Scalar products of the gluon momenta with $v$ are harmless, since they only depend on the gluon energy. Subsequently, the denominators involving $n_i \cdot (q_1 + q_2)$, are split with a Mellin-Barnes representation \begin{equation} \frac{1}{(n_i \cdot (q_1+q_2))^\alpha} = \frac{1}{\Gamma(\alpha)} \int_{-i \infty}^{+i \infty} \frac{dz}{2\pi i} \, \frac{1}{(n_i \cdot q_1)^{\alpha+z}(n_i \cdot q_2)^{-z}} \Gamma(\alpha+z) \Gamma(-z) \; , \end{equation} where the contour is chosen to separate the poles of the gamma functions. The angular integrations can be performed with the following formulae \cite{Somogyi:2011ir} \begin{equation} \begin{split} \int d\Omega_{{d-1}}(q)\,\frac{(q^0)^\alpha}{(n_{{i}} \cdot q)^\alpha}&=2^{2-\alpha-2\epsilon}\, \pi^{1-\epsilon}\,\frac{\Gamma(1-\alpha-\epsilon)}{\Gamma(2-\alpha-2\epsilon)} \; , \\ \\ \int d\Omega_{{d-1}}(q)\,\frac{(q^0)^{\alpha+\beta}}{(n \cdot q)^\alpha(\bar n \cdot q)^\beta} &=2^{2-\alpha-\beta-2\epsilon}\,\pi^{1-\epsilon}\, \frac{\Gamma(1-\alpha-\epsilon)\,\Gamma(1-\beta-\epsilon)}{\Gamma(1-\epsilon) \Gamma(2-\alpha-\beta-2\epsilon)} \; , \\ \\ \int d\Omega_{{d-1}}(q_2)\,\frac{(q_1^0)^\beta(q_2^0)^{\alpha+\beta}}{(n_i \cdot q_2)^\alpha(q_1 \cdot q_2)^\beta} &=2^{2-\alpha-\beta-2\epsilon}\,\pi^{1-\epsilon} \,\frac{1}{\Gamma(\alpha)\,\Gamma(\beta) \,\Gamma(2-\alpha-\beta-2\epsilon)}\\ &\!\!\!\! \!\!\!\! \!\!\!\! \times\int_{-i\infty}^{+i\infty}\frac{dz}{2\pi i}\,\Gamma(-z)\Gamma(z+\alpha) \Gamma(z+\beta) \Gamma(1-\alpha-\beta-\epsilon-z) \left(\frac{n_i \cdot q_1}{2 q_1^0} \right)^z \; . \end{split} \end{equation} The energy integrations in the resulting integrals can be performed directly in terms of the Euler beta function \begin{equation} \int dE_1 \, dE_2 \, \delta(\omega-E_1-E_2) \, E_1^\alpha \, E_2^\beta = \frac{\Gamma(\alpha+1)\Gamma(\beta+1)}{\Gamma(\alpha+\beta+2)} \omega^{\alpha+\beta+1} \; . \end{equation} After application of the Barnes' lemmas, we end up with at most one-fold Mellin-Barnes integrals, which can be resummed to hypergeometric functions by closing contours and taking residues. In practice, we have used the packages {\sc MB} \cite{Czakon:2005rk} and {\sc HypExp} \cite{Huber:2005yg} for the manipulation of Mellin-Barnes integrals and hypergeometric functions respectively. \begin{figure}[h] \begin{center} \includegraphics[width=50mm,angle=0]{Eikonal13.eps} \caption{\sf Interference diagrams contributing a hypergeometric function to $s_{{\Square}}^{(2)}$. \label{fig:Box}} \end{center} \end{figure} The results are given separately for three parts. The first part corresponds to the square of category (A) from Fig.~\ref{fig:EikonalRR}. The result contains a single hypergeometric function, due to the interference diagrams shown in Fig.~\ref{fig:Box}. We obtain \begin{equation} \begin{split} s_{{\Square}}^{(2)}&=-e^{2\gamma_{{E}}\epsilon}\frac{\Gamma^2(1-\epsilon)}{\Gamma(1-4\epsilon)}\\ &\times\left(C_A^2\,\left(\frac{2-19\epsilon+64\epsilon^2-74 \epsilon^3}{12 (1-2 \epsilon)^2\,(1-4 \epsilon)\,\epsilon^3 }+\frac{\Gamma (1+\epsilon)\,\Gamma (1+2 \epsilon)\,\Gamma (1-2 \epsilon)\,\Gamma (1-3 \epsilon)}{24 \epsilon^3\,\Gamma^2(1-\epsilon)}\right.\right.\\ &\left.\left.-\frac{1}{12 \epsilon^2 (1-2 \epsilon)}\,_3 F_2(1,1-2 \epsilon,1-\epsilon;2-2 \epsilon,1+\epsilon;1)\right)-\,C_A\,C_R\,\frac{3-14 \epsilon}{4 \epsilon^3\,(1-2 \epsilon)}+C_R^2\,\frac{2}{\epsilon^3}\right) \; . \end{split} \end{equation} The second part, due to the interference of categories (A) and (B) is the most complicated. The result contains several hypergeometric functions \footnote{A subset of the graphs evaluated here occurs in the case of singlet production, i.e. for the soft function for the Drell--Yan process given in \cite{Belitsky:1998tc}. The result from \cite{Belitsky:1998tc} contains an Appell function, which can be expressed in terms of a ${}_3F_2$ hypergeometric function: $F_2(1,1+\epsilon,-2\epsilon,2+\epsilon,1-2\epsilon;1,1)\,= \frac{1+\epsilon}{2\epsilon}\left(\frac{2\,\Gamma(1-\epsilon)\, \Gamma^3(1+\epsilon)}{\Gamma(1+2\epsilon)}-\,_3 F_2(1,1-\epsilon,-2 \epsilon;1-2 \epsilon,1-2 \epsilon;1)\right)$.} \begin{equation} \begin{split} s_{{\bigtriangleup}}^{(2)}&=-e^{2\gamma_{{E}}\epsilon}\frac{\Gamma^2(1-\epsilon)}{\Gamma(1-4\epsilon)}\\ &\times\left(C_A^2\left(\frac{3-\epsilon+2 \epsilon^2}{24\epsilon^3\,(1-2 \epsilon)^2 }+\frac{(1+\epsilon)\,\Gamma (1+\epsilon)\,\Gamma (1+2 \epsilon) \Gamma (1-2 \epsilon) \Gamma (1-3 \epsilon) }{12 \epsilon^3\,(1-2 \epsilon) \Gamma^2(1-\epsilon)}\right.\right.\\ &\left.\left.+\frac{2-\epsilon}{24 \epsilon^3\,(1-2 \epsilon)}\,_3 F_2(1,1-\epsilon,-2 \epsilon;1-2\epsilon,\epsilon;1)-\frac{1}{8 \epsilon^3}\,_3 F_2(1,1-\epsilon,-2 \epsilon;1-2 \epsilon,1-2 \epsilon;1)\right.\right.\\ &\left.\left.-\frac{1-\epsilon}{4 \epsilon^2\,(1-2 \epsilon)^2}\,_3 F_2(1,1-2 \epsilon,2-\epsilon;2-2 \epsilon,1+\epsilon;1)\right)+\,C_A\,C_R\,\left(\frac{1-\epsilon}{2 \epsilon^3\,(1-2 \epsilon)}\right.\right.\\ &\left.\left.+\frac{1}{4 \epsilon^3}\,_3 F_2(1,1-\epsilon,-2 \epsilon;1-2 \epsilon,1-2 \epsilon;1)\right)\right) \; . \end{split} \end{equation} The third part is given by the sum of the squares of categories (B) and (C). It is particularly simple, because it can be thought of as the ${\cal O}(\alpha_s)$ contribution with an insertion of the imaginary part of the gluon vacuum polarization on the gluon line. The result reads \begin{equation} \begin{split} s_{{\Circle}}^{(2)}&=-e^{2\gamma_{{E}}\epsilon}\frac{\Gamma^2(1-\epsilon)}{(3-2\epsilon)\,(1-2\epsilon)\,\Gamma(1-4\epsilon)}\\ &\times\left(C_A^2\,\frac{5-3 \epsilon}{4 \epsilon\,(1-4 \epsilon)}+\,C_A\,C_R\,\frac{5-3 \epsilon}{4 \epsilon^2}-\,C_A\,T_F\,n_f\,\frac{1-\epsilon}{\epsilon \,(1-4\epsilon)}-\,C_R\,T_F\,n_f\,\frac{1-\epsilon}{\epsilon^2}\right) \; . \end{split} \end{equation} \subsection{Real-virtual corrections} \noindent Let us now consider virtual corrections to the soft function. Since we are working in dimensional regularization, a Feynman integral does not vanish if the result of the integration can be represented in the form \begin{equation} \int d^d k f(k) \propto \Lambda^{a+b \epsilon} \; , \end{equation} where $a$ and $b$ are some real constants with $b \neq 0$, $\Lambda$ is a dimensionful parameter, and the proportionality coefficient does not contain any other dimensionful parameters with $\epsilon$ dependent exponents. In the case of purely virtual corrections involving external eikonal lines only, the result must have integer scaling with respect to all the available momentum parameters. This implies the vanishing of $b$, since the denominators of the Feynman integrands contain no other dimensionful parameters. In consequence, purely virtual corrections vanish. Non-vanishing contributions are obtained once at least one real particle, a gluon, is emitted, as for the real-virtual corrections we wish to discuss in this subsection. In this case, we have two possible types of dimensionful parameters, which are invariant under hard momentum rescaling, and can thus be used in place of $\Lambda$ above \begin{equation} \Lambda_1(p,q) = \frac{m}{p \cdot q} \; , \;\; \Lambda_2(p_1,p_2,q) = \frac{p_1 \cdot p_2}{p_1 \cdot q \; p_2 \cdot q} \; , \end{equation} where $q$ is the momentum of the external gluon, while $p$ with $p^2 = m^2$, $p_1$ and $p_2$ are the momenta of the eikonal lines. We can now consider two different types of emissions: from an eikonal line, or from a virtual gluon line. In the case of an emission from an eikonal line, the denominators of the Feynman integrands depend on the emitted gluon momentum through $p_1 \cdot (k + q)$ only, with $p_1$, $q$ and $k$ the momenta of the emitting eikonal line, the emitted gluon, and the loop integration respectively. Therefore, the result of the integration depends non-trivially only on $p_1 \cdot q$ (ignoring the numerator, which does not influence the $d$-dimensional scaling of dimensionful parameters), and thus either on $\Lambda_1(p_1,q)$ or $\Lambda_2(p_1,p_2,q)/\Lambda_1(p_2,q)$, where $p_2$ is the momentum of the other eikonal line integrated in the loop (if the integration involves two different eikonal lines). Such emissions will thus only give non-vanishing contributions, if either $p_1^2 \neq 0$ or $p_2^2 \neq 0$. Emissions from gluon lines give non-vanishing contributions either if the loop contains two different eikonal lines, in which case the result can be expressed through $\Lambda_2(p_1,p_2,q)$, or if the single eikonal line in the loop is massive, in which case the result is expressed through $\Lambda_1(p,q)$. The possible graphs generated according to these considerations are shown in Fig.~\ref{tab:EikonalRV}. \begin{figure}[h] \begin{center} \begin{tabular}{|cc|cc|} \multicolumn{1}{c}{\includegraphics[width=29mm,angle=0]{EikonalRV9.eps}}&\multicolumn{1}{c}{\includegraphics[width=29mm,angle=0]{EikonalRV10.eps}}&\multicolumn{1}{c}{\includegraphics[width=29mm,angle=0]{EikonalRV11.eps}}&\multicolumn{1}{c}{\includegraphics[width=29mm,angle=0]{EikonalRV13.eps}}\\ \hline &&&\\ \includegraphics[width=29mm,angle=0]{EikonalRV1.eps} &\includegraphics[width=29mm,angle=0]{EikonalRV2.eps}&\includegraphics[width=29mm,angle=0]{EikonalRV3.eps}&\includegraphics[width=29mm,angle=0]{EikonalRV4.eps}\\ \hline &&&\\ \includegraphics[width=29mm,angle=0]{EikonalRV5.eps} &\includegraphics[width=29mm,angle=0]{EikonalRV6.eps}&\includegraphics[width=29mm,angle=0]{EikonalRV7.eps}&\includegraphics[width=29mm,angle=0]{EikonalRV8.eps}\\ \hline &&\multicolumn{2}{c}{}\\ \includegraphics[width=29mm,angle=0]{EikonalRV14.eps} &\includegraphics[width=29mm,angle=0]{EikonalRV12.eps}&\multicolumn{2}{c}{}\\ \cline{1-2} \end{tabular} \end{center} \caption{\sf Complete set of non-vanishing real-virtual graphs contributing to the soft function. The pairs of graphs in boxes can be combined as explained in the text. \label{tab:EikonalRV}} \end{figure} \begin{figure} \begin{center} \includegraphics[width=80mm,angle=0]{EikonalId.eps} \end{center} \caption{\sf Real gluon emission (momentum $q$ and color index $a$) from an eikonal line in the presence of a virtual gluon (momentum $k$ and color index $b$).\label{fig:EikonalId}} \end{figure} The pairs of graphs in boxes in Fig.~\ref{tab:EikonalRV} can be combined using color algebra and eikonal identities. This is a simple case of eikonal exponentiation \cite{Mitov:2010rp, Gardi:2013ita} for multiple eikonal lines joined at one point. Consider for example the two graphs of Fig.~\ref{fig:EikonalId}. The color factor of the right graph can be represented as the sum of the color factor of the left graph and an additional contribution \begin{equation}\label{colorid} {\bf T}^{(\bm R) \, a} \, {\bf T}^{(\bm R) \, b} = {\bf T}^{(\bm R) \, b} \, {\bf T}^{(\bm R) \, a} + if^{abc} {\bf T}^{(\bm R) \, c} \; . \end{equation} If we ignore the second term on the right hand side of this equation, the two graphs will only differ in the kinematics. Their sum can then be written under the same integral sign and will contain the factor \begin{equation} \frac{1}{p \cdot q} \frac{1}{p \cdot (k + q)} + \frac{1}{p \cdot k} \frac{1}{p \cdot (k + q)} = \frac{1}{p \cdot q} \frac{1}{p \cdot k} \; . \end{equation} This is the basic eikonal identity, which shows that the two emissions can be completely factorized. Such a contribution corresponds to a product of a real emission and a pure virtual contribution. It thus vanishes by the arguments from the beginning of this subsection. In consequence, for each pair of emissions as in Fig.~\ref{fig:EikonalId}, it is sufficient to only consider the second graph with a modified color factor corresponding to the second term on the right hand side of Eq.~(\ref{colorid}). As far as the virtual integrals are concerned, it turns out that they are the same as those that were calculated in \cite{Bierenbaum:2011gg}. The complete list is \begin{center} \begin{tabular}{llll} $M_1$ : &$\int d^{d}k\,\frac{1}{[k^2]\,[(k+q)^2]\,[n_i\cdot k]}$ \; ,&$\int d^{d}k\,\frac{1}{[k^2]\,[(k+q)^2]\,[-v\cdot k]}$ \, ,&\\ &&&\\ $M_2$ : &$\int d^{d}k\,\frac{1}{[k^2]\,[-n_i\cdot k-n_i\cdot q]\,[-v\cdot k]}$ \, , & $\int d^{d}k\,\frac{1}{[k^2]\,[v\cdot k+v\cdot q]\,[n\cdot k]}$ \, , & \!\!\!\! \!\!\!\! \!\!\!\! $\int d^{d}k\,\frac{1}{[k^2]\,[v\cdot k+v\cdot q]\,[-v\cdot k]}$ \, , \\ &&&\\ $M_3$ : &$\int d^{d}k\,\frac{1}{[k^2]\,[(k+q)^2]\,[-n\cdot k-n\cdot q]\,[\overline{n}\cdot k]}$ \, , & $\int d^{d}k\,\frac{1}{[k^2]\,[(k+q)^2]\,[v\cdot k-v\cdot q]\,[n_i\cdot k]}$ \, , &\\ \end{tabular} \end{center} where the naming of the master integrals, $M_1$, $M_2$ and $M_3$, matches that of Ref.~\cite{Bierenbaum:2011gg}. Our integrals are all integrated to Euler gamma functions in $d$-dimensions, because we only have one massive line. The remaining phase space integration has the same complexity as in the ${\cal O}(\alpha_s)$ case. We use a subset of the formulae from the calculation of the double-real corrections, but never have to introduce any Mellin-Barnes integrations. Therefore, the final result does not contain any hypergeometric functions and is given by \begin{equation} \begin{split} s_{{\CIRCLE}}^{(2)}&=\frac{e^{2\gamma_{{E}}\epsilon}}{\Gamma(1-4\epsilon)}\left(C_{{A}}^2\,\left(\frac{\Gamma (1+ \epsilon) \Gamma (1+2 \epsilon) \Gamma (1-2 \epsilon) \Gamma (1-3\epsilon)}{8 \epsilon^3\,(1-2 \epsilon)}-\frac{\Gamma^3(1+\epsilon) \Gamma^3(1-\epsilon)}{4 \epsilon^3\,\Gamma (1+2 \epsilon)}\right.\right.\\ &\left.\left.+\frac{\left(1-2\epsilon+4 \epsilon^2\right)\,\Gamma^2(1+2\epsilon) \Gamma^2(1- \epsilon)}{8 \epsilon^3 (1-2 \epsilon)^2 \Gamma (1+4 \epsilon)}\right)+C_{{A}}\,C_{{R}}\frac{\Gamma^3(1+\epsilon) \Gamma^3(1-\epsilon)}{2 \epsilon^3\,\Gamma (1+2 \epsilon)}\right) \; . \end{split} \end{equation} \subsection{Renormalized result} \noindent After summing the contributions from the previous subsections and renormalizing according to Eq.~(\ref{eq:SR}), we obtain the following perturbative expansion of the soft function for color octet production at threshold in Mellin space \begin{equation} S(L) = 1 + \frac{\alpha_s}{\pi} \, S^{(1)}(L) + \left(\frac{\alpha_s}{\pi}\right)^2 \, S^{(2)}(L) + {\cal O}(\alpha_s^3) \; , \end{equation} with \begin{equation} \begin{split} S^{(1)}(L)&= C_{{A}} (L+1)+C_{{R}} \left(2 L^2+\frac{\pi ^2}{12}\right) \; , \\ S^{(2)}(L) &= C_{{A}}^2 \left(\frac{17 L^2}{12}+\left(\frac{\zeta (3)}{2}+\frac{151}{36}-\frac{\pi ^2}{12}\right)L-\frac{5 \zeta (3)}{8}+\frac{13 \pi ^4}{2880}+\frac{\pi ^2}{24}+\frac{223}{54}\right)\\ &+C_{{R}}\;C_{{A}} \left(\frac{29 L^3}{9}+\left(\frac{103}{18}-\frac{\pi ^2}{6}\right) L^2+ \left(-\frac{7 \zeta (3)}{2}+\frac{101}{27}+\frac{\pi ^2}{12}\right)L\right.\\ &\left.-\frac{11 \zeta (3)}{72}-\frac{\pi ^4}{48}+\frac{139 \pi ^2}{864}+\frac{607}{324}\right)\\ &+C_{{R}}^2 \left(2 L^4+\frac{\pi ^2 L^2}{6}+\frac{\pi ^4}{288}\right)\\ &+C_{{A}}\;T_{{F}}\;n_f \left(-\frac{L^2}{3}-\frac{11 L}{9}-\frac{40}{27}\right)\\ &+C_{{R}}\;T_{{F}}\;n_f \left(-\frac{4 L^3}{9}-\frac{10 L^2}{9}-\frac{28 L}{27}+\frac{\zeta (3)}{18}-\frac{5 \pi ^2}{216}-\frac{41}{81}\right) \; . \end{split} \end{equation} In case the initial state is a color octet as well, i.e. $C_R = C_A$, this result further simplifies to become \begin{equation} \begin{split} S_{{A}}^{(1)}(L)&= C_{{A}} \left(2 L^2+L+\frac{\pi ^2}{12}+1\right) \; , \\ S_{{A}}^{(2)}(L) &= C_{{A}}^2 \left(2 L^4+\frac{29 L^3}{9}+\frac{257 L^2}{36}+\left(\frac{857}{108}-3 \zeta (3)\right)L-\frac{7 \zeta (3)}{9}-\frac{37 \pi ^4}{2880}+\frac{175 \pi ^2}{864}+\frac{1945}{324}\right)\\ &+C_{{A}}\;T_{{F}}\;n_f \left(-\frac{4 L^3}{9}-\frac{13 L^2}{9}-\frac{61 L}{27}+\frac{\zeta (3)}{18}-\frac{5 \pi ^2}{216}-\frac{161}{81}\right) \; . \end{split} \end{equation} Of course, the logarithmic terms proportional to $L^n$, $n>0$, in these expressions can be derived directly from the RGE Eq.~(\ref{eq:RGE}). The new results are thus at $L=0$. \section{Conclusions and outlook} \noindent We have presented the result for the next-to-next-to-leading order soft function for color octet production at threshold. The primary use of this result is the derivation of the constants in the threshold expansion of NNLO QCD cross section for heavy-flavor pair-production. This will be one of the topics covered in an upcoming publication. Our results may have further applications, for example in the determination of threshold expansions of NNLO cross sections for other massive particle (e.g. squarks, gluinos) production in pairs, or as single fundamental states (e.g. color octet scalars). Essential is only the combined color configuration of the final state. The results could also be generalized to arbitrary representations, since the most complicated part of the calculation was the evaluation of the integrals. We have refrained from such a generalization in this publication. Since the result for the bare soft function is exact in $d$-dimensions, it can be used as part of a calculation of the soft function at higher orders. \section*{Acknowledgments} \noindent This research was supported by the German Research Foundation (DFG) via the Sonderforschungsbereich/Transregio SFB/TR-9 ``Computational Particle Physics''. The work of M.C. was supported by the DFG Heisenberg programme.
\section{Introduction} There are a few source-open programs for 5- and 6-point reductions: {LoopTools/FF} ($n\le 5, rank\le 4$) -- T. Hahn \cite{Hahn:1998yk,vanOldenborgh:1991yc2}, {Golem95} -- T. Binoth et al. \cite{Binoth:2008uq}, {PJFry} ($n\le 5$, rank $\le 5$) -- V. Yundin et al. \cite{yundin-phd-2012--oai:export,Fleischer:2011zzutphys}. Some of these packages need in addition a {library of scalar functions}: 't Hooft, Veltman \cite{'tHooft:1978xw}, {LoopTools/FF}, {QCDloop/FF} -- K. Ellis and G. Zanderighi \cite{Ellis:2007qk,vanOldenborgh:1991yc2} or {OneLOop} with complex masses -- van Hameren \cite{vanHameren:2010cp}. In most cases, these packages suffice to calculate one loop processes, however, there are at least two reasons for improvements. First, which is always desirable, speed improvements. This is important when calculations are included into precise measurements using Monte Carlo methods. As already discussed in \cite{guniaUstron2013}, available methods are at the edge of applications at low energy calculations. Second, not all of them are able to fulfill high demands concerning accuracy at very specific kinematic points; an example has been shown in \cite{Gluza:2012yz}. We are working on independent calculations for tensor contractions as an alternative to the PJFry reductions \cite{pjfry:2011-no-url}, based on work by Davydychev-Tarasov-Fleischer-Jegerlehner-Riemann-Yundin (DTFJRY) \cite{Davydychev:1991vautphys}, \cite{Tarasov:1996br}, \cite{Fleischer:1999hq}, \cite{Fleischer:2010sq}, \cite{Fleischer:2011nt}. First numerical studies have been discussed in \cite{Fleischer:2012adutphys}. In this material we report on analytic results for rank 4 and comment on improvements concerning the OLEC package \cite{olec-project-09utphys}. \section{Contractions for the 5-point functions with rank $R=4$} We present here for the first time results of contracted tensors of rank $R=4$ \begin{eqnarray}} % \newcommand{\bea} {\begin{eqnarray*}\label{T54} I_5^{\mu \nu \lambda \rho}=I_5^{\mu\nu\lambda} \cdot Q_0^{\rho} -\sum_{s=1}^{5} I_4^{\mu\nu\lambda, s} \cdot Q_s^{\rho}, \end{eqnarray}} % \newcommand{\end{eqnarray}} {\end{eqnarray*} for lower rank results, see \cite{Fleischer:2010sq}, \cite{Fleischer:2011nt}. After contraction with chords $q$ (differences of external momenta), we get: \begin{eqnarray}} % \newcommand{\bea} {\begin{eqnarray*} q_{a,\mu}q_{b,\nu}q_{c,\lambda}q_{d,\rho}~I_5^{\mu \nu \lambda \rho} ~ = ~ CE_{4,abcd} ~ = ~ -\frac{1}{2} CE_{3,abc} ~ Y_d + C_{5,abcd} . \end{eqnarray}} % \newcommand{\end{eqnarray}} {\end{eqnarray*} Here: \begin{eqnarray}} % \newcommand{\bea} {\begin{eqnarray*} \label{Q6} Q_{s}^{\rho}&=&\sum_{i=1}^{5} q_i^{\rho} \frac{{{s}\choose i}_5}{\left( \right)_5},~~~ {s}=0, \ldots , 5. \end{eqnarray}} % \newcommand{\end{eqnarray}} {\end{eqnarray*} The first term $q_{a,\mu}q_{b,\nu}q_{c,\lambda}~I_5^{\mu \nu \lambda}$ is known \cite{Fleischer:2012adutphys}, the second term has to be determined: \begin{equation} C_{5,abcd} = - \sum_{s=1}^5 q_{a\mu} q_{b\nu} q_{c\lambda} I_4^{\mu\nu\lambda,s} \frac{1}{2}\left(\delta_{ds}-\delta_{5s}\right), \end{equation} and it becomes: \begin{eqnarray} && C_{5,abcd} = \displaystyle\frac{1}{16}\,\, \Big\{ G^5 +\delta_{ab}\delta_{ac} \delta_{ad} G^d -I_1^{5abc} -I_1^{5abd} -I_1^{5acd} -I_1^{5bcd} +I_1^{abcd} \nonumber \\ & & -J_3^{a5} -J_3^{b5} -J_3^{c5} -J_3^{d5} +\,R^{5ab} +\,R^{5ac} +\,R^{5bc} +\,R^{5da} +\,R^{5db} +\,R^{5dc} \nonumber \\ & & +\,\delta_{bc}\delta_{bd}\left(J_3^{ad}-J_3^{5d}\right) +\,\delta_{ac}\delta_{ad}\left(J_3^{bd}-J_3^{5d}\right) +\,\delta_{ab}\delta_{ad}\left(J_3^{cd}-J_3^{5d}\right) \nonumber \\ & & +\,\delta_{ab}\delta_{ac}\left(J_3^{dc}-J_3^{5c}\right) +\,\delta_{ab}\delta_{cd} \tilde{J}_3^{db} +\,\delta_{ad}\delta_{bc} \tilde{J}_3^{dc} +\,\delta_{ac}\delta_{bd} \tilde{J}_3^{dc} \nonumber \\ & & +\,\delta_{ab} \left(\tilde{J}_3^{5b}-R^{b5c}-R^{bd5}+R^{bdc}\right) +\,\delta_{ac} \left(\tilde{J}_3^{5c}-R^{c5b}-R^{cd5}+R^{cdb}\right) \nonumber \\ & & +\,\delta_{ad} \left(\tilde{J}_3^{d5}-R^{d5b}-R^{d5c}+R^{dbc}\right) +\,\delta_{bc} \left(\tilde{J}_3^{5c}-R^{c5a}-R^{cd5}+R^{cda}\right) \nonumber \\ & & +\,\delta_{bd} \left(\tilde{J}_3^{d5}-R^{d5a}-R^{d5c}+R^{dac}\right) +\,\delta_{cd} \left(\tilde{J}_3^{d5}-R^{d5a}-R^{d5b}+R^{dab}\right) \nonumber \\ & & +\,\delta_{ab} \delta_{ad} J_4^d Y_c +\,\delta_{ac} \delta_{ad} J_4^d Y_b +\,\delta_{bc} \delta_{bd} J_4^d Y_a +\,\delta_{ab} \left(R^{bd}-R^{b5}\right) Y_c \nonumber \\ & & +\,\delta_{ac} \left(R^{cd}-R^{c5}\right) Y_b +\,\delta_{ad} \left(R^{dc}-R^{d5}\right) Y_b +\,\delta_{ad} \left(R^{db}-R^{d5}\right) Y_c \nonumber \\ & & +\,\delta_{bc} \left(R^{cd}-R^{c5}\right) Y_a +\,\delta_{bd} \left(R^{dc}-R^{d5}\right) Y_a +\,\delta_{bd} \left(R^{da}-R^{d5}\right) Y_c \nonumber \\ & & +\,\delta_{cd} \left(R^{da}-R^{d5}\right) Y_b +\,\delta_{cd} \left(R^{db}-R^{d5}\right) Y_a +\,\left( I_4^d-I_4^5\right) Y_a Y_b Y_c \nonumber \\ & & +\left( I_3^{cd} -I_3^{5c} -I_3^{5d} +R^5 +\delta_{cd} R^d \right) Y_a Y_b \nonumber \\ & & +\left( I_3^{bd} -I_3^{5b} -I_3^{5d} +R^5 +\delta_{bd} R^d \right) Y_a Y_c \nonumber \\ & & +\left( I_3^{ad} -I_3^{5a} -I_3^{5d} +R^5 +\delta_{ad} R^d \right) Y_b Y_c \nonumber \\ & & +\left( I_2^{bcd} -I_2^{5bc} -I_2^{5bd} -I_2^{5cd} -J_4^5 +R^{5b} +R^{5c} +R^{5d} \right) Y_a \nonumber \\ & & +\left( I_2^{acd} -I_2^{5ac} -I_2^{5ad} -I_2^{5cd} -J_4^5 +R^{5a} +R^{5c} +R^{5d} \right) Y_b \nonumber \\ & & +\left( I_2^{abd} -I_2^{5ab} -I_2^{5ad} -I_2^{5bd} -J_4^5 +R^{5a} +R^{5b} +R^{5d} \right) Y_c \Big\} , \end{eqnarray} where we have introduced: \begin{equation} J_3^{st} \equiv \frac{1}{{s t \choose s t }_5} \left\{ -{s \choose s }_5 I_3^{[d+],st}+{t s\choose 0 s}_5 R^{ts} - \sum_{u=1}^5 {t s\choose u s}_5 R^{tsu} \right\}, \end{equation} \begin{equation} \tilde{J}_3^{st} \equiv \frac{1}{{s t \choose s t }_5} \left\{ {s \choose t }_5 I_3^{[d+],st}+{s t\choose 0 t}_5 R^{ts} - \sum_{u=1}^5 {s t\choose u t}_5 R^{tsu} \right\}, \end{equation} \begin{equation} G^{s} \equiv \frac{1}{{s \choose s }_5} \left\{ -2 { \choose }_5 R^{[d+],s} + {s\choose 0}_5 J_4^{s} - \sum_{t=1}^5 {s \choose t}_5 J_3^{ts} \right\}. \end{equation} $J_4^s$ and $R^{[d+],s}$ are given in Eqs. (2.24) and (2.44) of \cite{Fleischer:2010sq}, respectively. For further abbreviations see (2.24), (2.49), (2.9), (2.17), (2.34), (2.41) of \cite{Fleischer:2010sq}. There also $R^s, R^{st},R^{tsu}$ are defined, $Y_a = Y_{a5} - Y_{55}, Y_{ab} = -(q_a-q_b)^2+m_a^2+m_b^2$. Some numerical results for a 5-point function with rank $R=4$ are added in \cite{olec-project-09utphys}. For scalar functions we use the OneLOop package and compare the results with LoopTools/FF. For the considered kinematic points full agreement has been obtained. The kinematics is that of the process $e^-e^+\to \mu^-\mu^+ \gamma$. Using an MC generator, we have checked thousands of points up to rank three. All of them agreed between OLEC and LT/OneLOop. For ranks 3 and the new results presented here we made checks also against the public version of Golem95, available at \verb+http://golem.hepforge.org/95/+. At the OLEC webpage \cite{olec-project-09utphys}, we added two sets of files with the output. In set I, the kinematics is chosen such that 3-point functions hit the IR singularities, while in set II the 3-point functions are slightly off the IR singularities. The Golem95 results were different from LT, OneLOop and OLEC, starting already from rank 3.\footnote{The problem in Golem95 v.1.2.1 has been settled in the meantime for the set I with changelog 128 (11 Oct 2013), {\tt https://golem.hepforge.org/trac/changeset/128}.} For rank four we give below such an example: \begin{verbatim} p1s = 1.1163688400000000E-002 p2s = 2.6109999999999998E-007 p3s = 0.0000000000000000 p4s = 2.6109999999999998E-007 p5s = 1.1163688400000000E-002 s12 = -0.70858278190000001 s23 = -1.5343299000000002E-003 s34 = -0.12851860429999998 s45 = -0.61023937949999996 s15 = 0.92668942420000000 m1s = 1.1163688361676107E-002 m2s = 0.0000000000000000 m3s = 2.6112003932088364E-007 m4s = 2.6112003932088364E-007 m5s = 0.0000000000000000 \end{verbatim} \begin{verbatim} The R=4 contractions, a,b,c,d=3,3,3,3 OLEC: ( -48094.1074 54542318 , -47802.08746 5035322 ) LoopTools: ( -48094.1074 65 , -47802.08746 05 ) The R=4 contractions, a,b,c,d=3,3,3,4 OLEC: ( -18463.1204 24842149 , -23446.4704 12257226 ) LoopTools: ( -18463.1204 31 , -23446.4704 09 ) The R=4 contractions, a,b,c,d=3,3,3,5 OLEC: ( 0.0000000000000000 , 0.0000000000000000 ) LoopTools ( 0.0000000000000000 , 0.0000000000000000 ) \end{verbatim} The last result with d=5 is a virtue of the construction of chords where $q_5=0$. \section{OLEC package, Fortran code} The idea of external contractions has been implemented for the first time in the C++ code OLEC for tensors up to rank 3 \cite{Fleischer:2012adutphys}, and basic examples are given at \cite{olec-project-09utphys}. In the meantime, a Fortran code has been written with the aim of further optimization. The calculation of contracted tensor integrals consists of two basic steps. The first and most time consuming step is a preparation of building blocks for a calculation - basic scalar integrals $(\sim 50\%)$ and so-called signed minors $(\sim 25\%)$. We cannot make much about scalar integral libraries, unless new independent developments appear. For some recent efforts in this direction, see \cite{Guillet:2013mta}. However, the {\it calculation} of signed minors can be improved, both the cache system and and the computation algorithm. The second step is {\it performing} of the contractions $(\sim 25\%)$, which also can be improved. During the initialization procedure scalar integrals and signed minors are calculated and stored in RAM memory. For efficient usage of memory appropriate data structures are constructed. E.g. for 5-point kinematics we have in total only 30 different basic scalar integrals while for signed minors situation is more complicated because all minors with up to 4 scratched rows and columns are needed. The ``natural'' storage model for such objects in the form of multidimensional arrays is not an optimal solution. For example, for minors of rank 3 (3 rows and columns are excluded) one has an array with 46656 elements but according to the minors' symmetry properties only 210 of them are different and non zero. For the purpose of minimal memory usage a linear storage model is chosen. In general minors of a given rank are stored in the memory as one--dimensional arrays and access to them can be done according to the following pattern: $minor_{rank}[AddressTable[func(i,j,k...)]]$, as it was first done in \cite{pjfry:2011-no-url}. For a given minor indices $i,j,k,...$ a function $func$ based on bitwise operations calculates an address of an element from the constant table $AddressTable$. This table is the same for all minors and contains positions of values of minors in array $minor_{rank}$. Finally the cache contains a few hundred of double precision numbers for all building blocks like minors, scalar integrals and auxiliary functions. \\ In the new version of the library the computation algorithm for signed minors has also been changed. In practical computations of one-loop cross--sections usually at least all contractions up to rank 3 are needed (like in QED). It means that minors of all ranks are needed as well. The procedure is iterative, cache is filled first from minors of the highest rank 4 (with 8 indices) which are used further to calculate minors of rank 3. And so on. All loops connected with minor indices during this procedure are unrolled and the highest level of optimization is applied for compilation (these are algebraical manipulations for which -O3 optimization does not spoil double accuracy). Unrolling and usage of optimalization for algebraic part of the package allows to decrease extremely the matrix algebra computation time. \section{Summary and Outlook} In recent years, the strategy described has been developed in a bunch of papers for explicit {analytical and recursive treatment} of { heptagon, hexagon and pentagon tensor integrals} of rank $R$ in terms of pentagons and boxes of rank $R-1$. A systematic derivation of expressions which are explicitly {free of inverse Gram determinants} {$()_5$} until pentagons of rank $R=5$ has been worked out. The numerical {package OLEC for contracted tensor integrals in C++ and Fortran \cite{olec-project-09utphys,AFGRprep} are under development and tests. Our preliminary benchmarks show that the OLEC library in the present form is faster than 5-point tensor reductions implemented in LoopTools/FF by about an order of magnitude (tested up to rank 3). Work in progress includes a correct treatment of small Gram determinant cases for the reductions (expansion in small parameters \cite{pjfry:2011-no-url} or using hypergeometric representations \cite{Fleischer:2003rm}), adding rank 4 (this paper) and rank 5 contractions for 5-point functions, programming contracted tensors for 6- and 7- point functions. \section*{Acknowledgments} Work supported by European Initial Training Network LHCPHENOnet PITN-GA-2010-264564. \providecommand{\href}[2]{#2}\begingroup
\section{Introduction} The goal of the \textit{kinetic data structure framework}, which was first introduced by Basch, Guibas and Hershberger~\cite{basch_data_1999}, is to provide a set of data structures and algorithms that maintain attributes (properties) of points as they move. At essentially any moment, one may seek efficient answers to certain queries (\emph{e.g.}, what is the closest pair?) about these moving points. Taken together, such a set of data structures and algorithms is called a \textit{kinetic data structure} (KDS). Kinetic versions of many geometry problems have been studied extensively over the past 15 years, \emph{e.g.}, kinetic Delaunay triangulation~\cite{Albers_voronoidiagrams,DBLP:journals/dcg/Rubin13}, kinetic point-set embeddability~\cite{DBLP:conf/gd/RahmatiZ12}, kinetic Euclidean minimum spanning tree~\cite{DBLP:conf/iwoca/RahmatiZ11,basch_data_1999}, kinetic closest pair~\cite{Agarwal:2008:KDD:1435375.1435379,basch_data_1999}, kinetic convex hull~\cite{basch_data_1999,Alexandron:2007:KDD:1219156.1219201}, kinetic spanners~\cite{Abam:2010:SEK:1630166.1630284,Karavelas:2001:SKG:365411.365441}, and kinetic range searching~\cite{Agarwal:2003:IMP:846156.846166}. Let $P$ be a set of $n$ points in the plane, and denote the position of each point $p$ by $p=(p_x,p_y)$ in a Cartesian coordinate system. In the kinetic setting, we assume the points are moving continuously with known trajectories, which may be changed to new known trajectories at any time. Thus the point set $P$ will sometimes be denoted $P(t)$, and an element $p=(p_x,p_y)$ by $p(t)=(p_x(t),p_y(t))$. For ease of notation, we denote the coordinate functions of a point $p_i(t)$ by $x_i(t)$ and $y_i(t)$. Throughout the paper we assume that all coordinate functions are polynomial functions of maximum degree bounded by some constant $s$. In this paper, we consider several fundamental proximity problems, which we define in more detail below. We design KDS's with better performance for some these problems, and we provide the first kinetic results for others. We introduce a simple method that underlies all these results. We briefly describe the approach in Section~\ref{sec:ourApproach}. Finding the nearest point in $P$ to a query point is called the \textit{nearest neighbor search} problem (or the \textit{post office} problem), and is a well-studied proximity problem. The \textit{all nearest neighbors} problem, a variant of the nearest neighbor search problem, is to find the nearest neighbor $q\in P$ to each point $p\in P$. The directed graph constructed by connecting each point $p$ to its nearest neighbor $q$ with a directed edge $\overrightarrow{pq}$ is called the \textit{nearest neighbor graph} (NNG). The \textit{closest pair} problem is to find a pair of points in $P$ whose separation distance is minimum; the endpoints of the edge(s) with minimum length in the nearest neighbor graph give the closest pair. For the set $P$, there exists a complete, edge-weighted graph $G(V,E)$ where $V=P$ and the weight of each edge is the distance between its two endpoints in the Euclidean metric. A \emph{Euclidean minimum spanning tree} (EMST) of $G$ is a connected subgraph of $G$ such that the sum of the edge weights in the Euclidean metric is minimum possible. The Yao graph~\cite{DBLP:journals/siamcomp/Yao82} and the Semi-Yao graph (or theta graph)~\cite{Clarkson:1987:AAS:28395.28402,Keil:1988:ACE:61764.61787} of a point set $P$ are two well-studied sparse proximity graphs. Both of these graphs are constructed in the following way. At each point $p\in P$, the plane is partitioned into $z$ wedges $W_0(p),...,W_{z-1}(p)$ with equal apex angles $2\pi/z$. Then for each wedge $W_i(p)$, $0\leq i\leq z-1$, the apex $p$ is connected to a particular point $q\in P\cap W_i(p)$. In the Yao graph, the point $q$ is the point in $P\cap W_i(p)$ with the minimum Euclidean distance to $p$; in the Semi-Yao graph, the point $q$ is the point in $P\cap W_i(p)$ with minimum length projection on the bisector of $W_i(p)$. From now on, unless stated otherwise, when we consider the Yao graph or the Semi-Yao graph, we assume $z=6$. With these definitions in mind, in Section~\ref{sec:ourApproach} we describe our approach. Before we can describe the main contributions and the kinetic results we obtain using our simple method, we need to review both the terminology of the KDS framework, which is described in Section~\ref{sec:KDSframework}, as well as the previous results, which are described in Section~\ref{sec:relatedwork}. \subsection{Our Approach}\label{sec:ourApproach} We provide a new, simple, and deterministic method for maintenance of all the nearest neighbors, the closest pair, the Euclidean minimum spanning tree (EMST or $L_2$-MST), the Yao graph, and the Semi-Yao graph. In particular, to the best of our knowledge our KDS's for these graphs are the first KDS's. The heart of our approach is to define, compute, and kinetically maintain supergraphs for the Yao graph and the Semi-Yao graph. Then we take advantage of the fact that (as we explain later) these graphs are themselves supergraphs of the EMST and the nearest neighbor graph, respectively. We define a supergraph for the Yao graph as follows. We partition a unit disk into six \textquotedblleft pieces of pie\textquotedblright~$\sigma_0,\sigma_2,...,\sigma_5$ with equal angles such that all $\sigma_l$, $l=0,...,5$, share a point at the center of the disk (see Figure~\ref{fig:disk_hexagon}(a)). Each piece of pie $\sigma_l$ is a convex shape. For each $\sigma_l$ we construct a triangulation as follows. Using the fact that, for a set $P$ of points, a Delaunay triangulation can be defined based on any convex shape~\cite{Chew:1985:VDB:323233.323264,Drysdale:1990:PAC:320176.320194}, we define a Delaunay triangulation $DT_l$ based on each piece of pie $\sigma_l$. The union of all of these Delaunay triangulations $DT_l$, $l=0,...,5$, which we call the \textit{Pie Delaunay graph}, is a supergraph of the Yao graph. Since the Yao graph, for $z\geq 6$, is guaranteed to contain the EMST, the Pie Delaunay graph contains the EMST. We define a supergraph for the Semi-Yao graph as follows. We partition a hexagon into six equilateral triangles $\Delta_0,\Delta_2,...,\Delta_5$ (see Figure~\ref{fig:disk_hexagon}(b)), and for each equilateral triangle $\Delta_l$ we define a Delaunay triangulation $DT_l$. The union of all of these Delaunay triangulations $DT_l$, $l=0,...,5$, which we call the \textit{Equilateral Delaunay graph}, is a supergraph of the Semi-Yao graph. We prove that the Semi-Yao graph is a supergraph of the nearest neighbor graph, which implies that the Equilateral Delaunay graph is a supergraph of the nearest neighbor graph. \begin{figure}[t!] \begin{center} \includegraphics[scale=1]{disk_hexagon.eps} \end{center} \caption{(a) Partitioning a unit disk into six pieces of pie. (b) Partitioning a hexagon into six equilateral triangles.} \label{fig:disk_hexagon} \end{figure} In the case that the Delaunay triangulation $DT_l$ is based on a piece of pie, the triangulation can easily be maintained over time. This leads us to a kinetic data structure for the union of the $DT_l$'s, \emph{i.e.}, the Pie Delaunay graph. Then we show how to use this sparse graph over time to give kinetic data structures for maintenance of the Yao graph and the EMST. Similarly, in the case that each $DT_l$ arises from an equilateral triangle, we obtain a kinetic data structure for the Equilateral Delaunay graph. Using the kinetic Equilateral Delaunay graph we give kinetic data structures for maintenance of the Semi-Yao graph, all the nearest neighbors, and the closest pair. \subsection{KDS Framework}\label{sec:KDSframework} Basch, Guibas and Hershberger~\cite{basch_data_1999} first introduced the \textit{kinetic data structure} (KDS) framework to maintain \textit{attributes}, \emph{e.g.}, the closest pair, of a set of $n$ moving points. This approach has been used extensively to model motion. They introduced four standard criteria to evaluate the performance of a KDS: \textit{efficiency}, \textit{responsiveness}, \textit{compactness}, and \textit{locality}. In the KDS framework, one defines a set of \textit{certificates} that together attest that the desired attribute holds throughout intervals of time between certain \textit{events}, described below. A certificate is a Boolean function of time, and it may have a failure time $t$. The certificate is valid until time $t$. A \textit{priority queue} of the failure times of the certificates is used to track the first time after the current time $t_c$ that a certificate will become invalid. When the failure time of a certificate with highest priority in the queue is equal to the current time $t_c$, the certificate fails, and we say that an \textit{event} occurs. Then we invoke an update mechanism to replace the certificates that become invalid with new valid ones, and apply the necessary changes to the data structures. Now we describe the four performance criteria: \begin{itemize} \item[1.] \textit{Responsiveness:} One of the most important KDS performance criteria is the processing time to handle an event. The KDS is \textit{responsive} if the response time of the update mechanism for an event is $O(\log^c n)$; $n$ is the number of points and $c$ is a constant. \item[2.] \textit{Compactness:} The compactness criterion concerns the total number of certificates stored in the KDS at any given time. If the number of certificates is $O(n\log^c n)$, the KDS is \textit{compact}. \item[3.] \textit{Locality:} If the number of certificates associated with a particular point is $O(\log^c n)$, the KDS is \textit{local}. Satisfaction of this criterion ensures that, for any point, if it changes its trajectory it participates in a small number of certificates, and therefore, only a small number of changes are needed in the KDS. \item[4.] \textit{Efficiency:} To count the number of events over time we make the assumption that the trajectories of the points are polynomial functions of bounded degree $s$. The efficiency of a KDS concerns the number of events in the KDS over time. To analyse the efficiency of a KDS one identifies two types of events. Some events do not necessarily change the attribute of interest (also called the \textit{desired} attribute) and may only change some internal data structures. Such events are called \textit{internal events}. Those events that change the attribute of interest are called \textit{external events}. If the ratio between the number of internal events and the number of external events is $O(\log^c n)$, the KDS is \textit{efficient}. The efficiency of a KDS can be viewed as measuring the fraction of events that are due to overhead. \end{itemize} \subsection{Other Related Work}\label{sec:relatedwork} \paragraph{Kinetic All Nearest Neighbors.} The nearest neighbor graph is a subgraph of the Delaunay triangulation and the Euclidean minimum spanning tree. Thus by maintaining either one of these supergraphs over time, all the nearest neighbors can also be maintained. In particular, by using the kinetic Delaunay triangulation~\cite{Albers_voronoidiagrams} or the kinetic Euclidean minimum spanning tree~\cite{DBLP:conf/iwoca/RahmatiZ11}, together with a basic tool in the KDS framework called the kinetic tournament tree~\cite{basch_data_1999}, we can maintain all the nearest neighbors over time. For both these two approaches, the number of internal events is nearly cubic in $n=|P|$. Since the number of external events for all the nearest neighbors is nearly quadratic, neither of these two approaches will give an efficient KDS as defined above. Agarwal, Kaplan, and Sharir~\cite{Agarwal:2008:KDD:1435375.1435379} presented the first efficient KDS for maintenance of all the nearest neighbors. For a set of points in the plane, their kinetic algorithm uses a $2$-dimensional \textit{range tree}. To bound the number of events in order to obtain an efficient KDS, they implemented the range tree by randomized search trees (\textit{treaps}). Their randomized kinetic approach uses $O(n\log^2 n)$ space and processes $O(n^2\beta_{2s+2}^2(n)\log^3 n)$ events, where $\beta_{s}(n)$ is an extremely slow-growing function. The expected time to process all events is $O(n^2\beta_{2s+2}^2(n)\log^4 n)$. In terms of the KDS performance criteria, their KDS is \textit{efficient}, \textit{responsive} (in an amortized sense), and \textit{compact}, but it is not \textit{local}. \paragraph{Kinetic Closest Pair.} For a set of points moving in $\mathbb{R}^2$, Basch, Guibas, and Hershberger~\cite{Basch:1997:DSM:314161.314435} presented a KDS to maintain the closest pair. Their kinetic algorithm uses $O(n)$ space and processes $O(n^2\beta_{2s+2}(n)\log n)$ events, each in $O(\log^2 n)$ time; their KDS is responsive, efficient, compact, and local. Basch, Guibas, and Zhang~\cite{Basch:1997:PPM:262839.262998} used a multidimensional range tree to maintain the closest pair. Their KDS uses $O(n\log n)$ space and processes $O(n^2\beta_{2s+2}(n)\log n)$ events, each in worst-case time $O(\log^2 n)$. Their KDS, which can be used for higher dimensions as well, is responsive, efficient, compact, and local. The same KDS with the same complexities as~\cite{Basch:1997:PPM:262839.262998} was independently presented by Agarwal, Kaplan, and Sharir~\cite{Agarwal:2008:KDD:1435375.1435379}; the KDS by Agarwal~\emph{et~al.}~supports point insertions and deletions. \paragraph{Kinetic EMST.} Fu and Lee~\cite{Fu:1991:MST:115128.115142} proposed the first kinetic algorithm for maintenance of an EMST on a set of $n$ moving points. Their algorithm uses $O(sn^4\log n)$ preprocessing time and $O(m)$ space, where $m$ is the maximum possible number of changes in the EMST from time $t=0$ to $t=\infty$. At any given time, the algorithm constructs the EMST in linear time. Agarwal~\emph{et~al.}~\cite{Agarwal:1998:PKM:795664.796402} proposed a sophisticated algorithm for a restricted kinetic version of the EMST over a graph where the distance between each pair of points in the graph is defined by a linear function of time. The processing time for each combinatorial change in the EMST is $O(n^{2\over 3}\log^{4\over 3} n)$; the bound reduces to $O(n^{1\over 2}\log^{3\over 2} n)$ for planar graphs. Their data structure does not explicitly bound the number of changes, but a bound of $O(n^4)$ is easily seen. For any $\epsilon>0$, Basch, Guibas, and Zhang~\cite{Basch:1997:PPM:262839.262998} presented a KDS for a $(1+\epsilon)$-EMST whose total weight is within a factor of $(1+\epsilon)$ of the total weight of an exact EMST. For a set of points in the plane, their KDS uses $O(\epsilon^{-1\over 2}n\log n)$ space and $O(\epsilon^{-1\over 2}n\log n)$ preprocessing time, and processes $O(\epsilon^{-1}n^3)$ events, each in $O(\log^2 n)$ time; their KDS works for higher dimensions. They claim that their structure can be used to maintain the minimum spanning tree in the $L_1$ and $L_\infty$ metrics. Rahmati and Zarei~\cite{DBLP:conf/iwoca/RahmatiZ11} improved the previous result by Fu and Lee~\cite{Fu:1991:MST:115128.115142}. In particular, Rahmati and Zarei presented an exact kinetic algorithm for maintenance of the EMST on a set of $n$ moving points in $\mathbb{R}^2$. In $O(n\log n)$ preprocessing time and $O(n)$ space, they build a KDS that processes $O(n^4)$ events, each in $O(\log^2 n)$ time. Their KDS uses the method of Guibas~\emph{et~al.}~\cite{conf/wg/GuibasM91} to track changes to the Delaunay triangulation, which is a supergraph of the EMST~\cite{O'Rourke:1998:CGC:521378}. Whenever two edges of the Delaunay triangulation swap their length order, their kinetic algorithm makes the required changes to the EMST. In fact, under an assumption we will explain soon, the number of changes in their algorithm is within a linear factor of the number of changes to the Delaunay triangulation~\cite{conf/wg/GuibasM91}. Rubin~\cite{DBLP:journals/dcg/Rubin13} proved that the number of discrete changes to the Delaunay triangulation is $O(n^{2+\epsilon})$, for any $\epsilon >0$, under the assumptions that ($i$) any four points can be co-circular at most twice, and ($ii$) either no ordered triple of points can be collinear more than once, or no triple of points can be collinear more than twice. Under these assumptions, the kinetic algorithm of Rahmati and Zarei processes $O(n^{3+\epsilon})$ events, which is within a linear factor of the number of changes to the Delaunay triangulation The kinetic approach by Rahmati and Zarei~\cite{DBLP:conf/iwoca/RahmatiZ11} can maintain the minimum spanning tree of a planar graph whose edge weights are polynomial functions of bounded degree; the processing time of each event is $O(\log^2 n)$. \paragraph{Kinetic Yao graph and Semi-Yao graph.} To the best of our knowledge there are no previous kinetic data structures for maintenance of the Semi-Yao graph and the Yao graph on a set of moving points. \subsection{Main Contributions and Results}\label{sec:mainResults} Based on the approach we described in Section~\ref{sec:ourApproach}, we obtain the results below. \paragraph{Kinetic All Nearest Neighbors and the Closest Pair.} We give a simple and deterministic kinetic algorithm for maintenance of all the nearest neighbors of a set $P$ of $n$ moving points in the plane, where the trajectory of each point is a polynomial function of at most constant degree $s$. Our KDS uses linear space and $O(n\log n)$ preprocessing time to construct the kinetic data structure, and processes $O(n^2\beta^2_{2s+2}(n)\log n)$ events with total processing time $O(n^2\beta^2_{2s+2}(n)\log^2 n)$. We also show how to maintain the closest pair over time. Our KDS for maintenance of the closest pair has the same complexities as the KDS for all the nearest neighbors; in particular, it uses $O(n)$ space and processes $O(n^2\beta^2_{2s+2}(n)\log n)$ events for a total processing time of $O(n^2\beta^2_{2s+2}(n)\log^2 n)$. Our KDS for the all nearest neighbors and the closest pair problems is efficient, responsive in an amortized sense, and compact. The compactness of the KDS implies that our KDS is local in an amortized sense. In particular, on average each point in our KDS participates in $O(1)$ certificates. Our \textit{deterministic} algorithm for maintenance of all the nearest neighbors in $\mathbb{R}^2$ is simpler and more efficient than the \textit{randomized} kinetic algorithm by Agarwal, Kaplan, and Sharir~\cite{Agarwal:2008:KDD:1435375.1435379}: both of these kinetic algorithms need a priority queue containing all certificates of the KDS (our priority queue uses linear space, but their priority queue uses $O(n\log^2 n)$ space). Our KDS uses a graph data structure for the Equilateral Delaunay graph and a constant number of tournament trees for each point, but their KDS uses a \textit{$2$d range tree} implemented by randomized search trees (treaps), a constant number of sorted lists, and in fact it maintains $O(\log^2 n)$ tournament trees for each point. In particular, \begin{itemize} \item we perform one-dimensional range searching, as opposed to the two-dimensional range searching of their work; \item the sparse graph representation allows us to obtain a linear space KDS, which improves the space complexity $O(n\log^2 n)$ of their KDS. Their KDS uses a $2$d range tree implemented by randomized search trees that in effect maintain a supergraph of the nearest neighbor graph with $O(n\log^2 n)$ candidate edges; \item in our kinetic algorithm, the number of changes to the Equilateral Delaunay graph when the points are moving is $O(n^2\beta_{2s+2}(n))$; this leads us to have total processing time $O(n^2\beta^2_{2s+2}(n)\log^2 n)$, which is an improvement of the total expected processing time $O(n^2\beta^2_{2s+2}(n)\log^4 n)$ of their randomized algorithm; \item on average each point in our KDS participates in a constant number of certificates, but each point in their KDS participates in $O(\log^2 n)$ certificates. \end{itemize} The certificates of our KDS for maintenance of the closest pair are simpler than the certificates of the previous kinetic algorithms by Basch, Guibas, and Hershberger (SODA'97)~\cite{Basch:1997:DSM:314161.314435}, Basch, Guibas, and Zhang (SoCG'97)~\cite{Basch:1997:PPM:262839.262998}, and Agarwal, Kaplan, and Sharir (TALG 2008)~\cite{Agarwal:2008:KDD:1435375.1435379}. \paragraph{Kinetic Yao Graph and Semi-Yao Graph.} We give the first kinetic data structures for maintenance of two well-studied sparse graphs, the Semi-Yao graph and the Yao graph. Our KDS processes $O(n^2\beta_{2s+2}(n))$ (resp. $O(n^3\beta_{2s+2}^2(n)\log n)$) events to maintain the Semi-Yao graph (resp. the Yao graph); each event can be processed in time $O(\log n)$ in an amortized sense. \paragraph{Kinetic EMST.} Our KDS for maintenance of the EMST uses $O(n)$ space, takes $O(n\log n)$ preprocessing time, and processes $O(n^3\beta^2_{2s+2}(n)\log n)$ events. The total cost to process all these events is $O(n^3\beta^2_{2s+2}(n)\log^2 n)$. Our KDS is responsive in an amortized sense, compact, and local on average. Our EMST KDS improves on the previous EMST KDS by Rahmati and Zarei~\cite{DBLP:conf/iwoca/RahmatiZ11}. Our KDS processes $O(n^3\beta^2_{2s+2}(n)\log n)$ events, whereas the KDS by Rahmati and Zarei processes $O(n^4)$ events. Table~\ref{table:RelatedWork} summarizes our results and compares them with the previous results. \begin{table}[t] \small \centering \begin{tabular}{ | c | p{2cm} || c | c | c | p{1.5cm} |} \cline{2-6} \multicolumn{1}{ c| }{} & problem & space & total number of events & proc. time per event& locality\\ \cline{1-6} \multirow{2}{*}{Basch~\emph{et~al.}~\cite{basch_data_1999}} & \multirow{2}{*}{closest pair} & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log^2 n)$ ~~~~[in wrc]} & $O(\log n)$ in wrc\\ \hline\hline \multirow{4}{*}{Basch~\emph{et~al.}~\cite{Basch:1997:PPM:262839.262998}} & \multirow{2}{*}{closest pair}& \multirow{2}{*}{$O(n\log n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log^2 n)$ ~~~~[in wrc]} & $O(\log n)$ in wrc\\ \cline{2-6} & \multirow{2}{*}{$(1+\epsilon)$-EMST}& \multirow{2}{*}{$O(\epsilon^{-1\over 2}n\log n)$} & \multirow{2}{*}{$O(\epsilon^{-1}n^3)$} & \multirow{2}{*}{$O(\log^2 n)$ ~~~~[in wrc]} & $O(\log n)$ in wrc\\ \hline\hline \multirow{4}{*}{Agarwal \emph{et~al.}~\cite{Agarwal:2008:KDD:1435375.1435379}} & \multirow{2}{*}{closest pair}& \multirow{2}{*}{$O(n\log n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log^2 n)$ ~~~~[in wrc]} & $O(\log n)$ in wrc\\ \cline{2-6} & all nearest neighbors& \multirow{2}{*}{$O(n\log^2 n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}^2(n)\log^3 n)$} & \multirow{2}{*}{$O(\log n)$ ~~~~[in amr]} & $O(\log^2 n)$ on avg\\ \hline\hline \multirow{2}{*}{Rahmati~\emph{et~al.}~\cite{DBLP:conf/iwoca/RahmatiZ11}} & \multirow{2}{*}{EMST} & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^4)$} & \multirow{2}{*}{$O(\log^2 n)$ ~~~[in wrc]} & $O(1)$ ~on avg\\ \hline\hline \multirow{10}{*}{{{\color{Mahogany}This Paper}}} & \multirow{2}{*}{closest pair} & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log n)$ ~~~~[in amr]} & $O(1)$ on ~avg\\ \cline{2-6} & all nearest neighbors& \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log n)$ ~~~~[in amr]} & $O(1)$ on ~avg \\ \cline{2-6} & \multirow{2}{*}{EMST} & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^3\beta^2_{2s+2}(n)\log n)$} & \multirow{2}{*}{ $O(\log n)$ ~~~~[in amr]} & $O(1)$ on ~avg \\ \cline{2-6} & \multirow{2}{*}{Yao graph} & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^3\beta^2_{2s+2}(n)\log n)$} & \multirow{2}{*}{$O(\log n)$ ~~~~[in amr]} & $O(1)$ on ~avg\\ \cline{2-6} & Semi-Yao graph & \multirow{2}{*}{$O(n)$} & \multirow{2}{*}{$O(n^2\beta_{2s+2}(n))$} & \multirow{2}{*}{$O(\log n)$ ~~~~[in amr]} & $O(1)$ on ~avg\\\hline \end{tabular} \vspace{+10pt} \caption{The comparison between our KDS's and the previous KDS's, for a set of $n$ points in the plane. The abbreviations amr, wrc, and avg stand for amortized, worst-case, and average, respectively.} \label{table:RelatedWork} \vspace{-20pt} \end{table} \subsection{Organization} As necessary background for our work, Section~\ref{sec:preliminary} reviews a basic tool, the \textit{kinetic tournament tree}, which is used in the kinetic data structure framework. Section~\ref{sec:ANN_CP} is organized as follows: Subsection~\ref{sec:ANN_CP_construction} gives the new method for computing all the nearest neighbors and the closest pair. In particular, it introduces our two new sparse graphs, the \textit{Semi-Yao graph} and the \textit{Equilateral Delaunay graph} (in fact we will show these graphs are the same). In Subsection~\ref{sec:kinetic_EDG}, we make a kinetic version of the Equilateral Delaunay graph, and then in Subsections~\ref{sec:kinetic_ANN} and~\ref{sec:kinetic_CP}, we show how to use it to maintain all the nearest neighbors and the closest pair. The organization of Section~\ref{sec:YG_EMST} is similar to that of Section~\ref{sec:ANN_CP}. Using a new sparse graph, which we call the Pie Delaunay graph, we provide our new method for constructing the Yao graph and the EMST in Subsection~\ref{sec:YG_EMST_construction}. Subsection~\ref{sec:kinetic_PDG} gives a KDS for maintenance of the Pie Delaunay graph, and Subsections~\ref{sec:kinetic:YG} and \ref{sec:kinetic_EMST} use this KDS to maintain the Yao graph and the EMST. Section~\ref{sec:conclusion} discusses the extensions of the presented kinetic data structures to higher dimensions and gives some open problems for continuing this research direction. \section{Preliminaries}\label{sec:preliminary} Let ${\cal O}=\{o_1,o_2,...,o_n\}$ be a set of $n$ moving objects in the plane, where the $y$-coordinate $y_i(t)$ of each object $o_i$ is a continuous function of time. Assuming $y_i(t)$ is a polynomial function of at most constant degree $s$, it follows from Theorem~\ref{the:totallyDFcomplexity} below that the number of all changes for the lowest object with respect to the $y$-axis, among the set of objects ${\cal O}$, is $\lambda_s(n)$. \begin{theorem}\label{the:totallyDFcomplexity}{\tt \cite{Agarwal:1995:DSG:868483}} The length of the lower envelope of $n$ totally-defined, continuous, univariate functions, such that each pair of them intersects at most $s$ times, is at most $\lambda_s(n)$. \end{theorem} Note that Theorem~\ref{the:totallyDFcomplexity} holds for totally-defined functions; there exists a similar result for partially-defined functions: \begin{theorem}\label{the:partiallyDFcomplexity}{\tt \cite{Agarwal:1995:DSG:868483}} The length of the lower envelope of $n$ partially-defined, continuous, univariate functions, such that each pair of them intersects at most $s$ times, is at most $\lambda_{s+2}(n)$. \end{theorem} Here, $\lambda_s(n)=n\beta_s(n)$ is the maximum length of Davenport-Schinzel sequences of order $s$ on $n$ symbols, and $\beta_s(n)$ is an extremely slow-growing function. In particular, \[ \lambda_s(n) = \begin{cases} {n}, & \text{for $s=1$ };\\ {2n-1}, & \text{for $s=2$ };\\ {2n\alpha(n)+O(n)}, & \text{for $s=3$ };\\ {\Theta(n2^{\alpha(n)})}, & \text{for $s=4$ };\\ {\Theta(n\alpha(n)2^{\alpha(n)})}, & \text{for $s=5$ };\\ {n2^{(1+o(1))\alpha^t(n)/t!}}, & \text{for $s\geq 6$}; \end{cases} \] here $t={\lfloor {(s-2)/2}\rfloor}$ and $\alpha(n)$ denotes the inverse Ackermann function~\cite{Pettie:2013:SBD:2493132.2462390}. For maintenance of the lowest object with respect to the $y$-axis among the set of moving objects $\cal O$ over time, we use a basic (kinetic) data structure called a \textit{kinetic tournament tree}~\cite{basch_data_1999,Agarwal:2008:KDD:1435375.1435379}. A kinetic tournament tree is a balanced binary tree $T$ such that the objects are stored at the leaves of the tree $T$ in an arbitrary order, and each internal node $v$ of the tree maintains the lowest object between its two children. In more detail, denote by $T_v$ the subtree rooted at internal node $v$ and denote by $P_v$ the set of objects stored at the leaves of $T_v$. The object stored at $v$ in the tournament tree is the lowest object among all the objects in $P_v$; this object is called the \textit{winner} of the subtree $T_v$. For each internal node $v$ of the tournament tree we define a \textit{certificate} to assert whether the left-winner (winner of the left subtree) or the right-winner (winner of the right subtree) is the winner for $v$. The failure time of the certificate corresponding to the internal node $v$ is the time when the winner at $v$ changes. All of the certificates together are stored in a \textit{priority queue}, with the failure times as the keys, to track the next time after the current time that a certificate will become invalid. When the certificate corresponding to an internal node $v$ fails, it may change some winners on the path from the parent of $v$ to the root. In some cases the winner of a node $v'$ on the path does not change, but the failure time corresponding to the certificate of the node $v'$ may change. Therefore, we must update the failure times of the certificates of the nodes on the path from the parent of $v$ to the root, and then we must replace the invalid certificates with new valid ones in the priority queue; this takes $O(\log^2 n)$ time, which implies that the KDS is \textit{responsive}. The number of internal events for all the internal nodes is $\sum_v\lambda_s(|P_v|)= O(\lambda_s(n)\log n)$. Since the number of external events, that is the number of changes to the root of the tournament tree, is $\lambda_s(n)$, the KDS is \textit{efficient}. The tournament tree uses linear space, which implies the KDS is \textit{compact}. Each object participates in $O(\log n)$ certificates, which means the KDS is \textit{local}. It is convenient for our purpose to make the tournament tree dynamic, to support point insertions and deletions; the dynamic version of the kinetic tournament tree is called a \textit{dynamic and kinetic tournament tree}. This dynamic and kinetic tournament tree can be implemented using a \textit{weight-balanced (BB($\alpha$)) tree}~\cite{NJRE:10.1137/0202005,Mehlhorn:1984}; see the construction of a dynamic and kinetic tournament tree in~\cite{Agarwal:2008:KDD:1435375.1435379}. Consider a sequence of $m$ insertions and deletions into a dynamic and kinetic tournament tree where the maximum size tree at any time is $n$ (assuming $m\geq n$). The following theorem gives the construction time and the processing time of a dynamic and kinetic tournament tree. \begin{theorem}\label{the:DynamicKineticTT}{\tt \cite{Agarwal:2008:KDD:1435375.1435379}} A dynamic and kinetic tournament tree on $n$ elements can be constructed in $O(n)$ time. The tournament tree generates at most $O(m\beta_{s+2}(n)\log n)$ events, for a total cost of $O(m\beta_{s+2}(n)\log^2 n)$. Processing an event takes $O(\log^2 n)$ time. \end{theorem} \section{All Nearest Neighbors and Closest Pair}\label{sec:ANN_CP} In this section we provide a sparse graph representation and show a new construction of the nearest neighbor graph. First, we introduce two new supergraphs of the nearest neighbor graph, namely the \textit{Semi-Yao graph} and the\textit{ Equilateral Delaunay graph} (EDG), and then we show that these graphs are in fact the same. Next, we show how to maintain the Equilateral Delaunay graph for moving points, and then we give simple KDS's for maintenance of all the nearest neighbors and the closest pair. \subsection{New Method for Computing All Nearest Neighbors and Closest Pair}\label{sec:ANN_CP_construction} Partition the plane into six \textit{wedges} (cones) $W_0,...,W_{5}$, each of angle $\pi/3$ with common apex at the origin $o$. For $0\leq l \leq 5$, let $W_l$ span the angular range $[(2l-1)\pi/6, (2l+1)\pi/6)$. Denote by $b_l$ the unit vector in the direction of the bisector ray of $W_l$. Let $W_l(p_i)$ denote the translate of wedge $W_l$ that moves the apex to point $p_i$, and let ${\cal V}_l(p_i)$ denote the intersection of $P$ with wedge $W_l(p_i)$: ${\cal V}_l(p_i)=P\cap W_l(p_i)$. Denote by $b_l(p_i)$ the unit vector emanating from $p_i$ in the direction of the bisector ray of $W_l(p_i)$; see Figure~\ref{fig:projection}(a). Observe that, in Figure~\ref{fig:projection}(a), since $p_i$ is the closest point to $p_j$, there are no other points of $P$ in the interior of the disc. Let $d(p_i,p_j)$ denote the distance between points $p_i$ and $p_j$. \begin{figure}[t!] \centering \includegraphics[scale=1]{projection.eps} \caption{ (a) Projection of the point $p_j$ to the bisector $b_0(p_i)$ of the wedge $W_0(p_i)$. (b) In-edges and out-edges of $p_j$.} \label{fig:projection} \end{figure} The following straightforward lemma is key for obtaining our kinetic data structure for the \textit{all nearest neighbors} and the \textit{closest pair} problems. Consider $p_j\in P$, and let $p_i$ denote the point of $P$ closest to $p_j$ and distinct from $p_j$. Let $W_l(p_i)$ denote the wedge of $p_i$ that contains $p_j$, and denote by $\hat{p}_j$ the projection of $p_j$ to the bisector $b_l(p_i)$ (see Figure~\ref{fig:projection}(a)). \begin{lemma}\label{the:CPidea}{\tt \cite{Agarwal:2008:KDD:1435375.1435379,basch_data_1999}} Point $p_j$ has the minimum length projection to $b_l(p_i)$, where the minimum is taken over ${\cal V}_l(p_i)$. That is, \begin{equation}\label{eq:MinCoordinate} d(\hat{p}_j,p_i)=\min\{d(\hat{p}_k,p_i) | p_k\in {\cal V}_l(p_i)\}. \end{equation} \end{lemma} Thus, Lemma~\ref{the:CPidea} gives a necessary condition for $p_i$ to be the nearest neighbor to $p_j$. We now use this lemma to define a super-graph of the nearest neighbor graph of $P$. To find the nearest neighbor for each point $p_j\in P$, we seek a set of candidate points ${\cal C}(p_j)=\{p_i|~p_i~and~p_j~satisfy~Equation~(1)\}$. From now on, when we say $p_j$ has the minimum $b_l$-coordinate inside the wedge $W_l(p_i)$, we mean that $p_j$ and $p_i$ satisfy Equation~(\ref{eq:MinCoordinate}). By connecting each point $p_i\in P$ to a point $p_j\in {\cal V}_l(p_i)$ with a directed edge $\overrightarrow{p_jp_i}$ from $p_j$ to $p_i$ whenever $p_j$ is the point with the minimum $b_l$-coordinate, among all the points in ${\cal V}_l(p_i)$, we obtain what we call the \textit{Semi-Yao graph} (SYG) of $P$~\footnote{This graph is called the $\theta_6$-graph in~\cite{DBLP:journals/dcg/KeilG92}, but we prefer to call it the Semi-Yao graph instead of the $\theta_6$-graph, because of its close relationship to the Yao graph~\cite{DBLP:journals/siamcomp/Yao82}}. The edge $\overrightarrow{p_jp_i}$ is called an \textit{in-edge} for $p_i$ and it is called an \textit{out-edge} for $p_j$. Each point in the Semi-Yao graph has at most six in-edges and has a set of out-edges; Figure~\ref{fig:projection}(b) depicts the in-edges and the out-edges of the point $p_j$. Denote by $S_{out}(p_j)$ the end points of the out-edges of $p_j$. From the above discussion, it is easy to see the following observation and lemma. \begin{observation} ${\cal C}(p_j)=S_{out}(p_j)$. \end{observation} \begin{lemma}\label{the:SYcontainsNNG} The Semi-Yao graph is a super-graph of the nearest neighbor graph. \end{lemma} From now on, when we say a convex set is \textit{empty}, we mean it has no point of $P$ in its interior. From Lemma~\ref{the:CPidea}, we obtain the following straightforward observation, which makes a connection to the Delaunay triangulations of the point set $P$. \begin{observation}\label{the:emptyTri} If $p_j$ has the minimum $b_l$-coordinate inside the wedge $W_l(p_i)$, then $p_i$ and $p_j$ touch the boundary of an empty equilateral triangle; $p_i$ touches a vertex and $p_j$ touches an edge of the triangle. \end{observation} A \textit{unit regular hexagon} is a regular hexagon whose edges have unit length; let $\varhexagon$ be the unit regular hexagon with center at the origin $o$ and vertices at $(\sqrt{3}/2,1/2)$, $(0,1)$, $(-\sqrt{3}/2,1/2)$, $(-\sqrt{3}/2,-1/2)$, $(0,-1)$, and $(\sqrt{3}/2,-1/2)$ (see Figure~\ref{fig:hexagon}(a)). Partition $\varhexagon$ into six equilateral triangles $\vartriangle_l$, $l=0,1,..,5$, and call any translated and scaled copy of $\vartriangle_l$ an \textit{$l$-tri} (see Figure~\ref{fig:hexagon}(b)). \begin{figure}[t] \begin{center} \includegraphics[scale=1]{hexagon.eps} \end{center} \caption{(a) Partitioning the unit regular hexagon into six equilateral triangles. (b) Some $0$-tri's.} \label{fig:hexagon} \end{figure} A Delaunay graph can be defined based on any convex shape, \emph{e.g.}, a square, a diamond, any triangle, or a piece of pie~\cite{Abam:2010:SEK:1630166.1630284,DBLP:conf/swat/RahmatiZ12,Drysdale:1990:PAC:320176.320194}. The Delaunay triangulation based on a convex shape is the maximal set of edges such that no two edges intersect except at common endpoints, and such that the endpoints of each edge lie on the boundary of an empty scaled translate of the convex shape. If the points are in \textit{general position}\footnote{The set of points $P$ is in general position with respect to a convex shape if it contains no four points on the boundary of any scaled translate of the convex shape.} the bounded faces of the Delaunay graph are triangles, and the Delaunay graph is called a Delaunay triangulation. Here we call the Delaunay triangulation constructed based on an equilateral triangle an \textit{Equilateral Delaunay triangulation} (EDT) There is a nice connection between the Semi-Yao graph and Equilateral Delaunay triangulations. In general, the Semi-Yao graph is the union of two Equilateral Delaunay triangulations~\cite{Bonichon:2010:CTD:1939238.1939265}. Next we describe this connection in a different, and in our view simpler, way than~\cite{Bonichon:2010:CTD:1939238.1939265}. Call an $l$-tri whose interior does not contain any point of $P$ an \textit{empty $l$-tri}. Denote by $EDT_l$ the Equilateral Delaunay triangulation based on the $l$-tri. The edge $p_ip_j$ is an edge of $EDT_l$ if and only if there is an empty $l$-tri such that $p_i$ and $p_j$ are on the boundary of the $l$-tri; Figure~\ref{fig:vor_del} depicts $EDT_0$ for a set of four points. Let ${\cal E}(G)$ be the set of edges of graph $G$; the set of vertices of $G$ is $P$. Since $\vartriangle_0$, $\vartriangle_2$, and $\vartriangle_4$ are translates of one another, and similarly for $\vartriangle_1$, $\vartriangle_3$, and $\vartriangle_5$, we have that ${\cal E}(EDT_0)={\cal E}(EDT_2)={\cal E}(EDT_4)$ and ${\cal E}(EDT_1)={\cal E}(EDT_3)={\cal E}(EDT_5)$. Thus, there are two different types of $l$-tri's. We define the \textit{Equilateral Delaunay graph} (EDG) to be the union of $EDT_0$ and $EDT_1$, \emph{i.e.}, $p_ip_j\in {\cal E}(EDG)$ if and only if $p_ip_j\in {\cal E}(EDT_0)$ or $p_ip_j\in {\cal E}(EDT_1)$. \begin{figure}[t] \centering \includegraphics[scale=0.4]{vor_del.eps} \caption{The Delaunay triangulation and the Voronoi diagram based on the $0$-tri, as produced by a program in~\cite{GeometryLabRolf}.} \label{fig:vor_del} \end{figure} The cell boundaries of a Voronoi diagram of a set $P$ of $n$ sites, based on a convex shape, consist of points where the convex-shaped waves emanating from the sites collide; to determine the Voronoi diagram of the set of four sites in Figure~\ref{fig:vor_del}, based on the $0$-tri, we use a program in~\cite{GeometryLabRolf}. Using divide and conquer algorithms by Chew and Drysdale~\cite{Chew:1985:VDB:323233.323264,Drysdale:1990:PAC:320176.320194}, \begin{theorem}\label{the:VD_DT_Con}{\tt \cite{Chew:1985:VDB:323233.323264,Drysdale:1990:PAC:320176.320194}} The Voronoi diagram and Delaunay triangulation of a set of $n$ sites based on a convex shape can be constructed in $O(n\log n)$ time. \end{theorem} Since each $\vartriangle_l$ is a convex shape, using the approaches of Chew and Drysdale, we can construct the corresponding Voronoi diagram/Delaunay triangulation in $O(n\log n)$ time. Then the following results. \begin{corollary}\label{the:EDT_Construction} The Equilateral Delaunay graph (EDG) can be constructed in $O(n\log n)$ time. \end{corollary} Let $p_ip_j\in {\cal E}(EDT_l)$. By definition there exists an empty $l$-tri such that $p_i$ and $p_j$ are on its boundary. By scaling down the $l$-tri, one of the $l$-tri vertices will be placed at $p_i$ or $p_j$; see Figures~\ref{fig:SY_EDT2}(b) and \ref{fig:SY_EDT2}(c). \begin{observation}\label{the:SqueezedTRI} If there is an empty $l$-tri such that $p_i$ and $p_j$ are on its boundary, then there is an empty $l$-tri with the same property such that either $p_i$ or $p_j$ is a vertex of the $l$-tri. \end{observation} The next lemma proves that the undirected Semi-Yao graph and the Equilateral Delaunay graph are equal to each other. \begin{lemma}\label{the:SY_EDT} Edge $p_ip_j\in {\cal E}(SYG)$ if and only if $p_ip_j\in {\cal E}(EDG)$. \end{lemma} \begin{proof} Let $p_ip_j$ be an edge of the undirected Semi-Yao graph such that $p_j$ has the minimum $b_l$-coordinate inside some wedge $W_l(p_i)$ (see Figure~\ref{fig:SY_EDT2}(a)). The bounded area created by the wedge $W_l(p_i)$ and the line through $p_j$ perpendicular to $b_l(p_i)$ is an $l$-tri. Therefore, for the edge $p_ip_j$, there exists an empty $l$-tri such that $p_i$ and $p_j$ are on its boundary. This implies that $p_ip_j$ is an edge of $EDT_l$. Let $p_ip_j\in {\cal E}(EDT_l)$. By the definition of $EDT_l$, there exists an empty $l$-tri such that $p_i$ and $p_j$ are on its boundary (see Figure~\ref{fig:SY_EDT2}(b)). By Observation~\ref{the:SqueezedTRI}, that is a rescaled $l$-tri such that $p_i$ and $p_j$ are on its boundary and such that one of the $l$-tri vertices is $p_i$ or $p_j$ (see Figure~\ref{fig:SY_EDT2}(c)); without loss of generality assume it is $p_i$. Point $p_j$ is inside the wedge $W_k(p_i)$, where $k\in\{l,(l+2)\bmod{6}, (l+4)\bmod{6}\}$. Point $p_j$ has the minimum $b_k$-coordinate inside the wedge $W_k(p_i)$; otherwise, there would be a point of $P$ inside the rescaled $l$-tri, which means that $p_ip_j\notin {\cal E}(EDT_l)$, a contradiction. Therefore, $p_ip_j\in {\cal E}(SYG)$. \end{proof} \begin{figure}[t!] \centering \includegraphics[scale=1]{SY_EDT2.eps} \caption{(a) The point $p_j$ has the minimum $b_0$-coordinate inside the wedge $W_0(p_i)$. (b) The $1$-tri corresponding to the edge $p_ip_j$ in $EDT_1$ does not contain any other points of $P$. (c) The point $p_j$ is inside the wedge $W_5(p_i)$ and has the minimum $b_5$-coordinate.} \label{fig:SY_EDT2} \end{figure} Now we can give the following result. \begin{theorem}\label{the:ANN_Construction} The all nearest neighbors and the closest pair problems in $\mathbb{R}^2$ can be solved in $O(n\log n)$ time. \end{theorem} \begin{proof} From Corollary~\ref{the:EDT_Construction} and Lemma~\ref{the:SY_EDT}, the Semi-Yao graph can be constructed in $O(n\log n)$ time. Since the number of edges in the Semi-Yao graph is at most $6n$, by traversing the Semi-Yao graph edges incident to each point, we can find all the nearest neighbors and the closest pair in linear time. \end{proof} \subsection{Kinetic Equilateral Delaunay Graph}\label{sec:kinetic_EDG} Since ${\cal E}(EDT_0)={\cal E}(EDT_2)={\cal E}(EDT_4)$ and ${\cal E}(EDT_1)={\cal E}(EDT_3)={\cal E}(EDT_5)$, to maintain the EDG, which is the union of $EDT_0$ and $EDT_1$, we need only to have kinetic data structures for $EDT_0$ and $EDT_1$. We describe how to maintain $EDT_0$; $EDT_1$ is handled similarly. The Delaunay triangulation $EDT_0$ is locally stable as long as the points are in general position. Note that we assume the set of points $P$ is in general position with respect to a $0$-tri; this means that no four or more points are on the boundary of any scaled, translated $0$-tri. When the points are moving, at a moment $t$ this assumption may fail. In fact for moving points, we make a further assumption: no four points are on the boundary of the $0$-tri throughout any positive interval of time. This ensures that the points are in general position over time except at some discrete moments. The number of these discrete moments over time is in the order of the number of changes to $EDT_0$, because the failure of the general position assumption is a necessary condition for changing the topological structure of $EDT_0$~\cite{Albers_voronoidiagrams}. When a point moves, $EDT_0$ can change only in the graph neighborhood of the point, and so the correctness of $EDT_0$ over time is asserted by a set of certificates. Our approach for maintenance of $EDT_0$ is a known approach also used in~\cite{Abam:2010:SEK:1630166.1630284,DBLP:conf/swat/RahmatiZ12,Agarwal:2010:KSD:1810959.1810984,Albers_voronoidiagrams} for maintenance of Delaunay triangulations based on convex shapes. Figure~\ref{fig:EDT_O}(a) depicts the $EDT_0$ of a set $P$ of points. Each edge on the boundary of the infinite face of $EDT_0$, like $p_ip_j$, is called a \textit{hull} edge; the other edges, like $p_{i'}p_{j'}$, are called \textit{interior} edges. Corresponding to these two types of edges, we define two types of certificates, \textit{NotInWedge} and \textit{NotInTri}, respectively. Below, we first we consider the interior edges and then the exterior edges. \begin{figure}[t!] \centering \includegraphics[scale=1]{EDT_O.eps} \caption{(a) The NotInTri certificate corresponding to the edge $p_{i'}p_{j'}$ certifies that $p_r$ is outside the $0$-tri of $p_{i'}$, $p_{j'}$, and $p_{r'}$. The NotInWedge certificates of the edge $p_ip_j$ certify that $p_{s_1}$, $p_{s_2}$, and $p_{s_3}$ are outside the corresponding $k$-wedge. (b) The changes to $EDT_0$ after $p_r$ moves inside the $0$-tri passing through $p_{i'}$, $p_{j'}$, and $p_{r'}$ and after $p_{s_1}$ moves inside the $k$-wedge of $p_ip_j$.} \label{fig:EDT_O} \end{figure} \paragraph{Interior Edges.} Each interior edge $p_{i'}p_{j'}\in EDT_0$ is incident to two triangles $p_{i'}p_{j'}p_{r'}$ and $p_{i'}p_{j'}p_r$ (see Figure~\ref{fig:EDT_O}(a)). For the triangle $p_{i'}p_{j'}p_{r'}$ (resp. $p_{i'}p_{j'}p_r$), there exists an empty $0$-tri, denoted by $\Delta^0_{r'}$ (resp. $\Delta^0_r$), such that $p_{i'}$, $p_{j'}$ and $p_{r'}$ (resp. $p_r$) are on the boundary of $\Delta^0_{r'}$ (resp. $\Delta^0_r$). For $p_{i'}p_{j'}$, we define a \textit{NotInTri} certificate certifying that $p_r$ (resp. $p_{r'}$) is outside $\Delta^0_{r'}$ (resp. $\Delta^0_r$). For sufficiently short time intervals, $p_r$ and $p_{r'}$ are the only points that can change the validity of edge $p_{i'}p_{j'}$ (see~\cite{Abam:2010:SEK:1630166.1630284,DBLP:conf/swat/RahmatiZ12,Agarwal:2010:KSD:1810959.1810984,Albers_voronoidiagrams}). Let $t$ be the time when the four points $p_{i'}$, $p_{j'}$, $p_{r'}$, and $p_r$ are on the boundary of a $0$-tri; at time $t^-$, $p_r$ (resp. $p_{r'}$) is outside $\Delta^0_{r'}$ (resp. $\Delta^0_r$). When $p_r$ (resp. $p_{r'}$) moves inside $\Delta^0_{r'}$ (resp. $\Delta^0_r$), at time $t^+$, this certificate fails and there is no empty $0$-tri such that $p_{i'}$ and $p_{j'}$ are on its boundary. Thus at time $t$, we have to delete the edge $p_{i'}p_{j'}$ and add the new edge $p_{r'}p_r$, because at time $t^+$ there exists an empty $0$-tri for $p_rp_{r'}$ (see Figure~\ref{fig:EDT_O}(b)). Also, we must define new certificates corresponding to the newly created triangles. \begin{figure}[t] \centering \includegraphics[scale=1.2]{k-wedge.eps} \caption{(a) A $0$-tri. (b) The $k$-wedges associated with the $0$-tri; edge $p_ip_j$ divides the $4$-wedge $\protect\overleftrightarrow{a_4oa_5}$ into the bounded area $\overline{op_ip_j}$ and the unbounded area $\protect\overleftrightarrow{a_4p_ip_ja_5}$.} \label{fig:k-wedge} \end{figure} \paragraph{Hull Edges.} By removing one of the $0$-tri edges and extending the other two edges to infinity, three types of wedges are created; call these wedges \textit{$k$-wedges}, for $k=\{0,2,4\}$, and denote them by $\overleftrightarrow{a_koa_{k+1}}$ (see Figure~\ref{fig:k-wedge}); the two sides $\overrightarrow{oa_k}$ and $\overrightarrow{oa_{k+1}}$ of the boundary of the $k$-wedge are parallel to the two corresponding sides of the wedge $W_k$. For a hull edge $p_ip_j$, there exists an empty $k$-wedge such that $p_i$ and $p_j$ are on the boundary. Each hull edge is incident to at most one triangle $p_ip_jp_{s_1}$, and adjacent to at most four other hull edges $p_ip_{s_2}, p_ip_{s_3}, p_jp_{s_4}$ and $p_jp_{s_5}$ on the boundary cycle of the infinite face; the point $p_{s_1}$ can be one of the points $p_{s_2}$ to $p_{s_5}$. The only points that can change the validity of the edge $p_ip_j$ over a sufficiently short time interval are the points $p_{s_i}$, $1\leq i\leq 5$. Therefore, we define at most four \textit{NotInWedge} certificates for the hull edge $p_ip_j$, certifying that the points $p_{s_i}$, $1\leq i\leq 5$, are outside the $k$-wedge (see Figure~\ref{fig:EDT_O}(a)). If $p_ip_j$ is adjacent to four other hull edges, this edge cannot be incident to a triangle, and if it is incident to a triangle, it cannot be adjacent to more than two other hull edges. Let $t$ be the time when three points $p_i$, $p_j$, and $p_{s_i}$ are on the boundary of the $k$-wedge; at time $t^-$, $p_{s_i}$ is outside the $k$-wedge. The hull edge $p_ip_j$ divides its corresponding $k$-wedge $\overleftrightarrow{a_koa_{k+1}}$ into a bounded area $\overline{op_ip_j}$ and an unbounded area $\overleftrightarrow{a_kp_ip_ja_{k+1}}$ (see Figure~\ref{fig:k-wedge}(b)). If $p_{s_i}$ moves inside the bounded area $\overline{op_ip_j}$ at time $t^+$, the NotInWedge certificate of $p_ip_j$ fails, and we must delete $p_ip_j$ from the hull edges at time $t$ and replace it with two edges incident to $p_{s_i}$. In Figure~\ref{fig:EDT_O}(a), if $p_{s_1}$ moves inside the bounded area $\overline{op_ip_j}$, then we replace the hull edge $p_ip_j$ with two edges $p_ip_{s_1}, p_{s_1}p_j$; in particular, the chain $[..., p_{s_2}p_i, p_ip_j, p_jp_{s_3},...]$ of hull edges changes to $[..., p_{s_2}p_i, p_ip_{s_1}, p_{s_1}p_j, p_jp_{s_3},...]$ when $p_{s_1}$ moves inside the $k$-wedge (see Figure~\ref{fig:EDT_O}(b)). When this event occurs the previous interior edges $p_ip_{s_1}$ and $p_{s_1}p_j$ become hull edges, and we must replace the previous certificates of these edges with new valid ones. If $p_{s_i}$ moves inside the unbounded area $\overleftrightarrow{a_kp_ip_ja_{k+1}}$, without loss of generality let $p_{s_i}$ be incident to $p_i$, we replace the hull edges $p_{s_i}p_i$ and $p_ip_j$ with $p_{s_i}p_j$. Then the previous hull edge $p_ip_j$ either is an edge of $EDT_0$, in which case we must define a valid certificate for it, or it is not, in which case we must delete it from $EDT_0$ and add a new edge $p_{s_i}p_{s_1}$, where $p_ip_j$ is incident to a triangle $p_ip_jp_{s_1}$; see Figure~\ref{fig:BadEvents}. (a, b, and c). \paragraph{\textbf{Consecutive Changes to EDT$_0$.}}\label{par:CC} In some cases, when a certificate fails, we must apply a \textit{sequence} of changes to $EDT_0$. These kinds of changes occur at incident triangles, and as we will see, they can be handled consecutively. When a NotInWedge certificate fails, we apply a sequence of edge insertions and edge deletions to $EDT_0$. In Figure~\ref{fig:BadEvents}(a), when $p_{s_2}$ moves inside the $k$-wedge of $p_ip_j$, we replace chain $p_{s_2}p_i, p_ip_j$ of hull edges with $p_{s_2}p_j$ (see Figure~\ref{fig:BadEvents}(b)), and then we apply a sequence of changes; the previous hull edge $p_ip_j$ is no longer an edge in ${\cal E}(EDT_0)$, because now the interior of its corresponding $0$-tri contains the point $p_{s_2}$, and so we replace it with the edge $p_{s_1}p_{s_2}$ (see Figure~\ref{fig:BadEvents}(c)). Finally, by checking the $0$-tri's of other incident triangles, we can obtain a set of valid edges for $EDT_0$ (see Figure~\ref{fig:BadEvents}(d)). \begin{figure}[t] \centering \includegraphics[scale=1]{BadEvents.eps} \caption{The consecutive changes to $EDT_0$ when $p_{s_2}$ moves inside the $k$-wedge of $p_ip_j$.} \label{fig:BadEvents} \end{figure} A similar scenario could happen when a NotInTri certificate fails. In Figure~\ref{fig:BadEvents}(d), if $p_i$ moves inside the $0$-tri of $p_{s_2}$, $p_{s'_2}$, and $p_{i'}$, we must apply a sequence of changes to $EDT_0$ that is the reverse of what we did above when the NotInWedge certificate failed. First we replace $p_{s_2}p_{s'_2}$ with $p_ip_{i'}$. Then we must replace $p_{s_2}p_{i'}$ with $p_ip_{i''}$, because $p_i$ is inside the $0$-tri of $p_{s_2}$, $p_{i'}$, and $p_{i''}$. By checking the $0$-tri's of other incident triangles we can obtain a valid set of edges for $EDT_0$; see Figure~\ref{fig:BadEvents}, read from $(d)$ to $(a)$. Therefore, after any change to $EDT_0$ we must check the validity of the incident triangles, which can be done easily. Theorem~\ref{the:EDT_Changes} below enumerates the changes to the Equilateral Delaunay graph (\emph{i.e.}, the Semi-Yao graph) when the points are moving and gives the time to process all these events. \begin{theorem}\label{the:EDT_Changes} The number of changes to the Equilateral Delaunay graph, when the points move according to polynomial functions of at most constant degree $s$, is $O(n^2\beta_{s+2}(n))$. The total processing time for all events is $O(n^2\beta_{s+2}(n)\log n)$. \end{theorem} \begin{proof} From Lemma~\ref{the:SY_EDT}, the Equilateral Delaunay graph changes if and only if the Semi-Yao graph changes. Fix a point $p_i$ and one of its wedges $W_l(p_i)$. Since the trajectory of each point $p_i(t)=(x_i(t),y_i(t))$ is defined by two polynomial functions of at most constant degree $s$, each point can insert into ${\cal V}_l(p_i)$ at most $s$ times. The $b_l$-coordinates of the points inserted into ${\cal V}_l(p_i)$ create at most $sn$ partial functions of at most constant degree $s$. From Theorem~\ref{the:partiallyDFcomplexity}, the minimum value of these $sn$ partial functions changes at most $\lambda_{s+2}(sn)$ times, which is equal to the number of all changes for the point with minimum $b_l$-coordinate among the points in ${\cal V}_l(p_i)$. Since $s$ is a constant, we have that $\lambda_{s+2}(sn)=O(\lambda_{s+2}(n))$. Thus the number of all changes for all points is $O(n\lambda_{s+2}(n))=O(n^2\beta_{s+2}(n))$. The number of certificates is in the order of the number of changes to $EDT_0$. When a change to $EDT_0$ occurs, we update the $EDT_0$ and replace the invalid certificate(s) with new valid one(s). The time to make a constant number of deletions/insertions into the priority queue is $O(\log n)$. Thus the total time to process all events is $O(n^2\beta_{s+2}(n)\log n)$. \end{proof} \subsection{Kinetic All Nearest Neighbors}\label{sec:kinetic_ANN} The Equilateral Delaunay graph (Semi-Yao graph) is a supergraph of the nearest neighbor graph. Let $Inc(p_i)$ be the set all edges incident to $p_i$ in the Semi-Yao graph. Over time, to maintain the nearest neighbor to each point $p_i$, we need to track the edge with the minimum length in $Inc(p_i)$. Using a dynamic and kinetic tournament tree (see Section~\ref{sec:preliminary}), we can maintain the edge with the minimum length among the edges in $Inc(p_i)$. For each $Inc(p_i)$, $i=1,2,...,n$, we construct a dynamic and kinetic tournament tree ${\cal T}_i$. The edges of $Inc(p_i)$ are stored at leaves of the tournament tree, and each of the internal nodes of the tree maintains the edge with the minimum length stored at its two children; the root of the tree maintains the edge with minimum length among all edges in $Inc(p_i)$. Let $n_i$ be the cardinality of the set $Inc(p_i)$. Consider a sequence of $m_i$ insertions and deletions into ${\cal T}_i$. From Theorem~\ref{the:DynamicKineticTT}, and the fact that the lengths of any two edges in $Inc(p_i)$ can become equal at most $2s$ times, the following results. \begin{lemma}\label{the:DKTT} The dynamic and kinetic tournament tree ${\cal T}_i$ of $n_i$ elements can be constructed in $O(n_i)$ time. The tournament tree ${\cal T}_i$ generates at most $O(m_i\beta_{2s+2}(n_i)\log n_i)$ events, for a total cost of $O(m_i\beta_{2s+2}(n_i)\log^2 n_i)$. \end{lemma} Now we can prove the following. \begin{corollary}\label{the:AllDKTT} All the dynamic and kinetic tournament trees ${\cal T}_i$'s can be constructed in $O(n)$ time. These dynamic and kinetic tournament trees generate at most $O(n^2\beta^2_{2s+2}(n)\log n)$ events, for a total cost of $O(n^2\beta^2_{2s+2}(n)\log^2 n)$. \end{corollary} \begin{proof} By Lemma~\ref{the:DKTT} all the dynamic and kinetic tournament trees ${\cal T}_i$, $i=1,...,n$, generate at most $O(\sum_{i=1}^{i=n}m_i\beta_{2s+2}(n_i)\log n_i)=O(\beta_{2s+2}(n)\log n\sum_{i=1}^{i=n}m_i)$ events. Since each edge is incident to two points, inserting (resp. deleting) an edge $p_ip_j$ into the Equilateral Delaunay graph causes two insertions (resp. deletions) into the tournament trees ${\cal T}_i$ and ${\cal T}_j$. Therefore, by Theorem~\ref{the:EDT_Changes}, the number of all insertions/deletions into the tournament trees is $\sum_{i=1}^{i=n}m_i=O(n^2\beta_{s+2}(n))=O(n^2\beta_{2s+2}(n))$. Hence, the number of all events is $O(n^2\beta^2_{2s+2}(n)\log n)$, and the total cost is $O(n^2\beta^2_{2s+2}(n)\log^2 n)$. \end{proof} Now we can prove the following theorem, which gives the results about our kinetic data structure for the all nearest neighbors problem. \begin{theorem}\label{the:KinecitNNG} Our kinetic data structure for maintenance of all the nearest neighbors uses linear space and $O(n\log n)$ preprocessing time. It handles $O(n^2\beta^2_{2s+2}(n)\log n)$ events with total processing time $O(n^2\beta^2_{2s+2}(n)\log^2 n)$. It is compact, efficient, responsive in an amortized sense, and local on average. \end{theorem} \begin{proof} Since $\sum_i n_i = n$, the total size of all the tournament trees ${\cal T}_i$, $i=1,...,n$, is $O(n)$. The number of all edges in the EDG is $O(n)$. For each edge in the EDG, we define a constant number of certificates. Furthermore, the number of all certificates corresponding to the internal nodes of all ${\cal T}_i$ is linear. Thus the KDS is compact. The ratio of the number of internal events $O(n^2\beta^2_{2s+2}(n)\log n)$ to the number of external events $O(n^2\beta_{2s})$ is polylogarithmic, which implies that the KDS is efficient. By Corollary~\ref{the:AllDKTT}, the ratio of the total processing time to the number of internal events is polylogarithmic, and so the KDS is responsive in an amortized sense. Since the number of all certificates is $O(n)$, each point participates in a constant number of certificates on average, which implies that the KDS is local on average. \end{proof} \subsection{Kinetic Closest Pair}\label{sec:kinetic_CP} The edge $p_ip_j$ with minimum length in the nearest neighbor graph gives the closest pair $(p_i,p_j)$. Since the Semi-Yao graph (EDG) is a supergraph of the nearest neighbor graph, to maintain the closest pair $(p_i,p_j)$ we need to maintain the edge with minimum length in the Semi-Yao graph. By constructing a dynamic and kinetic tournament tree, where the edges of the Semi-Yao graph are stored at the leaves of the dynamic and kinetic tournament tree, we can maintain the closest pair $(p_i,p_j)$ over time; the edge at the root of the dynamic and kinetic tournament tree gives the closest pair. The insertions and deletions into the dynamic and kinetic tournament tree occur when a change to the Semi-Yao graph occurs. Therefore, we can obtain the same results for maintenance of the closest pair over time as we obtained for maintenance of all the nearest neighbors in Theorem~\ref{the:KinecitNNG}: \begin{theorem}\label{the:KinecitCP} Our kinetic data structure for maintenance of the closest pair uses linear space and $O(n\log n)$ preprocessing time. It handles $O(n^2\beta^2_{2s+2}(n)\log n)$ events with total processing time $O(n^2\beta^2_{2s+2}(n)\log^2 n)$, and it is compact, efficient, responsive in an amortized sense, and local on average. \end{theorem} \section{Yao Graph and EMST}\label{sec:YG_EMST} Our approach for computing the Yao graph and the EMST is similar to the approach for computing all the nearest neighbors and the closest pair in Section~\ref{sec:ANN_CP_construction}. First we introduce a new supergraph of the Yao graph, namely the Pie Delaunay graph, then we show how to maintain the Pie Delaunay graph (PDG) over time, and finally, using the kinetic version of the Pie Delaunay graph, we provide a KDS for maintenance of the Yao graph and the EMST when the points are moving. \subsection{New Method for Computing the Yao Graph and the EMST}\label{sec:YG_EMST_construction} Consider a partition of a unit disk into six \textit{pieces of pie} $\sigma_0,...,\sigma_5$, each of angle $\pi/3$ with common apex at the origin $o$. For $0\leq l \leq 5$, let $\sigma_l$ span the angular range $[(2l-1)\pi/6, (2l+1)\pi/6)$, and call any translated and scaled copy of $\sigma_l$ an \textit{$l$-pie}; see Figure~\ref{fig:disk}. \begin{figure}[h] \begin{center} \includegraphics[scale=1]{disk.eps} \end{center} \caption{(a) Partitioning the unit disk into six pieces of pie. (b) Some $0$-pie's.} \label{fig:disk} \end{figure} We define a Delaunay triangulation, which we call a \textit{Pie Delaunay triangulation}, of the set $P$ of $n$ points, based on the convex shape $\sigma_l$. Denote by $PDT_l$ the Pie Delaunay triangulation based on the $l$-pie. For two points $p_i$ and $p_j$ in $P$, the edge $p_ip_j$ is an edge of $PDT_l$ if and only if there is an empty $l$-pie such that $p_i$ and $p_j$ are on its boundary. We define the \textit{Pie Delaunay graph} (PDG) to be the union of all $PDT_l$ for $i=0,...,5$; \emph{i.e.}, $p_ip_j$ is a PDG edge if and only if it is an edge in $PDT_l$, where $0\leq l\leq 5$. The next lemma follows from Theorem~\ref{the:VD_DT_Con}. \begin{lemma}\label{the:PDG_ConstructionTime} The Pie Delaunay graph (PDG) can be constructed in $O(n\log n)$ time. \end{lemma} For each point $p_i\in P$, partition the plane into six wedges $W_0(p),..., W_5(p)$ of angle $\pi/3$ where $p_i$ is the common apex of the wedges. For $0\leq l\leq 5$, let $W_l(p_i)$ span the angular range $[(2l-1)\pi/6, (2l+1)\pi/6)$ around $p_i$. The \textit{Yao graph} can be constructed by connecting the point $p_i$ to its nearest points inside the wedges $W_l(p)$ for all $i=0,...,5$. We denote the Yao graph of a set of $n$ points by YG, the set of its edges by ${\cal{E}}(YG)$, and the set of Pie Delaunay graph edges by ${\cal E}(PDG)$. The following lemma shows that the Pie Delaunay graph is a supergraph of the Yao graph (YG). \begin{lemma}\label{the:YG_SS_PDG} ${\cal E}(YG)\subseteq {\cal E}(PDG)$. \end{lemma} \begin{proof} Assume edge $p_ip_j\in {\cal E}(YG)$ and let $p_j$ to be the nearest point to $p_i$ inside the wedge $W_l(p_i)$; see Figure~\ref{fig:yao}. The two sides of the wedge $W_l(p_i)$ are parallel to the two corresponding sides of $\sigma_l$, so there is an empty $l$-pie such that $p_i$ and $p_j$ lie on its boundary. Therefore, $p_ip_j\in PDT_l$ and hence it is an edge of the Pie Delaunay graph. \end{proof} \begin{figure}[t] \centering \includegraphics[scale=1.3]{yao.eps} \caption{Nearest point to $p_i$ inside the wedge $W_l(p_i)$.} \label{fig:yao} \end{figure} Now we can state and prove the main result of this section. \begin{theorem}\label{the:YG_EMST_ConstructionTime} The Yao graph and the EMST can be constructed in $O(n\log n)$ time. \end{theorem} \begin{proof} The Pie Delaunay graph is the union of six Pie Delaunay triangulations, which implies that it has a linear number of edges. By Lemma~\ref{the:YG_SS_PDG}, the Pie Delaunay graph is a supergraph of the Yao graph. Thus by tracing over the edges incident to each point $p_i$, we can find the edge with minimum length inside each wedge $W_l(p_i)$, for $l=0,...,5$; this gives the Yao graph. Since the Pie Delaunay graph can be constructed in time $O(n\log n)$ (by Lemma~\ref{the:PDG_ConstructionTime}), the Yao graph can be constructed in time $O(n\log n)$. The Yao graph is a supergraph of the EMST~\cite{DBLP:journals/siamcomp/Yao82}. Thus the minimum spanning tree of the Yao graph is equal to the EMST. Since the cardinality of the set of edges in the Yao graph graph is at most $6n$, the EMST can be constructed using the Prim algorithm~\cite{Prim57} or the Kruskal algorithm~\cite{Kruskal1956} in time $O(n\log n)$. \end{proof} \subsection{Kinetic Pie Delaunay Graph}\label{sec:kinetic_PDG} Our KDS for maintenance of the Pie Delaunay graph is similar to the KDS for maintenance of the Equilateral Delaunay graph in Section~\ref{sec:kinetic_EDG}. The Pie Delaunay graph (PDG) is the union of all $PDT_l$, for $l=0,..,5$: ${\cal E}(PDG)=\bigcup_l {\cal E}(PDT_l)$. Here, we only provide a KDS for $PDT_0$; the other $PDT_l$, for $l=1,..,5$, are handled similarly. Similar to Section~\ref{sec:kinetic_EDG}, we call each edge that is not on the boundary of the infinite face of $PDT_0$ an \textit{interior edge} and the other edges on the boundary of the infinite face \textit{hull edges}, and corresponding to them we define two kinds of certificates, \textit{NotInCone} and \textit{NotInPie}, respectively. \paragraph{Interior Edges.} By definition, an interior edge $p_{i'}p_{j'}\in {\cal E}(PDT_0)$ is incident to two triangles of $PDT_0$ that together form a \textit{quadrilateral}. Let $p_{r'}$ and $p_r$ be the two other vertices of the quadrilateral. For the edge $p_{i'}p_{j'}$, we define a \textit{NotInPie} certificate which certifies that point $p_{r}$ (resp. $p_{r'}$) is outside the $0$-pie passing through $p_{i'}$, $p_{j'}$, and $p_{r'}$ (resp. $p_r$). When the certificate fails, we replace $p_{i'}p_{j'}$ by $p_{r}p_{r'}$. In general, when the certificates corresponding to an interior edge fails, we perform such an edge swap. \paragraph{Hull Edges.} Let $o$, $w_0$, and $w_1$ be vertices of a $0$-pie (see Figure~\ref{fig:NotInCone}(a)). Two of the edges on the boundary of the $0$-pie are line segments and one of them is an arc; denote the line segments by $\overline{ow_0}$ and $\overline{ow_1}$ and the arc by $\overline{w_0w_1}$. By removing one of them and extending the line segment(s) to infinity, a cone can be created. We call these cones \textit{$k$-cones}. By definition, the edge $p_ip_j$ is a hull edge of $PDT_0$ if and only if there exists an empty $k$-cone such that $p_i$ and $p_j$ are on its boundary. Consider the $k$-cone $ow_1w_0$ corresponding to the edge $p_ip_j$ where one of the endpoints $p_i$ lies on the half-line $\overrightarrow{w_0o}$ and the other point $p_j$ lies on the half-arc $\overrightarrow{w_0w_1}$ (see Figure~\ref{fig:NotInCone}(b)). Let $\overrightarrow{\tilde{w_1}\tilde{w_0}}$ be the half-line perpendicular to $\overrightarrow{w_1o}$ through $p_j$. For such a $k$-cone we assume that the line segment $\overrightarrow{w_1o}$ goes to infinity. This means that $w_1$ (resp. $w_0$) tends to $\tilde{w_1}$ (resp. $\tilde{w_0}$) and the $k$-cone approaches a right-angled wedge; see Figure~\ref{fig:NotInCone}(c). Each hull edge $p_ip_j$ is adjacent to at most four other hull edges, denoted by $p_ip_{s_2}$, $p_ip_{s_3}$, $p_jp_{s_4}$, $p_jp_{s_5}$, and incident to at most one triangle. Let $p_{s_1}$ be the third vertex of this triangle if it exists; $p_{s_1}$ can be one of the $s_i$ where $2\leq i\leq 5$. If $p_ip_j$ is adjacent to at most four other triangles, then it cannot be incident to a triangle. In particular, at any time, the number of points $p_{s_i}$ is at most four. Therefore, for the $k$-cone passing through $p_i$ and $p_j$, we define at most four \textit{NotInCone} certificates certifying that the $p_{s_i}$ are outside of the $k$-cone. Note that in the case that a $k$-cone approaches a right-angled wedge (see Figure~\ref{fig:NotInCone}(c)), the certificate of the hull edge $p_ip_j$ fails when a point either crosses the half-line $\overrightarrow{w_1o}$, or reaches the line-segment $\overline{\tilde{w_1}p_j}$, or crosses the half-line $\overrightarrow{p_j\tilde{w_0}}$. \begin{figure}[t] \centering \includegraphics[scale=1.2]{NotInCone.eps} \caption{(a) A $0$-pie. (b) Two $k$-cones corresponding to the hull edge $p_ip_j$. (c) The $k$-cone approaches a right-angled wedge as $o$ goes to infinity.} \label{fig:NotInCone} \end{figure} The changes that can occur to $PDT_0$ are similar to the changes to $EDT_0$ and can easily be handled; see the paragraph "Consecutive Changes to EDT$_0$" in Section~\ref{sec:kinetic_EDG} for more details. Next we state a theorem that enumerates the number of the combinatorial changes to the Pie Delaunay graph. \begin{theorem}\label{the:num_changes_PDG} The number of all changes (edge insertions and edge deletions) to the Pie Delaunay graph of a set of $n$ moving points with trajectories given by polynomial functions of at most constant degree $s$ is $O(n^3\beta_{2s+2}(n))$. \end{theorem} \begin{proof} Consider $PDT_0$. The number of hull-edge changes to $PDT_0$ is $O(n^3)$ as three points are involved in any hull change. Since $n^3 =O(n^3\beta_{2s+2}(n))$, we focus on the number of changes to the triangles of $PDT_0$. For each edge $p_ip_j$ of a triangle in $PDT_0$, four different cases are possible as shown in Figure~\ref{fig:com_changes}. It is easy to see for any triangle $\Delta$ in the $PDT_0$ that case (a) of Figure~\ref{fig:com_changes} may happen to one of its edges. We charge any change to $\Delta$ to this edge. Therefore, we consider the number of combinatorial changes to $PDT_0$ for an arbitrary edge $p_ip_j$ that satisfies case (a) of Figure~\ref{fig:com_changes}. As mentioned above, two edges of a $0$-pie are line segments $\overline{ow_0}$ and $\overline{ow_1}$ and one of them is an arc $\overline{w_0w_1}$. Let $C_{w_0w_1}$ be the cone whose sides are created by removing the arc $\overline{w_0w_1}$ of the $0$-pie and extending the two line segments to infinity; the wedge $C_{w_0w_1}$ is the area between two half-lines $\overrightarrow{ow_0}$ and $\overrightarrow{ow_1}$. Let ${\cal V}(C_{w_0w_1})$ be the set of all points inside the wedge $C_{w_0w_1}$. In Figure~\ref{fig:com_changes}(a), a change for triangle $p_ip_jp_r$ corresponding to $p_ip_j$ occurs in two cases: \\ \textit{Case (I).} For some $p_t\in {\cal V}(C_{w_0w_1})$, the length of the edge $op_t$ becomes smaller than the length of the edge $op_r$. Note that since the degree of each function describing each point's motion is at most $s$, each point of $P$ except $p_i$ and $p_j$ can move inside the cone $C_{w_0w_1}$ at most $s$ times. Summing over all points in $P$ there are $O(sn)$ insertions into ${\cal V}(C_{w_0w_1})$. The distance of these points from the apex $o$, in the $L_2$ metric, creates $O(sn)$ partial functions, and each pair of these functions intersects at most $2s$ times. Therefore, the number of combinatorial changes corresponding to an arbitrary edge $p_ip_j$ equals $\lambda_{2s+2}(sn)$, which is equal to the number of breakpoints in the lower envelope of $sn$ partial functions of at most degree $2s$ (see Theorem~\ref{the:partiallyDFcomplexity}). Since the maximum degree $s$ is a constant, $\lambda_{2s+2}(sn)=O(\lambda_{2s+2}(n))$. The number of all possible edges is $O(n^2)$, and therefore the number of combinatorial changes corresponding to all edges is $O(n^2\lambda_{2s+2}(n))$. \\ \textit{Case (II).} In addition to the above changes for the edge $p_ip_j$ in Case (I), there exist other changes that can occur when a point such as $p_{t'}$ passes through the segment $op_i$ or the segment $op_j$ and enters inside the area $op_ip_j$ (see Figure~\ref{fig:com_changes}(a)). Map each point $p_i=(x_i(t),y_i(t))$ to a point $p'_i=(u_i(t),v_i(t))$ in a new parametric plane where $u_i(t)=x_i(t)+\sqrt{3}y_i(t)$ and $v_i(t)=x_i(t)-\sqrt{3}y_i(t)$. Passing the point $p_{t'}$ through the segment $op_i$ or the segment $op_j$ means that the point $p_{t'}$ exchanges its $u$-coordinate or its $v$-coordinate with the $u$-coordinate or $v$-coordinate of $p'_i$ or $p'_j$. We call these changes \textit{swap-changes}. Observe that the total number of swap-changes for all cases is bounded by the number of all swaps between points in their ordering with respect to the $u$-axis and $v$-axis. The number of all the $u$-swaps and $v$-swaps between points is at most $O(n^2)$. Hence, the number of changes to the Pie Delaunay graph is $O(n^3\beta_{2s+2}(n))$. \end{proof} \begin{figure}[t] \centering \includegraphics[scale=0.9]{com.eps} \caption{Combinatorial changes for an arbitrary edge $p_ip_j$.} \label{fig:com_changes} \end{figure} After any change to the Pie Delaunay graph, we replace a constant number of (invalid) certificates from the priority queue with new valid ones, which takes $O(\log n)$ time. From the above discussion, together with Lemma~\ref{the:PDG_ConstructionTime} and Theorem~\ref{the:num_changes_PDG}, we obtain the following theorem. \begin{theorem}\label{the:pie_d} For a set of $n$ points in the plane with trajectories given by polynomial functions of at most constant degree $s$, there exists a KDS for maintenance of the Pie Delaunay graph that uses linear space, $O(n\log n)$ preprocessing time, and that processes $O(n^3\beta_{2s+2}(n))$ events with total processing time $O(n^3\beta_{2s+2}(n)\log n)$. \end{theorem} \subsection{Kinetic Yao Graph}\label{sec:kinetic:YG} To maintain the Yao graph, for each point $p_i\in P$, we must maintain the nearest points to $p_i$ inside the wedges $W_l(p_i)$, where $0\leq l\leq 5$. Since the Yao graph is a subgraph of the Pie Delaunay graph (by Lemma~\ref{the:SY_EDT}), to maintain the nearest points inside the wedges of $p_i$, we only need to track the edges of the Pie Delaunay graph incident to $p_i$ with minimum length inside the wedges $W_l(p_i)$ for all $l=0,...,5$. Let $Inc_l(p_i)$ be the set all edges of the Pie Delaunay graph incident to $p_i$ inside the wedge $W_l$. We store the edges of $Inc_l(p_i)$ at leaves of a dynamic and kinetic tournament tree ${\cal T}_{l,i}$ (see Section~\ref{sec:preliminary}). The root of ${\cal T}_{l,i}$ maintains the winner, the edge with minimum length among all edges in $Inc_l(p_i)$. Given the KDS of the Pie Delaunay graph and making an analysis similar to that of Corollary~\ref{the:AllDKTT} and Theorem~\ref{the:KinecitNNG}, the following theorem results. \begin{theorem}\label{the:kineticYG} The KDS for maintenance of the Yao graph uses $O(n)$ space, $O(n\log n)$ preprocessing time, and processes $O(n^3\beta^2_{2s+2}\log n)$ (internal) events with total processing time $O(n^3\beta^2_{2s+2}\log^2 n)$. It is compact, responsive in an amortized sense, and local on average, but it is not efficient. \end{theorem} For \textit{linearly} moving points in the plane, Katoh~\emph{et~al.}~\cite{Katoh:1992:MMS:1398516.1398835} showed that the number of changes to the Yao graph is $O(n\lambda_4(n))$. In the following theorem we bound the number of combinatorial changes to the Yao graph of a set of moving points whose trajectories are given by polynomial functions of at most constant degree $s$. For maintenance of the Yao graph, our KDS processes $O(n^3\beta^2_{2s+2}\log n)$ events, but the following theorem proves that the number of exact changes to the Yao graph is nearly quadratic, which explains why our KDS is not efficient. \begin{theorem}\label{the:num_changes_YG} The number of all changes to the Yao graph, when the points move with polynomial trajectories of at most constant degree $s$, is $O(n^2\beta_{2s+2}(n))$. \end{theorem} \begin{proof} Consider the point $p_i\in P$ and one of its wedges $W_l(p_i)$. Each of the other points in $P$ can be moved inside the wedge $W_l(p_i)$ at most $s$ times, and so there exist $O(sn)$ insertions into the wedge $W_l(p_i)$. The distance of these points from $p_i$ creates $O(sn)$ partial functions; each pair of these functions intersects at most $2s$ times. By Theorem~\ref{the:partiallyDFcomplexity}, the edge with minimum length changes at most $\lambda_{2s+2}(sn)=O(\lambda_{2s+2}(n))$ times. Hence, the number of all changes to the Yao graph of a set of $n$ moving points is $O(n\lambda_{2s+2}(n))$. \end{proof} \begin{remark} Using an argument similar to that for the KDS we obtained for the Yao graph in the $L_2$ metric, a KDS for the Yao graph in the $L_1$ and $L_\infty$ metrics can be obtained. Denote by $\square$ the \textit{unit square} with corners at $(0,0)$, $(1, 0)$, $(0, 1)$, and $(1, 1)$ in a Cartesian coordinate system, and call any translated and scaled copy of $\square$ an \textit{SQR}. The edge $p_ip_j$ is an edge of the Delaunay triangulation based on an SQR in the $L_\infty$ metric if and only if there is an empty SQR such that $p_i$ and $p_j$ are on its boundary (\emph{i.e.}, the interior of SQR contains no point of $P$). Abam~\emph{et~al.}~\cite{Abam:2010:SEK:1630166.1630284} showed how to maintain a Delaunay triangulation based on a diamond. Each SQR is a diamond, so using their approach applies. The Delaunay triangulation where the triangulation is based on an SQR in the $L_\infty$ metric can be maintained kinetically by processing at most $O(n\lambda_{s+2}(n))$ events, each in amortized time $O(\log n)$. The Delaunay triangulation based on an SQR is a supergraph for the Yao graph in the $L_\infty$ metric. Therefore, we can have a KDS for the Yao graph in the $L_\infty$ metric that uses $O(n)$ space, $O(n\log n)$ preprocessing time, and that processes $O(n^2\beta^2_{s+2}(n)\log n)$ events, each in amortized time $O(\log n)$. The Delaunay triangulation in the $L_1$ metric can be constructed/maintained analogously, by rotating all points $45$ degrees around the origin and constructing/maintaining the Delaunay triangulation in the $L_\infty$ metric. \end{remark} \subsection{Kinetic EMST}\label{sec:kinetic_EMST} Our kinetic approach for maintaining the EMST is based on the fact that the EMST is a subgraph of the Yao graph, where the number of the wedges around each point in the Yao graph is greater than or equal to six~\cite{DBLP:journals/siamcomp/Yao82}. Let $L$ be a list of the Yao graph edges (which in fact are stored at the roots of the dynamic and kinetic tournament trees ${\cal T}_{l,i}$, for each point $p_i\in P$ and $l=1,...,6n$), sorted with respect to their Euclidean lengths. A change to the EMST may occur when two edges in $L$ change their ordering. For each two consecutive edges in $L$, we define a certificate certifying the respective sorted order of the edges. Whenever the ordering of two edges in this list is changed, we apply the required changes to the EMST KDS. Therefore, to update the EMST when the points are moving, we must track the changes to $L$. There exist two types of changes to $L$: $(a)$ edge insertion and edge deletion from $L$, and $(b)$ a change in the order of two consecutive edges in $L$. The following discusses how to handle these two types of events. \paragraph{Case (a):} As soon as an edge is deleted from $L$ a new one is inserted. Both the deleted edge and the inserted edge are in the same dynamic and kinetic tournament tree, and both of them have a common endpoint; see Figure~\ref{fig:InsDel}. Call the deleted edge and the inserted edge $p_ip_j$ and $p_ip_r$, respectively, and denote by ${\cal T}_{i,l}$ the dynamic and kinetic tournament tree that contains $p_ip_j$ and $p_ip_r$. The deleted edge $p_ip_j$ can be one of the EMST edges at time $t^-$ and if so, we have to find a new edge to repair the EMST at time $t^+$. The following lemma proves that this new edge is $p_ip_r$. \begin{figure}[t!] \centering \includegraphics[scale=0.9]{InsDel.eps} \caption{The edge connecting two subtrees $T_1(P_1,E_1)$ and $T_2(P_2,E_2)$: (a) At time $t^-$, $|p_ip_r|>|p_ip_j|>|p_jp_r|$ and the edge connecting $T_1$ and $T_2$ is $p_ip_j$. (b) At time $t^+$, $|p_ip_j|>|p_ip_r|>|p_jp_r|$ and the edge connecting $T_1$ and $T_2$ is $p_ip_r$.} \label{fig:InsDel} \end{figure} \begin{lemma} Let $p_ip_j$ be the winner of the dynamic and kinetic tournament tree ${\cal T}_{i,l}$. Suppose $p_ip_j\in {\cal E}(EMST)$ at time $t^-$ and let $p_ip_r$ be the winner of ${\cal T}_{i,l}$ at time $t^+$. Then ($i$) at time $t^-$, $p_ip_r\notin {\cal E}(EMST)$, and ($ii$) at time $t^+$, $p_ip_r\in {\cal E}(EMST)$ and $p_ip_j\notin {\cal E}(EMST)$. \end{lemma} \begin{proof} Deleting an edge $p_ip_j$ from EMST creates two subtrees $T_1(P_1,E_1)$ and $T_2(P_2,E_2)$. Let $p_i\in P_1$ and $p_j\in P_2$; see Figure~\ref{fig:InsDel}. At time $t^-$, since $p_ip_j\in {\cal E}(EMST)$, $|p_ip_r|>|p_ip_j|>|p_jp_r|$, and $\angle p_jp_ip_r \leq \pi/3$, we have that $p_r\in P_2$. This can be concluded by contradiction. Thus ($i$) at time $t^-$, $p_ip_r\notin {\cal E}(EMST)$. The proof that $p_ip_j\notin {\cal E}(EMST)$ at time $t^+$ is analogous to the proof for ($i$). Therefore, at time $t^+$, the EMST is the union of two trees $T_1$ and $T_2$ and the edge $p_ip_r$. \end{proof} \paragraph{Case (b):} Let $path(e)$ be the simple path in the EMST between the endpoints of edge $e$ and let $|e|$ be the Euclidean length of $e$. A change in the sorted list $L$ corresponds to a pair of edges $e$ and $e'$ in ${\cal E}(YG)$ such that at time $t^-$, $|e|<|e'|$, and at time $t^+$, $|e|>|e'|$. Thus at time $t$, $e$ may be replaced by $e'$ in the EMST. It is easy to see the following. \begin{observation} The EMST changes if and only if at time $t^-$, $|e|<|e'|$, $e\in {\cal E}(EMST)$, $e'\notin {\cal E}(EMST)$, $e\in path(e')$, and at time $t^+$, $|e|>|e'|$. \end{observation} Such events can be detected and maintained in $O(\log n)$ time per operation using the \textit{link-cut tree data structure} of Sleator and Tarjan~\cite{Sleator:1983:DSD:61337.61338}. Given a KDS for maintenance of the Yao graph, the following bounds the number of events for maintaining the EMST. \begin{lemma}\label{the:num_changes_EMST} Given a Yao graph KDS for a set of $n$ points moving with polynomial trajectories of constant maximum degree $s$, there exists a KDS for maintenance of the EMST that processes $O(n^3\beta_{2s+2}(n))$ events. \end{lemma} \begin{proof} The set of Yao graph edges is a superset of the set of the EMST edges, and any change in the order of consecutive edges in the sorted list $L$ of the Yao graph edges may change the EMST. More precisely, any edge insertion/deletion in the Yao graph implies an insertion/deletion into $L$, and each insertion may cause $O(n)$ changes in the EMST. From Theorem~\ref{the:num_changes_YG}, the number of all insertions and deletions into the sorted list $L$ is $O(n^2\beta_{2s+2}(n))$. Thus the number of events that our KDS processes is $O(n^3\beta_{2s+2}(n))$. \end{proof} The KDS for maintenance of the EMST uses the Pie Delaunay graph KDS and the Yao graph KDS. From the above discussion and Theorems~\ref{the:pie_d} and \ref{the:kineticYG}, the following results. \begin{theorem}\label{the:kineticEMST} The KDS for maintenance of the EMST uses linear space and requires $O(n\log n)$ preprocessing time. The KDS processes $O(n^3\beta^2_{2s+2}(n)\log n)$ events, each in amortized time $O(\log n)$. The KDS is compact, responsive in an amortized sense, and local on average. \end{theorem} \section{Discussion and Open Problems}\label{sec:conclusion} We have provided a kinetic data structure for the all nearest neighbors problem for a set of moving points in the plane. We have applied our structure to maintain the closest pair as the points move. Comparison of our algorithm with the algorithm of Agarwal~\emph{et~al.}~\cite{Agarwal:2008:KDD:1435375.1435379} shows that in $\mathbb{R}^2$, our deterministic algorithm is simpler and more efficient than their randomized algorithm for maintaining all the nearest neighbors. In $\mathbb{R}^3$, the number of edges of the Equilateral Delaunay graph is $O(n^2)$, and so for maintenance of all the nearest neighbors, our kinetic approach needs $O(n^2)$ space. By contrast, the randomized kinetic data structure by Agarwal~\emph{et~al.}~\cite{Agarwal:2008:KDD:1435375.1435379} uses $O(n\log^3n)$ space. Thus, for higher dimensions ($d\geq 3$), their approach is asymptotically more efficient, but the simplicity of our algorithm may make it more attractive. In higher dimensions, our deterministic method of maintaining the Equilateral Delaunay graph, does not satisfy all four kinetic performance criteria. Thus, finding a deterministic kinetic algorithm for maintenance of all the nearest neighbors in higher dimensions, and that satisfies the performance criteria, is a future direction. We have also provided a KDS for maintenance of the EMST and the Yao graph on a set of $n$ moving points. Our KDS for maintenance of the EMST processes $O(n^3\beta^2_{2s+2}(n)\log n)$ events, which improves the previous $O(n^4)$ bound of Rahmati and Zarei~\cite{DBLP:conf/iwoca/RahmatiZ11}. The kinetic algorithm of Rahmati and Zarei results in a KDS having $O(n^{3+\epsilon})$ events, for any $\epsilon >0$, under the assumptions that ($i$) any four points can be co-circular at most twice, and ($ii$) either no ordered triple of points can be collinear more than once, or no triple of points can be collinear more than twice. Our kinetic approach further improves the upper bound $O(n^{3+\epsilon})$ under the above assumptions. A tight upper bound is not known. Our KDS can also be used to maintain an $L_1$-MST and an $L_\infty$-MST. By defining the Pie Delaunay graph and the Yao graph in $\mathbb{R}^d$, our kinetic approach can be used to give a simple KDS for the EMST in higher dimensions, but this approach does not satisfy all the performance criteria. For linearly moving points in the plane, Katoh~\emph{et~al.}~\cite{Katoh:1992:MMS:1398516.1398835} proved an upper bound of $O(n^32^{\alpha(n)})$ (resp. $O(n^{5/2}\alpha(n)$) for the number of combinatorial changes of the EMST (resp. $L_1$-MST and $L_\infty$-MST), where $\alpha(n)$ is the inverse Ackermann function. The upper bound was later proved to $O(\lambda_{ps+2}(n)n^{2-1/(9.2^{ps-3}})\log^{2/3}n)$ for the $L_p$-MST in $\mathbb{R}^d$, where the coordinates of the points are polynomial functions of constant maximum degree $s$~\cite{chan__2003}; for $p=2$ and $s=1$, this formula gives the first improvement $O(n^{25/9}2^{\alpha(n)}\log^{2/3}n)$ over Katoh~\emph{et~al.}'s $O(n^32^{\alpha(n)})$ bound. An even better bound $O(n^{8/3}2^{\alpha(n)}\log^{4/3}n)$ can be obtained by combining the results of Chan~\cite{chan__2003} with those of Marcus and Tardos~\cite{Marcus:2006:IRS:1142905.1142913}. Finding a tight upper bound for the number of combinatorial changes of the EMST, and finding a KDS for the EMST in $\mathbb{R}^d$ that processes a sub-cubic number of events are other future directions. \bibliographystyle{splncs03}
\section{Introduction} Merely a few years ago the computational power on the order of TFLOPS was an attribute of a supercomputer. Today, the latest commodity Graphics Processing Units (GPU) are capable of more than 5 TFLOPS and have much lower energy requirements than a set of Central Processing Units (CPUs) of comparable performance -- a result achieved by employing a massively parallel architecture with hundreds of cores on a single device. In order to take advantage of the parallel hardware, appropriate algorithms have to be developed, and this generally needs to be done on a problem-specific basis. Ideally, the problem naturally decomposes into a large number of (semi-)independent tasks whose results can be combined in a simple way, as in e.g. numerical Monte-Carlo solution of a stochastic differential equation~\cite{Januszewski2010}, parameter space studies of dynamical systems, or event simulation and reconstruction in particle physics~\cite{halyo2013gpu}. In recent years, the lattice Boltzmann method (LBM) emerged as an interesting alternative to more established methods for fluid flow simulations. Originally developed as an extension of lattice gas automata, nowadays LBM stands on well-established theoretical foundations and serves as a method of choice for many researchers in various fields~\cite{ChenDoolen1998, Aidun2010rev}, while still retaining its relative simplicity. Perhaps its most important feature in practice is its suitability for parallel architectures -- the algorithm is executed on a regular lattice and typically only interactions between nearest neighbor nodes are necessary. This paper discusses various issues related to the implementation of the LBM on GPUs, and presents a concrete and flexible solution, taking a new overall approach to the problem of LB software design. Our \textit{Sailfish} code has been under development since April 2009 and is publicly available under an open source license. With support for both single and binary fluid simulations, a wide variety of boundary conditions, and calculations on both a single GPU and multiple GPUs, it is, to the best of our knowledge, the most comprehensive open source LBM code for GPUs. The text is organized as follows. First, a short overview of the discussed lattice Boltzmann models is presented. Then, the primary design principles of the Sailfish project are laid out, and various LB GPU implementations are compared. This is followed by performance measurements on diverse GPU hardware, both on single machines and in clusters. In the next section, the code is validated on 4 standard test cases. The final section concludes the paper and provides some directions for future work. \section{Lattice Boltzmann methods overview} In this section, \emph{all} models implemented in Sailfish as of September 2013 will be briefly overviewed. We start with the basic concepts and nomenclature of LBM. We then describe single fluid models: single-relaxation time LBGK and regularized dynamics, multiple relaxation times and the entropic LBM, and multi-component models: the Shan-Chen model and the free energy model. Finally we give a short overview of our implementation of body forces and various boundary conditions. We refer the reader to the reviews \cite{ChenDoolen1998, Aidun2010rev} for a more detailed discussion of the LBM. In an LB simulation space is discretized into a regular Cartesian grid, where each node represents a small pocket of fluid. The state of the fluid is encoded by a particle distribution function $f_i$ where $i$ spans a set of discrete velocity vectors $\{ \vec{e}_i \}$ indicating the allowed directions of mass movement between the nodes of the lattice. $f_i$ are often also called \textit{mass fractions}. LB lattices are typically named using the DxQy scheme, where $x$ indicates the dimensionality of the lattice, and $y$ is the number of discrete velocity vectors. \figref{lattices} shows some common LB lattices. \begin{figure} \centering \includegraphics[width=\textwidth]{lattices.pdf} \caption{Sample lattices used for LB simulations: D2Q9 (left panel), D3Q15 (middle panel), and D3Q19 (right panel). Sailfish additionally implements D3Q27 and D3Q13~\cite{PhysRevE.63.066702}.} \label{fig:lattices} \end{figure} Macroscopic fluid fields, such as density ($\rho$) or velocity ($\vec{u}$) are defined as moments of the distribution function: \begin{equation} \rho = \sum_i f_i, \qquad \rho \vec{u} = \sum_i f_i \vec{e}_i \label{eq:lb_macro} \end{equation} In the simplest case the system dynamics is described by: \begin{equation} f_i(\vec{x} + \vec{e}_i, t + 1) - f_i(\vec{x}, t) = \frac{1}{\tau}\left(f_i - f_i^{\mathrm{eq}} \right)(\vec{x}, t) \label{eq:lbgk} \end{equation} where $\left\{ f_i^{\mathrm{eq}}(\vec{x}, t), i = 1, \ldots, y \right\}$ is a set of equilibrium distributions which are functions of the macroscopic fields ($\rho$, $\vec{u}$) at node $\vec{x}$ at time $t$, $\tau$ is a relaxation time related to the kinematic viscosity $\nu$ via $\tau = (1 + 6 \nu) / 2$, and where the positions and time are expressed in the lattice unit system, i.e. the time unit represents a single simulation step and the lattice nodes are separated by 1 distance unit along the primary axes. This LB model is often referred to as Lattice Bhatnagar-Gross-Krook (LBGK), which owes its name to the BGK operator from kinetic theory \cite{bgk54}, and by which the collision operator on the right hand side of \eqref{eq:lbgk} is inspired. \begin{table} \caption{Weights $w_i$ for different lattices implemented in Sailfish.} \centering \begin{tabular}{l | c c c c} \hline \multirow{2}{*}{Lattice} & \multicolumn{4}{c}{Link length} \\ & 0 & 1 & $\sqrt{2}$ & $\sqrt{3}$ \\ \hline D2Q9 & 4/9 & 1/9 & 1/36 & - \\ D3Q13 & 1/2 & - & 1/24 & - \\ D3Q15 & 2/9 & 1/9 & - & 1/72 \\ D3Q19 & 1/3 & 1/18 & 1/36 & - \\ D3Q27 & 8/27 & 2/27 & 1/54 & 1/216 \\ \hline \end{tabular} \label{tbl:bgkweights} \end{table} When $f_i^{\mathrm{eq}}$ takes the form: \begin{equation} f_i^{\mathrm{eq}} = \rho w_i \left( 1 + 3 \vec{e}_i \cdot \vec{u} + \frac{9}{2} (\vec{e}_i \cdot \vec{u} )^2 - \frac{3}{2} u^2 \right) \label{eq:equilibrium} \end{equation} where $\{w_i\}$ is a set of weights (see Table~\ref{tbl:bgkweights}), the Navier-Stokes equations can be recovered from \eqref{eq:lbgk} in the limit of low Mach numbers through Chapman-Enskog expansion \cite{ChenDoolen1998} or expansion in a Hermite basis \cite{ShanHe1998}. It can be shown that \eqref{eq:lbgk} is second-order accurate in space and time in the bulk of the fluid~\cite{ChenDoolen1998}. It should be noted that \eqref{eq:lb_macro} and the right hand side of \eqref{eq:lbgk} are fully local, i.e. only information from node $\vec{x}$ is necessary to calculate the system state at the next step of the simulation. When data is exchanged between nodes, the process is limited to the nearest neighbors (left hand side of \eqref{eq:lbgk}). This makes the LB method ideally suited for implementation on massively parallel computer architectures. Conceptually, \eqref{eq:lbgk} is often decomposed into two steps: relaxation (evaluation of the collision operator, right hand side) and propagation (movement of mass fractions to neighboring nodes, left hand side). While such a decomposition is often suboptimal from the point of view of a software implementation, it still provides a useful mental framework to discuss LB algorithms, and we will use these two names throughout the text. Many extended LB models have been proposed to (a) to improve numerical stability, or (b) simulate more complex systems. A comprehensive discussion of such models would be prohibitively long, so here we limit ourselves to a short overview of models that are currently implemented within the Sailfish framework. In the first group, we discuss the multiple-relaxation times (MRT) model, the entropic LB model (ELBM) and the Smagorinsky large eddy model. In the second group, Sailfish implements two multicomponent fluid models: the Shan-Chen model \cite{ShanChen1993}, and the free energy model \cite{swift-binary}. \subsection{Single fluid models with enhanced stability} One practical problem with the LBGK approximation is the limited range of viscosities at which the model remains stable, which directly translates into a limited range of Reynolds numbers $\mathrm{Re} = L u / \nu$ at which fluid flows can be simulated ($L$ being the spatial extent of the domain expressed as a number of nodes). With $u$ limited to about $0.05$ by the low Mach number requirement for most LB models where the speed of sound $c_s = 1/\sqrt{3}$ and the stable range of numerical viscosities at these speeds being about $(10^{-3}, 1/6)$, the highest attainable Reynolds number assuming $L = 300$ is on the order of $\mathrm{Re} = 15000$. In practice, the range of Reynolds numbers and viscosities is often even narrower due to instabilities introduced by complex geometry or boundary conditions. To address this problem, a number of models exhibiting enhanced numerical stability have been developed. \subsubsection{Regularized LBM} Perhaps the simplest modification of the LBGK model aiming at improving its stability is the regularized model~\cite{latt_regularized}. The basic idea of this approach is to replace the nonequilibrium part of the distributions $f_i^{\mathrm{neq}} = f_i - f^\mathrm{eq}_i$ with a new first order regularized value $f^{\mathrm{reg}}$ derived from the non-equilibrium stress tensor $\Pi_{\alpha \beta}^{\mathrm{neq}} = \sum_i f_i^{\mathrm{neq}} (e_{i\alpha} e_{i \beta} - \delta_{\alpha \beta} c_s^2)$ as $f^{\mathrm{reg}} = \frac{w_i}{2 c_s^4} (e_{i\alpha} e_{i\beta} - c_s^2 \delta_{\alpha \beta}) \Pi_{\alpha \beta}^{\mathrm{neq}}$. After this regularization step, the collision and propagation proceed as in the standard LBGK scheme. \subsubsection{Multiple relaxation times} If the particle distributions $f_i$ are considered to form vectors $\ket{f}$, the collision operator in the LBGK approximation \eqref{eq:lbgk} can be interpreted as a diagonal matrix $S = \omega I$, with $\omega = 1/\tau$ acting on the vector difference $\ket{f^\mathrm{eq}(\vec{x}, t)} - \ket{f(\vec{x}, t)}$. The basis of the distribution vector can be changed using an invertible matrix $M$: $\ket{m} = M \ket{f}$. In particular, the matrix $M$ can be chosen such that $\ket{m}$ is a vector of moments of the particle distributions $f_i$. The idea of MRT is to perform the collision in the moment basis, and then change the basis to recover the original form of the particle distributions for the streaming step \cite{DhumieresMrt2002}. The MRT collision operator can be written as: \begin{equation} M^{-1} \hat{S} \left( \ket{m^{\mathrm{eq}}(\vec{x}, t)} - \ket{m(\vec{x}, t)}\right) \label{eq:mrt_collision} \end{equation} where the collision matrix $\hat{S} = M S M^{-1}$ is diagonal with its elements $s_i = 1 / \tau_i$ representing the inverses of relaxation times for individual moments. Collision matrix elements corresponding to conserved moments, such as density or momentum are set to 0. Some of the remaining elements are related to kinematic and bulk viscosities, but the other ones can be tuned to increase the numerical stability of the simulation without changing its physical results. We refer the reader to the paper of D'Humieres et al.~\cite{DhumieresMrt2002} for a detailed discussion of optimal choices for relaxation times, and the form of the matrix $M$. \subsubsection{Entropic LBM} In the entropic LB model, a discrete H-function $H(f) = \sum_i f_i \ln (f_i / w_i)$ is defined and the equilibrium distribution function is defined as the extremum of $H(f)$ under constraints of mass and momentum conservation: \begin{equation} f_i^{\mathrm{eq}} = \rho w_i \prod_{\alpha=1}^D \left( 2 - \sqrt{1 + u_{\alpha}^2} \right) \left( \frac{2 u_{\alpha} + \sqrt{1 + 3u_{\alpha}^2}}{1 - u_{\alpha}} \right)^{e_{i \alpha}} \end{equation} where $D$ is the dimensionality of the lattice. The relaxation process is also modified to include a new dynamically adjusted parameter $\alpha$: \begin{equation} f_i(\vec{x} + \vec{e}_i, t + 1) - f_i(\vec{x}, t) = \omega_0 \frac{\alpha}{2} \left(f_i - f_i^{\mathrm{eq}} \right)(\vec{x}, t) \label{eq:elbm} \end{equation} with $\omega_0 = 1 / \tau$. $\alpha$ is evaluated at every simulation step as the solution of the $H$-function monotonicity constraint: \begin{equation} H(f) = H\left(f - \alpha (f - f^\mathrm{eq})\right) \label{eq:alpha_cond} \end{equation} This procedure guarantees unconditional stability of the numerical scheme, as the $H(f)$ function can be proven to be a Lyapunov function for the system. When $\alpha = 2$ as is the case when the system is close to equilibrium, \eqref{eq:elbm} has the same form as \eqref{eq:lbgk}. The corrections resulting from H-theorem compliance can be both positive and negative, temporarily increasing and decreasing the effective local viscosity of the fluid~\cite{ansumali_minimal_2007,chikatamarla_entropic_2006}. \subsubsection{Smagorinsky subgrid model} The idea behind the Smagorinsky subgrid model is to locally modify the fluid viscosity by adding an eddy viscosity term dependent on the magnitude of the strain rate tensor $S$: \begin{equation} \nu = \nu_0 + \nu_t \label{eq:smagorinsky} \end{equation} where $\nu_0$ is the baseline viscosity of the fluid and $\nu_t = \tau_t / 3$ is an eddy viscosity, the form of which depends on the subgrid model used in the simulation. In the Smagorinsky model, the relaxation time $\tau_t$ can be calculated using the momentum flux tensor $Q_{\alpha \beta} = \sum_i e_{i \alpha} e_{i \beta} (f_i - f_i^{\mathrm{eq}})$: \begin{equation} \tau_t = \frac{1}{2} \left( \sqrt{\tau_0^2 + 4 c_s^{-4} C_S^2 (Q_{\alpha \beta} Q_{\alpha \beta})^{1/2}} - \tau_0 \right) \label{eq:taut} \end{equation} where $C_S$ is the Smagorinsky constant, which for LB models can be taken to be $0.1$. This effectively abandons the single relaxation time approximation by making it a spatially and temporally varying quantity dependent on the local gradients of the fluid velocity \cite{yu_dns_2005}. \subsection{Multi-fluid models} The multi-fluid models discussed here are all diffuse interface models, without any explicit interface tracking -- the dynamics of the interface emerges naturally from the interactions between the fluid components. All models in this class use an additional lattice (whose mass fractions we will call $g_i$) to represent the second fluid species and nonlocal interactions between the lattices in the collision process. \subsubsection{Shan-Chen} In the Shan-Chen model, both lattices are used to represent fluid components, with the 0-th moment of $f_i$ and $g_i$ representing the density of the first (A) and second fluid components (B), respectively. The equilibrium function and relaxation schemes remain unchanged for both lattices, but an additional coupling term in the form of a body force: \begin{equation} \vec{F}_A(\vec{x}) = G \psi_A(\vec{x}) \sum_{i} w_i \psi_B(\vec{x} + \vec{e}_i) \vec{e}_i \label{eq:sc_force} \end{equation} is introduced into \eqref{eq:lbgk} (see Section~\ref{sec:body_forces} for a short overview of ways of adding body forces to an LB simulation), where $\psi_B(\vec{x})$ is a pseudopotential function, dependent on the density $\rho_B$ at node $\vec{x}$ and $G$ is a coupling constant. A similar term $\vec{F}_B$ of the form \eqref{eq:sc_force} with $\psi_A$ replaced by $\psi_B$ and vice versa is added to the collision operator for the second component. A commonly used pseudopotential function is $\psi(\rho) = 1 - e^{-\rho}$. The velocity of the fluid becomes a weighted average of the first moments of the distribution functions: $\vec{u} = \frac{\rho_A \vec{u}_A / \tau_A + \rho_B \vec{u}_B / \tau_B}{\rho_A / \tau_A + \rho_B / \tau_B}$, where $\vec{u}_A$, $\vec{u}_B$ and $\rho_A$, $\rho_B$ are velocities and densities computed respectively from $f_i$ and $g_i$ using \eqref{eq:lb_macro}, and $\tau_A$, $\tau_B$ are relaxation times for the two fluid components. \subsubsection{The free energy model} The free energy model is based on a Landau free energy functional for a binary fluid \cite{PhysRevE.78.056709}. In the LB realization, the 0-th moment of $f_i$ represents the density $\rho$ of both components, while the 0-th moment of $g_i$ is a smoothly varying order parameter $\phi$, with $-1$ indicating a pure first component and $1$ indicating a pure second component: \begin{equation} \rho = \sum_i f_i \qquad \phi = \sum_i g_i \qquad \rho \vec{u} = \sum_i \vec{e}_i f_i. \label{eq:free_energy_macro} \end{equation} The macroscopic dynamics of the fluid is described by the system of equations \begin{align} \begin{split} \partial_t \rho + \partial_{\alpha} (\rho u_{\alpha}) &= 0 \\ \partial_t (\rho u_{\alpha}) + \partial_{\alpha}(\rho u_{\alpha} u_{\beta}) &= -\partial_{\alpha} P_{\alpha \beta} + \partial_{\alpha} \{ \nu \rho (\partial_{\beta} u_{\alpha} + \partial_{\alpha} u_{\beta})\} \\ \partial_t \phi + \partial_{\alpha} (\phi u_{\alpha}) &= M \nabla^2 \mu \label{eq:free_energy_dynamics} \end{split} \end{align} where $M$ is a mobility parameter, $\mu$ is the chemical potential and the pressure tensor $P_{\alpha \beta}$ is defined as: \begin{equation} P_{\alpha \beta} = \left(p_0 - \kappa \phi \nabla^2 \phi - \frac{\kappa}{2} |\nabla \phi |^2 \right) \delta_{\alpha \beta} + \kappa \partial_{\alpha} \phi \partial_{\beta} \phi, \end{equation} and $p_0 = \rho/3 + a\left(\phi^2 / 2 + 3 \phi^4/4 \right)$ is the bulk pressure. In the corresponding LB scheme, distributions on both lattices are relaxed and streamed using the standard LBGK scheme \eqref{eq:lbgk} using the relaxation times $\tau_\rho = \tau_B + \frac{\phi + 1}{2} (\tau_A - \tau_B)$ and $\tau_\phi$, respectively. The following equilibrium functions \cite{PhysRevE.78.056709} are used to recover \eqref{eq:free_energy_dynamics} in the macroscopic limit: \begin{align} f_i^{\mathrm{eq}} &= w_i \left(p_0 - \kappa \phi \nabla^2 \phi + e_{i \alpha} u_\alpha \rho + \frac{3}{2} \left[ e_{i \alpha} e_{i \beta} - \frac{\delta_{\alpha \beta}}{3} \right] \rho u_{\alpha} u_{\beta} \right) + \kappa w_i^{\alpha \beta} \partial_{\alpha} \phi \partial_{\beta} \phi \\ g_i^{\mathrm{eq}} &= w_i \left( \Gamma \mu + e_{i \alpha} u_{\alpha} \phi + \frac{3}{2} \left[ e_{i \alpha} e_{i \beta} - \frac{\delta_{\alpha \beta}}{3} \right] \phi u_{\alpha} u_{\beta} \right) \label{eq:free_energy_eq} \end{align} where $\Gamma$ is a tunable parameter related to mobility via $M = \Gamma (\tau_\phi - 1/2)$, and the chemical potential $\mu = a( -\phi + \phi^3) -\kappa \nabla^2 \phi$. $\kappa$ and $a$ are constant parameters related to the surface tension $\gamma = \sqrt{8\kappa a / 9}$ and interface width $\xi = 2 \sqrt{2\kappa/a}$. The values of the weights $w_i$ and $w_i^{\alpha \beta}$ can be found in the paper by Kusumaatmaja and Yeomans~\cite{kusumaatmaja_contact_2008}. In order to minimize spurious currents at the interface between the two fluids, we use optimized gradient and Laplacian stencils \cite{PhysRevE.77.046702}. \subsection{Boundary conditions} Sailfish implements various boundary conditions, which taken together make it possible to model a wide range of physical situations: \begin{itemize} \item periodic boundary conditions (used to simulate a system spatially infinite along a given axis), \item no-slip (solid walls): half-way bounce-back, full-way bounce-back (both with configurable wetting when used with the free energy model), Tamm-Moth-Smith~\cite{chikatamarla_entropic_2013}, \item density/pressure: Guo's method, Zou-He's method, equilibrium distribution, regularized~\cite{PhysRevE.77.056703}, \item velocity: Zou-He's method, equilibrium distribution, regularized~\cite{PhysRevE.77.056703}, \item outflow: Grad's approximation~\cite{grad_outflow}, Yu's method~\cite{yu2005improved}, Neumann's, nearest-neighbor copy, ,,do nothing''~\cite{junk2008}. \end{itemize} \begin{figure} \centering \includegraphics[width=\textwidth]{bb.pdf} \label{fig:bb} \caption{The bounce-back scheme illustrated for the D2Q9 lattice. (1) full-way bounce-back, (2) half-way bounce-back. The bottom row of nodes within each group represents the bounce-back nodes, while the top row represents normal fluid. Black nodes are ,,dry'' (do not represent fluid, do not undergo relaxation), while grey nodes are ,,wet''. The dashed line represents the effective location of the no-slip boundary condition. For simplicity, only distributions originating from the top left node are shown in subsequent time steps. The steps are: (a) initial state, time $t$ (b) streaming, time $t+1$ (c) relaxation, time $t+1$ (arrows ending with disks represent the post-relaxation state), (d) streaming, time $t+2$.} \end{figure} The bounce-back method is a simple idea originating from lattice gas automatons -- a particle distribution is streamed to the wall node, and scattered back to the node it came from (see \figref{bb}). In the full-way bounce-back scheme the wall nodes do not represent any fluid (,,dry'' nodes) and are only used to temporarily store distributions before scattering. This approach is slightly easier to implement, but has the downside of the distributions needing two time steps to reach back to the originating fluid node (one step to reach the wall node, then another to be reflected back and streamed to the fluid node). The physical wall is effectively located between the last fluid node and the wall node. In contrast, in the half-way bounce-back the wall nodes do represent fluid, the wall is located beyond the wall node (away from the fluid domain), and the distributions are reflected in the same time step in which they reach the wall node~\cite{Ladd94A}. For other types of boundary conditions, the problem is more complex, as the mass fractions pointing into the fluid domain are undefined (there are no nodes streaming to them in these directions). Various schemes have been proposed to work around this problem. In the equilibrium approach, the missing incoming distributions are assumed to have the same values as those opposite to them (bounce-back), and the post-collision distributions are defined to be $f_i^{\mathrm{eq}}(\rho, \vec{u})$ with one of $\{\rho, \vec{u}\}$ being specified as part of the boundary condition, and the other calculated from the local distribution function. Zou's and He's method \cite{zouhe} uses the idea of the bounce-back of the non-equilibrium parts of the distributions to calculate the missing mass fractions. The regularized boundary conditions also use the non-equilibrium bounce-back idea, but only for an intermediate step to calculate the 2-nd moment of the non-equilibrium distribution, which is then used to reset all mass fractions at the boundary node via a regularization procedure. We refer the reader to the excellent paper by Latt et al.~\cite{PhysRevE.77.056703} for a detailed overview of these and other boundary conditions. The Grad's approximation method uses a formula dependent on the density, velocity and the pressure tensor $P_{\alpha \beta} = \sum_i e_{i\alpha} e_{i \beta} f_i$ to replace all distributions on the boundary node. The three macroscopic quantities can be taken to be those from the boundary node at the previous time step (less precise, fully local) or extrapolated from neighboring fluid nodes (more precise, nonlocal). Other outflow methods, such as the Neumann's boundary implementation or Yu's method, are also nonlocal and use a simple extrapolation scheme. The simplest methods (nearest-neighbor copy, do nothing) do not require extrapolation, but also give less control over the resulting macroscopic fields. \subsection{Body forces} \label{sec:body_forces} Sailfish implements two popular ways of adding a body force $\vec{F}$ to the simulation: Guo's method \cite{GuoForce}, and the Exact Difference Method (EDM) \cite{Kupershtokh2009965}. In both schemes, the actual fluid velocity $\vec{v}$ can be computed as \begin{equation} \rho \vec{v} = \sum_i \vec{e}_i f_i + \vec{F} / 2 = \rho \vec{u} + \vec{F} / 2, \label{eq:edm_momentum} \end{equation} and a force term $F_i$ is added to the right hand side of \eqref{eq:lbgk}. Guo et al. analyzed a number of popular schemes and concluded that the optimal choice for $F_i$ in an LBGK simulation is \begin{equation} F_i = \left(1 - \frac{1}{2\tau} \right) w_i \left[ \frac{\vec{e}_i - \vec{v}}{c_s^2} + \frac{\vec{e}_i \cdot \vec{v}}{c_s^4} \vec{e}_i \right] \cdot \vec{F}. \label{eq:guo_force} \end{equation} In the EDM, the body force term added to the collision operator is equal to the difference of equilibrium distribution functions computed using momentum after and before the action of the force $\vec{F}$: \begin{equation} F_i = f_i^{\mathrm{eq}}(\rho, \vec{u} + \Delta\vec{u}) - f_i^{\mathrm{eq}}(\rho, \vec{u}) \label{eq:edm_force} \end{equation} with $\Delta\vec{u} = \vec{F}/\rho$. The advantages of the EDM are its lack of any spurious terms in the macroscopic Navier-Stokes equations and its applicability to any collision operator (not only LBGK). \section{Software design and algorithms} Computational software is traditionally developed in one of two different paradigms. \textit{Prototype} code is written with relatively little effort in a high level environment such as Matlab or Python for exploratory or teaching purposes. Due to the relatively low performance of these environments, this type of code is unsuitable for large scale problems. In contrast, \textit{production} code is written in lower level languages such as C++ or Fortran in a higher effort process, which can sometimes span many years and involve teams of developers. The resulting programs are more efficient, but also longer, less readable, and potentially difficult to maintain. In our implementation of the lattice Boltzmann method for GPUs released under the name \textit{project Sailfish}, we took a hybrid approach. We use the Python programming language with a template-based code generation module (using the Mako language) and a computer algebra system (Sympy) to generate code in CUDA C or OpenCL. We chose Python because it is a very expressive language, with bindings to many system libraries and great support for GPU programming via the PyCUDA and PyOpenCL packages~\cite{klockner2012pycuda}. In Sailfish, Python is used for setting up the simulation (initial conditions, boundary conditions, selecting an LB model), for simulation control, communication (e.g. between compute nodes in a cluster) and input/output (loading geometry, saving simulation results). We also employ the NumPy package to perform matrix operations efficiently. \begin{lstlisting}[label=lst:equilibrium,language=Python,caption=BGK equilibrium function defined using Sympy expressions. \texttt{S} is a class with predefined symbols. \texttt{grid} is a class representing the lattice used for the simulation. Note that this code works for all lattices and dimensions. The discrete velocity vectors are stored in symbolic form in the \texttt{grid.basis} list. This allows for e.g. simple computation of dot products. Numerical coefficients in the formulas are also stored in symbolic form as rational expressions (e.g. \texttt{Rational(3, 2)}) instead of floating point values. This makes them exact and allows further symbolic simplification.] from sympy import Rational def bgk_equilibrium(grid): out = [] for ei, weight in zip(grid.basis, grid.weights): out.append(weight * (S.rho + S.rho * (3 * ei.dot(grid.v) + Rational(9, 2) * (ei.dot(grid.v))**2 - Rational(3, 2) * grid.v.dot(grid.v)))) return out \end{lstlisting} Mathematical formulas, such as \eqref{eq:lb_macro}, \eqref{eq:equilibrium}, \eqref{eq:free_energy_eq} are stored in the form of Sympy expressions. We decided to do this after noticing that computer code for numerical calculations is often very repetitive, e.g. \eqref{eq:equilibrium}, when evaluated on a D3Q19 lattice, would expand to 19 long lines of code, even though the initial formula easily fits in one line (see \lstref{equilibrium}). This approach makes the code easier to read (the formulas can also be automatically converted to LaTeX expressions), allows for automated consistency checks (e.g. one can easily, in a single line of code, verify that the 0th moment of $f^\mathrm{eq}$ is $\rho$), enables easy experimentation with code optimization techniques (formulas can be reorganized e.g. for speed or precision of floating-point operations), and enables code reuse (e.g. the formula for equilibrium or the bounce-back rule is written down once, but can generate code for various types of lattices, both in 2D and 3D, see \lstref{bb} for an example). Finally, the Mako template language is used to render the formulas into low-level CUDA/OpenCL code. The generated code is fairly small, contains comments, and is automatically formatted, making it suitable for instructional purposes as well as for one-off, ad-hoc modifications. This is in stark contrast to large, multimodular codes where the architecture of the system can be difficult to understand without extensive documentation. Sailfish was designed from scratch and optimized for modern massively parallel architectures, such as GPUs. We wanted to achieve high flexibility of the code, ease of use when running or defining new simulations, and not take any performance hits compared to code written directly in CUDA C or OpenCL. In order to achieve that, we decided to use run-time code generation techniques. This provides some isolation from low-level hardware details, important in the case of GPUs which are still evolving rapidly and changing with every new GPU architecture. It also makes it possible to generate optimized code on a case-by-case basis and to automatically explore parameter spaces to find optimal solutions, thus saving programmer time and increasing their productivity. With many details, such as the number of compute units, size of on-chip memory, speed of access patterns to on-chip and off-chip memory, host-device latency and bandwidth, memory bandwidth to computational performance ratio directly impacting the performance of the code, experimentation and microbenchmarking are necessary to find combinations of parameters that work well. We also considered other metaprogramming approaches to code generation, such as domain-specific languages, and templates in C++. We deemed the first solution to have too much overhead, and decided against the latter one since the expanded code cannot be saved for inspection and modification, and there were no open source computer algebra libraries providing a level of flexibility and sophistication comparable to Sympy. \begin{lstlisting}[multicols=2,label=lst:bb,escapechar=\!,language=C,morekeywords={endfor,in,\$},caption=Full-way bounce-back rule. Left column: source code in Mako. Right column: generated CUDA C code for the D2Q9 lattice.] ${device_func} inline void bounce_back(Dist *fi) float t; opp_i = grid.idx_opposite[i] b = grid.idx_name[opp_i] t = fi->${a}; fi->${a} = fi->${b}; fi->${b} = t; } !\vfill\columnbreak! __device__ inline void bounce_back(Dist * fi) { float t; t = fi->fE; fi->fE = fi->fW; fi->fW = t; t = fi->fN; fi->fN = fi->fS; fi->fS = t; t = fi->fNE; fi->fNE = fi->fSW; fi->fSW = t; t = fi->fNW; fi->fNW = fi->fSE; fi->fSE = t; } \end{lstlisting} \subsection{High-level simulation architecture} Sailfish takes extensive advantage of objected-oriented programming techniques for modularization and code reuse purposes. Each simulation is defined using two classes -- a simulation class, and a geometry class (see \lstref{ldc_example}). The simulation domain can be divided into cuboid subdomains, which do not need to fill the parts of the domain that do not contain any fluid (this makes it possible to handle complex geometries). The geometry class defines initial and boundary conditions for a single subdomain. The simulation class derives from a base class specific to the LB model used such as \texttt{LBFluidSim} (single component simulations) or \texttt{LBBinaryFluidFreeEnergy} (binary fluids using the free energy model). The simulation class can optionally also add body forces or define custom code to be run after selected steps of the simulation (e.g. to display a status update, check whether the steady state has been reached, etc). The base simulation class specifies the details of the used LB model, such as the form of the equilibrium distribution function (in symbolic form, stored as a Sympy expression), number and names of macroscopic fields (density, velocity, order parameter, etc), and names of GPU code functions that need to be called at every simulation step. When executed, every Sailfish simulation starts a controller process which parses any command line parameters and reads configuration files, decides how many computational nodes and GPUs are to be used (in case of distributed simulations), and how the subdomains are going to be assigned to them. It then starts a master process on every computational node (see \figref{system_arch}), which in turn starts a subdomain handler process for every subdomain assigned to the computational node. We use subprocesses instead of threads in order to avoid limitations of the Python runtime in this area (the global interpreter lock preventing true multithreading), as well as to simplify GPU programming. The subdomain handlers instantiate the simulation and geometry classes for their respective subdomains. The system state is initialized in the form of Numpy arrays describing the initial macroscopic fields, and these are then used to initialize the distribution functions on the GPU (using the equilibrium distribution function). Optionally, a self-consistent initialization procedure \cite{luo-initial} can be performed to compute the density field from a known velocity field. Once the system is fully initialized, the simulation starts running, with the subdomain handlers exchanging information in a peer-to-peer fashion as determined by the connectivity of the global geometry. \begin{lstlisting}[label=lst:ldc_example,language=Python,caption=2D Lid-driven cavity example in Sailfish.] from sailfish.subdomain import Subdomain2D from sailfish.node_type import NTFullBBWall, NTZouHeVelocity from sailfish.controller import LBSimulationController from sailfish.lb_single import LBFluidSim class LDCSubdomain(Subdomain2D): max_v = 0.1 def boundary_conditions(self, hx, hy): wall_map = (hx == self.gx-1) | (hx == 0) | (hy == 0) self.set_node((hy == self.gy-1) & (hx > 0) & (hx < self.gx-1), NTZouHeVelocity((self.max_v, 0.0))) self.set_node(wall_map, NTFullBBWall) def initial_conditions(self, sim, hx, hy): sim.rho[:] = 1.0 sim.vx[hy == self.gy-1] = self.max_v class LDCSim(LBFluidSim): subdomain = LDCSubdomain if __name__ == '__main__': LBSimulationController(LDCSim).run() \end{lstlisting} \begin{figure} \centering \includegraphics[width=0.9\textwidth]{system_overview.png} \caption{High-level architecture of a distributed Sailfish simulation. The simulation is divided into 4 subdomains. The controller process uses the execnet Python library to start machine master processes on two computational nodes, which then spawn children processes to handle individual subdomains. The master and subdomain handlers use the ZeroMQ library to communicate.} \label{fig:system_arch} \end{figure} \subsection{GPU architecture overview} Modern GPUs are massively parallel computational devices, capable of performing trillions of floating-point operations per second. We will now briefly present the architecture of CUDA-compatible devices as a representative example of the hardware architecture targeted by Sailfish. Other devices not supporting CUDA but supporting OpenCL are based on the same core concepts. CUDA devices can be grouped into three generations, called Tesla, Fermi, and Kepler -- each one offering progressively more advanced features and better performance. The CUDA GPU is organized around the concept of a streaming multiprocessor (MP). Such a multiprocessor consists of several scalar processors (SPs), each of which is capable of executing a thread in a SIMT (Single Instruction, Multiple Threads) manner. Each MP also has a limited amount of specialized on-chip memory: a set of 32-bit registers, a shared memory block and L1 cache, a constant cache, and a texture cache. The registers are logically local to the scalar processor, but the other types of memory are shared between all SPs in a MP, which allows data sharing between threads. Perhaps the most salient feature of the CUDA architecture is the memory hierarchy with 1-2 orders of magnitude differences between access times at each successive level. The slowest kind of memory is the host memory (RAM). While the RAM can nowadays be quite large, it is separated from the GPU by the PCIe bus, with a maximum theoretical throughput in one direction of 16 GB/s (PCI Express 3.0, x16 link). Next in line is the global device memory of the GPU, which is currently limited to several gigabytes and which has a bandwidth of about 100-200~GB/s. Global memory accesses are however high-latency operations, taking several hundred clock cycles of the GPU to complete. The fastest kind of memory currently available on GPUs is the shared memory block residing on MPs. It is currently limited in size to just 48~kB (16~kB on Tesla devices), but has a bandwidth of ca~1.3~TB/s and a latency usually no higher than that of a~SP register access. The above description readily suggests an optimization strategy which we will generally follow in the next section and which can be summarized as: move as much data as possible to the fastest kind of memory available and keep it there as long as possible, while minimizing accesses to slower kinds of memory. When memory accesses are necessary, it also makes sense to try to overlap them with independent computation, which can then be executed in parallel effectively hiding the memory latency. From the programmer's point of view, CUDA programs are organized into kernels. A kernel is a function that is executed multiple times simultaneously on different MPs. Each instance of this function is called a thread, and is assigned to a single scalar processor. Threads are then grouped in one-, two- or three-dimensional blocks assigned to multiprocessors in an 1-1 manner (1 block –- 1 MP). The blocks are organized into a one- or two-dimensional grid. The size and dimensionality of the grid and blocks is determined by the programmer at the time of kernel invocation. Knowledge of the grid position, and the in-block position makes it possible to calculate a thread ID that is unique during the kernel execution. Within a single block threads can synchronize their execution and share information through the on-chip shared memory. Synchronization between blocks is not supported in any way other than serializing the execution of kernels, and through atomic operations on global memory. \subsection{Low-level LB algorithms and data structures on GPUs} The great potential of modern GPUs as a hardware platform for simulating both two-dimensional \cite{tolke-cuda-twod} and three-dimensional flows \cite{tolke-GPU} with the lattice Boltzmann method was quickly realized after the initial release of the CUDA programming environment in 2007. Recently, GPUs have also been used to investigate more complex, two-dimensional multicomponent flows \cite{PhysRevE.80.066707}. There are three main factors that need to be carefully optimized in order to fully utilize the computational power of modern GPUs: memory access patterns, register utilization, and overlap of memory transfers and arithmetic operations. The first factor has received the most attention in the literature, but in this work we show that the remaining two are also important. The most important data structure used in an LB simulation are the particle distributions $f_i$. They are stored in the global GPU memory in a Structure of Arrays (SoA) fashion, effectively forming a 4D or 3D array (for 3D and 2D simulations, respectively), with the following index ordering: $(q, z, y, x)$, where $q$ is the number of discrete velocities in the lattice. An Array of Structures (AoS) approach, while elegant from the point of view of object-oriented programming, is completely unsuitable for GPU architectures due to how global reads and writes are performed by the hardware. Global memory accesses are performed by thread warps (32 threads) in transactions of 32, 64, or 128 bytes. On Fermi devices, accesses cached in the L1 cache are serviced with 128-byte transactions (the size of a full L1 cache line), while those cached in the L2 cache are serviced with 32-byte transactions. In order to attain good bandwidth utilization, the memory has to be accessed in contiguous blocks so that all bytes in a transaction are used for meaningful data. The memory location also has to be naturally aligned, i.e. the first address in the transferred segment must by a multiple of the segment's size. In Sailfish, we run all kernels in 1-dimensional thread blocks spanning the X axis of the subdomain, with a typical block size being 64, 128 or 192 nodes. Each thread handles a single node of the subdomain. Due to the layout of the distributions in global memory, we issue $q$ read requests to load a full set of mass fractions in a thread block. Each request results in a fully utilized memory transaction. To ensure natural alignment of transactions, the X dimension of the distributions array in global memory is padded with unused nodes so that $x$ is a multiple of 32 or 16, for single and double precision floating point numbers, respectively. In addition to the distributions array, we also store a node type map (see Section~\ref{sec:boundary}), and macroscopic field arrays in global memory, all following the same memory layout and padding as described above. A simple LB simulation in Sailfish repeatedly calls a single CUDA kernel called \texttt{CollideAndPropagate}, which implements both the collision and propagation step (see e.g. \eqref{eq:lbgk}), as well as any boundary conditions. A high-level summary of this kernel is presented in \algref{cnp}. \begin{algorithm*} \caption{A GPU kernel to perform one step of a lattice Boltzmann simulation, using shared memory for propagation in the $x$ direction. $\mathrm{offset}(x, y, z)$ is a function computing a linear index in a 1D array corresponding to a shift in the 3D subdomain space. $i$ is a thread index within the local block of threads (\texttt{threadIdx.x} in CUDA C). $e_{km}$ is the $m$-th component of the $k$-th discrete velocity vector.} \label{alg:cnp} \singlespacing \begin{algorithmic}[1] \State allocate shared memory array $\mathit{buf}_k[\mathit{blockSize}]$ \State compute global array index $i_G$ for the current node $i$ \State load and decode node type from global memory to $nt$ \State load distributions from global memory to $f$ \State compute macroscopic variables $\rho$ and $\vec{v}$ \If{$nt$ is a boundary node} \State apply boundary conditions \EndIf \If{output requested} \State save $\rho$ and $\vec{v}$ in global memory arrays \EndIf \State compute the equilibrium distribution $f^\mathrm{eq}$ using $\rho$ and $\vec{v}$ \State relaxation: $f \gets f + \frac{1}{\tau} (f^\mathrm{eq} - f)$ \For{$k$ such that $e_{k x} = 0$} \Comment{Propagation in directions orthogonal to X.} \State write $f_k$ to global memory at $i_G + \mathrm{offset}(0, e_{k y}, e_{k z})$ \EndFor \If{$i \ge 1$} \Comment{Propagation backward in the X direction.} \For{$k$ such that $e_{k x} = -1$} \State $\mathit{buf}_k[i-1] \gets f_{k}$ \EndFor \Else \For{$k$ such that $e_{k x} = -1$} \State write $f_k$ to global memory at $i_G + \mathrm{offset}(e_{k x}, e_{k y}, e_{k z})$ \EndFor \EndIf \State synchronize threads within the block \If{$i < \mathit{blockSize} - 1$} \For{$k$ such that $e_{k x} = -1$} \State write $\mathit{buf}_k[i]$ to global memory at $i_G + \mathrm{offset}(0, e_{k y}, e_{k z})$ \EndFor \For{$k$ such that $e_{k x} = 1$} \Comment{Propagation forward in the X direction.} \State $\mathit{buf}_k[i+1] \gets f_{k}$ \EndFor \Else \For{$k$ such that $e_{k x} = 1$} \State write $f_k$ to global memory at $i_G + \mathrm{offset}(e_{k x}, e_{k y}, e_{k z})$ \EndFor \EndIf \State synchronize threads within the block \If{$i > 0$} \For{$k$ such that $e_{k x} = 1$} \State write $\mathit{buf}_k[i]$ to global memory at $i_G + \mathrm{offset}(0, e_{k y}, e_{k z})$ \EndFor \EndIf \end{algorithmic} \end{algorithm*} Two basic patterns of accessing the distributions have been described in the literature, commonly called AB and AA~\cite{bailey_2009}. In the AB access pattern, there are two copies of the distributions in global memory (A and B). The simulation step alternates between reading from A and writing to B, and vice versa. In the AA access pattern, only a single copy of the distributions array is stored in global memory. Since there are no guarantees about the order of execution of individual threads, care has to be taken that a thread reads from and writes to exactly the same locations in memory. This is illustrated in~\figref{memory}. A third access pattern that is used in practice is the so-called \emph{indirect addressing}. In this mode, in order to access data for a node, its address has to be read first. This causes overhead both in storage (need to store the addresses) and in memory accesses, but can be very useful for complex geometries where the active nodes are only a tiny fraction of the volume of the bounding box. Indirect addressing can be combined with both AA and AB access patterns. For a dense subdomain using indirect addressing, the performance can be 5-25\% lower than when using direct addressing, with C2050 showing better results with the AA access pattern than in AB, and the GTX 680 exhibiting the opposite tendency. The exact performance is however necessarily geometry-dependent, and as such it is not discussed further here. \begin{figure} \centering \includegraphics{memory.pdf} \label{fig:memory} \caption{Memory layouts supported by Sailfish. In (a)-(d), dark arrows represent distributions used for calculations on the central node (middle square in every panel). In the AB layout, there are two copies of the lattice stored in memory. Data is read from the first lattice (a), and propagated into the second lattice (c). In the AA layout, there is only one lattice copy. Data is read from (a), stored as (b), and in the following step, read from (d) and stored as (c). In indirect addressing, there is a dense array of pointers to the actual distributions, which are stored in a continuous array (e). Dark nodes with black circles represent active (fluid) nodes. White nodes represent areas outside of the simulation domain which do not have corresponding distributions.} \end{figure} The propagation step of the LB algorithm shifts the distributions by $\pm 1$ node in all directions -- when this happens for the X axis, it results in misaligned reads/writes. To reduce the impact of misaligned writes, Sailfish utilizes shared memory to shift data in the X direction within the thread block (see~\algref{cnp}). We performed numerical tests to evaluate the impact of various data access patterns on the overall performance of the simulation. Earlier works \cite{Obrecht2010,ObrechtMemory} indicate that the cost of unaligned writes is significantly higher than the cost of unaligned reads, and that a propagate-on-read strategy results in up to 15\% performance gain. Our experiments confirmed this on older GT200 hardware (GTX 285), however we failed to replicate this effect on Fermi and Kepler devices (Tesla C2050, K10, K20), where the performance of both implementations was nearly identical (typically, with a few \% loss for propagate-on-read). This is most likely caused by hardware improvements in the Fermi and later architectures. We also compared the performance for the AB and AA access patterns. On Tesla-generation devices (GTX 285, Tesla C1060), the AA memory layout results in a slightly higher performance, and is therefore clearly preferred. On newer devices the AB scheme is typically a little faster, but the performance gains over the AA scheme are minor ($<10\%$), and as such the AB scheme is only preferable when ample GPU memory is available. \subsection{Boundary condition handling} \label{sec:boundary} \begin{figure} \centering \includegraphics[width=\textwidth]{encoding.pdf} \caption{Node parameter and type encoding into a single 32-bit unsigned integer to be used in the node type map. Bitfields (node type, node parameter index, ...) have adjustable size which can vary between simulations.} \label{fig:encoding} \end{figure} Boundary conditions are handled in Sailfish with the help of a \emph{node type map} -- an unsigned 32-bit integer array stored in the global GPU memory. Each entry in this array contains encoded information about the node type (fluid, unused, ghost, type of boundary condition to use), orientation (vector normal to the boundary, pointing into the fluid) and a parameter ID. The parameter ID is an index to a global array of values used by boundary conditions (e.g. densities, velocities). The encoding scheme uses variable-size bitfields, which are dynamically chosen for every simulation depending on the usage of different boundary conditions in a subdomain (see \figref{encoding}). Time-dependence is supported for all types of boundary conditions. When a value changing in time is required, it is typically specified in the form of a Sympy expression. This expression is then transformed into a GPU function and assigned an ID, analogous to the parameter ID for static boundary conditions. \subsection{Multicomponent models} Models with more than one distribution function, such as the Shan-Chen model or the free energy model introduce a nonlocal coupling via one or more macroscopic fields. To minimize the amount of data read from global memory, we split the simulation step into two logical parts, implemented as two GPU kernels. In the first kernel, \texttt{ComputeMacroFields}, we load the distributions and compute the macroscopic field value for every field which needs to be accessed in a nonlocal manner in the collision kernel. The collision step is implemented similarly to single fluid models, and the nonlocal quantities (Shan-Chen force, gradient and Laplacian of the order parameter in the free energy model) are computed by directly accessing the macroscopic field values in global memory. We considered three approaches to realizing the collision and propagation part of the algorithm for multifluid models. In the first variant, we used a single collision kernel, which loaded both distributions into registers and ran the collision and propagation step for both lattices. In the second variant, we tried to speed-up the calculations of nonlocal values by binding the macroscopic fields to GPU textures and accessing them through these bindings. In the last variant, we split the collision process into two separate kernels, one for every lattice. The second variant yielded minimal speed-ups (on the order of a few percent) on old Tesla-generation devices, which however did not carry over to Fermi and Kepler ones. We abandoned the idea as it introduced unnecessary complexity to the code. The third approach using split kernels proved to be the most efficient one. We were able to obtain 5.4\% (D2Q9, free energy) -- 53\% (D3Q19, free energy) speed-ups as compared to a monolithic kernel, mainly due to the kernels using fewer registers and being able to achieve higher occupancy, which hid the minor overhead introduced by a slightly increased number of global memory accesses. As a side benefit, this approach also resulted in simpler code, so we decided to standardize on it in the Sailfish framework. We also expect it to scale well to models requiring more than 2 lattices. \subsection{Distributed simulations} In Sailfish, the mechanisms that support distributed simulations are very similar to those supporting multiple subdomains on a single GPU or many GPUs on one computational node. Every subdomain has a layer of \emph{ghost} nodes around it. These nodes do not participate in the simulation, but are used for data storage for mass fractions leaving the subdomain or for macroscopic fields of neighboring nodes located in remote subdomains. Once mass fractions are streamed into the ghost nodes, we run additional CUDA kernels to collect the data leaving the subdomain into a linear buffer in global memory, which is then transferred back to the host and sent to remote nodes using a network link. The subdomain is also split into two areas called \emph{bulk} and \emph{boundary}, which are simulated via two separate kernel calls. The boundary area is defined as all nodes belonging to CUDA thread blocks where at least one node touches the subdomain boundary. The simulation step is first performed for nodes in the boundary area, so that communication with remote subdomain runners can start as soon as possible and can overlap in time with the simulation of the bulk of the subdomain. As a further optimization, we limit the data sent between computational nodes exactly to the mass fractions that actually need to be sent (e.g. if two nodes are connected along the X axis, then only the mass fractions corresponding to discrete velocities with a positive X component will be transferred from the left subdomain to the right one). We also make it possible to optionally compress the data before sending it to the remote node, which can be helpful if the simulation is run on multiple machines with slow network links between them, or when only a small fraction of nodes on the subdomain interface plane is active. \subsection{Impact of single precision} Many GPUs are significantly faster when calculations are done in single precision as opposed to the standard double precision which is typically used in scientific and engineering applications. The speed-up factor can vary between 10 and 2, depending on the device model and generation. The performance gap is smaller in newer devices (e.g. Kepler generation). Earlier works \cite{Kuznik20102380,Obrecht2011} used the lid-driven cavity benchmark to verify that single precision calculations produce satisfactory results. Here, we use the 2D Taylor-Green decaying vortex flow -- a benchmark problem with known analytical solution -- to study the accuracy of our LB implementation in both single (SP) and double (DP) precision. The simulations are done using the LBGK relaxation model on a D2Q9 lattice. For single precision, both the standard formulation and the round-off minimizing formulation \cite{DhumieresMrt2002} (SRO) are tested. The Taylor-Green vortex can be described by the following system of equations: \begin{align} \begin{split} u_x(t) &= -u_0 \cos(x) \sin(y) e^{-2 \nu t} \\ u_y(t) &= u_0 \sin(x) \cos(y) e^{-2 \nu t} \\ \rho(t) &= \frac{\rho}{4} \left(\cos(2x) + \cos(2y) \right) e^{-4 \nu t} \end{split} \label{eq:taylor_green} \end{align} where $u_0$ is a velocity constant and $x, y \in [0;2\pi)$. \eqref{eq:taylor_green} can be easily verified to satisfy the incompressible Navier-Stokes equations. We performed multiple runs of the simulation on a $256^2$ lattice with periodic boundary conditions in both directions, varying the viscosity $\nu$ but keeping the Reynolds number constant at $\mathrm{Re} = 1000$. Each simulation was run until it reached a specific point in physical time, corresponding to $10^6$ iterations at $u_0 = 0.0005$. We use the L2 norm of the velocity and density field difference to characterize the deviation of the numerical solution from the analytical one: \begin{equation} \epsilon = \sum_{\mathrm{nodes}}\sqrt{\frac{(u_x - \hat{u}_x)^2 + (u_y - \hat{u}_y)^2}{u_x^2 + u_y^2}} \label{eq:velocity_dev} \end{equation} where $\hat{u}_x$, $\hat{u}_y$ is the numerical solution. \begin{figure} \centering \includegraphics[width=\textwidth]{precision.pdf} \caption{Solution error for the Taylor-Green vortex test case with the LBGK model on a D2Q9 lattice, in single precision (SP), single precision with round-off minimization (SRO) and double precision (DP). Left panel: velocity error. Right panel: density error.} \label{fig:precision} \end{figure} The results presented in~\figref{precision} illustrate interesting differences between double precision and single precision calculations. In double precision, the error stays constant until $u_0 \approx 0.02$ and raises quadratically for higher values of velocity, as could be expected from the $O(\mathrm{Ma}^2)$ accuracy of the LBGK model. In single precision however, lower speeds lead to higher error values. The interplay between model accuracy and numerical round-off leads to a sweet spot around $u_0 \approx 0.05$, where the errors are minimized. For higher speeds, there is no difference between double and single precision. This result shows that the conventional wisdom that lower speeds always lead to better precision is not true when applied to single precision simulations. Since many practical simulations are run at velocities $0.05$ and higher for reasons of efficiency, it also explains why in many cases no differences are observed between single and double precision results. The density error shows that the round-off minimization model generates significantly better results than the standard single precision implementation, and as such should be preferred in cases where the accuracy in the density field is important. \section{Performance} All performance figures in the this section are quoted in MLUPS (Millions of Lattice-site Updates per Second). The tests were run using CUDA Toolkit 5.0, PyCUDA 2013.1.1 on 64-bit Linux systems. \subsection{Comparison of different models} \begin{figure} \centering \includegraphics[width=\textwidth]{model_benchmark.pdf} \caption{Performance comparison of different models using the D3Q19 lattice, AB memory access pattern. Used acronyms: ELBM: entropic LBM, SC: Shan-Chen, FE: free energy. All GPUs had ECC disabled. ELBM used intrinsic functions as described in section~\ref{sec:intrinsic}. Both logical GPUs were used for the K10 tests. Left panel: single precision. Right panel: double precision.} \label{fig:perf_models} \end{figure} \begin{figure} \centering \caption{Performance of an LBGK simulation with the AA memory access pattern as a function of lattice type. The lower performance for the K10 simulation for D3Q13 and D3Q15 lattices is caused by overhead in copying the data between the two GPUs through the host (both logical GPUs were used for the K10 tests). Test case: Kida vortex (see text). Left panel: single precision. Right panel: double precision.} \includegraphics[width=\textwidth]{memory_grid.pdf} \label{fig:memory_grid} \end{figure} Single fluid models were benchmarked using a lid-driven cavity test case, with a $254^3$ lattice at $\mathrm{Re} = 1000$, using full bounce-back walls and equilibrium velocity boundary conditions for the lid. To illustrate the impact of the entropy-stabilized time-stepping, the ELBM solver was also tested at $\mathrm{Re} = 10000$. The binary fluid models were benchmarked using a simple spinodal decomposition test case, where a uniform mix of both fluid components fills the whole simulation domain ($254 \times 100 \times 100$) and periodic boundary conditions are enabled in all directions. Whenever the domain size was too large to fit on a specific GPU we tested, we reduced the size in the Z direction until it fit. We find our results to be comparable or better to those reported by Habich et al.~\cite{Habich2012}, who used hardware very similar to ours (Tesla C2050 and C2070 differ only in memory size, according to NVIDIA specifications). The performance of single precision simulations can be shown to be limited by global memory bandwidth. We find that our code runs at $\sim80\%$ of the theoretical bandwidth as reported in NVIDIA whitepapers, and close to 100\% of the real bandwidth measured by sample code from the NVIDIA SDK. Double precision simulations run at $\sim60\%-80\%$ of the theoretical maximum. They are limited by the double precision instruction throughput on the GPU on Fermi-class hardware and by memory bandwidth in Kepler hardware. Overall, we find that the memory bandwidth is a reliable indicator of expected performance for single precision simulations across all three generations of NVIDIA devices (see also~\figref{memory_grid}, which shows that the simulation performance is inversely proportional to the lattice connectivity $Q$ in single precision). For double precision simulations on Fermi hardware, we have found increasing the L1 cache size to 48 kB, disabling L1 cache for global memory accesses and replacing division operations by equivalent multiplication operations to have a large positive impact on the performance of the code ($\sim 3.5$ speed-up in total). The impact of the L1 cache is understandable if one considers the fact that double precision codes use significantly more registers. This causes the compiler to spill some of them over to local memory, accesses to which always go via the L1 cache. The larger the size of the unused part of that cache, the more operations can execute without actually accessing the global memory. This optimization strategy does not apply to Kepler-class devices, where both in single and double precision we found that disabling the preference for L1 cache for the main computational kernel had a positive impact on performance. We also tested our code on lower end GPUs (mobile versions used in laptops) with compute capability 3.0, where we found a significant speed up (up to 40\%) by using shuffle operations for in-warp propagation, and limiting shared memory for data exchange between warps only. This optimization strategy does not work with the higher end GPUs discussed in this paper, where the performance is limited by global memory bandwidth already. Overall, we unsurprisingly find that more recent GPUs perform noticeably better. The K10 is an interesting option if only single precision is required as it delivers the highest overall performance at a level higher than 1.3 GLUPS per board with D3Q19 and simple fluid models. For double precision, the K20x card being the most advanced NVIDIA GPU available on the market when this paper is written, is a clear winner performance-wise. \subsection{Scaling on GPU clusters} While the computational power provided by a single GPU is impressive, practical simulations often require large domains, and for these the total size of GPU memory (a few GBs) is an important limitation. The natural solution of this problem is to run a distributed simulation using multiple GPUs, which can be physically located in a single computer or multiple computers in a network. In order to measure performance of distributed simulations, we ran a 3D duct flow test case (periodic boundary conditions in the streamwise Z direction, bounce-back walls at other boundaries) on the Zeus cluster (part of the PLGRID infrastructure), consisting of computational nodes with 8 M2090 GPUs and interconnected with an Infiniband QDR network. \subsubsection{Weak scaling} The first test we performed measured weak scaling, i.e. code performance as a function of increasing size of the computational domain. The domain size was $254 \times 127 \times 512 \cdot N$, where $N$ is the number of GPUs. We used the D3Q19 lattice, the AA memory access pattern, a CUDA block size of 128, and single precision. \figref{weak_scaling} shows excellent scaling up to 64 GPUs, which was the largest job size we were able to run on the cluster. The 1.5\% efficiency loss takes place as soon as more than 1 subdomain is used, and does not degrade noticeably as the domain size is increased. This small efficiency loss could be further minimized by employing additional optimization techniques, such as using peer-to-peer copies when the GPUs are located on the same host. \begin{figure} \centering \includegraphics[width=\textwidth]{weak_scaling.pdf} \caption{Weak scaling properties of the Sailfish code. The simulation was run on nodes with 8 M2090 GPUs, with one subdomain used for every GPU. Test case: duct flow. Left panel: absolute performance values. Right panel: efficiency fraction.} \label{fig:weak_scaling} \end{figure} \subsection{Strong scaling} The second test we ran measured strong scaling, i.e. code performance with a constant domain size, as a function of increasing number of GPUs. We used a $254 \times 127 \times 1664$ geometry (largest domain size that fit within the memory of a single M2090 GPU) and other settings as in the weak scaling test, and divided the domain into equal-length chunks along the Z axis as more GPUs were used for the simulation. The results of this test are presented in \figref{strong_scaling}. A slightly worse performance is visible in comparison to the weak scaling test, but even when the simulation is expanded to run on 8 GPUs, only a 3.5\% efficiency loss can be observed. It should also be noted, that there is a minimum domain size below which performance quickly degrades due to the overhead of starting kernels on the GPUs. This size can be seen in \figref{min_domain_size} to be about 14\% of the GPU memory or 8.2 M lattice nodes. \begin{figure} \centering \includegraphics[width=\textwidth]{strong_scaling.pdf} \caption{Strong scaling properties of the Sailfish code. One subdomain was used for every M2090 GPU. Test case: duct flow. Left panel: absolute performance values. Right panel: efficiency fraction.} \label{fig:strong_scaling} \end{figure} \begin{figure} \centering \includegraphics[width=0.5\textwidth]{single_domain.pdf} \caption{Duct flow simulation performance as a function of used memory fraction of a single M2090 GPU. The vertical line shows the point at which the computational capabilities of the GPU are saturated. The geometry was the same as in the strong scaling test and the memory fill fraction was controlled by varying the extent of the domain in the Z direction. Very similar behavior is observed for K10 and K20 GPUs (not shown here).} \label{fig:min_domain_size} \end{figure} \subsection{Further optimization with intrinsic functions} \label{sec:intrinsic} CUDA GPUs provide an alternative hardware implementation of various transcendental functions such as the exponent, logarithm or trigonometric functions. These functions, known as intrinsic functions, are faster but less precise than their normal counterparts, especially when their arguments do not fall within a narrow range specified for each function. We analyzed the impact of these functions on the performance and precision of the LB models that can take advantage of them, namely the Shan-Chen model with a non-linear pseudopotential and the entropic LBM. With ELBM the use of intrinsic functions, together with the FMAD (fused multiply-add) instruction yields a speed-up of $\sim43$\% without any noticeable impact on the correctness of the results (in terms of global metrics such as the total kinetic energy, enstrophy and the kinetic energy spectrum). While testing these optimizations, we also found that the FTZ (denormalize to 0) option of the CUDA compiler causes the ELBM simulation to crash. With the proposed optimizations, the performance of ELBM is at 72\% of the LBGK performance, making it a very interesting alternative for some simulations. For the Shan-Chen model with a nonlinear pseudopotential we saw 17-20\% speed-ups in 2D and 3D, with relative changes in the density fields smaller than 1.5\% after 10000 steps. Unfortunately, the same approach does not yield speed-ups in double precision, as most intrinsic functions are available in single precision only. \section{Validation} In order to validate our implementation of the LBM, we performed simulations for four classical computational fluid dynamics test cases and compared our results with those published in the literature. \subsection{Lid-driven cavity} The lid-driven cavity geometry consists of a cube cavity with a face length $L$. The geometric center of the cavity is located at the origin of the coordinate system. The cube face at $x = - L/2$ moves tangentially in the y-direction with a constant velocity $v$, while all other faces are no-slip walls. We carried out simulations of this problem at $\mathrm{Re} = 1000$ with various LB models (BGK, MRT, regularized BGK, ELBM) using a $201^3$ D3Q19 lattice, full-way bounce-back for no-slip walls, and the regularized velocity boundary condition with $u_y = 0.05$ for the moving wall. Our results (both in single and double precision) agree with those published by Albensoeder et al.~\cite{Albensoeder2005536} (see \figref{ldc-comparison}) \begin{figure} \centering \includegraphics[width=0.5\textwidth]{ldc_comparison.pdf} \caption{Lid-driven cavity velocity profiles $u_y(x, 0, 0)$ and $u_x(0, y, 0)$. Round dots: data from Tables~5 and~6 in Albensoeder et al.~\cite{Albensoeder2005536} Solid line: results from Sailfish simulations on a $201^3$ lattice, after $2 \cdot 10^5$ steps using LBGK, MRT, regularized BGK and ELBM in single and double precision. The results from all Sailfish simulations are in agreement to within the width of the line on the plot. LB results are rescaled using: $\vec{u} = \vec{u}_\mathrm{LB} / 0.05$.} \label{fig:ldc-comparison} \end{figure} \subsection{Kida vortex} The Kida vortex flow is a free decay from the initial conditions~\cite{Kida1985}: \begin{align*} u_x(\vec{x}, 0) &= u_0 \sin x \left( \cos 3 y \cos z - \cos y \cos 3 z \right) \\ u_y(\vec{x}, 0) &= u_0 \sin y \left( \cos 3 z \cos x - \cos z \cos 3 x \right) \\ u_z(\vec{x}, 0) &= u_0 \sin z \left( \cos 3 x \cos y - \cos x \cos 3 y \right) \end{align*} defined on a cubic domain with face length $2 \pi$, with periodic boundary conditions in all directions. To validate our code, we performed simulations for $\mathrm{Re} = N u_0 / \nu = 4000$, $1.28 \cdot 10^4$, and $1.28 \cdot 10^5$ and compared them with results published in \cite{ChikatamarlaKidaDNS} and \cite{PhysRevE.75.036712}, respectively. The simulations were run using $u_0 = 0.05$ on a $700^3$ grid ($350^3$ for $\mathrm{Re} = 4000$), using both single and double precision (with no noticeable difference between them). The $\mathrm{Re} = 4000$ and $\mathrm{Re} = 1.28 \cdot 10^4$ cases were investigated using the LBGK, MRT, regularized LBGK, Smagorinsky-LES and entropic models. At $\mathrm{Re} = 1.28 \cdot 10^5$, only simulations using the entropic model and the Smagorinsky subgrid model (with $C_S = 0.1$) remained stable. During the simulation time kinetic energy $E = \frac{1}{2 V} \int d^3 x \vec{v}^2$ and enstrophy $\Omega = \frac{1}{2 V} \int d^3 x (\vec{\nabla} \times \vec{v})^2$, where $V$ is the volume of the simulation domain, were tracked directly on the GPU for optimal efficiency. Vorticity was computed using the first order central difference scheme in the bulk of the fluid and forward/backward differences at subdomain boundaries. \begin{figure} \centering \includegraphics[width=\textwidth]{kinetic_energy_enstrophy.pdf} \caption{(a) evolution of normalized kinetic energy, (b) evolution of normalized enstrophy, (c) evolution of normalized enstrophy at $\mathrm{Re} = 12800$ for various collision models, (d) kinetic energy spectrum for selected collision models and Reynolds numbers. For panels (a)-(c) time is rescaled assuming a domain size of $(2 \pi)^3$ and $u_0 = 1$.} \label{fig:ke-ens} \end{figure} For $\mathrm{Re}=4000$ all four models gave the same results (\figref{ke-ens}). At $\mathrm{Re} = 1.28 \cdot 10^4$ some minor differences are visible, particularly in the evolution of enstrophy. Its peak value is slightly underpredicted by both models that locally modify effective viscosity (Smagorinsky, ELBM) (see \figref{ke-ens}(c)). At $\mathrm{Re} = 1.28 \cdot 10^5$, the differences are more pronounced and we observe that the Smagorinsky model underpredicts the absolute value of peak enstrophy. The kinetic energy spectrum shown on \figref{ke-ens}(d) was computed as $E(k) = \sum_{k \leq k' < k+1} \hat{u}(k)^2$, for $k = 0, 1, 2, \ldots$. A good agreement is visible in comparison to the Kolmogorov scaling $k^{-5/3}$, especially for the high Reynolds number cases. All collision models lead to similar spectra, with ELBM at $\mathrm{Re} = 4000$ predicting a slightly higher value around $k = 10$ than LBGK or other models, and with ELBM keeping a slightly flatter spectrum for high $k$ values at $\mathrm{Re} = 1.28 \cdot 10^5$. In all cases the simulation results show the same features as those discussed in previous papers on this topic~\cite{Kida1985,ChikatamarlaKidaDNS,PhysRevE.75.036712}. \subsection{Binary Poiseuille flow} To verify the binary fluid models we consider a 2D Poiseuille flow in the $x$-direction. No-slip walls are imposed at $y = \pm L$ using the half-way bounce-back boundary conditions, and periodic boundary conditions are used in the $x$-direction. A body force $G$ drives the flow. In the core flow region ($|y| \leq L / 2$) a fluid of viscosity $\nu_1$ is placed, while in the boundary flow region ($L / 2 < |y| < L$) the viscosity of the fluid is $\nu_0$. The analytical solution for this case can be expressed as: \begin{equation} u_x(y) = \begin{cases} \frac{G}{2 \rho \nu_0} \left(L^2 - y^2\right) &\mbox{if } L / 2 < |y| < L \\ \frac{G L^2}{8} \left( \frac{3}{\rho \nu_0} + \frac{1}{\rho \nu_1} \left( 1 - 4 \frac{y^2}{L^2} \right) \right) &\mbox{if } |y| \leq L/2. \end{cases} \label{eq:poiseuille_solution} \end{equation} We run the simulation on a $64 \times 256$ grid with $\nu_0 = 1/6$, and $\nu_1 = 1/24$. The simulation starts with $u_x = 0$ in the whole domain and we let the flow reach the stationary state on its own. The free energy simulation was run with $\Gamma = 25$, $\kappa = 10^{-4}$, and $A = 32 \cdot 10^{-4}$, while for the Shan-Chen model, $G = 1.5$ was used. The parameters were chosen to ensure that the interface remains as sharp as possible without destabilizing the simulation. The Exact Difference Method was used to introduce body forces. The Shan-Chen model permits residual mixing between the fluid components, so the effective density and viscosity were calculated as $\rho = \rho_0 + \rho_1$ and $\nu \rho = \rho_0 \nu_0 + \rho_1 \nu_1$, respectively. \figref{poiseuille_comparison} illustrates the good agreement of the simulation results with \eqref{eq:poiseuille_solution}. \begin{figure} \centering \includegraphics[width=\textwidth]{poiseuille_comparison.pdf} \caption{Comparison between simulated and theoretical velocity profiles for the binary Poiseuille flow. Left panel: Results for the free energy model. Right panel: Shan-Chen results. The effective viscosity ratio was $3.7985$ due to mixing between the two fluid components. The free energy model shows better agreement with the theoretical profile due to a thinner interface.} \label{fig:poiseuille_comparison} \end{figure} \subsection{Capillary waves} In order to verify the binary fluid models also in a dynamic case, we simulate the decay of the amplitude of a capillary wave. The boundary conditions of the system are the same as in the binary Poiseuille flow case, but we now use a larger $512 \times 512$ lattice in order to accommodate higher frequency waves. The region $y > 0$ is filled with one component (A) and the region $y < 0$ is filled with another component (B). For simplicity, we choose both components to have the same viscosities $\nu = 1/18$. The interface between the two fluids is initialized to a sinusoidal curve of wavelength $\lambda$ chosen such that an integer number of wavelengths fits exactly in the simulation domain. The interface is then allowed to relax naturally with no external forces, resulting in a damped capillary wave. At each timestep of the simulation, the height of the interface at $x = L - \lambda / 4$ is recorded. In order to recover the frequency of the wave, an exponentially damped sine function is fit to the interface height data. For the Shan-Chen model, we used $G = 0.9$ and for the free energy model we used $A = 0.02$, $\kappa = 0.04$, and $\Gamma = 0.8$. \begin{figure} \centering \includegraphics[width=0.5\textwidth]{capillary.pdf} \caption{Theoretical and measured dispersion relation $\omega(k)$ for the capillary wave for the free energy and Shan-Chen models.} \label{fig:capillary} \end{figure} As expected \cite{Langaas_Yeomans,chin_binary_shan_chen}, the dispersion relation shows a power law form $\omega \propto k^{3/2}$ for both the Shan-Chen and free energy models (see \figref{capillary}). \section{Conclusions} In the previous sections we have demonstrated our Sailfish code as a flexible framework for implementing lattice Boltzmann models on modern graphics processing units. With novel optimization techniques for complex models it provides a very efficient tool for a wide range of simulations. We hope that our observations collected while running the presented benchmarks will serve as a guideline in the choice of both LB models and computational hardware for users of Sailfish and of other similar codes. For single precision simulations, we advocate a careful choice of parameters and correctness testing via comparisons to similar test cases in double precision. While all of our benchmark problems did not show any noticeable differences between single and double precision, the Taylor-Green test case clearly demonstrates that these do exist and can significantly impact the results if the simulation is in the slow velocity regime. Whenever possible, the round-off minimizing model should be used to reduce precision losses without any impact on performance. While the capabilities of Sailfish are already quite extensive, much work remains to be done. Among the most important remaining tasks, we mention ongoing efforts to implement a hierarchical lattice, allowing for local grid refinement and a fluid-structure interaction model based on the immersed boundary method. Since the code is freely available under an open source license, we would like to invite the reader to participate in its development and contribute new enhancements according to their interests. \subsection{Acknowledgments} This work was supported by the TWING project co-financed by the European Social Fund, as well as in part by PL-Grid Infrastructure. M.J. thanks S.~Chikatamarla and I.~Karlin for useful discussions and sharing reference data from their simulations. The authors would also like to thank NVIDIA for providing hardware resources for development and benchmarking. \bibliographystyle{elsarticle-num}
\section{Introduction} One of the very rare exactly solvable non-equilibrium systems is the totally asymmetric simple exclusion process (TASEP), see \cite{Mallick_rev} and references therein. The model is defined on a 1d discrete lattice with $L$ sites that are either occupied by a single particle or empty. In the latter case they can be thought of as occupied by a single hole. In its version with parallel update the TASEP is equivalent to a special case of the Nagel-Schreckenberg model for traffic flow \cite{NaSch}. Here all bonds are updated simultaneously while in the original formulation, the random-sequential update, only one particle move can occur per infinitesimal time-step. In general, particles enter the lattice on site $1$ with probability $\alpha$ and leave the lattice from site $L$ with probability $\beta$. In between particles on sites $l$ ($l=1,2,\dots, L-1$) may move with probability $p$ if the target site $l+1$ is empty. For later use we introduce also the symbol $q=1-p$ which denotes the probability that any of the possible moves is not executed. Note that for random-sequential update those probabilities are replaced by rates. In the thermodynamic limit $L\rightarrow \infty$ there are three different phases: a low-density phase, a high-density phase and a maximum-current phase. Here, the maximum-current phase is special, since all physical quantities become independent of the input and output probabilities $\alpha$ and $\beta$ -- the system behaves as if $\alpha=\beta=p$. The average density in the maximum-current phase is $1/2$. In the limit $p\rightarrow 0$ only one move per time-step occurs and one recovers the well-studied TASEP in continuous time.\\ While for random updating many results are known there is a definite lack of results for the parallel updating scheme \cite{Blythe}. One reason is that the structure of the exact solution \cite{ERS}, \cite{DeGier} appeared to be much more involved. Based on a simplified formulation of the exact solution \cite{woelki_par} the present article tries to close this gap a little more since especially the parallel update is important for practical modeling of traffic \cite{NaSch}. In this article we calculate as the main result the weight for $N$ particles and $M$ holes in the open system. For continuous time this quantitiy has been calculated by Derrida et. al. twenty years ago in \cite{Derrida_2class}. It further has a natural interpretation as the normalization of a related system on a ring with one second-class particle. This normalization is given by so-called Narayana numbers \cite{Blythe_dyck}. We will derive the analogous results for the parallel case here. Recently a connection between equilibrium lattice walks and the normalization of the TASEP in different variants has been established. One of those interpretations involves so-called Motzkin paths. A Motzkin path is a path defined on the triangular lattice. It starts at $(0,0)$ and ends at $(n,0)$ with never going below the horizontal axis. Possible are the steps $\{U,L,D\}$, where $U=(1,1)$ is an up-step, $L=(1,0)$ is a level step and $D=(1,-1)$ is a down step. Let $M(n)$ denote the set of those Motzkin paths. Then their number $|M(n)|$ is given by the $n$-th Motzkin number. The problem can also be formulated as a random walk in 1D that starts and ends at site 0 (or 1). If the horizontal axis represents time, the two-dimensional walk corresponds to the trajectory of the walker. A level step means that the walker has not moved during the time-step. An up-step (down-step) means that the walker increases (lowers) its coordinate by one. A Brownian excursion \cite{Derrida_2class}, \cite{Derrida_Brownian} is a special Motzkin path that never even touches the horizontal axis apart from the first and last vertex. It can simply be constructed from a Motzkin path by adding an up-step at the beginning and a down-step at the end \cite{Blythe_dyck}. In this paper we will find a Motzkin path interpretation of the generating function for $N$ particles on an open lattice of $L$ sites. For this interpretation we use a coloring of those different steps as it was considered in \cite{Schork}.\\ The paper is organized as follows. First, we consider the generating function of the open TASEP and show how it is related to the generating function of weighted Motzkin paths. After taking the thermodynamic limit we obtain expressions for the generating function of weighted Motzkin paths at given length. We calculate the TASEP-generating function for the weight of configurations with given number of particles. Then we see how this function is interpreted in terms of a coloring of the corresponding Motzkin path. We write an expression for the weight that the open TASEP system contains $N$ particles. Finally we introduce a second-class particle dynamics on the ring that has conserving dynamics and relate it to the one obtained for the TASEP in the thermodynamic limit. \section{The TASEP and the matrix ansatz} The matrix representation that Derrida et. al. \cite{Derrida_matrix} presented for the known recursion relations of the TASEP came as a very elegant and compact formulation of a non-equilibrium steady state. Since then, this technique has become very successful to calculate many stationary properties for the TASEP and related models, see \cite{Blythe} for a recent review. The authors of \cite{Derrida_matrix} found that the stationary weight for a lattice configuration $(\tau_1, \dots, \tau_L)$ of the TASEP with random-sequential update can be written as $F(\tau_1,\dots, \tau_{L})=\langle W| \prod_{i} (\tau_i D+(1-\tau_i)E )|V\rangle.$ In this notation the matrix $D$ represents occupied sites ($\tau=1$) and the matrix $E$ represents empty sites $\tau=0$. The boundary vectors $\langle W|$ and $|V\rangle$ ensure that the resulting matrix product is reduced to a scalar $F$. Those weights $F$ are stationary if the matrices and vectors involved satisfy the set of relations $DE=D+E$, $\langle W|E=\alpha^{-1}\langle W|$ and $D|V\rangle=\beta^{-1}|V\rangle$, now widely known as the DEHP-algebra. The matrices and vectors are half-infinite and there are different representations that have certain combinatorial interpretations \cite{Brak}. Note that $\langle 1|(D+E)^L|1 \rangle$ is given by Catalan numbers. It was shown in \cite{Derrida_2class} for the case of a single second-class particle on the ring that the weight for a fixed number of particles $N$ and $L-N$ holes is given by \begin{equation} \label{ra} [x^Ny^{L-N}]\langle 1|(xD+yE)^{L}|1 \rangle = \frac{1}{L+1}{L+1 \choose N}{L+1 \choose N+1}. \end{equation} Here $x$ and $y$ are the fugacities for particles and holes respectively. The combinatorial term on the right-hand side is known as the Narayana numbers \cite{Blythe_dyck}. Among many other applications they are known to count Dyck paths -- those are Motzkin paths without level steps. In \cite{Derrida_Brownian} this formula (or more precisely the equivalent formula of \cite{Mallick}) was used for the TASEP with open boundaries, too. The reason why in both models the weights for $N$ particles are the same is that the DEHP algebra holds for both models under the same Ansatz. Now to the parallel-update TASEP. The exact stationary state is known to be of a matrix form, too \cite{ERS,DeGier}. From the findings in \cite{woelki_par} it follows that the weight of a configuration in the maximum-current phase can be written as \begin{equation} \label{ans} F(\tau_1,\dots, \tau_{L})= q^{-W(\{\tau\})}\langle W| \prod_{i}( \tau_i D+(1-\tau_i)E) |V\rangle \end{equation} with $q=1-p$. In this paper we interpret $W(\{\tau\})$ as the number of particle-hole domain walls in the configuration $\{\tau\}$. This is one possible choice following \cite{woelki_par}. Again the vectors $\langle W|$ and $|V\rangle$ reflect the boundaries and the matrices $D$ and $E$ represent particles and holes respectively but are different from those for the random-sequential case. The operators obey the relations found by Evans et. al. in \cite{ERS} \begin{eqnarray} \label{WE} \langle W|E &=& q \langle W|,\\ \label{DE} DE &=& q\left[D+E+p \right],\\ \label{DV} D|V\rangle &=& q |V\rangle \end{eqnarray} which generalize the DEHP algebra. It was shown in \cite{ERS} that the normalization can be expressed as \begin{equation} \label{norm} Z_L=z_L+pz_{L-1}, \quad{\rm with}\; z_L=\langle 1|(D+E+p)^L|1 \rangle = \sum\limits_{t=0}^{L}\frac{1}{L+1}{L+1 \choose t}{L+1 \choose t+1}q^t. \end{equation} In contrast to (\ref{ra}) the term on the right-hand side is the $(L+1)$th Narayana polynomial. Note that a possible representation is \begin{eqnarray} \label{repres} E=\left(\begin{array} {rrrrr} q & 0 & 0 & 0 & \dots\\ \sqrt{q} & q & 0 & 0 & \dots\\ 0 & \sqrt{q} & q & 0 & \dots\\ 0 & 0 & \sqrt{q} & q & \dots\\ \dots & \dots & \dots & \dots & \dots \end{array} \right),\; D=\left(\begin{array} {rrrrr} q & \sqrt{q} & 0 & 0 & \dots\\ 0 & q & \sqrt{q} & 0 & \dots\\ 0 & 0 & q & \sqrt{q} & \dots\\ 0 & 0 & 0 & q & \dots\\ \dots & \dots & \dots & \dots & \dots \end{array} \right) \end{eqnarray} along with $\langle W|=\langle 1|$ and $|V\rangle =|1 \rangle$. This representation is a direct generalization of one known representation for the continuous-time case $q\rightarrow 1$ \cite{DeGier}. \section{The TASEP and fixed number of particles} To calculate the weight for $N$ particles on a lattice of size $L$, our aim is to study $\mathcal{Z}_L(x,y)=\sum_{\{\tau\}} F_L(\{\tau\})\delta_{\sum\tau,N}x^Ny^{L-N}$. For details of the following steps, the reader is referred to Appendix \ref{prelim}. First, we note (\ref{ZS}) that $\mathcal{Z}_L(x,y)$ is given by \begin{equation} \label{recm} \mathcal{Z}_L(x,y)=m_L( x,y) + p(x+y)m_{L-1}(x,y) + p^2xym_{L-2}(x,y). \end{equation} Our intention is to study the generating function \begin{equation} \label{Mxy} \mathcal{M}(x,y,\lambda,\mu) :=\sum\limits_{L} m_L(x,y,\mu) \lambda^L = \sum_{n\geq 0}(\lambda\mu)^n \langle 1|\left(xD+yE+\frac{p}{q}xy\lambda DE\right)^n |1 \rangle =: \sum_{n\geq 0}\langle 1|\tilde{C}^n|1 \rangle \mu^n \end{equation} which follows from (\ref{ab}). The formula for the generating function shows its two 'faces'. On the one hand $\mathcal{M}$ has an interpretation as contributing to the grand-canonical normalization $\sum_L\mathcal{Z}_L\lambda^L$ for TASEPs of size $L$ (in terms of the TASEP fugacity $\lambda$) and on the other hand $\mathcal{M}$ will be interpreted as the generating function of Motzkin paths of length $n$ with fugacity $\mu$. We could also take $\lambda=1$ since the exponent of $\lambda$ can already be worked out by replacing $x\rightarrow x\lambda$ and $y\rightarrow y\lambda$. However at this stage it will turn out useful. \subsection{Motzkin path interpretation} The use of the explicit representation (\ref{repres}) yields the tri-diagonal matrix \begin{eqnarray} \label{tildeC} \tilde{C} = \left(\begin{array} {lllll} w_l(x,y,\lambda) & w_u(x,y,\lambda) & 0 & 0 & \dots\\ w_d(x,y,\lambda) & w_l(x,y,\lambda) & w_u(x,y,\lambda) & 0 & \dots\\ 0 & w_d(x,y,\lambda) & w_l(x,y,\lambda) & w_u(x,y,\lambda) & \dots\\ \dots & \dots & \dots & \dots & \dots \end{array} \right). \end{eqnarray} This is the actual starting point of the problem with (\ref{Mxy}) being interpreted in terms of a weighted Motzkin path with the transition matrix (\ref{tildeC}) with $w_l$, $w_u$ and $w_d$ being the weights for level-steps, up-steps and down-steps, respectively. Related tri-diagonal matrices appear in the TASEP with a second-class particle \cite{Derrida_2class} and in the partially asymmetric exclusion process \cite{Blythe_PASEP} for example. For the first element of the $n$-th power of this matrix we find due to the tri-diagonal structure \cite{Derrida_2class} \begin{equation} \label{mn} \langle 1 | \tilde{C}^n | 1 \rangle = \sum\limits_{r=0}^{[n/2]}\frac{1}{r+1}{n \choose 2r}{2r \choose r}w_l^{n-2r}w_u^rw_d^r. \end{equation} with \begin{equation} \label{lud} w_l(x,y,\lambda)= q(x+y)\lambda+pxy\lambda^2(1+q), \quad\quad w_u(x,y,\lambda)=\sqrt{q}x\lambda+pxy\lambda^2\sqrt{q}, \quad\quad w_d(x,y,\lambda)=\sqrt{q}y\lambda+pxy\lambda^2\sqrt{q}. \end{equation} Note that for random update (\ref{mn}) turns into the well-known formula of \cite{Derrida_2class} and can be extracted to (\ref{ra}). Hence (\ref{mn}) is the total weight for all Motzkin paths of length $n$ with weights (\ref{lud}). From the first-return decomposition one finds \cite{Schork} that $\mathcal{M}$ fulfills $\mathcal{M}=1+w_l(x,y,\lambda)\mu\mathcal{M}+w_u(x,y,\lambda)w_d(x,y,\lambda)\mu^2\mathcal{M}^2$. We see that the polynomials $w_u(x,y,\lambda)$ and $w_d(x,y,\lambda)$ only appear as a product $w_uw_d$. This is because a path from height $0$ to $0$ has the same number of up- and down-steps. Hence up-/down-step pairs have weight $w_{ud}:=w_u w_d$ and the generating function reads \cite{Derrida_2class}, \cite{Schork} in terms of the fugacity $\mu$ counting Motzkin steps: \begin{equation} \label{Mxy2} \mathcal{M}(w_l,w_{ud},\mu)=\frac{1-w_l\mu-\sqrt{(1-w_l\mu)^2-4w_{ud}\mu^2}}{2w_{ud}\mu^2}. \end{equation} In Appendix \ref{Eq}, $\mathcal{M}$ is calculated alternatively, following calculations in \cite{woelki_par}. This is done without use of an explicit representation just by manipulation of the general matrix $\tilde{C}$ in (\ref{Mxy}) to calculate $(1-\tilde{C})^{-1}$ and finally taking the matrix element. The resulting expression which involves a function that appears in \cite{Blythe_dyck} for several lattice paths is shown to agree with (\ref{Mxy2}).\\ The radius of convergence of $\mathcal{M}(\mu)=\sum_{n\geq 0}\langle 1|\tilde{C}^n|1 \rangle \mu^n$ reads with (\ref{Mxy2}) $\mu_*=(w_l + 2\sqrt{w_{ud}})^{-1}$ and the thermodynamic contribution of $\langle 1|\tilde{C}^n|1 \rangle$ can be calculated by standard techniques \cite{Blythe}. Here, we find \begin{equation} \langle 1|\tilde{C}^n|1 \rangle \sim \frac{(w_l+2\sqrt{w_{ud}})^{n+3/2}}{2\sqrt{\pi}n^{3/2}w_{ud}^{3/4}}. \end{equation} The average number of level steps fulfills $\langle n_l \rangle = w_l\;d/dw_l\; \log m_n$. This leads for $n$ large to $\rho_l = w_l/(w_l + 2\sqrt{w_{ud}})$ and $\rho_{u} = \rho_d = \sqrt{w_{ud}}/(w_l + 2\sqrt{w_{ud}})$. In the following we consider paths at finite length. A realization of a path of length $n$ with $r$ up-down pairs has the total weight $w_l^{n-2r}w_d^r w_u^r$. Using the explicit expressions (\ref{lud}) and expanding leads to the weight \begin{equation} w_l^{n-2r}w_d^r w_u^r = \sum\limits_{s=0}^{r}\sum\limits_{t=0}^{r}\sum\limits_{c=0}^{n-2r}{n-2r \choose c}(1+q)^{c}(qx+qy)^{n-2r-c}q^r{r\choose s}x^{r-s}{r \choose t} y^{r-t} \times (pxy)^{c+s+t}\lambda^{n+c+s+t}. \end{equation} Now working out the power of $pxy$ and inserting into (\ref{mn}) leads \begin{table} \begin{center} \begin{tabular}{lll} \begin{tabular}{c|c|c} Red Step & Weight & Total number\\\hline $U_r$ & $\sqrt{q}x$ & $r-s$\\ $D_r$ & $\sqrt{q}y$ & $r-t$\\ $L_r$ & $q(x+y)$ & $L-2B-(2r-s-t)$ \end{tabular} & \quad\quad\quad\quad & \begin{tabular}{c|c|c} Black Step & Weight & Total number\\\hline $U_b$ & $\sqrt{q}$ & $s$\\ $D_b$ & $\sqrt{q}$ & $t$\\ $L_b$ & $(1+q)$ & $B-s-t$ \end{tabular} \end{tabular} \end{center} \caption{Enumeration of the (2,2,2)-colored Motzkin path with red (left table) and black steps (right table).} \label{tab} \end{table} by comparison with (\ref{ab}) to $m_n=\sum_{c\geq 0}(pxy)^c f_{n,c} \lambda^{n+c}$ with \begin{equation} \label{fab} f_{n,c}= \sum\limits_{r=0}^{[n/2]}\frac{1}{r+1}{n \choose 2r}{2r \choose r}\sum\limits_{s=0}^{r}\sum\limits_{t=0}^{r}{n-2r \choose c-s-t}{r\choose s}{r \choose t}(1+q)^{c-s-t}(qx+qy)^{n-2r-c+s+t}q^r x^{r-s} y^{r-t}. \end{equation} \subsection{Application to the TASEP} To apply this formula to the TASEP, we first remark that the TASEP system size $L$ obviously enters as $L=n+c$. The result for the TASEP then reads (\ref{mncompl}) \begin{equation} \label{mL} m_L=\sum\limits_{B\geq 0}(pxy)^B \mu^{L-B} f_{L-B,B}, \end{equation} where the $f_{L-B,B}$ are \begin{equation} \label{f} f_{L-B,B}= \sum\limits_{r=0}^{(L-B)/2}\frac{1}{r+1}{L-B \choose 2r}{2r \choose r}\sum\limits_{s,t=0}^{r}{L-B-2r \choose B-s-t}{r\choose s}{r \choose t}(1+q)^{B-s-t}(qx+qy)^{L-2B-2r+s+t}q^r x^{r-s} y^{r-t}. \end{equation} What we actually did in working out (\ref{f}) is dividing the three weights $w_l(x,y,\lambda)$, $w_u(x,y,\lambda)$ and $w_d(x,y,\lambda)$ in two terms each. $w_l$ was split in $q(x+y)\lambda$ and $(1+q)pxy\lambda^2$, $w_u$ in $\sqrt{q}x\lambda$ and $\sqrt{q}pxy\lambda^2$ and $w_d$ equivalently. The following section shows how that appears naturally in coloring Motzkin paths. \subsection{Coloring of the Motzkin path} A $(u,l,d)$-colored Motzkin path \cite{Schork} has up-steps, level-steps and down-steps in $u$, $l$ and $d$ colors respectively. Splitting the weights (\ref{lud}) into two parts suggests to use two colors which results in a $(2,2,2)$-colored Motzkin path. We distinguish each of the two distinct steps by the colors red and black, see table \ref{tab}. $f_{L-B,B}$ is the number of (2,2,2)-colored weighted Motzkin paths with $L$ steps, $B$ of those being black steps. Further $m_L$ is the generating function of this ensemble of paths weighting each ensemble with $(pxy)^B$ (a fugacity $p$ for each black step as well as a factor $xy$). There is no need for $\lambda$ to the explanation thus we set here $\lambda=1$.\\ To actually count the paths we have to split the contribution $(x+y)$ to red level steps further: a $(2,3,2)$-colored Motzkin path is the most appropriate interpretation. We take a continuous red line $L_{r1}$ and a red dashed line $L_{r0}$ as further coloring. Fig. \ref{allowed} shows the possible steps. All red up-steps are drawn as continuous lines and all down-steps as a dashed line while black steps are drawn as dotted lines, for a better reading in black and white. Figure \ref{motzkin232} shows an example of a (2,3,2)-colored Motzkin path of length $7$ (see figure capture). Note that the path could also touch the horizontal axis.\\ \begin{figure} \centering \includegraphics[height=1.4cm]{motzkin_steps} \caption{The possible steps in the (2,3,2)-colored Motzkin path.} \label{allowed} \end{figure} Let us consider the density of steps that are associated with weight $x$. Those are continuous red level steps, red up-steps and black steps. We can rescale the radius of convergence $\mu_*=(w_l+2\sqrt{w_{ud}})^{-1}$ of $\mathcal{M}=\sum_n \langle 1 | \tilde{C}^n | 1 \rangle \mu^n = \sum_{n}\mu^n\sum_{b=0}^{n}(pxy)^b f_{n,b}\lambda^{n+b}$ to one, solve for $x$ and identify the physical root. Then we calculate the density of steps associated with $x$, namely $\rho = n^{-1} x\;d/dx\; \log \langle 1 | \tilde{C}^n | 1 \rangle = -x(\lambda)/\lambda/x'(\lambda)$. The resulting expression can be written as a series \begin{equation} \rho(\lambda) = 1-\sqrt{\frac{\lambda}{q}} + \sum\limits_{k\geq 1}\left(\frac{p\lambda}{q}\right)^k\sum\limits_{j=0}^{k}p^{k-j}\left[\frac{1}{p}{2k\choose 2j+1} - \sqrt{\frac{\lambda}{q}}{2k+1\choose 2j+1}\right]. \end{equation} The case $\rho=1$ corresponds to $\lambda=0$ and the density decreases monotonically to $0$ where $\lambda=q^{-1}$. The density takes the value $1/2$ for $\lambda=(2-p-2\sqrt{1-p})/p^2=(1-\sqrt{q})^2/p^2$. Note that the right-hand side is equal to (\ref{zsimeq}) and equals the asymptotic value of the lattice fugacity \cite{ERS}, \cite{Blythe_dyck}. \subsection{The weight for N particles in TASEP} The function $f_{L-B,B}$ can be interpreted with (\ref{mncompl}) and (\ref{ZS}) as contributing to the weight for configurations of length $L$ with at least $B$ particle-hole domain-walls. Choosing $B$ of $W$ available particle-hole domain-walls happens naturally at every update step, where a fraction $B$ of $W$ bonds adds to the flow. The $(2,3,2)$ colored Motzkin path in figure \ref{motzkin232} has $B=5$ and is one realization of a TASEP configuration with $W=6$ since the red peak at $x=9$ contributes a domain wall, too. Note that in the TASEP the case $B=0$ corresponds to the $0$ order in $p$. Then one has $q=1$ and the continuous-time result is recovered.\\ We continue by working out the weight $Z_{M,N}$ for $N$ particles and $M$ holes in TASEP that reads due to (\ref{recm}) \begin{equation} \label{ZMN} Z_{M,N}=s_{M,N}+ps_{M-1 N}+ps_{M N-1}+p^2s_{M-1 N-1}. \end{equation} The probability that the TASEP contains $N$ particles then is Prob$(\sum_i\tau_i=N)=Z_{M,N}/Z_{M+N}$. Here, the denominator is the normalization constant (\ref{norm}). First one shall expand $(qx+qy)^{L-2B-2r+s+t}$. The weight of a path with $B$ black steps and $r$ up-down pairs of a total of $L-B$ steps then reads \begin{equation} \mathcal{W}(B,L-B,r) = \sum\limits_{s,t,k}{L-B-2r \choose B-s-t}{r\choose s}{r \choose t}{L-2B-2r+s+t \choose k}(1+q)^{B-s-t}q^{L-2B-r+s+t} p^B x^{B+r-s+k} y^{L-B-r+s-k}. \end{equation} Now introducing the particle number $N$ and picking out the coefficient of $x^Ny^{L-N}$ yields \begin{equation} \{x^Ny^{L-N}\}\mathcal{W}(B,L-B,r) = \sum\limits_{s,t}{L-B-2r \choose B-s-t}{r\choose s}{r \choose t}{L-2B-2r+s+t \choose N-B-r+s}(1+q)^{B-s-t}q^{L-2B-r+s+t} p^B. \end{equation} Then one obtains \begin{equation} s_{L-N,N}=\sum\limits_{B}\sum\limits_{r=0}^{(L-B)/2}\frac{1}{r+1}{L-B \choose 2r}{2r \choose r} \{x^Ny^{L-N}\}\mathcal{W}(B,L-B,r). \end{equation} Note that one has due to the particle-hole symmetry that $s_{L-N,N}=s_{N,L-N}$. Further the $z_L$ entering the normalization (\ref{norm}) read $z_L=\sum_N \left(s_{L-N,N} + ps_{L-N,N-1}\right)$. \section{Second-class particle dynamics} In \cite{woelki_par} it was argued that second-class particle dynamics in the parallel TASEP are not as natural to define as for the generic random update. In this section a model system is presented that mimics second-class particle dynamics and for which preliminary results were published in \cite{woelki_cargo}. We change the boundary conditions and consider a closed periodic chain with one of the particles carrying a cargo. The unit of the particle and the cargo is referred to as the second-class particle. Under the parallel update every particle (with or without cargo) moves forward with probability $p$ if the target site is empty. In the most simple case the cargo is carried to the next site. However if the site behind it is occupied, the cargo can actively jump onto this particle which happens independently with probability $p$. Thus the cargo can jump to the left while its carrier moves at the same time to the right. This dynamics reminds a bit of someone standing on a train either standing still and moving with the train or actively jumping to the wagon behind. While in the continuous-time case a 'move' corresponds to an interchange of occupation numbers $\tau_i\tau_{i+1}\rightarrow \tau_{i+1}\tau_i$ obviously the situation here is more complex. One has the transitions \begin{eqnarray} \label{rel1} 10 &\rightarrow 01, & \text{with probability } p,\\ 020 &\rightarrow X02, & \text{with probability } p,\\ 120 &\rightarrow 210, & \text{with probability } pq,\\ &\rightarrow 102, & \text{with probability } pq,\\ &\rightarrow 201, & \text{with probability } p^2,\\ \label{rel6} 121 &\rightarrow 21X, & \text{with probability } p, \end{eqnarray} with $X$ being either $0$ or $1$. In the thermodynamic limit the generating function $\mathcal{M}_2$ for the second-class particle process reads \begin{equation} \label{simeq} \mathcal{M}_2 \simeq \mathcal{M} \frac{p^2\gamma}{q(1-\sqrt{q})^2}, \text{ with } \mathcal{M}=\frac{\gamma}{(1-\gamma)\sqrt{qw_{ud}}}. \end{equation} The expression for $\mathcal{M}$ is the corresponding one for the open-boundary TASEP and $\gamma$ is defined in (\ref{def_z}), see Appendix \ref{AppC}. \begin{figure} \centering \includegraphics[height=4.5cm]{motzkin232alt} \caption{Example of a (2,3,2)-colored Motzkin path of length $14$. The corresponding TASEP size is $19$.} \label{motzkin232} \end{figure} In the thermodynamic limit, only the denominator $(1-\gamma)\sqrt{qw_{ud}}$ of $\mathcal{M}$ in (\ref{simeq}) is physically relevant. The other factors contribute just a finite number of excursions to the corresponding Motzkin path \cite{Blythe_dyck}. Therefore the second-class particle process has the same thermodynamic physics as the open TASEP. Note that for random update both expressions for $\mathcal{M}$ and $\mathcal{M}_2$ are even the same. This is due to the fact that for random update the weights for $N$ particles on a lattice with $L$ sites are the same for open boundaries and a second-class particle on the ring.\\ Reference \cite{woelki_cargo} shows the equivalence of this process and the one studied in \cite{woelki_par}. It focuses on the thermodynamic limit and simplifies some results of \cite{woelki_par}. The velocity of the 2nd-class particle is shown to be calculated through \cite{woelki_par} \begin{equation} \label{2nd_v1} v=p(1-\rho_+)(1-p\rho_-)-p\rho_-. \end{equation} Here $\rho_-$ and $\rho_+$ are the densities directly behind and in front of the 2nd-class particle $2$. The dynamics of the 2nd-class particle is obtained from (\ref{rel1}-\ref{rel6}). It moves either forward if it has a hole in front while at the same time there is no particle directly behind that simultaneously catches the cargo: first term in (\ref{2nd_v1}). Or it moves backwards if it has a particle behind: second term in (\ref{2nd_v1}). With the help of Appendix \ref{asy2} one finds \cite{woelki_cargo} that the neighboring densities are \begin{eqnarray} \label{2nd_dens} \rho_-=&\dfrac{1}{p}\dfrac{p\rho-J}{1-2J}, \quad\quad\quad 1-\rho_+=&\left( \dfrac{J}{p\rho}\right)^2. \end{eqnarray} which yields \begin{equation} \label{2nd_v} \frac{v(\rho)}{p}=\dfrac{1-2\rho}{1-2J}. \end{equation} Hence the velocity of the 2nd-class particle vanishes only at half filling $\rho=1/2$. Note that this has the form of a group velocity thus the 2nd-class particle is travelling with the velocity of the density disturbance. This feature should of course be ensured by a 2nd-class particle dynamics and it is a further underpinning of the fact that the model is able to describe the TASEP properly on hydrodynamic scale. \section{Conclusion} \begin{figure} \centering \includegraphics[height=2.0cm]{situation120} \caption{Example of a local configuration 120: shown are two chains, the cargo (star) moves in the upper chain; usual particles move in the lower chain. The cargo particle is always carried by a usual particle; therefore it can move independently to the left with probability $p$ (while its carrier either moves to the right ($p^2$) or not ($pq$)); if not, it stays on its carrier with probability $q$ which means that it either moves to the right with its carrier ($pq$) or stays with it on the current site ($q^2$).} \label{sit120} \end{figure} To conclude, we have studied the parallel TASEP with open boundaries where every particle move occurs with probability $p$. This case defines the physics of the maximum-current phase for all $p\in (0,1)$. The main result is the generating function of arbitrary particle numbers $N$ in a system of size $L$ for parallel update. We established an analogy to weighted Motzkin paths and derived an equivalence to ($u,l,d$)-colored Motzkin paths appearing in \cite{Schork}. Already known interpretations of TASEP normalizations with colored Motzkin paths \cite{Blythe_PASEP}, \cite{Duchi} for random-sequential update correspond in this terminology to $(1,2,1)$-colored paths, i.e.\ where only level steps appear in two colors. Here, we found the most natural interpretation as $(2,3,2)$-colored paths. The reason for this increase in the necessary number of colours is the complexity of the parallel update. It is known, that driven-diffusive systems under parallel dynamics typically have a quartic matrix algebra \cite{ERS}, \cite{woelki_par}, \cite{arita} rather than a quadratic one as under random update, ordered-sequential or sublattice parallel update \cite{Blythe}. In fact, we could alternatively use $(2,4,2)$-colored paths for the parallel update so that in contrast to the other updates the number of colours is doubled. This nicely reflects the increase from a quadratic to a quartic algebra. It is expected that the structure found here also holds for similar models with parallel update while the other possible updates mentioned above lead to a ($1,2,1$)-colored path. In the $(2,3,2)$ Motzkin path, red steps are associated with particles or holes while black steps correspond to particle-hole pairs. Therefore paths with the same length do not necessarily correspond to TASEP configurations of the same system size. One could alternatively represent particle-hole pairs by a succession of two steps, see also Motzkin paths with higher rank considered in \cite{Schork}. However those paths would not end at height $0$, in general. The fact that a fraction of the particle-hole domain walls corresponds to Motzkin steps in a different colour somewhat reflects the fact that under parallel update such a fraction contributes to the flow. In this paper it was shown how known lattice-path interpretations for the parallel TASEP \cite{Blythe_dyck}, \cite{Blythe} are mapped onto the present one that keeps information of the particle number. Closely related to those problems concerning a given number of particles is the second-class particle process \cite{Derrida_2class}. For parallel update such a process is not as straightforward to define \cite{woelki_par}, though. Here we presented one possible choice of second-class particle dynamics in terms of a cargo process \cite{woelki_cargo} for which the generating function shares the thermodynamics with the generating function that has been derived here for the TASEP.\\ Since the parallel TASEP is equivalent to the Nagel-Schreckenberg model with maximal velocity 1, the results also apply to a bottleneck situation in traffic flow. The average density is $1/2$, however, the density fluctuates around this value. With the help of the generating function obtained in this paper one finds the probability that the road section contains any density of cars. Further investigations shall consider the fluctuations of the particle number as in \cite{Derrida_Brownian} using the findings of the present paper. This is planned to be done in the near future. Additionally it would be interesting to compare the joint current-density distribution with the one for random-sequential update \cite{Blythe}. To finally generalize present results to arbitrary $\alpha$ and $\beta$ one proceeds as in \cite{Mallick}: Denote by $H_{N,M}$ the sum of all possible products of $N$ $D$-matrices and $M$ $E$-matrices (with fugacities $x$ and $y$) in which particle-hole domain-walls are weighted by $q^{-1}$. One observes that $H_{N,M}$ obeys the recursion $H_{N,M}=xH_{N,M-1}D+yH_{N,M-1}E+pq^{-1}xyH_{N-1,M-1}DE$. In order to obtain the weight for $N$ particles and $M$ holes one considers $\langle n|H_{N,M}|m\rangle$ and obtains a recursion that can be solved. Once having $\langle n|H_{N,M}|m\rangle$ one uses the representation of the boundary vectors $\langle W|=\sum_{n\geq 1}((p-\alpha)/\alpha\sqrt{q})^{n-1}\langle n|$, $|V\rangle=\sum_{m\geq 1}((p-\beta)/\beta\sqrt{q})^{m-1}| m\rangle$ to find $\langle W|H_{N,M}|V\rangle$. The result is no longer a Motzkin path but a sum over lattice paths. Walks that start at height $n$ and end at height $m$ each contain an additional factor $((p-\alpha)/\alpha\sqrt{q})^{n-1}((p-\beta)/\beta\sqrt{q})^{m-1}$.
\section{Introduction} Liquid argon time projection chambers (LArTPCs) offer fine-grained tracking and precise calorimetry, which are ideal for the study of neutrino interactions and search for rare phenomena such as proton decay. The ArgoNeuT (Argon Neutrino Teststand) experiment at Fermilab \cite{Anderson:2012vc} was the first experiment that utilizes a LArTPC exposed to a low-energy neutrino beam (neutrino energies in the 0.5-10.0 GeV range). In this article, I will discuss the latest results from ArgoNeuT. \section{ArgoNeuT} ArgoNeuT consists of a vacuum insulated cryostat for ultra-pure liquid argon containment, in which is mounted a time projection chamber (TPC). The cryostat volume is 550 L (0.77 t). The TPC is 47 cm wide (drift direction), 40 cm high and 90 cm long (neutrino beam direction), corresponding to an active volume of 170 L. A high voltage is applied to the cathode to produce a uniform electric field of 500 V/cm inside the TPC. The electron drift velocity is 1.59 mm/$\mu$s at this nominal field, which leads to a maximum drift time of 295 $\mu$s. The ionization electrons are read out by two wire planes (induction and collection planes). ArgoNeuT took data in the NuMI (Neutrinos at the Main Injector) beam from September 2009 to February 2010. ArgoNeuT was located approximately 1.5 m upstream of the MINOS Near Detector (MINOS-ND), with the TPC centered 26 cm below the center of the NuMI on-axis beam. The MINOS-ND was used as a spectrometer to measure the momentum and charge of uncontained long-track muons from charged current (CC) neutrino interactions in the ArgoNeuT. The physics run consisted of about two weeks of neutrino-mode running and four-and-a-half months of anti-neutrino mode running. \section{Inclusive CC cross sections} Inclusive neutrino cross sections have been measured on a variety of targets, however, measurements on the liquid argon target are sparse. Ref.~\cite{Anderson:2011ce} presents the first measurement of inclusive muon neutrino CC differential cross sections on argon using the ArgoNeuT data taken in the {\it neutrino-mode} beam. We are currently finalizing similar measurements of inclusive muon neutrino and anti-neutrino CC differential cross sections on argon using the ArgoNeuT data taken in the {\it anti-neutrino-mode} beam. This corresponds to data from the $1.20\times10^{20}$ protons on target (POT), for which MINOS and ArgoNeuT were both up and running over the 4.5 month physics run. Neutrinos in the anti-neutrino beam originate from the decay of unfocused $\pi^{+}$s and have a broad energy spectrum. Their interactions comprise almost half of all neutrino/anti-neutrino-induced events in the detector, which enables the measurements of both $\nu_{\mu}$ and $\bar{\nu}_{\mu}$ CC differential cross sections using this data sample. The differential cross sections are measured in terms of muon angle and muon momentum. Neutrino events are reconstructed in the framework of the LArSoft automated reconstruction software package. The muon angle is measured in the TPC. The muon momentum and the sign of the muon charge are measured in the magnetized MINOS-ND. After three dimensional track formation in ArgoNeuT, an attempt is made to match the tracks that leave the ArgoNeuT TPC with muons that have been reconstructed in MINOS and have a hit within 20 cm of upstream face of the detector. The neutrino event's reconstructed vertex is required to be inside of the ArgoNeuT fiducial volume. The track matching criteria along with a requirement that the reconstructed and matched MINOS track is negatively (positively) charged represent the only other selection criteria for CC $\nu_{\mu}$s ($\bar{\nu}_{\mu}$s) in this analysis. The neutral current (NC) background is highly suppressed by the track matching requirement since the probability of reconstructing tracks from hadrons entering MINOS is low. The remaining NC background and wrong-sign (WS) background are estimated using a simulated sample. The WS background is defined as negative (positive) muons from the CC interactions matched to positive (negative) muons in MINOS because of the misreconstruction of the muon sign in MINOS. Neutrino-induced crossing muons that pass the fiducial volume requirement due to inefficiency in vertex reconstruction are removed through handscanning of events. Figure~\ref{recotruth} shows the comparisons of reconstructed quantities versus true quantities for simulated $\nu_{\mu}$s and $\bar{\nu}_{\mu}$s after event selection. The event reconstruction is unbiased and reliable even for complicated events such as deep inelastic scatterings (DIS). \begin{figure} \begin{center} \includegraphics[width=30pc]{recotruth.pdf} \end{center} \caption{\label{recotruth}Comparisons of reconstructed quantities (vertex x, y, z, muon angle and muon momentum) versus true quantities for simulated $\nu_{\mu}$s and $\bar{\nu}_{\mu}$s after event selection.} \end{figure} Figure~\ref{ccinc} shows distributions of muon angle and muon momentum for $\mu^{-}$'s and $\mu^{+}$'s after full event selection compared with {\sc genie} \cite{Andreopoulos:2009rq} predictions. Both data and {\sc genie} total predictions are normalized to unit area. The {\sc genie} simulation describes the shape of data distributions well. The NC and WS backgrounds are negligible. The $P_{\mu^{-}}$ distribution is broader than the $P_{\mu^{+}}$ distribution because of the broader $\nu_{\mu}$ energy spectrum. The $\nu_{\mu}$ sample is dominated by DIS interactions while the $\bar{\nu}_{\mu}$ sample has approximately equal contributions from quasielastic scattering (QE), resonance production (RES) and DIS. The final differential cross sections after correcting for detector acceptance and smearing, selection efficiency, and accounting for neutrino flux will be released soon. \begin{figure} \begin{center} \includegraphics[width=15pc]{pmum.pdf} \includegraphics[width=15pc]{thmum.pdf} \includegraphics[width=15pc]{pmup.pdf} \includegraphics[width=15pc]{thmup.pdf} \end{center} \caption{\label{ccinc}Distributions of muon angle and muon momentum for $\mu^{-}$'s and $\mu^{+}$'s compared with {\sc genie} predictions. The {\sc genie} prediction is broken down as the NC (neutral current) background, WS (wrong-sign) background, QE (quasielastic) CC, RES (resonance) CC and DIS (deep inelastic scattering) CC. Both data and {\sc genie} total predictions are normalized to unit area.} \end{figure} \section{Topological cross sections} Conventionally measurements of exclusive channels are performed by looking for a certain event topology in the detector. For example, an event with a muon and a proton is a candidate for $\nu_{\mu}$ CCQE interaction. However, nuclear effects play an important role in neutrino-nucleus interactions. Intranuclear rescattering, a.k.a final state interaction (FSI), and possible effects of correlation between target nucleons can change event topology drastically. For example, the proton produced in a $\nu_{\mu}$ CCQE interaction can be absorbed before it exits the argon nucleus so the event only has a muon, which does not look consistent with a CCQE interaction. On the other hand, a pion produced in a resonance production can be absorbed so the event has a muon and a proton, which looks consistent with a CCQE interaction. Conventional exclusive-channel measurements rely on neutrino generators (e.g. {\sc genie}) to correct for nuclear effects, which makes the results model dependent. In ArgoNeuT we developed a different approach: instead of measuring a specific channel, we measure the cross section for a well-defined event topology \cite{Palamara:2013maa}. Such measurements provide useful information for both neutrino cross sections and nuclear effects. A first topological analysis is currently developed by the ArgoNeuT experiment, namely $1\mu+Np+0\pi$, where we select events with one muon, any number of protons but without pions. The analysis takes two steps. First a similar event selection to the inclusive CC cross section measurements is applied to select events with muons. The selected events are then scanned to remove pion background and reconstruct the proton kinematics. The particle identification is done taking advantage of the detailed calorimetry information provided by the LArTPC. The particle species is determined by the energy deposition information as a function of residual range along the track trajectory if the particle stops inside the TPC. The kinetic energy is measured by summing the charge collected on each wire after correcting for electron lifetime and recombination effects. The threshold to reconstruct a proton in ArgoNeuT is $T_{p} = 21$ MeV, where $T_{p}$ is the proton kinetic energy. Figure~\ref{0pi} shows the distributions of proton multiplicity for $1\mu+Np+0\pi$ sample in comparison with {\sc genie} predictions. The {\sc genie} predictions have longer tails compared to ArgoNeuT data. This information is useful to improve the neutrino event generators. \begin{figure} \begin{center} \includegraphics[width=15pc]{nunpro.pdf} \includegraphics[width=15pc]{anunpro.pdf} \end{center} \caption{\label{0pi} Distributions of proton multiplicity for $1\mu+Np+0\pi$ sample compared with {\sc genie} predictions. Right panel shows $\nu_{\mu}$ distributions and left panel shows $\bar{\nu}_{\mu}$ distributions. Both data and {\sc genie} predictions are normalized to unit area.} \end{figure} \section{Conclusions} ArgoNeuT was the first LArTPC placed in a low energy neutrino beam and it provides invaluable experience for future experiments that employ LArTPCs. We have measured the $\nu_{\mu}$ and $\bar{\nu}_{\mu}$ CC inclusive cross sections on argon. We also developed techniques to measure topological cross sections, which have yields results that are useful for understanding nuclear effects in neutrino-nucleus interactions. Currently there are many ongoing ArgoNeuT analyses. We are searching for events with one muon and two protons where the two protons go back-to-back. This would be an indication for nucleon-nucleon correlation in neutrino scattering. Other analyses include a search for neutral hyperon production, electron/photon identification, measurements of NC $\pi^{0}$ cross sections, nuclear deexcitation $\gamma$s, coherent pion production, and $\nu_{e}$ measurements. \section*{References}
\section{Introduction} Wilson loops are among the most interesting operators in any gauge theory. Their expectation values can serve as order parameters for the different phases of gauge theory. However, for generic four-dimensional gauge theories, the analytic evaluation of these expectation values for generic contours is currently out of reach. The situation improves for gauge theories with additional symmetries and Wilson loops with suitably chosen contours: for conformal theories with an AdS dual, it is possible to evaluate the vev of Wilson loops \cite{Rey:1998ik, Maldacena:1998im} with a variety of contours \cite{Berenstein:1998ij, Ishizeki:2011bf}, and in a variety of representations \cite{Drukker:2005kx, Yamaguchi:2006tq, Gomis:2006sb}. A second tool to compute vevs of Wilson loops is integrability, either of the dual string world-sheet \cite{Drukker:2005cu}, or of the planar limit of ${\cal N}=4$ SYM \cite{Drukker:2012de}. There have been also applications of the relations among certain 4d susy gauge theories and 2d CFTs \cite{Alday:2009aq} to the evaluation of Wilson loops \cite{Alday:2009fs}. Last but not least, it is also possible to use localization techniques to evaluate vevs of Wilson loops. In this regard one of the most remarkable results is due to Pestun \cite{Pestun:2007rz}, who showed that for ${\cal N}=2$ super Yang-Mills theories, localization arguments reduce the evaluation of the expectation value of Euclidean half-BPS circular Wilson loops to a matrix model computation. For the particular case of ${\cal N}=4$ SYM, this had been anticipated in \cite{Erickson:2000af, Drukker:2000rr}. Localization techniques have also been applied to the evaluation of the vacuum expectation value of 't Hooft loops \cite{Gomis:2011pf}, loops preserving less supersymmetry \cite{Pestun:2009nn} and 2-point functions of Wilson loops and local operators \cite{Giombi:2012ep}. The arguments in \cite{Pestun:2007rz} that reduce the evaluation of the vev of the Wilson loop to a matrix model computation are independent of the representation of the gauge group entering the definition of the Wilson loop. However, so far, the evaluation of the resulting matrix model integral has been carried out exactly only for a Wilson loop of ${\cal N}=4$ SYM in the fundamental representation \cite{Drukker:2000rr}, yielding a strikingly simple result in terms of a generalized Laguerre polynomial, \begin{equation} \vev{W_\Box(g)}=\frac{1}{N}L_{N-1}^1\left(-g \right)e^{\frac{g}{2}} \label{exactfund} \end{equation} where we have defined $g=\lambda/4N$. The main goal of this work is to evaluate the relevant matrix model integrals to obtain the vev of half-BPS circular Wilson loops of ${\cal N}=4$ U(N) SYM in arbitrary irreducible representations. Half-BPS Wilson loops in higher rank representations have already been studied using holography by means of D-branes \cite{Drukker:2005kx, Yamaguchi:2006tq, Gomis:2006sb}. They have also been studied by solving the matrix model integrals in the large N limit for various representations \cite{Yamaguchi:2006tq, Hartnoll:2006is, Okuyama:2006jc} and also at strong coupling for arbitrary representations and gauge groups \cite{Gomis:2009ir}. Quite interestingly, this last reference managed to turn those results into non-trivial explicit checks of S-duality for ${\cal N}=4$ SYM. Let us now explain what we have accomplished in the present work. The evaluation of the vev of the Wilson loop in the representation ${\cal R}$ of U(N) amounts to compute the vev of the insertion of $tr_{\cal R} e^X$ into the Gaussian Hermitian matrix model ($X$ is the Hermitian matrix being integrated over); after diagonalization in the matrix model, this insertion boils down to the Schur polynomial associated to the representation ${\cal R}$\footnote{In appendix B we have collected the definitions and some basic facts of the different basis of symmetric functions that appear in this work.}, as a function of the exponentials of the eigenvalues $x_i$, $$ \vev{W_{\cal R}(g)}=\frac{1}{\dim \mathcal{R}}\vev{S_{\cal R}(e^{x_1},\dots,e^{x_N})}_{m.m.} $$ Schur polynomials form a basis of the space of symmetric polynomials of $N$ variables, but as it turns out, they seem not to be the most convenient basis to carry out the integrals. Something similar was already encountered in the computation of the vev of Wilson loops in two-dimensional QCD \cite{Gross:1993yt}, where it was proposed to carry out the computations in the basis of multiply wound Wilson loops (which corresponds to the power sum symmetric functions, see appendix B), related to the Schur basis by Frobenius formula\footnote{See \cite{Gross:1998gk} for a similar discussion in the context of ${\cal N}=4$ SYM.}. Similarly, in this work, we manage to compute the exact vevs of insertions in the basis of monomial symmetric functions, from which one can recover the vevs in the Schur basis by a linear transformation. A couple of basics points of this linear transformation will help us to understand the structure of the answer we get for the vevs of Wilson loops. First, the Schur polynomial $s_{\cal R}$ can be written in terms of monomial polynomials $m_\tau$ with the same weight as ${\cal R}$ ({\it i.e.} the number of boxes of the associated Young diagrams is the same) and second, after imposing certain ordering among the partitions of $n$ (see appendix B), the matrix for this linear transformation is upper triangular when written using the ordering mentioned above, so only $\tau \leq {\cal R}$ representations have a non-zero contribution. The structure of the answer we get is then \begin{equation} \vev{W_{\cal R}(g)}=\frac{1}{\dim \mathcal{R}} \sum_{\tau \leq {\cal R}} K_{{\cal R} \tau} P_\tau(g) e^{\sum _i \tau_i^2\frac{g}{2}} \label{genvev} \end{equation} where $K_{{\cal R} \tau}$ are positive integers (the {\it Kostka numbers}) that realize the linear transformation, and $P_\tau(g)$ are polynomials that we compute explicitly. Each polynomial multiplies an exponential, with exponent given by the sum of the squares of the elements of the partition $\tau$, $\sum _i \tau_i^2$ (these exponentials were found already in \cite{Gomis:2009ir}). The discussion above describes the generic case, but this general picture simplifies drastically for the particular case of Wilson loops in the antisymmetric representation of U(N). The reason is that Schur functions for the antisymmetric representation are already monomial symmetric functions, so in this case no change of basis is needed, and the final answer is given by a polynomial times an exponential, similar to the result found for the fundamental representation, eq. (\ref{exactfund}). We can write the resulting polynomial in a couple of ways: the matrix integral spits it out as a sum of products of generalized Laguerre polynomials, which is a straightforward but not terribly illuminating expression. We have managed to rewrite it as a single polynomial in $g$, and for the $k$-th antisymmetric representation we obtain a result of the form \begin{equation} \vev{W_{{\cal A}_k}(g)} = e^{\frac{g k}{2}}\; \sum_{j=0}^{k(N-k)} d_j(k,N)\frac{g^j}{j!} \hspace{1cm}d_j(k,N)\in \mathbb{N} \label{genantivev} \end{equation} where $d_j(k,N)$ are positive integers for which we derive a precise combinatorial formula, that we evaluate for specific values of $j$. An alternative and very succinct presentation of these results can be given in terms of the generating function of the vevs of these Wilson loops, $$ \vev{F_A(t)}=\sum_{k=0}^Nt^{N-k}{N\choose k}\vev{W_{{\cal A}_k}} $$ for which we obtain $$ \vev{F_A(t)}=|t+A(g)e^{\frac{g}{2}}| $$ where $A(g)$ is an $N\times N$ matrix with generalized Laguerre polynomials as entries, $A_{ij}=L_{i-1}^{j-i}(-g)$. The result for the fundamental representation, eq. (\ref{exactfund}), follows by taking the trace of this matrix. Besides its intrinsic interest, having explicit exact results for these Wilson loops can have a number of applications that we will discuss in the last section of the paper. In the present note, we will use these results to discuss the exact Bremsstrahlung functions for the corresponding heavy probes. For any heavy probe coupled to an arbitrary four-dimensional conformal field theory, the Bremsstrahlung function determines many relevant properties, like the total radiated power \cite{Correa:2012at, Fiol:2012sg} and the momentum diffusion coefficient \cite{Fiol:2013iaa} of an accelerated probe. For the specific case of 1/2 BPS heavy probes of ${\cal N}=4$ SYM it was shown in \cite{Correa:2012at} that the Bremsstrahlung function can be derived from the vev of the 1/2 BPS circular Wilson loop, \begin{equation} B(\lambda,N)_{\cal R}=\frac{1}{2\pi^2}\lambda\partial_\lambda \hbox{ log}\vev{W_{\cal R}} \label{introbrem} \end{equation} so once we have these vevs in a given representation, it is straightforward to obtain the Bremsstrahlung function. It was observed in \cite{Fiol:2012sg, Fiol:2013iaa} that the Bremsstrahlung function for a heavy probe in the fundamental representation is, for fixed N, a rational function of the 't Hooft coupling, and it becomes linear both at weak and a strong coupling. In the light of the general structure presented in (\ref{genvev}) for generic representations and in (\ref{genantivev}) for the antisymmetric representation, it follows from (\ref{introbrem}) that while the linearity in the 't Hooft coupling at weak and strong coupling (for fixed N) is common to all Bremsstrahlung functions, exact Bremsstrahlung functions are rational functions of the coupling only when there is a single exponential in the vev of the Wilson loop, {\it i.e.}, for antisymmetric representations. The outline of the paper is as follows. In section 2 we start by considering the vevs of Wilson loops in antisymmetric representations. We solve the matrix integral explicitly and discuss some properties of the generating function for these vevs. We point out that the resulting vevs admit a certain expansion with positive integer coefficients. The combinatorial formulas for these coefficients involve a number of ingredients suggestive of an interpretation in terms of fermions on a lattice, but we have not managed to come up with a satisfactory physical realization of these integers. In section 3 we turn to arbitrary representations; we first consider a perturbative expansion in $g$ for the matrix model integral, at finite N. We point out that this expansion involves evaluating {\it shifted Schur functions} \cite{okounkov}; we manage to compute at every order in $g$ the part of the coefficient that is of highest degree in the Casimir invariants of ${\cal R}$, and obtain a quite simple result involving the second Casimir of ${\cal R}$, $$ \vev{W_{R}(g)}=\sum_{n=0}^\infty \left [\left(\frac{c_2({\cal R})}{2}\right)^n+\dots\right ] \frac{g^n}{n!} $$ We then turn to computing the exact expectation value for a basis of symmetric functions, the monomial symmetric functions. In section 4 we discuss the Bremsstrahlung function for the corresponding heavy probes, using the results obtained in the previous sections. We conclude in the last section by briefly mentioning possible directions for future research. We have included three appendices: two with brief summaries on skew Young tableaux and on symmetric functions, and a third one with an alternative proof of one of the results in the main text. \section{Wilson loop in antisymmetric representation} In this work we are concerned with some specific non-local operators in ${\cal N}=4$ SYM. As argued in the seminal works \cite{Rey:1998ik, Maldacena:1998im}, in ${\cal N}=4$ SYM it is natural to generalize the usual Wilson loop to include scalar fields. Locally BPS Wilson loops are then determined by a representation ${\cal R}$ and a contour ${\cal C}$ \begin{equation} W_{\cal R}[{\cal C}]=\frac{1}{\hbox{dim }{\cal R}}\hbox{Tr}_{\mathcal R}{\cal P}\hbox{exp } \left(i \int_{\cal C} (A_\mu \dot x^\mu +|\dot x|\Phi_i \theta^i)ds \right) \end{equation} We have fixed the overall normalization of the Wilson loop by the requirement that at weak coupling, $\vev{W_{\cal R}}=1+{\cal O}(g)$. We will restrict ourselves to half BPS Wilson loops. The simplest case is an infinite straight line, with an arbitrary representation. For any representation, this Wilson loop has a trivial vev, $\vev{W_{\cal R}}=1$, due to the large amount of supersymmetry. By a special conformal transformation, this Wilson loop can be mapped to a circular Wilson loop in Euclidean signature (or in Lorentzian signature, to a loop with hyperbolic contour \cite{Branding:2009fw, Fiol:2011zg}). This conformal mapping does not, however preserve the value of the vev. It was argued in \cite{Erickson:2000af, Drukker:2000rr} and later proven in \cite{Pestun:2007rz} that the vev of 1/2 BPS circular Wilson loops of ${\cal N}=4$ SYM can be computed exactly by means of a Gaussian Hermitian matrix model. After diagonalization, the partition function of this matrix model is given by \begin{equation} {\cal Z}= \int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} \label{partfunc} \end{equation} The vev of the 1/2 BPS Wilson loop in an arbitrary representation ${\cal R}$ is given by the expected value of the Schur polynomial of ${\cal R}$ \begin{equation} \vev{W_{\cal R}(g)}=\frac{1}{\hbox{dim }{\cal R}} \frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 S_{{\cal R}}\left(e^{x_1},\dots,e^{x_N}\right) e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} \end{equation} In the next section we will discuss this integral for arbitrary irreducible representations ${\cal R}$. It turns out that the case of antisymmetric representations is particularly simple and the results are most explicit, so we will start with it. For the k-antisymmetric representation, the Schur polynomial is given by the k-th {\it elementary symmetric function} $e_k$\footnote{This can be seen for instance from the dual Jacobi-Trudi identity, see appendix B.}, so \begin{equation} \vev{W_{{\cal A}_k}(g)}=\frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 e^{x_1+\dots+x_k} e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} \label{wilsonmm} \end{equation} From this integral, eq. (\ref{wilsonmm}), it is straightforward to relate the vevs of the Wilson loops for the k-th and the (N-k)-th antisymmetric representations. To do so, complete the squares for the $x_1,\dots,x_k$ eigenvalues in (\ref{wilsonmm}), and then change variables $\tilde x_i=x_i-g$. Except for the $x_i$-independent exponents generated by completing squares, the resulting integral is the one that yields the vev of the Wilson loop in the $(N-k)$-th representation, so we arrive at the relation \begin{equation} \vev{W_{{\cal A}_k}(g)}e^{-\frac{kg}{2}}=\vev{W_{{\cal A}_{N-k}}(g)}e^{-\frac{(N-k)g}{2}} \label{particlehole} \end{equation} For future reference, we define the following generating function for the elementary symmetric functions $e_k$\footnote{This generating function is closely related to the usual one (see appendix B), $E(t)=\sum_{k=0}^N e_k(y)t^k=\prod_{i=1}^N (1+y_it)$. Indeed, $F_A(t)=t^NE(1/t).$}, $$ F_A(t)=\sum_{k=0}^N e_kt^{N-k}=\prod_{i=1}^N\left(t+e^{x_i}\right) $$ so its expectation value serves as the generating function for the expectation values of Wilson loops in antisymmetric representations, \begin{equation} \vev{F_A(t)}=\sum_{k=0}^Nt^{N-k}{N\choose k}\vev{W_{{\cal A}_k}(g)} \end{equation} To compute the integral (\ref{wilsonmm}) we will apply the method of orthogonal polynomials (see e.g. \cite{Marino:2004eq} for a recent introduction), following and generalizing the approach in \cite{Drukker:2000rr}. In a nutshell, we introduce a family of polynomials $p_n(x)$ orthogonal with respect to the measure $dx \; e^{-\frac{x^2}{2g}}$. In the case at hand these polynomials are closely related to the Hermite polynomials $$ p_n(x)\equiv \left(\frac{g}{2}\right)^{\frac{n}{2}}H_n\left(\frac{x}{\sqrt{2g}}\right) $$ since \begin{equation} \int \frac{dx}{2\pi}\; p_m(x)p_n(x)\; e^{-\frac{x^2}{2g}}= n!g^n\sqrt{\frac{g}{2\pi}}\delta_{mn}\equiv h_n \delta_{mn} \label{orthopol} \end{equation} It is straightforward to prove that the partition function (\ref{partfunc}) admits a very simple expression \begin{equation} {\cal Z}=N! \prod_{k=0}^{N-1} h_k \label{partfunch} \end{equation} The relevance of these orthogonal polynomials for the computation at hand becomes apparent when we realize that we can substitute the Vandermonde determinant in (\ref{wilsonmm}) by a determinant of orthogonal polynomials, $$ \vev{W_{{\cal A}_k}(g)}= \frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} |p_{i-1}(x_j)|^2 e^{x_1+\dots+x_k} e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} $$ We now expand the determinants of orthogonal polynomials in terms of permutations of its matrix elements $$ |p_{i-1}(x_j)|^2=\sum _{\sigma_1 \in S_N} (-1)^{\epsilon(\sigma_1)} \prod _{k_1=1}^N p_{\sigma_1(k_1)-1}(x_{k_1}) \sum _{\sigma_2 \in S_N} (-1)^{\epsilon(\sigma_2)} \prod _{k_2=1}^N p_{\sigma_2(k_2)-1}(x_{k_2}) $$ The crucial point in the argument is that for the eigenvalues $x_{k+1},\dots,x_N$ the integrals in (\ref{wilsonmm}) are not modified by the insertion of the Wilson loop operator, so due to the orthogonality of the polynomials, for a given $\sigma_1\in S_N$, the only $\sigma_2$s that survive integration are those for which $\sigma_2(m)=\sigma_1(m)$ for $m>k$. This means that in order to contribute to the full matrix model integral, $\{\sigma_2(1),\dots,\sigma_2(k)\}$ must be a permutation of $\{\sigma_1(1),\dots,\sigma_1(k)\}$. Let's call this permutation $\mu$. The integral is now $$ \vev{W_{{\cal A}_k}(g)}= \frac{1}{N!}\sum _{\sigma \in S_N} \sum _{\mu \in S_k} (-1)^{\epsilon(\mu)} \int \prod_{i=1}^k \frac{dx_i}{2\pi} \frac{p_{\sigma(i)-1}(x_i) p_{\mu(\sigma (i))-1}(x_i)}{h_{\sigma(i)-1}} e^{-\frac{x_i^2}{2g}+x_i} $$ where we already performed the integrals over the $x_{k+1},\dots,x_N$ eigenvalues using (\ref{orthopol}), and substituted the partition function ${\cal Z}$ using (\ref{partfunch}). The remaining integral can carried out explicitly \cite{gradshteyn} and it is given in term of generalized Laguerre polynomials, \begin{equation} L_n^\alpha(-g)=\sum_{j=0}^n {n+\alpha \choose n-j} \frac{g^j}{j!} \label{thelaguerres} \end{equation} so we arrive at \begin{equation} \vev{W_{{\cal A}_k}(g)}= \frac{e^{\frac{kg}{2}}}{N!}\sum _{\sigma \in S_N} \sum _{\mu \in S_k} (-1)^{\epsilon(\mu)}\prod_{m=1}^k L_{\sigma(m)-1}^{\mu(\sigma(m))-\sigma(m)}(-g) \label{antivev} \end{equation} As a check, if we set $k=1$, $\mu$ is the identity, and out of the $N!$ permutations in ${\cal S}_N$, $(N-1)!$ have the same $\sigma(1)$ so $$ \vev{W_{{\cal A}_1}(g)}=\frac{e^{\frac{g}{2}}}{N!}(N-1)! \sum_{n=0}^{N-1}L_n^0(-g)=\frac{1}{N}L_{N-1}^1(-g) e^{\frac{g}{2}} $$ recovering the result (\ref{exactfund}) of \cite{Drukker:2000rr}. Before we proceed, it is important to notice that the result (\ref{antivev}) can be very succinctly stated in terms of the generating function for the antisymmetric representation. Define the $N\times N$ matrix $$ A(g)_{ij}= L_{i-1}^{j-i}(-g) $$ where $i,j=1,\dots,N$. The expression (\ref{antivev}) is then equivalent to \begin{equation} \vev{F_A(t)}=\sum_{k=0}^Nt^{N-k}{N\choose k}\vev{W_{{\cal A}_k}(g)}=\left |t+A(g)e^{\frac{g}{2}}\right | \label{vevoffa} \end{equation} For notation purposes, it is very convenient to define the polynomial factor of $\vev{W_{{\cal A}_k}(g)}$ in (\ref{antivev}) as \begin{equation} \vev{W_{{\cal A}_k}(g)}=\frac{1}{{N\choose k}}P_k(g) e^{k\frac{g}{2}} \label{defthep} \end{equation} The polynomials $P_k(g)$ have as generating function, $$ \sum _{k=0}^NP_k(g)t^{N-k}=\left |t+A(g)\right | $$ The contribution of the exponential factor in (\ref{vevoffa}) is easy to keep track of, and in what follows we will mostly discuss properties of $|t+A(g)|$ rather than $|t+A(g)e^{\frac{g}{2}}|$. A first property of $|t+A(g)|$ that is not manifest from the definition of $A$, is that it is a palindromic polynomial in $t$, that is, $P_k(g)=P_{(N-k)}(g)$. This follows by construction from the relation (\ref{particlehole}), and can also be proven from the definition of the matrix $A$ (see appendix C). A second property of $|t+A(g)|$ is that $P_k(g)$ can be written as a linear combination of the monomials $g^i/i!$ with positive integer coefficients. That is, we have \begin{equation} P_k(g)=\sum_{j=0}^{k(N-k)} d_j(k,N)\frac{g^j}{j!} \hspace{1cm}d_j(k,N)\in \mathbb{N} \label{definethed} \end{equation} It is easy to argue that the coefficients $d_j(k,N)$ are integers: the definition of the generalized Laguerre polynomials (\ref{thelaguerres}) implies that each entry in the matrix $A_{ij}(g)$ is a linear combination of terms $g^j/j!$ with integer coefficients, and in computing the determinant $|t+A|$, products, sums and subtractions of integers give integers; furthermore $$ \frac{g^i}{i!}\frac{g^j}{j!}={i+j \choose i}\frac{g^{i+j}}{(i+j)!} $$ so we can conclude that the coefficients of $|t+A(g)|$ are linear combinations of the monomials $g^i/i!$ with integer coefficients. The proof that these coefficients are all positive will take a little more effort, and we postpone it for a moment. Although we won't dwell in this direction, it is possible to promote $t$ and $g$ to complex variables and consider $|t+A(g)|$ as a spectral curve \cite{Hartnoll:2006is}. The fact that $|t+A(g)|$ is a palindromic polynomial in $t$ implies that roots come in pairs $t_i, 1/t_i$ (except $t_i=-1$, that appears unpaired for $N$ odd)\footnote{If, in analogy with the analysis of the spectral curve of classical strings in $AdS_5\times S^5$ (see \cite{SchaferNameki:2010jy} for a review), we define quasimomenta $p_j$ by $t_j=e^{ip_j}$, this $\bZ_2$ involution translates into the quasi-momenta coming in pairs $(p_j,-p_j)$.}. For $g$ real, if $t_i$ is a root, so is $t_i^*$. From the fact that all coefficients in $P_k(g)$ are real and positive, we learn that for $g>0$, the roots of $|t+Ae^{g/2}|$ can't be positive real numbers. Numerical experimentation suggests the following picture: for $g>0$ all roots are real and negative; at $g=0$ all eigenvalues are equal to -1, and as $g\rightarrow +\infty$, half of them tend to $-\infty$ as powers of $g$, while the other half are their pairs $1/t_i$ and tend to zero. This is consistent with the observation of \cite{Hartnoll:2006is} that at large N, the discrete zeros coalesce on a branch cut along the negative real axis. It is easy to compute $|t+A(g)|$ at linear order in $g$. For $g=0$, the matrix $A$ is upper triangular, so the g-independent term in the determinant is immediately computed to be $(1+t)^N$. At linear order in $g$, the matrix is no longer upper triangular, there are non-zero matrix elements immediately below the diagonal, $A_{i,i+1}=g$. It is nevertheless still straightforward to compute the determinant to this order by evaluating minors, and the final result is $$ |t+A(g)|=(1+t)^N+{N \choose 2}t(t+1)^{N-2}g+{\cal O}(g^2) $$ From this we deduce the first terms in the polynomial entering the vevs of the Wilson loops, and expanding the exponential, find the coefficient at order $g$, \begin{equation} \vev{W_{k,N}}=\frac{1}{{N\choose k}} \left({N\choose k}+{N\choose 2}{N-2 \choose k-1}g+\dots\right)e^{\frac{gk}{2}} =\left(1+\frac{c_2({\cal A}_k)}{2}g+\dots\right) \label{linearvev} \end{equation} In the next section we will prove that for an arbitrary representation ${\cal R}$, the term linear in $g$ has coefficient $c_2({\cal R})/2$. We have derived a formula that gives the vevs of the Wilson loops in terms of a matrix that has generalized Laguerre polynomials as entries, eq. (\ref{antivev}). We are now going to derive a formula for the individual coefficients. In particular, we will recover the results in (\ref{linearvev}) and prove that the coefficients $d_j(k,N)$ in (\ref{definethed}) are positive. Starting with equation (\ref{antivev}), the first observation is that for any given $\sigma \in S_N$, there are $(N-k)!$ permutations $\tilde \sigma\in S_N$ such that $\tilde \sigma (1)=\sigma(1),\dots, \tilde \sigma(k)=\sigma(k)$, and they all contribute the same to the sum in (\ref{antivev}). In fact, the same is true if $\{\tilde \sigma(1),\dots,\tilde \sigma(k)\}$ is a permutation of $\{\sigma(1),\dots,\sigma(k)\}$ so \begin{align*} P_k(g) & =& \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1} & \sum _{\mu \in S_k} (-1)^{\epsilon(\mu)}\prod_{m=1}^k L_{\sigma(m)}^{\mu(\sigma(m))-\sigma(m)}(-g)= \\ & = & \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1} & \sum _{\mu \in S_k} (-1)^{\epsilon(\mu)}\prod_{m=1}^k \sum_{n_m=0}^{\sigma_m}{\sigma_{\mu(m)} \choose \sigma_m-n_m}\frac{g^m}{n_m!} \end{align*} To proceed, it is convenient to define $\tau_m=\sigma_m-n_m$. With the understanding that $\frac{1}{a!}=0$ for $-a\in \mathbb{N}$ we can extend the range of $\tau_m$ to $N-1$ $$ P_k(g)=\sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1} \sum_{\tau_1,\dots, \tau_k=0}^{ N-1} \left(\prod_{n=1}^k \frac{g^{\sigma_n-\tau_n}}{(\sigma_n-\tau_n)!}\right) \sum _{\mu \in S_k} (-1)^{\epsilon(\mu)}\prod_{m=1}^k {\sigma_{\mu(m)}\choose \tau_m} $$ In the expression above, the last sum is antisymmetric in $\tau_m$, so we can restrict the sum over ${\tau_m}$ to k-tuples of different $\tau_i$. We write it as a sum over ordered k-tuples and its permutations, \begin{align*} P_k(g )& = & \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1\atop 0\leq \tau_k<\dots <\tau_1\leq N-1} g^{\sum_m \sigma_m-\tau_m} & \sum_{\nu\in S_k} \prod_{m=1}^k\frac{1}{(\sigma_m-\tau_{\nu_m})!} \sum_{\mu \in S_k} (-1)^{\epsilon(\mu)} \prod_{n=1}^k {\sigma_{\mu_n} \choose \tau_{\nu_n}} \\ & = &\sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1 \atop 0\leq \tau_k<\dots <\tau_1\leq N-1} g^{\sum_m \sigma_m-\tau_m} & \sum_{\nu\in S_k} \prod_{m=1}^k\frac{1}{(\sigma_m-\tau_{\nu_m})!} \epsilon_{\nu_1 \dots \nu_k} \sum_{\mu \in S_k} (-1)^{\epsilon(\mu)} \prod_{n=1}^k {\sigma_{\mu_n}\choose \tau_n} \\ & = & \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1 \atop 0\leq \tau_k<\dots <\tau_1\leq N-1} g^{\sum_m \sigma_m-\tau_m} & \prod_{m=1}^k \frac{\tau_m!}{\sigma_m!} \left | {\sigma_i \choose \tau_j} \right |^2 \end{align*} where in the second line we used the properties of reordering the rows of a determinant. Collecting all the terms with $g^n$, we arrive then at a formula for the coefficients $d_n(k,N)$ in (\ref{definethed}) that makes manifest that they are positive, \begin{equation} d_n(k,N)=n! \sum_{0\leq \sigma_k<\dots \sigma_1\leq N-1 \atop 0\leq \tau_k<\dots \tau_1\leq N-1} \prod_{m=1}^k \frac{\tau_m!}{\sigma_m!} \left | {\sigma_i \choose \tau_j} \right |^2 \delta _{\sum (\sigma_m-\tau_m),n} \label{bcoeff} \end{equation} As a check, if we set $k=1$ we arrive at $$ d_n(1,N)= \sum_{\sigma=n}^{N-1}{\sigma \choose \sigma -n}= {N \choose n+1} $$ reproducing the expansion of $P_1(g)=L_{N-1}^1(-g)$. \subsection{Relation with skew Young tableaux and free fermions} We have just derived an expression, eq. (\ref{bcoeff}), for the coefficients $d_n(k,N)$ of the polynomial $P_k(g)$ in (\ref{definethed}). We would like to recast it in terms of a sum over skew Young diagrams (see appendix A for a brief introduction). By pulling common factors out of the binomial determinants, we can write these coefficients in various ways, \begin{align} d_n(k,N)& =& n! \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1\atop 0\leq \tau_k<\dots <\tau_1\leq N-1} & \left(\prod_{m=1}^k \frac{\tau_m!}{\sigma_m!} \right) \left | {\sigma_i \choose \tau_j} \right |^2 \delta _{\sum_{m=1}^k (\sigma_m-\tau_m),n}= \nonumber \\ & =& n! \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1 \atop 0\leq \tau_k<\dots <\tau_1\leq N-1} & \left(\prod_{m=1}^k \frac{\sigma_m!}{\tau_m!} \right) \left | \frac{1}{(\sigma_i - \tau_j)!} \right |^2 \delta _{\sum_{m=1}^k (\sigma_m-\tau_m),n} = \nonumber \\ & =& n! \sum_{0\leq \sigma_k<\dots <\sigma_1\leq N-1 \atop 0\leq \tau_k<\dots <\tau_1\leq N-1} & \left | {\sigma_i \choose \tau_j} \right | \left | \frac{1}{(\sigma_i - \tau_j)!} \right | \delta _{\sum_{m=1}^k (\sigma_m-\tau_m),n} \label{dcoeff} \end{align} An important observation is that the sum above can be restricted to pairs of k-tuples such that $\sigma_i\geq \tau_i$ for $i=1,\dots,k$. The reason is that if for some $j$ it happens that $\sigma_j<\tau_j$, the matrix with binomial coefficients in (\ref{bcoeff}) has a zero block in the upper right corner. The full determinant is then the product of determinants of the diagonal blocks, but the determinant of the lower diagonal block is zero, since it has a zero row. In what follows, it is understood that the sum in (\ref{dcoeff}) is restricted to $\sigma_i\geq \tau_i$. Now, for every pair of k-tuples $\{\sigma\},\{\tau\}$, we define $$ \lambda_i=\sigma_i-k+i\hspace{1cm} \mu_i=\tau_i-k+i \hspace{1cm}i=1,\dots,k $$ It is easy to see that $\lambda_i\geq \lambda_{i+1}$ and $\mu_i\geq \mu_{i+1}$, so $\{\lambda\}$ and $\{\mu\}$ are partitions. It also follows that $\lambda_i\leq N-k$ and $\mu_i\leq N-k$ so the corresponding Young diagrams can be both fitted in a rectangle with $(N-k)\times k$ boxes. If we denote by $L(m,n)$ be the set of all Young diagrams that fit into a board of $m$ rows and $n$ columns \cite{stanalgebra}, we have just argued that the partitions $\lambda$ and $\mu$ have Young diagrams in $L(k,N-k)$. Finally, it is also easy to prove that $\lambda$ and $\mu$ are partitions satisfying $\lambda_i\geq \mu_i$, so the Young diagram of $\mu$ fits inside the Young diagram of $\lambda$, and it makes sense to consider its complement, the skew Young diagram $\lambda/\mu$, see table \ref{tableskew} in appendix A. The Kronecker delta in (\ref{dcoeff}) suggests that the coefficients $d_n(k,N)$ can be rewritten in terms of skew Young diagrams with $n$ boxes. An important step in this direction is to recognize the determinant in the second line of (\ref{dcoeff}) as the one that appears in the formula (\ref{aitkenform}) giving the number $f_{\lambda/\mu}$ of standard Young tableaux for the skew diagram $\lambda/\mu$, \begin{equation} d_n(k,N)= \frac{1}{n!} \sum_{\mu \subseteq \lambda \in L(k,N-k), \atop| \lambda|-|\mu|=n} \prod_{m=1}^k \frac{(\lambda_m+k-m)!}{(\mu_m+k-m)!} f_{\lambda/\mu}^2 \label{combid} \end{equation} The various expressions we have derived for the coefficients $d_n(k,N)$, either the original formulas (\ref{dcoeff}) or the one just derived, eq. (\ref{combid}), involve ingredients that have a number of combinatorial interpretations, and in particular can be interpreted as counting paths of fermions in different lattices. The binomial determinant appearing in the first line of (\ref{combid}) was given a beautiful interpretation as counting the number of $k$ non-intersecting ({\it i.e.} fermionic) paths in a 2d-lattice, with the $k$-tuples $\{\sigma\}$ and $\{\tau \}$ giving respectively the initial and final conditions for the $k$ paths \cite{gessel}. However, if we try to pursue this interpretation, we don't know what meaning we should assign to the prefactor in the first line of (\ref{combid}). A second possibility is to try to interpret these coefficients in terms of time-dependent processes for fermions on a 1d lattice. The first step in this direction is to map each Young diagram to a configuration of fermions (see {\it e.g.} \cite{zinn-justin}). A standard Young tableau of skew shape $\lambda/\mu$ is then interpreted as a time-dependent process, with $\mu$ being the initial configuration, $\lambda$ being the final configuration, and the labeling of the boxes indicating the order in which they appear \cite{zinn-justin, stanalgebra}. $f_{\lambda/\mu}$ counts the number of ways to evolve from $\mu$ to $\lambda$, but again we don't know of a clear combinatorial interpretation for the prefactor in (\ref{combid}). Perhaps the best way of summarizing our lack of a simple combinatorial interpretation for these coefficients is the third line in (\ref{combid}); there the two determinants with interpretations outlined above appear, but as we have seen each one hints at a different physical realization. \subsection{Evaluating the coefficients} We have derived a formula for the coefficients $d_j(k,N)$, eq. (\ref{bcoeff}). The index $j$ runs from $0$ to $k(N-k)$, and for arbitrary values of it, it seems doubtful that the sum can carried out explicitly. We will now evaluate these coefficients for a few values of $j$, close to the endpoints of its range. As a consistency check, all the explicit results we obtain satisfy $d_n(k,N)=d_n(N-k,N)$. An important ingredient in the evaluation is that as argued in the previous subsection, we can restrict to pairs of k-tuples such that $\sigma_i\geq \tau_i$ for $i=1,\dots,k$. For $n=0$, both k-tuples have to be identical to contribute: $\sigma_i=\tau_i$. In this case the matrix with entries ${\sigma _i\choose \tau_j}$ is lower triangular, the determinant in (\ref{bcoeff}) is 1, as well as the prefactor, so $d_0$ is given by the number of k-tuples, $$ d_0={N\choose k} $$ Alternatively, in the language of skew Young diagrams, $n=0$ corresponds to the case of $\lambda=\mu$ and $d_0$ is just counting the number of Young diagrams that fit into a rectangle with $(N-k)\times k$ boxes, which is precisely ${N \choose k}$ \ (proposition 6.3 in \cite{stanalgebra}). For $n=1$, given a k-tuple $\tau_i$, the only k-tuples $\sigma_i$ that contribute are those where all the $\sigma_i=\tau_i$, except for precisely one element $\sigma_j=\tau_j+1$. For each of those cases the matrix with entries ${\sigma _i\choose \tau_j}$ is lower triangular, the determinant is $\sigma_j$ and the contribution in each case is $\sigma_j$. It remains to count how many such pairs of k-tuples there are, which is easily seen to be ${N-2 \choose k-1}$. Adding all contributions we obtain $$ d_1={N\choose 2}{N-2\choose k-1} $$ These two computations reproduce the result obtained for the Wilson loop by expanding $|t+A(g)|$ to linear order in $g$, eq. (\ref{linearvev}). For $n=2$, there are two types of contributions. There are contributions from pairs of k-tuples when all $\sigma_i=\tau_i$ except for a single $\sigma_j=\tau_j+2$. There are also contributions from pairs of k-tuples when $\sigma_m=\tau_m$ except for two $\sigma$s, $\sigma_i=\tau_i+1$ and $\sigma_j=\tau_j+1$, with $i<j$. It is convenient to treat separately the cases where $\tau_j=\sigma_i$ (in which case the matrix fails to be lower triangular) and the case $\tau_j>\sigma_i$. By arguments very similar to the ones in the previous cases we arrive at $$ d_2=\frac{N!}{12 (k-1)!(N-k-1)!}\left(3k(N-k)-N-1\right) $$ This coefficient allows us to write the perturbative expansion of the antisymmetric Wilson loop to order $g^2$, \begin{equation} \vev{W_{{\cal A}_k}}=1+\frac{c_2({\cal A}_k)}{2}g+\left(\frac{1}{4}c_2({\cal A}_k)^2-\frac{N+1}{12}\left(c_2({\cal A}_k)-c_1({\cal A}_k) \right) \right) \frac{g^2}{2!}+\dots \label{antigtwo} \end{equation} Having computed the first three coefficients $d_j(k,N)$, we turn to the other end of the range, when $j$ is close to $k(N-k)$. For $n=k(N-k)$, there is only one term that contributes: $\sigma_i=N-k-1+i$, $\tau_j=j-1$. The determinant is 1, as can be proven by induction on $k$, for $N$ fixed. Therefore $$ d_{k(N-k)}(k,N)= \left( k(N-k)\right)!\; \frac{0! 1!\dots (k-1)!}{(N-1)! (N-2)!\dots (N-k)!} $$ For $n=k(N-k)-1$, there are two terms that contribute. The first one has $\sigma_i=N-k-1+i$, $\tau_j=0,1,\dots,k-2,k$; the corresponding determinant is $N-k$. The second term has $\sigma_i=N-k-1,N-k+1,\dots,N-1$ and $\tau_j=j-1$; the corresponding determinant is $k$. Adding these two terms one obtains $$ d_{k(N-k)-1}(k,N)= \left( k(N-k)\right)!\; \frac{0! 1!\dots (k-1)!}{(N-1)! (N-2)!\dots (N-k)!}N $$ \section{Arbitrary Representations} In this section we will perform two different types of computations, both regarding Wilson loops in arbitrary representations ${\cal R}$ of the group U(N). First we will consider a perturbative expansion in $g$ for the vev of Wilson loops for arbitrary representation and at finite N. We will show that the Schur function evaluated on exponentials of eigenvalues admits an expansion in terms of Schur functions evaluated on eigenvalues, whose vevs are known exactly for the Gaussian Hermitian matrix model. The coefficients of this expansion turn out to be essentially {\it shifted Schur functions} \cite{okounkov}, evaluated on the components of the highest weight $\lambda$ of the representation ${\cal R}$. In order to bring the resulting expressions closer to familiar quantities, we would have to write these shifted Schur polynomials in terms of the Casimir invariants of the representation, $c_p({\cal R})$. This is quite easy at order $g$, as we show, but it already becomes quite cumbersome at higher orders. Nevertheless, if we settle for finding the highest degree Casimir combination appearing at each order $g^n$, this turns out to be a solvable problem with an extremely simple answer. We then switch gears and compute the exact expectation value of a complete basis of the space of symmetric polynomials. The most convenient one turns out to be the basis of monomial symmetric functions; from these one can then recover the vevs of Wilson loops by a linear transformation. \subsection{Perturbative computation} For a generic representation ${\cal R}$, the vev of the Wilson loop is given by the following integral, \begin{equation} \vev{W_{\cal R}(g)}=\frac{1}{\dim\mathcal{R}}\frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 S_{{\cal R}}\left(e^{x_1},\dots,e^{x_N}\right) e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} \end{equation} We want to address the evaluation of this integral for arbitrary representation ${\cal R}$, for finite $N$ and perturbatively in $g$. By a rescaling of the eigenvalues $x_i\rightarrow \sqrt{g} x_i$ we learn that an expansion in $g$ amounts to expanding the exponentials of eigenvalues in the Schur function $S_{{\cal R}}\left(e^{x_1},\dots,e^{x_N}\right)$: to compute the term at order $g^n$ we need to expand the exponentials up to terms $x^{2n}_i$. The key ingredient to expand Schur functions around unity is the following formula (see section I.3, example 10 in \cite{macdonald}), $$ s_\lambda(1+y_1,\dots,1+y_N)=\sum _{\mu} b_{\lambda _\mu} s_{\mu}(y_1,\dots,y_N) $$ where $b_{\lambda \mu}$ is given by a determinant of binomial coefficients, $$ b_{\lambda \mu}=\left | {\lambda_i+N-i \choose \mu_j+N-j} \right | $$ Let's start by considering the computation to one-loop, {\it i.e.} to order $g$. At this stage, we prefer not to resort to some of the heavier machinery that we will introduce for higher loops, and keep the computation intuitive. We need to expand the Schur function $S_{\cal R}(e^{x_1},\dots,e^{x_N})$ to order $x_i^2$, and the answer is \ytableausetup{boxsize=6pt} \begin{align} S_{\cal R}(e^{x_1},\dots,e^{x_N}) & = b_{{\cal R}\emptyset}+ b_{{\cal R}\; \ydiagram{1}}(x_1+\dots +x_N)+\left(\frac{1}{2}b_{{\cal R}\; \ydiagram{1}}+b_{{\cal R}\; \ydiagram{2}}\right)\left(x_1^2+\dots+x_N^2\right)+ \nonumber \\ & + \left(b_{{\cal R}\; \ydiagram{2}}+b_{{\cal R}\; \ydiagram{1,1}}\right)\sum_{i<j}x_i x_j+\dots \label{expandschur} \end{align} The next step is to evaluate the binomial determinants $b_{{\cal R} \mu}$ that appear in (\ref{expandschur}). We will express the results in terms of the Casimir invariants of the representation ${\cal R}$, written as polynomials of the components of its highest weight $\lambda$. The generating function of the Casimirs of ${\cal R}$ can be written in terms of $\sigma_i=\lambda_i+N-i$ as \cite{barut} \begin{equation} G(z)=\sum c_p(\sigma)z^p=\frac{1}{z} \left[1-\prod_{i=1}^N\left(1-\frac{z}{1-\sigma_i z}\right)\right] \label{genercasi} \end{equation} To this order only the first two Casimirs can appear, $$ c_1({\cal R})=|\lambda|=\sum_i \lambda_i $$ $$ c_2({\cal R})=\sum_i \lambda_i(\lambda_i+N+1-2i) $$ We obtain \begin{align} b_{{\cal R}\emptyset}=\hbox{dim }{\cal R}, & & b_{{\cal R}\; \ydiagram{2}}=\frac{\hbox{dim }{\cal R}}{2n(n+1)}\left( c_2({\cal R})+c_1({\cal R})^2-(n+1)c_1({\cal R})\right) \nonumber \\ b_{{\cal R}\; \ydiagram{1}}=\hbox{dim }{\cal R}\frac{c_1({\cal R})}{N} & & b_{{\cal R}\; \ydiagram{1,1}}=\frac{\hbox{dim }{\cal R}}{2n(n-1)}\left(- c_2({\cal R})+c_1({\cal R})^2+(n-1)c_1({\cal R})\right) \label{thebdets} \end{align} The last ingredient we need are the vevs $\vev{x_1+\dots+x_N}$, $\vev{x_1^2+\dots+x_N^2}$ and $\vev{\sum_{i<j} x_i x_j}$. These vevs can be easily computed with the method of orthogonal polynomials, making repeated use of the recurrence relations among the orthogonal polynomials \cite{Marino:2004eq}, but an ever easier way to compute them is to write them as linear combinations of the known n-point functions $\vev{\hbox{tr}X^2}$ and $\vev{\hbox{tr}X\hbox{tr}X}$. Either way, we obtain, \begin{equation} \vev{x_1+\dots+x_N}=0, \; \; \vev{x_1^2+\dots+x_N^2}=gN^2, \hspace{1cm} \vev{\sum_{i<j} x_i x_j}=-{N \choose 2}g \label{easyvevs} \end{equation} Plugging the results for the coefficients (\ref{thebdets}) and the vevs (\ref{easyvevs}) back in the expansion (\ref{expandschur}), we finally obtain, \begin{equation} \vev{W_{\cal R}(g)}=1+\frac{c_2(\cal R)}{2}g+{\cal O}(g^2) \label{genrepatg} \end{equation} In principle the method presented above can be used at higher loops. To compute efficiently the coefficients $b_{{\cal R} \mu}$ it is very useful to realize that they are essentially given by {\it shifted Schur functions}, $s_\mu^*({\cal R})$ \cite{okounkov} \begin{equation} b_{{\cal R} \mu}=\hbox{dim}{\cal R} \prod_{i=1}^N \frac{(N-i)!}{(\mu_i+N-i)!} s_\mu^*({\cal R}) \label{shiftedschur} \end{equation} The reason this connection is useful is that shifted Schur functions have many properties that generalize the properties of ordinary Schur functions; in particular, they admit a combinatorial definition in terms of reverse Young tableaux \cite{okounkov} which we have found quite efficient when it comes to actually computing them. We have carried out the computation of all necessary shifted Schur functions needed at order $g^2$ (i.e. corresponding to Young diagrams with up to 4 boxes). Unfortunately, the expressions obtained are long and far from illuminating, and we haven't found an efficient way to rewrite the coefficients $b_{{\cal R}\mu}$ in terms of the Casimir invariants that can appear: $c_1({\cal R}),\dots, c_4({\cal R})$. In this work we will settle for a more modest problem, which is the following: assign a degree $p$ to the Casimir invariant $c_p({\cal R})$, corresponding to the highest power of $\lambda_i$ that appears in $c_p({\cal R})$, see (\ref{genercasi}). It follows that a product of Casimirs $c_{p_1} c_{p_2}$ has degree $p_1+p_2$, and it is easy to see that at order $g^n$ the highest degree that can appear is $2n$. For instance, at order $g^2$, the expansion of the Schur function $S_{\cal R}(e^{x_1},\dots,e^{x_N})$ will involve $c_4({\cal R}), c_3({\cal R})c_1({\cal R}),c_2({\cal R})^2,c_2({\cal R})c_1({\cal R})^2$ and $c_1({\cal R})^4$, plus terms of lower degree ({\it e.g.} $c_3({\cal R})$ or $c_1({\cal R})^2$). Our aim in what follows it to compute the term with highest degree possible ({\it i.e.} degree $2n$) in the coefficient of $g^n$, at every order $n$ in the $g$ expansion. The first observation is that the only coefficients $b_{{\cal R} \mu}$ that contribute to the term with degree $2n$ are those with $\mu$ being a partition of $2n$: $\mu$ with a smaller number of boxes can only contribute to lower degree terms. By (\ref{shiftedschur}), this amounts to considering shifted Schur functions $s_\mu^*({\cal R})$, with $\mu$ being a partition of $2n$. Furthermore, each $s_\mu^*({\cal R})$ can be written as a sum of the ordinary Schur function $s_\mu({\cal R})$ plus lower degree polynomials \cite{okounkov} that don't contribute to the term we are considering. To recapitulate, we have argued that the degree $2n$ term at order $g^n$ is given by \begin{equation} \hbox{dim}{\cal R} \sum_{|\mu|=2n} \prod_{i=1}^N \frac{(N-i)!}{(\mu_i+N-i)!} s_\mu({\cal R}) \vev{s_\mu(x)} \label{sumone} \end{equation} In order to proceed, we will now make use that $\vev{s_\mu(x)}$ is known exactly for the Gaussian Hermitian matrix model \cite{DiFrancesco:1992cn}. The result of \cite{DiFrancesco:1992cn} can be written as \begin{equation} \vev{s_{\mu}(x)}=\frac{1}{2n!} \#[2^n] \chi_{\mu}[2^n] \prod_{i=1}^N\frac{(\mu_i+N-i)!}{(N-i)!} \label{exactschur} \end{equation} where $[2^n]$ is the conjugacy class in ${\cal S}_{2n}$ with $n$ disjoint 2-cycles, and $\#[2^n]$ gives the number of elements in this conjugacy class, see eq. (\ref{elemconj}). When we plug (\ref{exactschur}) into (\ref{sumone}) the fractions of products cancel out. We can now write $s_\mu({\cal R})$ in terms of power sum polynomials, see eq. (\ref{frobenius}), $$ s_{\mu}({\cal R})=\frac{1}{2n!} \sum_\nu \#[\nu] \chi_\mu[\nu]p_\nu({\cal R}) $$ and make use of the orthogonality of the characters to simplify the coefficient (\ref{sumone}) as $$ \hbox{dim }({\cal R})\frac{\#[2^n]}{(2n)!}p_{[2^n]}({\cal R}) $$ Now, $p_{[2^n]}({\cal R})$ differs from $c_2({\cal R})$ only in lower degree terms, so for the purpose of computing the highest degree term, we can replace $p_{[2^n]}({\cal R})\rightarrow c_2({\cal R})$. Plugging the value for $\#[2^n]=(2n!)/(n! 2^n)$ - see eq. (\ref{elemconj}) - we arrive at the main result of this section, $$ \vev{W_{R}}=\sum_{n=0}^\infty \left [\left(\frac{c_2({\cal R})}{2}\right)^n+\dots\right ] \frac{g^n}{n!} $$ To reiterate, the dots stand for terms that we are missing at every order in $g^n$, that are of degree in Casimirs lower than $2n$, see (\ref{antigtwo}) for their explicit expression in the antisymmetric representation. These terms that we haven't computed come from different sources: first there are the contributions from shifted Schur functions with $|\mu|<n$; for instance, at order $g$ we see in (\ref{expandschur}) that $b_{{\cal R}\; \ydiagram{1}}$ also contributes, and in fact will contribute to every order. Second, when considering the shifted Schur functions with $|\mu|=n$, we replaced them by ordinary Schur functions, since they differ by lower degree polynomials. \subsection{Exact results in the monomial basis} The Schur polynomials are the characters of irreducible representations of U(N), and they form a basis of the space of symmetric functions of $N$ variables, but it turns out that they don't form the most convenient basis to perform the integrals above. This is similar to what was encountered in the study of two-dimensional QCD \cite{Gross:1993yt}. The route taken there is to use Frobenius formula to relate the Schur polynomial to products of $\hbox {tr }U^k$ (i.e. multiply wound Wilson loops). In mathematical language this corresponds to changing basis from the Schur one to the {\it power sum symmetric functions}, $p_\lambda$. In the case at hand $\vev{p_k}$ is immediate to compute by a simple rescaling of the integral considered in \cite{Drukker:2000rr}, $$ \vev{p_k(g)}=\frac{1}{N}L_{N-1}^1(-k^2g)e^{k^2\frac{g}{2}} $$ However, to evaluate the integral in a full basis of the ring of symmetric polynomials, we would have to compute $\vev{p_\lambda}$ for arbitrary partitions $\lambda$, and when facing this problem, we end up having to compute $\vev{m_\lambda}$. For this reason, we consider the integrals above in the basis of {\it monomial symmetric polynomials} $m_\lambda$. These results can then be used to write down the vevs of Schur polynomials, since they are related by a linear transformation $s_\lambda=\sum _\tau K_{\lambda \tau} m_\tau $, where the coefficients $K_{\lambda \tau}$ are positive integers, the so-called {\it Kostka numbers} (see appendix B). The vev of monomial symmetric functions is \begin{equation} \vev{m_\tau}=\frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 m_{{\tau}}\left(e^{x_1},\dots,e^{x_N}\right) e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} \end{equation} since $m_\tau$ is a symmetric polynomial, and the rest of the integral is also symmetric, each monomial in $m_\tau$ gives the same contribution to the integral, so we can restrict to just one of them. Denote by $\ell(\tau)$ the number of non-zero entries of $\tau$, and by $\alpha_i(\tau)$ the number of times the number $i$ appears in the partition $\tau=(\tau_1,\tau_2,\dots)$. Then the number of monomials in $m_\tau$ is $$ {N \choose \ell(\tau)} {\ell(\tau) \choose \alpha_1(\tau),\alpha_2(\tau),\dots} $$ so $$ \vev{m_\tau}={N \choose \ell(\tau)} {\ell(\tau) \choose \alpha_1(\tau),\alpha_2(\tau),\dots} \frac{1}{{\cal Z}}\int \prod_{i=1}^N \frac{dx_i}{2\pi} \prod_{i<j} |x_i-x_j|^2 e^{\tau_1 x_1+\dots + \tau_{\ell(\tau)} x_{\ell(\tau)}} e^{-\frac{1}{2g}\sum _{k=1}^N x_k^2} $$ Now the argument proceeds along the same lines as in the previous section, for the antisymmetric representation. After substituting the Vandermonde determinant for a determinant of orthogonal polynomials, we expand these determinants. For the eigenvalues $x_{\ell(\tau)+1},\dots, x_N$ the insertion of the operator does not suppose any change, and we can carry out the integrals as always. The new integral that appears can again be computed explicitly \cite{gradshteyn} $$ \int \frac{dx}{2\pi} p_i(x)p_j(x) e^{\tau x} e^{-\frac{1}{2g}x^2}= \sqrt{\frac{g}{2\pi}} g^j i! \tau^{j-i}L_i^{j-i}(-\tau^2g)e^{\frac{g}{2}\tau^2} $$ so we arrive at \begin{equation} \vev{m_\tau}=\frac{1}{N!}{N \choose \ell(\tau)} {\ell(\tau) \choose \alpha_1(\tau),\alpha_2(\tau),\dots} \sum _{\sigma \in S_N} \sum _{\mu \in S_{\ell(\tau)}} (-1)^{\epsilon(\mu)} \prod_{m=1}^{\ell (\tau)} L_{\sigma(m)-1}^{\mu(\sigma(m))-\sigma(m)}(-\tau_m^2g) e^{\tau_m^2\frac{g}{2}} \label{monovev} \end{equation} We can also introduce a generating functional for these vevs. It contains the antisymmetric representation as a particular case, but is more complicated. Define an infinite family of $N\times N$ matrices, labelled by $n\in N$ $$ A_{ij}^{(n)}(g)=n^{j-i}L_{i}^{j-i}(-n^2g) $$ The vevs (\ref{monovev}) have then the following generating function $$ \vev{\prod_{i=1}^N\left(t+\sum_{j=1}^\infty y_j e^{j x_i}\right)}= \left |t+\sum_{j=1}^\infty y_j A^{(j)}(g) e^{j^2\frac{g}{2}}\right | $$ Finally, as in the antisymmetric case, an explicit formula for the $\vev{m_{\tau}}$ can be found in terms of determinants of binomial coefficients: \begin{equation} \vev{m_{\tau }(g)} = \frac{e^{\frac{g}{2}\sum_{i} \tau_{i}^{2}} }{\prod_{i} \alpha_{i}!} \sum_{c \in S_{\ell(\tau)}} \sum_{\mu \subseteq \lambda \in L(\ell,N-\ell)} \prod_{m=1}^\ell \frac{(\lambda_m+\ell-m)!}{(\mu_m+\ell-m)!} \left|\frac{\tau_{c_{i}}^{\lambda_{i}-\mu_{j}+j-i}}{(\lambda_{i}-\mu_{j}+j-i)!}\right|^2 g^{\sum \lambda_m-\mu_m} \label{thebigsum} \end{equation} We have then managed to evaluate exactly the vevs of these symmetric functions in the Gaussian Hermitian matrix model. The vevs of the Schur polynomials are then linear combinations of these, with the transformation matrix given by {\it Kostka numbers}, see appendix B. The general form of the vevs of Wilson loops is then \begin{equation} \vev{W_{\cal R}(g)}=\frac{1}{\dim \mathcal{R}} \sum_{\tau \leq {\cal R}} K_{{\cal R} \tau} P_\tau(g) e^{\sum _i \tau_i^2\frac{g}{2}} \label{generalvev} \end{equation} The polynomials in (\ref{generalvev}) can be read from either (\ref{monovev}) or (\ref{thebigsum}). A relevant question, already answered in \cite{Gomis:2009ir}, is what is the largest exponent in (\ref{generalvev}), and it is immediate to see that it corresponds to $\sum _i \tau_i^2$ where $\tau$ is the highest weight of the representation ${\cal R}$. \section{Bremsstrahlung functions} Having devoted the previous sections to the computation of the exact vevs of half-BPS circular Wilson loops of ${\cal N}=4$ SYM, we would like to use the results obtained to discuss some properties of the associated heavy probes. We will do that by considering the corresponding Bremsstrahlung functions. The Bremsstrahlung function can be defined as the small angle limit of the cusp anomalous dimension, and for any heavy probe coupled to any four dimensional conformal field theory, it determines a number of interesting properties, like the two-point function of displacement operators \cite{Correa:2012at}, or the total radiated power \cite{Correa:2012at, Fiol:2012sg} and the momentum diffusion coefficient \cite{Fiol:2013iaa} of an accelerated probe. Therefore computing the Bremsstrahlung function for probes of different conformal field theories is an interesting but in general difficult question. For $1/2$ BPS probes in ${\cal N}=4$ SYM, the situation is more favorable, since it was argued in \cite{Correa:2012at} that for these probes the Bremsstrahlung function can be derived from the vev of the corresponding circular Wilson loop \begin{equation} B_{\cal R}(\lambda,N)=\frac{1}{2\pi^2}\lambda\partial_\lambda \hbox{ log}\vev{W_{\cal R}} \label{brefun} \end{equation} The argument leading to this relation is independent of the representation, so we can put to use our results for $\vev{W_{\cal R}}$. Since our results are most explicit for the antisymmetric representation, let's start with this case. For the antisymmetric representations, the vev of the Wilson loop is a polynomial in $g$ times an exponential, see eqs. (\ref{antivev}) or (\ref{defthep}). From this simple fact, it follows that when taking the logarithmic derivative in (\ref{brefun}) the final answer is a rational function in the coupling, $$ B_{{\cal A}_k}^{U(N)}= \frac{\lambda}{16\pi^2 N} \frac{\sum_{j=0}^{k(N-k)} \frac{2d_{j+1}+kd_j}{j!}\left(\frac{\lambda}{4N}\right)^j} {\sum_{j=0}^{k(N-k)}\frac{d_j}{j!} \left(\frac{\lambda}{4N}\right)^j} $$ with the understanding that $d_{k(N-k)+1}=0$. For fixed $N$, both at weak and at strong 't Hooft coupling, the Bremsstrahlung function is linear in the 't Hooft coupling $$ B_{{\cal A}_k}^{U(N)}= \frac{c_2({\cal A}_k) }{16\pi^2 N}\lambda \hspace{1cm} \lambda \ll 1 $$ $$ B_{{\cal A}_k}^{U(N)}= \frac{k}{16\pi^2 N} \lambda \hspace{1cm} \lambda \gg 1 $$ Let's now briefly discuss the case of general representations. Now $\vev{W_{\cal R}}$ is given by a linear combination of $\vev{m_\tau}$, so it is a sum of polynomials times exponentials, eq. (\ref{generalvev}). Since in general the exponents in these exponentials are different, it follows from (\ref{brefun}) that the corresponding Bremsstrahlung functions are no longer rational in the coupling. On the other hand, it also follows from (\ref{brefun}) that it is still true that for fixed $N$, both at weak and at strong 't Hooft coupling, the Bremsstrahlung function is linear in the 't Hooft coupling. The weak coupling result can be read off from (\ref{genrepatg}). In the large coupling limit, the coefficient of $\lambda$ is given by the largest exponent in the exponentials which as pointed out after (\ref{generalvev}) (see also (\cite{Gomis:2009ir}) for a representation ${\cal R}$ with partition $\tau$ this exponent is $(\sum_i \tau_i^2)g/2$, so $$ B_{{\cal R}}^{U(N)}= \frac{c_2({\cal R}) }{16\pi^2 N}\lambda \hspace{1cm} \lambda \ll 1 $$ $$ B_{{\cal R}}^{U(N)}= \frac{\sum_i \tau_i^2}{16\pi^2 N}\lambda \hspace{1cm} \lambda \gg 1 $$ \section{Outlook} In closing, we would like to point out future directions and potential applications of the results obtained here. A possible direction for further research is the use of our results as a concrete example of emergent spacetime out of a matrix model. Half-BPS Wilson loops have associated certain ten-dimensional solutions of IIB supergravity, that are given by a fibration over a Riemann surface \cite{D'Hoker:2007fq}. The relation between the classical geometry and the large N limit of the matrix model has been elucidated in \cite{Okuda:2008px}. It will be interesting to explore if having exact results for the vevs of these Wilson loops for any $N$ and $\lambda$ can lead to a deeper understanding of the emergence of these spacetimes. Finally, in recent years there has been a fruitful interplay between matrix models and topological strings. For the model at hand, it was conjectured in \cite{Bonelli:2008rv} that antisymmetric Wilson loops of ${\cal N}=4$ are dual to amplitudes of the open topological string of Berkovits-Vafa \cite{Berkovits:2007rj}. It would be interesting to derive the results presented here by a direct computation of the relevant topological string amplitudes, and also to give an interpretation of positive integer coefficients $d_j$ in terms of enumerative geometry of the string background. \section{Acknowledgements} We would like to thank Marcos Mari\~no for enjoyable conversations. BF would also like to thank the Theory Division at CERN for hospitality during part of this work. The research of BF is supported by MEC FPA2010-20807-C02-02, CPAN CSD2007-00042, within the Consolider-Ingenio2010 program, and AGAUR 2009SGR00168. GT is supported by an FI scholarship by the Generalitat the Catalunya.
\section{Introduction} \ Magnetism is one of important subjects in condensed matter physics because of fundamental question of what the origin of magnets is and applications to the modern technology. The magnetism should be explained from the electromagnetic interaction. On the other hand, there are other fundamental forces in nature, such as the strong interaction acting on quarks and gluons with color degrees of freedom. Here, we report a new state of matter which we call ``color magnetism," that may appear in non-Abelian vortex lattices when high density quark matter is rotating. Quark matter at finite temperature and/or density is likely to have a rich phase structure due to a variety of internal symmetries \cite{Fukushima:2010bq}. At very high baryon number densities and low temperatures, quark matter is expected to exhibit a color superconductivity where condensation of quark Cooper pairs breaks color symmetry as well as flavor and the baryon number symmetries. For three colors and three massless flavors, the order parameter of the quark pair condensation is given by \cite{Alford:1998mk,Alford:2007xm} \beq \Phi_{a \alpha}=\epsilon_{abc}\epsilon_{\alpha \beta \gamma} \langle \psi^T_{b \beta}C\gamma_5 \psi_{c \gamma}\rangle=\Delta \delta_{a\alpha}, \eeq where $ a,b,c$ ($\alpha,\beta,\gamma$) represent the flavor (color) indices, and $\Delta$ is the quark pairing gap. The symmetry breaking pattern, ${\rm SU}(3)_{\rm C}\times {\rm SU}(3)_{\rm L}\times {\rm SU}(3)_{\rm R} \times {\rm U}(1)_{\rm B}\rightarrow {\rm SU}(3)_{\rm C+L+R}$, shows that this phase retains only a simultaneous rotation of color and flavors as the residual symmetry, thus is called color-flavor locking (CFL) phase. We denote the suffix L+R by F in the following. A possible candidate for system where the CFL phase takes place might be the core of very heavy compact stars, such as neuron stars. If the CFL phase develops in the rotating neutron star, vortices are created because of the superfluid nature associated with the $U(1)_{\rm B}$ breaking in the CFL phase. The most fundamental vortex in the CFL phase is a non-Abelian vortex \cite{Balachandran:2005ev,Eto:2009kg}, see Ref.~\cite{Eto:2013hoa} as a review. The solution is of the form \beq \Phi=\Delta {\rm diag} \ltk e^{i\theta}f(r), g(r), g(r) \rtk \eeq where $(\theta, r)$ are the polar coordinates for a vortex extended along the $z$ direction, and $f$ and $g$ are profile functions with suitable boundary conditions. Since the winding factor $e^{i\theta}$ is generated by both of the color $SU(3)_{\rm C}$ and $U(1)_{\rm B}$ rotations, the non-Abelian vortex carries a color-magnetic flux along the vortex core, and $1/3$ winding of the $U(1)_{\rm B}$ phase as well. Therefore, the non-Abelian vortex possesses nature of both superconducting and superfluid vortices, associated with breakings of the $SU(3)_{\rm C}$ color gauge symmetry and the baryon number $U(1)_{\rm B}$ symmetry, respectively. As the most remarkable property, the non-Abelian vortex has localized zero modes; the advent of the non-Abelian vortex further breaks the residual symmetry, ${\rm SU}(3)_{\rm C+F}\rightarrow [{\rm SU}(2)\times U(1)]_{\rm C+F}$, and thus there appear the associated Nambu-Goldstone (NG) modes which are identified as ${\mathbb C}P^2$ orientation modes \cite{Nakano:2007dr,Eto:2009bh}; \beq \frac{{\rm SU}(3)_{\rm C+F}}{[{\rm SU}(2)\times U(1)]_{\rm C+F}} \simeq {\mathbb C}P^2. \eeq Since bulk side far from vortex core is approaching to the uniform CFL phase, the ${\mathbb C}P^2$ modes are normalizable, i.e., well localized along the vortex core. A different orientation of the ${\mathbb C}P^2$ modes corresponds to a different color of magnetic fluxes. Since the ${\mathbb C}P^2$ modes are transformed under $SU(3)_{\rm C+F}$ rotation as the fundamental representation, there are three eigenstates which we denote by a three-component complex variable $\phi$ as homogeneous coordinates of the ${\mathbb C}P^2$ space, where the $\phi$ satisfies a constraint $\phi^\dagger\phi=1$, and its overall phase factor is redundant by definition. Since the ${\mathbb C}P^2$ modes reside along the vortex line which we put along the $z$-axis, its low-energy effective theory on the vortex world sheet is obtained if the ${\mathbb C}P^2$ modes are promoted to $1+1$ dimensional fields $\phi(z,t)$ through the moduli space approximation \cite{Manton:1981mp}. It results in a nonlinear ${\mathbb C}P^2$ sigma model \cite{Eto:2009bh,Eto:2009tr}, \beq {\mathcal L}_{{\mathbb C}P^2}=\sum_{\mu=0,3} K_\mu \ldk \partial^\mu \phi^\dagger \partial_\mu \phi +\lk \phi^\dagger \partial^\mu \phi \rk \lk \phi^\dagger \partial_\mu \phi \rk \rdk \label{eq:CPN} \eeq where $K_{0,3}$ are called the K\"ahler class given by the integration in terms of the profile functions $f$ and $g$. However, the above ${\mathbb C}P^2$ model in 1+1 dimensions is gapped through quantum corrections \cite{Gorsky:2011hd,Eto:2011mk}, because of the Coleman-Mermin-Wagner theorem \cite{Coleman:1973ci,PhysRevLett.17.1133} forbidding the presence of NG modes or a long range order in 1+1 dimensions. In this Letter, we show that this is only the case with a single vortex. When there is a vortex lattice consisting of a huge number of vortices, as expected in rotating quark matter such as in the core of neutron stars, there exist ${\mathbb C}P^2$ NG modes as gapless excitations in 3+1 dimensions whose dispersion relation is linear and anisotropic. This is due to interactions between vortices depending on ${\mathbb C}P^2$ orientations \cite{Auzzi:2007wj}. We discuss transition temperature between ordered and disordered phases and find that non-Abelian vortex matter form a color ferromagnet. Apart from dense quark matter, non-Abelian vortices with ${\mathbb C}P^{N-1}$ modes were found in supersymmetric $U(N)$ gauge theories with $N$ flavors \cite{Hanany:2003hp,Auzzi:2003fs} (see \cite{Eto:2006pg,Shifman:2007ce,2009supersymmetric} as a review). The low-energy effective theory of a vortex is the ${\mathbb C}P^{N-1}$ model which is in the same form with Eq.~(\ref{eq:CPN}) if we interpret $\phi$ as a complex $N$ vector. Supersymmetric theories usually admit Bogomol'nyi-Prasado-Sommerfield (BPS) vortices among which no static interaction exist. In this case, multiple vortices stably exist at arbitrary separations \cite{Eto:2005yh,Eto:2006cx}. On the other hand, here we are interested in non-BPS vortices among which static force exist \cite{Auzzi:2007wj}. In the following, we consider general $N$. \section{Non-Abelian vortex lattice} We discuss a color ferromagnetism of ${\mathbb C}P^{N-1}$ on a vortex lattice system. In rotating systems of CFL quark matter, multiple non-Abelian vortices are created and may form a vortex lattice. In general, structure of the vortex lattice is determined by vortex-vortex interaction. In the present case of non-Abelian vortices, the interaction is mediated by massless $U(1)_{\rm B}$ phonons \cite{Nakano:2007dr,Nakano:2008dc}, and also by scalars and gluons \cite{Hirono:2010gq} which are massive, where the scalars are amplitude fluctuations of the gap function $\Phi$. If the vortex lattice is dilute, i.e., the inter vortex distance is large, the vortex-vortex interaction is dominated by the $U(1)_{\rm B}$ phonons which provide a long-range repulsion in the same manner of usual superfluid vortices. Thus the structure of the dilute vortex lattice must be the Abrikosov's triangular lattice \cite{Nakano:2007dr,Nakano:2008dc}. We first define the system of our interest more precisely: vortex lattice has the triangle structure of spacing $L$ in the $x$-$y$ plane, and these vortices are set parallel to each other along the $z$ direction. Once locations of vortices are fixed on the triangular lattice sites by the repulsive force, relative orientation of the ${\mathbb C}P^{N-1}$ modes between neighboring vortices are determined solely by the interaction mediated by massive scalars and gluons. The ${\mathbb C}P^{N-1}$ dependent vortex-vortex interaction has already been derived in the previous study \cite{Eto:2013hoa}, and $V_{\rm int}^{\left<i,j\right>}$, the interaction energy per unit length in the $z$-direction, is of the form \beq V_{\rm int}^{\left<i,j\right>} /\Delta^2 = \sum_A G_{\left<i,j\right>,A} \lk \phi_i^\dagger T_A \phi_i\rk \lk \phi_j^\dagger T_A \phi_j\rk, \label{cp2int1} \eeq where $\left<i,j\right>$ represents sites of two adjacent vortices, $G_{\left<i,j\right>,A}$ is the interaction strength between them, and $T_{A}$, $A=1,\cdots,N^2-1$, are generators of $SU(N)$. The above expression is the leading order result under assumption that well separated two vortices are located parallel and their relative ${\mathbb C}P^{N-1}$ orientations are constant along vortex line. $G_{\left<i,j\right>,A}$ includes contributions from gluon-exchange and scalar-exchange potentials. The $U(1)_{\rm B}$ phonon does not feel the ${\mathbb C}P^{N-1}$ orientation, thus gives no contribution. In the context of quark matter, the gluon exchange part of the above ${\mathbb C}P^{N-1}$ dependent interaction was devised \cite{Eto:2013hoa} from the dual transformation which was used \cite{Hirono:2010gq} to describe a topological interaction between non-Abelian vortex and quasiparticles such as gluons and phonon in the bulk space, while, at present, we do not have a such systematic derivation for coupling between the scalar fields and non-Abelian vortex. Nevertheless, a seminal work \cite{Auzzi:2007wj} on numerical simulation on the potential energy between two non-BPS non-Abelian vortices provides implication of an asymptotic form of full $G_{\left<i,j\right>,A}$ including both contributions. With inter vortex distance $r$ being defined and isotropy in the color space $A$, the asymptotic form of $G(r)=G_{\left<i,j\right>,A}$ would be given by \cite{Auzzi:2007wj} \beq G(r)\simeq -C_1 B_0(m_{\chi} r)+C_2 B_0(m_g r) \label{af1} \eeq where $C_1$ and $C_2$ are positive numerical factors, and $B_0(r)$ the modified Bessel function, $B_0(mr)\sim \sqrt{\pi/2mr}e^{-mr}$ for $mr\gg 1$. In the above equation, the first term corresponds to massive scalar contributions and the second to massive gluon contributions, and it should be noted that the massive gluons and scalars give the opposite sign of interaction. For sufficiently large lattice spacing $L$, contribution of the lightest particle dominates the interaction. The ${\mathbb C}P^{N-1}$ modes live in each vortex site, and their dynamics is described by the following effective Hamiltonian, \beq && H = \int dz\: \sum_{\left<i,j\right>,A} \Big[ -J_{xy} S_{i,A} S_{j,A} \nn && \hspace{1cm} + K_3 \{ |\partial_z \phi_i|^2 + (\phi_i^\dagger \partial_z \phi_i)^2 \} \Big] \label{eq:cpn-magnet} \\ && \quad S_{i,A}:= \phi_i^\dagger T_A \phi_i, \quad J_{xy} := \Delta^2 G(L) \eeq where $S_{i,A}$ are ``spin" variables and $J_{xy}$ is the interaction between neighboring vortices. If $\phi$ is uniform in the $z$-direction, this Hamiltonian reduces to the Heisenberg model for $N=2$. In this case, $J_{xy}>0$ and $J_{xy}<0$ correspond to ferromagnets and anti-ferromagnets, respectively. We use these terminologies for general $N$. The model with $N \geq 3$ further provides higher-dimensional nontrivial spin system as a new statistical model. The case of $N=3$ corresponds to color magnetism in rotating dense quark matter discussed here. First, the K\"{a}hler class $K_{0,3}$ is estimated as $K_0 \simeq 3K_3 \simeq 3\mu^2/\Delta^2$ at very high density \cite{Eto:2009tr}. The interaction $J_{xy}$ is determined through the masses of particles which vortices exchange. At very high densities, the masses are given as $m_s=2m_\chi \sim 2\Delta , \; m_g \sim g \mu$, where $m_{s,\chi}$ corresponds to the singlet (adjoint) part of scalar fields in the residual $SU(3)_{\rm C+F}$ symmetry, $\mu$ is the baryon number chemical potential, and $g$ is coupling constant of strong interaction. At very high densities, i.e., very large $\mu$, $g$ and $\Delta$ are exponentially small with respect to $\mu$, therefore, one expects $m_g\gg m_s, m_\chi$. Thus, \beq G(L)\simeq -C_1 \sqrt{\frac{\pi}{2m_\chi L}}e^{-m_\chi L}. \eeq Therefore, in the case of rotating CFL quark matter, because of the sign of the interaction above, the nearest neighbor interaction of ${\mathbb C}P^{N-1}$ modes is found to align their orientations (color fluxes), leading to a color ferromagnetism. An estimation of $J_{xy}$ is given as follows. Supposing the present system is inside of a neutron star of the radius $R$ with rotation period $P_{\rm rot}$, we can valuate a ballpark of the number of vortex as $N_v\simeq 1.9 \times 10^{19}\lk\frac{1{\rm ms}}{P_{{\rm rot}}}\rk \lk\frac{\mu/3}{300{\rm MeV}}\rk \lk\frac{R}{10{\rm km}}\rk$, from which the lattice spacing is estimated \cite{Hirono:2012ki} as $L\equiv \lk\frac{\pi R^2}{N_{v}}\rk^{1/2} \simeq 4.0\times 10^{-6}{\rm m} \lk\frac{P_{{\rm rot}}}{1{\rm ms}}\rk^{1/2}\lk\frac{300{\rm MeV}}{\mu/3}\rk^{1/2}$. A set of values of $C_1\sim 10$, $L\simeq 4.0\times 10^{-6}$m, and the gap energy $\Delta\sim 100$ MeV for low temperatures ~\cite{Alford:1998mk,Hessels:2006ze} gives $\log\lk J_{xy}/\Delta^2\rk \sim -10^9$, which is quite small in comparison to the coupling in the $z$ direction. To have $J_{xy}/\Delta^2=O(1)$, we need to put the system very close to the critical point $(T_{\mathrm{c}}^{\mathrm{CFL}} - T)/T_{\mathrm{c}}^{\mathrm{CFL}} \sim 10^{-18}$, where we have used a mean-field result $\Delta \simeq 2.16 T_{\mathrm{c}}^{\mathrm{CFL}} \sqrt{1-T/T_{\mathrm{c}}^{\mathrm{CFL}}}$~\cite{Iida:2003cc}, or we have to set an asymptotically high density. Here let us point out that the smallness of the effective coupling $J_{xy}$ in the above estimation is based upon multiple approximations, i.e., the moduli space approximation and an asymptotic form of the interaction (\ref{af1}), both of which work against correlations among orientions in the $x$-$y$ plane. What we claim here is that the present study provides the qualitative argument on how orientational modes make order of color magnetism, which is common feature for general non-BPS vortex lattice systems \cite{Auzzi:2007wj}. So actual correlations amog orientations in the $x$-$y$ plane are more measurable, at least, for higher vortex densities. \section{Color magnons in color ferromagnets} Low-energy excitations in the ordered phase have an important role in determination of, e.g., thermodynamic and transport properties of the system. The present system exhibits a color ferromagnetism, thus we expect that there appear massless excitations like magnons in the ordinary ferromagnetism, which correspond to fluctuations of the ${\mathbb C}P^{N-1}$ modes around an orientation spontaneously set in color ferromagnetic phase. To investigate dynamics of these excitations, we take a long wavelength limit of the interaction between neighboring ${\mathbb C}P^{N-1}$ modes, Eq.~(\ref{cp2int1}), which can be described by a continuum limit of the interaction. By use of Fierz transformation, $\sum_{A} \lk T_A\rk^a_b \lk T_A\rk^c_d$ $=$$\delta^a_d\delta^c_b/2-\delta^a_b\delta^c_d/2N$ ($a,b,c,d=1,\cdots,N$), we can take a continuum limit of each bond of the triangle lattice in Eq.~(\ref{eq:cpn-magnet}), up to irrelevant constants, \beq &&\sum_{A} (\phi_{i}^\dagger T_A \phi_i) \ (\phi_{j}^\dagger T_A \phi_j) \nn \rightarrow && - \frac{L^2}{2}\ldk |\nabla\phi|^2+\lk\phi^\dagger\nabla\phi\rk^2 +{\mathcal O}\lk \nabla^4\rk \label{eq:CPN-lattice} \rdk, \eeq for a pair of the nearest neighbors $i,j$. Then, we sum all bonds to obtain \beq && -\sum_{\left<i,j\right>,A} J_{xy}\lk \phi_i^\dagger T_A \phi_i\rk \lk \phi_j^\dagger T_A \phi_j\rk \nn &\rightarrow& \tilde{K}_{xy} \int {\rm d}x{\rm d}y \sum_{i=x,y} \ldk |\partial_i \phi|^2 +\lk \phi^\dagger \partial_i \phi \rk^2 \rdk, \quad \eeq with a coupling constant being defined by $\tilde{K}_{xy}=J_{xy}(3L^2/2)(2/L^2\sqrt{3})(1/2)$$=$$\sqrt{3}J_{xy}/2$. Incorporating the effective theory in the $z$ and temporal directions with redefined couplings $\tilde{K}_{0,3}$$=$$2K_{0,3}/L^2\sqrt{3}$, where $2/L^2\sqrt{3}$ is the vortex density in the $x$-$y$ plane, we obtain a low-energy effective theory for the magnon modes in 3+1 dimension, $L_{\rm eff}= \int {\rm d}x{\rm d}y{\rm d}z{\mathcal L}_{\rm eff}$ with \beq {\mathcal L}_{\rm eff} = \sum_{\mu=0}^3 \tilde{K}_\mu \ldk \partial^\mu \phi^\dagger \partial_\mu \phi +\lk \phi^\dagger \partial^\mu \phi \rk \lk \phi^\dagger \partial_\mu \phi \rk \rdk \label{eq:spin-Lagrangian} \eeq which is a ${\mathbb C}P^{N-1}$ model with spatially anisotropic couplings; $\tilde{K}_1=\tilde{K}_2=\tilde{K}_{xy}\neq \tilde{K}_{3}$. Now let us study the low-energy excitations of this model. To this end, we can set an orientation of the ground state as $\phi_0=\lk 1,0,\cdots,0 \rk$ without loss of generality. The color magnon corresponds to fluctuations around the orientation, $\phi=U(\delta_a)\phi_0=\phi_0+\tilde{\phi}+O(\tilde{\phi}^2)$, where $U(\delta_a)\in SU(N)_{\rm C+F}$ and $\tilde{\phi}:=i X_a \delta_a \phi_0=\lk i\delta_0, -\delta_1+i \delta_2, \cdots, -\delta_{2N-3}+i \delta_{2(N-1)} \rk^T$. Parameters $\delta_a$ are assigned to broken generators $X_a$ of the dimension of $SU(N)/SU(N-1)$, and are identified as the color magnons. The fluctuation $\delta_0$ is eliminated by overall $U(1)$ operation, thus will be redundant. Plugging these expression into the effective Lagrangian, we have dynamics of the color magnons as \beq L_{\rm eff}&=& \sum_{\mu=0}^3 \tilde{K}_\mu {\rm Tr} \ldk (1-\phi_0 \phi_0^\dagger) \partial^\mu \tilde{\phi} \partial_\mu \tilde{\phi}^\dagger \rdk +O(\tilde{\phi}^3), \nn &=& \sum_{\mu=0}^3 \sum_{a=1}^{2(N-1)} \tilde{K}_\mu \partial^\mu \delta_a \partial_\mu \delta_a +O(\delta_a^3). \eeq Then, the dispersion relation can be obtained as \beq \omega_p^2 = (\tilde{K}_{xy}(p_x^2+p_y^2)+\tilde{K}_zp_z^2)/\tilde{K}_0, \eeq which shows that the color magnons propagate in the $x$, $y$ directions with velocity $c_{xy}^2=\tilde{K}_{xy}/\tilde{K}_0$ and the $z$-direction with $c_{z}^2=\tilde{K}_{3}/\tilde{K}_0$. In the case of $N=3$ for the CFL quark matter, $c_{z}^2=K_{3}/K_0\simeq1/3$, and $c_{xy}^2=\Delta^4G(L)L^2\sqrt{3}/2\mu^2 \simeq a_2 t^{7/4} e^{-a_1\sqrt{t}}$ where $t=(T_{\mathrm{c}}^{\mathrm{CFL}}-T)/T_{\mathrm{c}}^{\mathrm{CFL}}$, which is obtained from a mean-field result at very high densities. This formula gives the maximal value of $c_{xy}$ with respect to temperature, which is achieved at $t=(7/2a_1)^2$. For typical values of parameters, $a_1 \sim 10^{9}$, and $a_2\sim 10^{12.5}$, $c_{xy}\sim 10^{-9} (=0.3 {\rm m/s})$ at most, giving 9 hours oscillation for a neutron star with $R=10$km. The color magnons are massless excitations propagating with a linear dispersion (of type I) independently from the other NG modes, e.g., Kelvin/Tkachenko modes associated with the breaking of translational invariance, which have a quadratic dispersion (of type II). For a single non-Abelian vortex system, where the effective theory is given by the 1+1 dimensional ${\mathbb C}P^{N-1}$ model, there is no ordered phase because of strong quantum fluctuation effects. For the present vortex lattice system, however, an ordered phase can be expected because the fluctuations are suppressed by propagating along the $xy$-directions in the long-wavelength limit, and the mean-field analysis becomes better in such a 3+1 dimensional system compared to lower dimensional systems. Within the mean-field analysis, the critical temperature $T_{\mathrm{c}}^{\mathrm{order}}$, below which the vortex lattice system is ordered, can be estimated as (see Appendix A for derivation) \begin{align} T_{\mathrm{c}}^{\mathrm{order}} \sim \frac{J_{xy}}{k_{\mathrm{max}}} + K_3 k_{\mathrm{max}}, \end{align} where the first and second terms correspond to the mean-field critical temperature along the $xy$ and $z$-directions, respectively, and $k_{\mathrm{max}}$ is the maximum wavenumber taken into account along the $z$-direction. In the continuum limit $k_{\mathrm{max}} \to \infty$ for the $z$-direction, the transition temperature diverges, which implies the ordered phase and the existence of the corresponding color magnon excitations in the whole temperature region. However, the divergent temperature is a mean-field artifact, and the disordered phase may be expected at the finite temperature within the effect beyond mean-field analysis. \section{Color anti-ferromagnets} Although the color magnetism with $J_{xy}>0$ is only the physical system with $N\geq 3$ known thus far, it may be worth to mention basic properties of color anti-ferromagnets with $J_{xy}<0$. We denote the order parameter space (OPS) for anti-ferromagnets by ${\cal M}^{N=2}_{\rm AF}$. We then find (see Appendix B for derivation) \beq && {\cal M}^{N=2}_{\rm AF} \simeq SU(2)/{\mathbb Z}_2 \simeq SO(3)\\ && {\cal M}^{N=3}_{\rm AF} \simeq SU(3)/U(1)^2 \eeq The OPS for $N=2$, $SO(3)$, is well-known for Heisenberg anti-ferromagnets \cite{Kawamura:1984,Okubo:2010} while the second is a so-called flag manifold of rank two. The first and second homotopy groups of the OPS are important for the presence or absence of a Kosterlitz-Thouless (KT) phase transition \cite{Kawamura:1984}. They are \beq \pi_1({\cal M}^{N=2}_{\rm AF}) \simeq {\mathbb Z}_2 , && \pi_2({\cal M}^{N=2}_{\rm AF}) \simeq 0 \\ \pi_1( {\cal M}^{N=3}_{\rm AF}) \simeq 0 , && \pi_2( {\cal M}^{N=3}_{\rm AF}) \simeq {\mathbb Z} \oplus {\mathbb Z} . \eeq The first homotopy groups show that the KT transition can happen only for $N=2$. The second homotopy groups imply the existence of two-dimensional Skyrmions (lumps) extended as strings for $N = 3$. \section{Summary and Discussion} As a generalization of the ordinary magnetism, we have found color magnets in lattices consisting of non-Abelian vortices with ${\mathbb C}P^{N-1}$ internal orientations. While the case of $N=2$ reduces to Heisenberg (anti-)ferromagnets, rotating dense quark matter provides an example of color ferromagnets in the $N=3$ case. \footnote{ In response to an external magnetic field other vortex solution can be found, where gluons are condensed due to a chromomagnetic instability \cite{Ferrer:2006ie,Ferrer:2007uw}. Since the non-Abelian vortex we have discussed here generates a sponteneous magnetic field if the electromagnetism is introduced \cite{Vinci:2012mc}, investigation of their interplay is of interest as a future work. } For color ferromagnets, we have found color magnons as gapless NG modes propagating in 3+1 dimensions with the dispersion relation anisotropic in the direction along the rotating axis and the orthogonal plane, and have estimated the transition temperature in the mean-field approximation. For color anti-ferromagnets, we have found the OPS and relevant homotopy groups. When we take into account the electromagnetic interactions $U(1)_{\rm EM}$, the ${\mathbb C}P^2$ modes of a non-Abelian vortex are electrically charged and the interaction between a vortex and the electromagnetic field is described by a $U(1)_{\rm EM}$ gauged ${\mathbb C}P^2$ model \cite{Hirono:2012ki}. Hence, the ${\mathbb C}P^2$ modes affect the electric conductivity in addition to the thermal conductivity. There should be anisotropy on these conductivities, which may affect evolution of neutron stars such as the cooling process of the star, etc. As for electric conductivity, it was predicted that electro-magnetic waves along a vortex lattice decay and therefore the CFL phase acts as a polarizer \cite{Hirono:2012ki}. Another effect of the electromagnetic interactions is a mixing between $U(1)_{\rm EM}$ and the $\lambda_8$ component of strong interaction. This induces a tension difference among the ${\mathbb C}P^2$ degenerated vortices \cite{Vinci:2012mc}. This effect can be incorporated by an effective potential in the ${\mathbb C}P^2$ model. A lattice formulation of the ${\mathbb C}P^{N-1}$ model was proposed by using a $U(1)$ gauge field on link variables \cite{Campostrini:1992ar,Campostrini:1992it}. Our model in Eq.~(\ref{eq:CPN-lattice}) provides a simpler lattice formulation without gauge fields. \section*{Acknowledgements} This work is supported in part by Grant-in-Aid for Scientific Research (Grant No. 22740219 (M.K.),No. 24740166 (E.N.), and No. 25400268 (M.N.)) and the work of M.N. is also supported in part by the `Topological Quantum Phenomena'' Grant-in-Aid for Scientific Research on Innovative Areas (No. 25103720) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan.
\section{MZ Interferometry sequence} A standard MZ interferometer sequence is constructed as follows, each part of which is labeled with roman numerals corresponding to its depiction in Fig~\ref{Config}~(b). First, a cold cloud of Rubidium-87 atoms with a narrow momentum width is loaded into a horizontal optical waveguide, which effectively confines them to move freely in only the longitudinal direction ($x$ in Fig.~\ref{Config}). \textbf{i.} An optical standing wave formed from two counter-propagating beams each of wavevector $\pm \mathbf{k}$ is then used as a matter-wave diffraction grating, beam-splitting the atoms along two trajectories which are separated in momenta by the Bragg condition, $\Delta \mathbf{p}_x=2n\hbar \mathbf{k}$, where $n$ is the diffraction order. \textbf{iv.} After a time $T$, another Bragg-diffraction pulse is applied which diffracts the stationary state to $2n\hbar \mathbf{k}$ at the same time as diffracting the $2n\hbar \mathbf{k}$ state to be stationary. This effectively swaps the momenta of each state, so it is the atom-optical equivalent of a mirror. \textbf{vi.}~When a total time $2T$ has elapsed and both trajectories are overlapping once again, a final Bragg-diffraction beam-splitter pulse is applied which recombines the two momentum states. This final optical standing wave can have an arbitrary phase offset $\phi_\frac{\pi}{2}$ from the initial optical standing wave \textbf{i}, which amounts to an $x$-displacement of the peaks and troughs in optical intensity. \begin{figure*} \begin{center} \includegraphics[width=2\columnwidth]{GMcDonaldFig2.eps} \caption{(a) A BEC is formed in a cross beam dipole trap (red), before it is released into the horizontal waveguide. The interferometer is formed by optical pulses from the Bragg/Bloch beam (blue). Inset: The potential (not to scale) experienced by the atoms when both the waveguide and the optical lattice are on. (b) Top: The space-time trajectory of each arm of the interferometer (total time $2T$) is shown for both a $10{\hbar}\mathbf{k}$ MZ configuration (purple dashed lines), and the CAB sequence (green solid lines) for the same time $T$. The vertical axis represents position along the horizontal waveguide in the experiment. Bottom: Our pulse sequence for the CAB interferometer. $10{\hbar}\mathbf{k}$ Bragg diffraction pulses (purple gaussians) are the beam splitter ($\frac{\pi}{2}$, outside) and mirror ($\pi$, centre) pulses. $2n_{b}{\hbar}\mathbf{k}$ Bloch lattice accelerations present in the CAB sequence (green trapezoids) act only upon one arm of the interferometer at a time. (c) The relative population in one of the output states of an MZ with $T=0.2$ms is measured as the phase of the final beamsplitter pulse $\phi_\frac{\pi}{2}$ is scanned. Each realisation of the experiment is shown as a red circle, and a sinusoidal fit to the data of the form $N_{rel}=\frac{\mathcal{V}}{2}\cos(\Phi+n\phi_{\frac{\pi}{2}})+c$ (shown in blue) is used to extract the interferometric phase $\Phi$, which is the acceleration signal.} \label{Config} \end{center} \end{figure*} \section{CAB interferometry sequence} Our CAB interferometer sequence is built upon this standard MZ configuration. \textbf{i.} After the first Bragg-diffraction beam-splitter there are two momentum states in the waveguide, one stationary and one with momentum $\mathbf{p}_x=2n\hbar \mathbf{k}$ in the positive-$x$ direction. \textbf{ii.} Over a time $T_{r}$ we selectively load the stationary momentum state into a stationary optical lattice with a potential energy depth of $15E_{r}$, where the photon recoil energy is given by $E_{r}=({\hbar}k)^2/2m$. We accelerate the lattice in the negative-$x$ direction via $n_b$ Bloch oscillations \citep{PeikBloch} over a time $T_{b}$, each of which imparts $2\hbar \mathbf{k}$ momentum. This amounts to a constant acceleration rate of $\Delta \mathbf{a}_{b}=\frac{2n_{b}\hbar \mathbf{k}}{mT_{b}}$ applied to the state loaded into the lattice. The state selectivity is accomplished because the optical lattice is moving fast enough with respect to the $\mathbf{p}=2n\hbar \mathbf{k}$ momentum state that this state experiences just the time-averaged lattice potential, which imparts no acceleration. \textbf{iii.} After a negligible time $T_{f}$ we decelerate the state until it is stationary again. \textbf{iv.} The Bragg-diffraction mirror pulse is applied a time $T$ after the beginning of the interferometer, which swaps the momenta of each state. \textbf{v.} The Bloch lattice acceleration and deceleration sequence is now applied to the other arm of the interferometer. \textbf{vi.} The two momentum states are recombined by a final Bragg-diffraction beam-splitter. It should be noted that the CAB sequence is actually a combination of both configuration (b) and (c) depicted in Figure 1. To obtain an interferometric fringe like that shown in Fig.~\ref{Config}~(c), we allow the two final momentum states to separate before taking an absorption image to count the number of atoms $N_i$ in each final state $i$. We calculate $N_{rel}=\frac{N_1}{N_1+N_2}$ to remove the effect of run-to-run fluctuations in total atom number of $\approx15\%$. As the phase $\phi_\frac{\pi}{2}$ of the final beam-splitter is scanned, a fringe will be obtained of the form $N_{rel}=\frac{\mathcal{V}}{2}\cos(\Phi+n\phi_{\frac{\pi}{2}})+c$ as shown in Fig.~\ref{Config}~(c). The phase offset $\Phi$ of this fringe is our acceleration signal. The acceleration dependence of this phase offset is given by the space-time area $\mathcal{A}_{\text{CAB}}=\int \Delta \mathbf{x}\, dt$ of our interferometer, \begin{align} \Phi_{CAB}&=\frac{m}{\hbar}\mathbf{a}\cdot\mathbf{\mathcal{A}}_{CAB}\nonumber\\ &=2\left(n+n_{b}\cdot \frac{T_{b}+T_{f}}{T}\right)\mathbf{k\cdot a}T^2 \label{eq:phase} \end{align} which reduces to the phase offset of an MZ configuration $\Phi_{MZ}=2n\mathbf{k\cdot a}T^2$ when $n_{b}=0$. In order to obtain the $T^3$ scaling, we must look at what happens when the interferometer time $T$ is increased, keeping constant the relative acceleration $\Delta \mathbf{a}_{b}$ which we apply via the Bloch lattice. In this case $n_{b}=T_{b}/\tau$, where the constant $\tau=\frac{2{\hbar}k}{m|\Delta \mathbf{a}_{b}|}$ is the period for one Bloch oscillation. Assuming $T_{r}$ and $T_{f}$ are much smaller than $T_{b}$, then we have $T_{b}\rightarrow\frac{T}{2}$, and Eq.~(\ref{eq:phase}) becomes \begin{align} \label{eq:T3phase} \Phi_{CAB}&=2n\mathbf{k\cdot a}T^2+\frac{\mathbf{k\cdot a}T^3}{2\tau}\\ &=\Phi_{MZ}+\frac{\mathbf{k\cdot a}T^3}{2\tau}\nonumber \end{align} which explicitly shows the extra $T^3$ scaling in acceleration sensitivity achievable with this configuration. We experimentally test Eq.~(\ref{eq:phase}) by first measuring $\Phi_{MZ}$ for a standard MZ configuration with $n=5$, $n_{b}=0$ (blue triangles on Fig.~\ref{MoneyShot}) in order to extract the external acceleration $\mathbf{a}$ due to a tilt in the waveguide. From this we calculate via Eq.~\ref{eq:phase} what $\Phi_{CAB}$ will be for the CAB sequence (red dashed line). We then measure $\Phi_{CAB}$ for $T=0.642$ms, $n_{b}=1$ through to $T=1.042$ms, $n_{b}=21$ (red circles) and our experimental parameters were set such that $T=2n_{b}\tau+0.622$ms with a Bloch oscillation period of $\tau=10\mu$s. The excellent agreement shown in Fig.~\ref{MoneyShot} between the predicted phase for the CAB sequence which was deduced from the MZ measurements, and the measurements of $\Phi_{CAB}$ validate Equation (\ref{eq:phase}) and demonstrate a $T^3$ scaling in phase sensitivity to acceleration. \begin{figure}[!htp] \centering{} \includegraphics[width=.9\columnwidth]{GMcDonaldFig3Revised.eps} \caption{Here we experimentally demonstration the faster sensitivity scaling. The interferometric phase offset $\Phi$ is measured for each fringe as the interferometer time $T$ is increased. The blue solid line is a fit to the MZ phase $\Phi_{MZ}$ (blue triangles) of the form $\Phi_{MZ}=2n\mathbf{k\cdot a}T^2$ to extract the acceleration $a$ along the waveguide (due to a slight tilt) for the day the data was taken. This acceleration is then used to predict $\Phi_{CAB}$ for the CAB sequence by equation (\ref{eq:phase}) (red dashed line). Experimental measurement of $\Phi_{CAB}$ for the CAB sequence (red circles) increases faster than $T^2$, as predicted by Eq. (\ref{eq:phase}). Uncertainties in $\Phi$ extracted from each fringe are one standard deviation confidence intervals.} \label{MoneyShot} \end{figure} We now turn to yet higher scalings with respect to $T$. The maximum adiabatic acceleration rate $\Delta \mathbf{a}_{b}$ using a Bloch lattice increases quadratically with lattice depth \citep{PeikBloch}, and therefore also increases quadratically with available laser power. This means that a Bloch-based constant acceleration configuration can achieve a larger space-time area for constant $T$ and a given laser power than an equivalent sequential-Bragg configuration \citep{Kasevich102hk}, in which the momentum transferrable in each Bragg diffraction pulse increases as the square root of the available laser power \citep{SzigetiBragg}. However, in the CAB scheme lattice depth and acceleration rate are limited by the instantaneous velocity separation of the two clouds \citep{80hkPRA}. This is because if the lattice is too deep it will also bind the other momentum state, and thus our acceleration will no longer be state selective. Therefore, the best possible use of available laser power would require the application of a constantly increasing relative acceleration $\Delta\frac{d\mathbf{a}_{b}}{dt}$ (known as constant jerk) between the two states, while commensurately increasing the optical lattice depth. This would give a a space-time area of $\mathcal{A}_{T^4}=\frac{1}{24}\Delta\frac{d\mathbf{a}_{b}}{dt} T^4$, producing an interferometer with phase $\Phi=\Phi_{MZ}+\frac{m}{\hbar}\mathbf{a}\cdot\mathcal{A}_{T^4}$. At the point at which lattice depth is limited by available laser power, the maximum acceleration rate will become constant again, and so the scaling will revert to $T^3$. Generalising to arbitrary scalings in $T$ would be possible if an interferometer were developed with a constant $n$-th derivative of displacement (for $n\geq 1$). It would have a space time area of \begin{align} \label{eq:generalisation} \mathcal{A}_{T^{n+1}}&=\frac{1}{n!\,2^{n-1}}\left[ \left(\frac{d}{dt}\right)^{n} \Delta \mathbf{x}\right]T^{n+1} \end{align} and therefore a $T^{n+1}$ scaling in acceleration sensitivity. Although there is no practical advantage beyond constant jerk in the present system, it is possible that an analogous scheme will be developed in the future which is not limited by laser power, and could practically take advantage of these higher scalings. For instance, one can envisage positionally dependent trapping potentials (e.g. dipole traps) accelerating each cloud in opposite directions once they are separated in space. The choice of Bragg~\cite{80hkPRA} and not Raman~\cite{CladeLMT} beamsplitters in this configuration allows both momentum states to be in the same magnetic internal state throughout the interferometer, eliminating a systematic phase shift. The use of a BEC source cloud presents its own systematic phase shift, the density-dependent mean-field shift~\cite{KyleCoherence}. However, this shift can be reduced arbitrarily by lowering the cloud's density via the delta-kick-cooling process~\cite{80hkPRA}, or removed entirely by turning off mean-field interactions through the use of a Feshbach resonance~\cite{SoltionArxiv}. To separate the effects of acceleration due to gravity, the optical potential and the magnetic field gradients possibly present in the experiment, one could use a magnetically insensitive internal state~\cite{SelfWaveguide} or compare the phase shift across different isotopes simultaneously~\cite{DualSpecies}. In the future, optical-lattice intensity-noise reduction in this system is possible by using multiple overlaid Bloch lattices to address both momentum states separately, and accelerate them in opposite directions at the same time~\citep{HMullerBlochBraggBloch}. This will cancel the a.c. Stark shift due to Bloch laser intensity noise, as each arm simultaneously experiences the Bloch lattice. In the present CAB sequence the a.c. Stark shift is only cancelled at a later time in the interferometer, so with this improvement the interferometer will become less sensitive to such fluctuations. However, as each arm of the interferometer only experiences half the laser intensity, this technique will quarter the maximum acceleration rate of each state, due to it quadratic scaling with intensity \citep{PeikBloch}. This results in a halving of the relative acceleration rate $\Delta\mathbf{a}_{b}$, and hence a halving of the signal $\Phi$. Other noise reductions in order to enhance signal-to-noise can be achieved by reducing mechanical vibrations~\cite{80hkPRA}, evacuating the optical path and locking-out optical-lattice-laser frequency fluctuations~\cite{GbThesis}. In summary we have demonstrated a novel configuration for a cold-atom interferometer in which acceleration sensitivity scales as $T^3$. This CAB configuration is realised using an optical Bloch lattice to subject one arm of the interferometer at a time to an additional constant acceleration. The additional $T^3$ scaling in sensitivity to the external inertial acceleration $\mathbf{a}$ allows this CAB configuration to have increased sensitivity to accelerations measured with a given interferometer time $T$ than what is possible with a Mach-Zehnder configuration. This CAB configuration will therefore be useful in increasing the phase sensitivity to accelerations at any given frequency, without requiring any increase in available laser power. This technique can therefore be immediately applied to greatest effect in navigation and inertial sensors which are currently under development, and in proposed schemes for gravitational wave detection. \acknowledgments The authors would like to thank Hannah Keal for her assistance moving electronic equipment out of the lab, and Paul Altin for baking the vacuum system. We gratefully acknowledge the support of the Australian Research Council Discovery program. C.C.N.K. would like to acknowledge financial support from CNPq (Conselho Nacional de Desenvolvimento Cientifico e Tecnologico). J.E.D. would like to acknowledge financial support from the IC postdoctoral fellowship program.
\section[1]{\emph{#1}.---} \def \bs #1{\boldsymbol{#1}} \def\tocite#1{({\color{blue} {#1}})} \begin{document} \title{Can CMB Lensing Help Cosmic Shear Surveys? } \author{Sudeep Das} \affiliation{High Energy Physics Division, Argonne National Laboratory, 9700 S Cass Avenue, Lemont, IL 60439} \affiliation{Berkeley Center for Cosmological Physics, Berkeley, CA 94720} \author{Josquin Errard} \affiliation{Computational Cosmology Center, Lawrence Berkeley National Laboratory, Berkeley, CA 94720} \author{David Spergel} \affiliation{Peyton Hall, Ivy Lane, Princeton University, Princeton, NJ 08544} \date{\today} \begin{abstract}{Yes! Upcoming galaxy shear surveys have the potential to significantly improve our understanding of dark energy and neutrino mass {\it if} lensing systematics can be sufficiently controlled. The cross-correlations between the weak lensing shear, galaxy number counts from a galaxy redshift survey, and the CMB lensing convergence can be used to calibrate the shear multiplicative bias, one of the most challenging systematics in lensing surveys. These cross-correlations can significantly reduce the deleterious effects of the uncertainties in multiplicative bias. } \end{abstract} \maketitle \section{Introduction} The large scale structure in the universe gravitationally deflects the light from distant galaxies inducing weak coherent distortions of their images. This ``cosmic shear" signal \cite{hoekstra_jain_2008,refregier_2003}, depends sensitively on the expansion rate of the universe as well the growth of structure with time, making it a potentially rich probe of the nature of dark energy, the validity of general relativity (GR) on cosmological scales, and the neutrino mass sum~\cite{1987tasi.rept.....A,2006Msngr.125...48P,2010MNRAS.404..110D,1999ApJ...514L..65H,2002PhRvD..65f3001H,2013arXiv1301.1037D,2013arXiv1309.5383A}. In the next few years, we expect a flood of data relevant to cosmic shear studies from ongoing, upcoming and planned galaxy surveys, such as PanSTARRS~\footnote{http://pan-starrs.ifa.hawaii.edu/}, the Subaru HyperSuprimeCam (HSC) survey~\footnote{http://sumire.ipmu.jp/en/3358}, the Dark Energy Survey (DES)~\footnote{http://www.darkenergysurvey.org/survey/}, KIDS~\footnote{http://kids.strw.leidenuniv.nl}, LSST~\footnote{www.lsst.org}, Euclid~\footnote{http://www.euclid-ec.org/},~WFIRST~\footnote{http://wfirst.gsfc.nasa.gov}, BigBoss~\footnote{http://bigboss.lbl.gov/}, etc. The weak lensing signal from these surveys can in principle pin down cosmological parameters to unprecedented precision, but realizing their full potential imposes stringent requirements on the control of systematics~\cite{2008MNRAS.391..228A, 2006MNRAS.366..101H}. \par Weak lensing systematics mainly stem from four sources: the inability to accurately measure galaxy shapes due to instrumental and atmospheric effects, the uncertainty in the distance to the background galaxies (photometric redshift errors), intrinsic alignment of galaxies due to the galaxy formation process, and the theoretical uncertainties in dark matter clustering on small scales. Shape measurement errors fall mainly into two categories -- additive and multiplicative biases in the deduced shear signal, both of which can be redshift dependent in general. The multiplicative bias (which multiplies the true shear signal with an unknown multiplicative factor) is particularly notorious because being redshift dependent it can be degenerate with the growth of structure, significantly degrading cosmological parameter constraints and inducing large parameter biases~\cite{2008MNRAS.391..228A, 2006MNRAS.366..101H}. \par A multiplicative bias in the shear signal can in principle be calibrated by cross correlating the observed shear signal with a measurement of the projected dark matter field that does not suffer from such multiplicative uncertainty. The gravitational lensing of the cosmic microwave background (CMB) provides a promising route to such a solution, as we describe in this Letter. Recently, Vallinotto (2012)~\cite{2012ApJ...759...32V} has proposed using the cross-correlation of CMB lensing with cosmic shear as a method to control the multiplicative bias. It is an useful method for controlling a redshift dependent multiplicative bias, as long as the growth function is assumed to be standard. However, a non-standard growth function can be hidden by a redshift dependent multiplicative bias even when using this cross-correlation technique. We propose that by additionally using cross-correlations with spectroscopic galaxy surveys such degeneracy can be further broken, significantly improving the constraints on dark energy, deviation from GR, and neutrino mass. \\ \section{Controlling Shear Multiplicative Bias with Cross-correlations} We begin by writing the weak lensing convergence as the projected matter overdensity field $\delta$: \begin{equation} \kappa(\hat{\mathbf{n}})=\frac32 \ensuremath{\Omega_{\mathrm{m}}} H_{0}^{2}\int d\eta ~d_{A}^{2} (\eta) \frac{g(\eta)}{a(\eta)}\delta(d_A(\eta) \hat{\mathbf{n}},\eta), \eeq where the kernel \beq g(\eta)=\frac{1}{d_A(\eta)}\int_{\eta}^{\infty}d\eta' ~W_b(\eta') \frac{d_A(\eta'-\eta)}{d_A(\eta') } \eeq depends on the normalized source distribution in comoving distance $\eta$: $W_b(\eta)$. Here $d_A(\eta)$ is the comoving angular diameter distance, $a(\eta)$ is the scale factor, while $\ensuremath{\Omega_{\mathrm{m}}}$ and $H_{0}$ represent the present values of the matter density parameter and the Hubble parameter, respectively. For the CMB we can approximate the source distribution function as a screen at the last scattering distance $\eta_0$: $W_{b}(\eta) \simeq \delta_D(\eta - \eta_{0})$. \par Next we consider a tracer population with a known redshift distribution $W_f(\eta)$: \beq \Sigma(\hat{\mathbf{n}})=\int d\eta ~W_f(\eta) \delta_g(\eta \hat{\mathbf{n}},\eta), \eeq where $\delta_g$ represents the fractional tracer overdensity that is related to the underlying matter density field via a scale and redshift dependent bias factor: $\delta_{g}(\mathbf k,\eta) = b(k,\eta) \delta(\mathbf k,\eta)$. CMB lensing estimators lets us reconstruct the convergence field, $\kappa_{CMB}$ between us and the last scattering surface. The observed CMB lensing-tracer cross-correlation depends on the bias of the galaxy distribution, $b_\ell(\eta)$ and the underlying cosmology, \beq C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\Sigma}=\frac{3}{2}{\ensuremath{\Omega_{\mathrm{m}}} H_{0}^{2}} \int d\eta ~b_\ell(\eta) W_f(\eta) \frac{\ensuremath{g_{\mathrm{\scriptscriptstyle{CMB}}}}(\eta)}{a(\eta)} P\left(\frac{\ell}{d_A},\eta\right), \eeq where we have used the Limber approximation, the orthogonality of spherical harmonics, and employed the shorthand notation, $b_\ell(\eta) \equiv b(\frac{\ell}{d_A},\eta)$.\par The weak lensing-tracer cross-correlation is similarly computed, but it depends also on the shear multiplicative bias, $m$: \beq C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}\Sigma}= m ~ \frac{3}{2}{\ensuremath{\Omega_{\mathrm{m}}} H_{0}^{2}} \int d\eta ~ b_\ell(\eta) W_f(\eta) \frac{\ensuremath{g_{\mathrm{\scriptstyle opt}}}(\eta)}{a(\eta)} P\left(\frac{\ell}{d_A},\eta\right). \eeq Note that if the tracer redshift distribution is narrow, then the ratio between the two reduces to a product of the multiplicative bias times the geometric distance ratio: \begin{equation} \frac{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}} \Sigma}} {C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\Sigma}}= m\frac{\ensuremath{g_{\mathrm{\scriptstyle opt}}}(\eta)}{\ensuremath{g_{\mathrm{\scriptscriptstyle{CMB}}}}(\eta)} \end{equation} As Vallinotto (2012)~\cite{2012ApJ...759...32V} noted, we can also estimate this multiplicative bias by looking at the ratio of the convergence power spectra: \begin{equation} \frac{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}}{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\kcmb}} = m \frac{\int d\eta\ d_A^2(\eta)\frac{\ensuremath{g_{\mathrm{\scriptscriptstyle{CMB}}}}(\eta) \ensuremath{g_{\mathrm{\scriptstyle opt}}}(\eta)}{a^2(\eta)} P(\frac{\ell}{d_A},\eta)}{\int d\eta\ d_A^2(\eta)\frac{\ensuremath{g_{\mathrm{\scriptscriptstyle{CMB}}}}^2(\eta)}{a^2(\eta)} P(\frac{\ell}{d_A},\eta)}\end{equation} or $C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}/{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}\kgal}}$ which has a different dependence on cosmological parameters than the tracer-lensing ratio.\\ \begin{figure*} \includegraphics[width=\columnwidth]{w0_wa_all_bias_actpol_hsc_boss_wgg.pdf} \includegraphics[width=\columnwidth]{nu_sig8_all_bias_actpol_hsc_boss_wgg.pdf} \caption{\emph{Left:} Marginalized 68\% confidence contours on the dark energy equation of state parameters using various combinations of observables. The outer (orange) dashed ellipse is from the addition of the deflection field power spectrum from CMB lensing reconstruction to the tomographic shear power spectra from optical lensing, when the shear multiplicative biases are marginalized over along with other cosmological parameters, including neutrino mass sum. The outer (orange) dotted ellipse is the same as above, but with the biases held fixed at their fiducial values. The outer solid (orange) ellipse results from the addition of the CMB lensing - optical lensing cross-correlations to the above, with the multiplicative biases marginalized over (and thereby self-calibrated). The inner (green) dashed contour results from adding, without any cross-correlation, the CMB lensing, the optical lensing, and the galaxy angular clustering measurements with both the shear biases and the galaxy linear biases marginalized over. The inner (green) dotted contour results from holding these biases fixed, while the inner (solid) ellipse is the result of including cross-correlations between CMB lensing, optical lensing, and galaxy angular clustering, with all the biases marginalized over. The resulting constraints on the biases are displayed in Table~\ref{biasTable}. \emph{Right:} The interpretation is the same as left panel, except here we show the constraints on the neutrino mass sum and $\sigma_8$. Note that in this case, uncertainties in $w_0$, $w_a$ are always marginalized over. Complementary constraints are summarized in Table ~\ref{paramsTable}. See text for survey details. \label{ellipses}} \end{figure*} \section{A Fisher Matrix Approach} We study the effectiveness of the cross-correlation statistics in controlling systematics and their effects on cosmological parameters using a Fisher Matrix approach. We assume $\ensuremath{N_{\mathrm{\scriptstyle opt}}}$ tomographic redshift bins for the weak lensing survey and $\ensuremath{N_{\mathrm{\scriptstyle f}}}$ redshift slices constructed from the photometric galaxy survey. The data vector consists of the possible auto and cross spectra between weak lensing, galaxy angular clustering, and CMB lensing: \beqn \nonumber C_\ell^{XY} &\equiv& \left\{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\kcmb}, \underbrace{\{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}\}}_{\ensuremath{N_{\mathrm{\scriptstyle opt}}}\ \mathrm{spectra}}, \underbrace{ \{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\Sigma}\}}_{\ensuremath{N_{\mathrm{\scriptstyle f}}}\ \mathrm{spectra}}, \right.\\ &&\nonumber \left. \underbrace{\{ C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}\kgal}\}}_{\ensuremath{N_{\mathrm{\scriptstyle opt}}} (\ensuremath{N_{\mathrm{\scriptstyle opt}}} +1)/2\ \mathrm{spectra}}, \underbrace{\{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}\Sigma}\}}_{\ensuremath{N_{\mathrm{\scriptstyle opt}}} \ensuremath{N_{\mathrm{\scriptstyle f}}}\ \mathrm{spectra}},\right. \\ && \left. \underbrace{\{C_{\ell}^{\Sigma\Sigma}\}}_{\ensuremath{N_{\mathrm{\scriptstyle f}}} (\ensuremath{N_{\mathrm{\scriptstyle f}}}+1)/2\ \mathrm{spectra}} \right\} . \eeqn The Fisher Matrix is constructed as, \beq F_{\alpha\beta} = \sum_\ell \sum_{\substack{ \left\{X, Y, W, Z \right\} \in \\ \left\{ \ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}, \ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}, \Sigma \right\}}} \left[ \frac{\partial C_\ell^{XY}}{\partial \alpha}\left( \mathbf{Cov}_\ell^{{XY, WZ}}\right)^{-1}\frac{ \partial C_\ell^{WZ}}{\partial \beta}\right] \eeq where $\alpha, \beta$ run over cosmological as well as nuisance parameters (e.g. multiplicative biases), and the covariance matrix is computed under the assumption of Gaussian random fields: \beq \mathbf{Cov}_\ell^{{XY, WZ}} = \frac{1}{(2\ell+1) f_{\mathrm{sky}}} \left( \tilde C_{\ell}^{XW} \tildeC_{\ell}^{YZ} + \tilde C_{\ell}^{XZ} \tildeC_{\ell}^{YW} \right) \eeq where $f_{\mathrm{sky}}$ is the fraction of sky covered by the overlapping experiments, and $\tilde C_{\ell} \equiv C_{\ell}+ N_\ell$ includes the cosmological signal and noise. The noise for the CMB lensing reconstruction $N_\ell^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}}$ is computed using the quadratic estimator technique \cite{hu/okamoto:2002}. Weak lensing noise in a tomographic bin gets contributions from the shape noise $N_\ell^{\kappa_{opt}} = {\ave{\gamma^2}^{1/2}/\bar n}$ and additive shear errors, following the implementation described in~\cite{2013arXiv1301.1037D}. Lastly, we assume that the noise in the angular correlation of tracers in a redshift bin is purely shot noise dominated: $N^{\Sigma}_\ell= {1}/{\bar n_f}$, where $\bar n_f$ is the number of tracers per steradian in that bin. \\ \section{Surveys} For a near term application of our proposed method we consider the cross-correlations between the weak lensing signal from a HSC-like survey, the CMB lensing from an ACTPol-like experiment and galaxies from a BOSS-like survey. We assume that HSC, the CMB ground based, and the BOSS-like surveys have the same footprint of 4000 deg$^2$ on the sky. For the ACTPol-like survey, we assume a depth of 5 (7) $\mu$K-arcmin in temperature (polarization) and a $1.4$ arcmin beam. The HSC weak lensing sources are assumed to be distributed in redshift as \begin{eqnarray} \centering n(z) = \frac{3 N_g}{2 z_0^3 } \ z^2 \exp\left[-\left(\frac{z}{z_0}\right)^{3/2}\right] \end{eqnarray} with $z_0 = 0.69$ corresponding to a median redshift of unity, and a source density of $N_g$ = 35 galaxies per arcmin$^2$. We assume a photometric redshift error distribution of $\sigma(z) = 0.03 (1+z)$, and an intrinsic shape noise of $\langle \gamma^2\rangle ^{1/2} = 0.4$. We consider three tomographic bins from $0< z <0.7$, $0.7<z<1.5$, $1.5<z<4.0$. We treat the multiplicative bias $m = 1+ \alpha$ in each bin as an independent parameter with empirically motivated fiducial values of $\alpha$= 0.008, 0.01, and 0.02. As the focus of the study is the multiplicative bias, we do not treat the additive shear errors for reasons of clarity. For the BOSS-like survey, we assume three redshift bins $0.3<z<0.4$, $0.4<z<0.5$ and $0.5<z<0.6$ with the linear bias parameter of 2.0 in each bin (which are marginalized over), and a total galaxy density of $0.011$ per arcmin$^2$. \par For the Fisher analysis, we consider the standard set of $\Lambda$CDM parameters with fiducial values: $\Omega_b h^2= 0.02258$, $\Omega_c h^2 = 0.1093$, $\Omega_\Lambda = 0.734$, $n_s = 0.96$ and $\sigma_8 = 0.8$, plus a massive neutrino component $\Omega_\nu h^2 = 0.001596$ corresponding to $\sum m_\nu = 0.15$ eV, and the dark energy equation of state parameterized via $w(z) = w_0 + w_a z/ (1+z) $. We vary these parameters as well as the tracer galaxy biases in each spectroscopic bin, and the shear multiplicative biases in each tomographic bin. We do not include any CMB power spectra information in this analysis.\\ \section{Results} We first confine ourselves to the optical lensing and CMB lensing convergence fields only, and consider only the internal spectra so that the data vector only consists of $\{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\kcmb}$, $C_{\ell}^{{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}_i {\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}_j}\}$. We perform two parameter estimation studies: one with the multiplicative biases fixed at their fiducial value [CMBL+ optL (fix bias)] and the other letting these biases vary and be thereby self-calibrated [CMBL+ optL (self cal)]. Note that although we do not consider the cross-correlation between CMB lensing and optical lensing as an observable yet, such cross-correlation terms do appear in the covariance matrix. We cannot simply add the CMB lensing and optical lensing Fisher matrices here, as the two effects are correlated. Fig.~\ref{ellipses} and Table~\ref{paramsTable} show the resulting constraints on the dark energy equation of state and the neutrino mass sum for these two scenarios (note that we always marginalize over all other parameters in these plots, and complementary cases are summarized in the table). Letting the multiplicative shear float significantly degrades both the dark energy and neutrino figures of merit. Next, we expand our data vector to include the CMB lensing-optical lensing cross-correlation: $\{C_{\ell}^{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}\kcmb}$, $C_{\ell}^{{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}_i {\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}_j}$, $C_{\ell}^{{\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{{CMB}}}}}}_i {\ensuremath{\kappa_{\mathrm{\scriptscriptstyle{opt}}}}}_j}\}$ and let the biases vary [CMBL $\times$ optL (self cal with c.c.)]. This shows the power of the cross-correlations --- including the cross-correlation leads to self calibration of the shear multiplicative biases to sufficient accuracy so that the parameter uncertainty ellipses shrink to give almost the same figures of merit as the case where the multiplicative biases were held fixed. The corresponding marginalized constraints on the multiplicative biases are shown in Table~\ref{biasTable}. Note that $\sim$10\% deviations of the multiplicative biases from unity can be calibrated by this method. The constraints will improve significantly in future with larger and deeper weak lensing/CMB lensing surveys, e.g.~\cite{2013arXiv1309.5383A}. \par Next, we include the three BOSS-like galaxy number density bins in the analysis. First, we look at the constraints avoiding any cross-correlations , once fixing both shear and galaxy biases [CMBL + optL + gal (fix bias)], and then letting the biases self calibrate [CMBL + optL + gal (self cal)]. As in the previous case, letting biases float significantly degrades all constraints. There are two interesting things to note here: first, just adding galaxy clustering information significantly improves dark energy and neutrino mass. Second, and related to the above, just by having these overlapping data sets self calibrates the galaxy biases to 3-4\% This is not surprising because both CMB lensing and optical lensing constrains growth (and $\sigma_8$) and galaxy biases. For comparison, if we only had the BOSS-like survey, then the Fisher analysis predicts galaxy bias errors of $(0.23, 0.24, 0.28 )$. Next we introduce all possible cross-correlations between the CMB lensing, optical lensing, and the galaxy density fields [CMBL $\times$ optL $\times$ gal (self cal w/ c. c.)]. We observe a significant improvement the shear multiplicative bias constraints in this case, with the deviation from unity constrained down to $0.5\%$ in the highest redshift bin. We also notice an appreciable improvement in the constraints on the galaxies bias parameters. These improvements propagate to cosmological constraints, and one can get significantly closer through this cross-correlation procedure to the dark energy (neutrino) figures of merit where the biases are artificially held fixed. Having a spectroscopic sample breaks the degeneracy between multiplicative bias and growth that limits the use of lensing correlations alone \cite{2012ApJ...759...32V}. \begin{table} \caption{Estimated marginalized 1-$\sigma$ error on the shear multiplicative bias parameter $m_i= (1+\alpha_i)$ in the tomographic bin $i$, and the galaxy bias parameter $b_j$ in spectroscopic bin $j$ for various ways of combining data sets.\label{biasTable}} \center \begin{tabular}{l|l|l|l|l|l} \hline \hline Bias & Fiducial & CMBL & CMBL & CMBL & CMBL \\ parameter & value & + optL & $\times$ optL & + optL & $\times $ optL \\ & & & & + gal. & $\times $ gal \\ \hline \hline $\alpha_1$ & 0.008 & 0.058 & 0.026 & 0.047 & 0.021\\ $\alpha_2$ & 0.014 & 0.063 & 0.010 & 0.054 & 0.008\\ $\alpha_3$ & 0.020 & 0.063 & 0.007 & 0.053 & 0.005\\ \hline $b_1$ & 2.000 & - & - & 0.070 & 0.053 \\ $b_2$ & 2.000 & - & - & 0.058 & 0.044 \\ $b_3$ & 2.000 & - & - & 0.073 & 0.055 \\ \hline \hline \end{tabular} \end{table} \begin{table*} \caption{Estimated 1-$\sigma$ error on chosen cosmological parameters: $\left\{ \sigma_8, \Sigma m_\nu, w_0, w_a \right\}$ \emph{marginalized} over the usual $\Lambda$CDM parameters. Numbers with (without) parenthesis are obtained with fixed (varying) bias parameters. A dash means that the constraint has not been marginalized over this parameters\label{paramsTable}.} \center {\tiny \begin{tabular}{||c|c|ccc|ccc|ccc|ccc||} \hline \hline & \textbf{Fiducial} & \multicolumn{3}{c|}{\textbf{CMBL}} & \multicolumn{3}{c|}{\textbf{CMBL}} & \multicolumn{3}{c}{\textbf{CMBL}} & \multicolumn{3}{|c||}{\textbf{CMBL}} \\ & \textbf{value} & \multicolumn{3}{c}{\textbf{+ optL} }& \multicolumn{3}{|c}{$\times$ \textbf{optL}} & \multicolumn{3}{|c}{\textbf{+ optL}} & \multicolumn{3}{|c||}{$\times $ \textbf{optL}} \\ & & \multicolumn{3}{c|}{ }& \multicolumn{3}{c|}{ } & \multicolumn{3}{c|}{\textbf{+ gal.}} & \multicolumn{3}{c||}{$\times $ \textbf{gal}} \\ \hline \hline $\sigma \left( \sigma_8 \right)$ & 0.80 & 0.0211 & 0.0524 & 0.0842 &0.0172 & 0.0238 & 0.0777 & 0.0199 & 0.0470 & 0.0642 & 0.0158 & 0.0203 & 0.0498 \\ & & (0.0144) & (0.0160) & (0.0704) & (0.0133) & (0.0149) & (0.0683) & (0.00484) & (0.0114) & (0.0381) & (0.0112) & (0.0120) & (0.0464) \\ \hline $\sigma\left( \Sigma m_\nu \right) [{\rm meV}]$ & 150 & 305 & 327 & 483 & 208 & 311 & 411 & 273 & 288 & 414 & 191 & 257 & 350 \\ & & (156) & (313) & (424) & (142) & (275) & (380) & (128) & (184) & (199) & (138) & (240) & (327) \\ \hline $\sigma\left( w_0 \right)$ & -1.0 & - & 0.288 & 0.668 & - & 0.132 & 0.607 & - & 0.256 & 0.489 & - & 0.103 & 0.381 \\ & & & (0.0983) & (0.542) & & (0.0921) & (0.523) & & (0.0677) & (0.346) & & (0.0779) & (0.359) \\ \hline $\sigma\left( w_a \right)$ & 0.0 & - &- & 2.16 & - & - & 1.70 & - & - & 1.63 & - & - & 1.14 \\ & & & & (1.55) & & & (1.49) & & & (0.932) & & & (1.06) \\ \hline \hline \end{tabular}} \end{table*} \section{Conclusions} The world astrophysics community is about to embark on several major weak lensing surveys. Over the coming decade, we are likely to invest several billion dollars in missions among whose primary goals is to use weak lensing measurements to determine the nature of dark energy, the geometry of the universe and to constrain the nature of dark matter. Multiplicative biases may well limit the scientific utility of these surveys. In this letter, we show that by combining optical weak lensing measurements with CMB lensing measurements and a spectroscopic survey, we can calibrate the amplitude of the multiplicative bias in the optical lensing survey. This will likely noticeably improve our ability to measure cosmological parameters. In particular, it is remarkable that the Fisher cross terms brings enough information to significantly shrink uncertainties on parameters estimation. For the upcoming HSC survey, the combination of CMB lensing measurements from an ACTPol-like experiment and BOSS galaxy spectroscopy should enable a calibration of the multiplicative bias at the $0.5-2.0\%$ level, and reduce the error bar on cosmological parameters by $\sim 20-25\%$. In the near future, stage-III CMB experiments should be able to make accurate enough CMB lensing measurements to be able to calibrate optical weak lensing surveys like LSST at the $0.25\%$ level. \acknowledgments{The Berkeley Weak Lensing code developed by Sudeep Das, Reiko Nakajima, Roland de Putter, and Eric Linder has been used extensively in this study. We thank Gil Holder, Alberto Vallinotto and Benjamin Joachimi for useful comments and discussions. }
\section{Introduction\label{sec:intro}} The cosmological and astrophysical observations suggest that 27\% of the energy density of the universe is in the form of dark matter (DM) \cite{Komatsu:2010fb, Ade:2013zuv}. The most promising candidate for DM is so-called weakly interacting massive particle (WIMP). In that case we may study its nature through creation at accelerators such as LHC, the scattering with ordinary matter, or the pair annihilation into ordinary standard model (SM) particles including photon~\cite{Zeldovich:1980st}. The current DM density of the universe is related to the annihilation cross section at the decoupling temperature as \begin{eqnarray} \Omega_{\rm DM} h^2 = \frac{3 \times 10^{-27} {\rm cm^3/s}} {\langle \sigma v \rangle_{\rm th}}.\label{eq:cs-fermi} \end{eqnarray} Recently the analysis of FermiLAT gamma-ray data showed that there may be some peak near 130 GeV, which can be interpreted as the annihilation of DM~\cite{Bringmann:2012,Weniger:2012}. This interpretation requires the annihilation cross section to be about 4\% of freeze-out cross section: \begin{eqnarray} \langle \sigma v \rangle_{\gamma\gamma} = 0.042 \langle \sigma v \rangle_{\rm th} =0.042 {\rm ~pb}\, c. \label{eq:sigmav_Fermi} \end{eqnarray} Since we expect the annihilation into photons apparently come from loop-induced process whose cross section is estimated to be \begin{eqnarray} \frac{ \langle \sigma v \rangle_{\gamma\gamma}}{ \langle \sigma v \rangle_{\rm th}} = \left(\frac{\alpha_{\rm em}}{4\pi} \right)^2 \sim 10^{-7}, \end{eqnarray} the observation calls for some non-conventional models. There are many attempts to explain the Fermi-LAT data with DM annihilation or decay by many authors \cite{recent,non-SUSY-1, non-SUSY-2, non-SUSY-3,Toma:2013bka, Giacchino:2013bta,BKS}. In this paper we introduce a new mechanism that provides a possible explanation for the Fermi-LAT anomaly on the basis of a model recently proposed by Weinberg~\cite{Weinberg:2013kea}. This model is originally suggested in order to explain the possible deviation in the effective neutrino number, $\Delta N_{\rm eff}=0.36 \pm 0.34$ at the 68~\% confidence level from the Planck, WMAP9 polarization and ground-based data~\cite{Ade:2013zuv}, although it is not very significant. We will just use the central value of the deviation from now on. The author introduces a complex scalar field charged under a global $U(1)_X$ symmetry in the hidden sector. All the SM particles are neutral under $U(1)_X$ and they interact with the hidden sector via the renormalizable Higgs portal interaction~\cite{Higgs_portal}. The Goldstone boson (GB) after spontaneous symmetry breaking can contribute to the relativistic energy density. Moreover he showed that a fermion in the hidden sector can be introduced in such a way that it can carry odd parity after $U(1)_X$ is broken down to $Z_2$. So the hidden fermion could be a promising DM candidate. Some authors analyze the Weinberg model in another aspects as well as LHC phenomenology ~\cite{Garcia-Cely:2013nin, Cheung:2013oya}. To give the observed Fermi-LAT gamma-ray line signal, the DM is assumed to have mass around 214 GeV. The correct thermal relic density is achieved by the DM interaction with the light scalar which decays dominantly into the GB. It turns out that the same channel can accommodate the Fermi-LAT data when the light scalar decays subdominantly into neutral pions. This paper is organized as follows. In Section 2, we define our model based on Weinberg model. In Section 3, we discuss the possibility to explain the Fermi-LAT observation retaining consistency with various other DM phenomenology. We conclude in Section 4. \textcolor{red}{ } \section{ Model} In this section we set up a model, in which two new fields charged under global $U(1)_X$ symmetry $\chi$ and $\Psi_\pm$ are introduced in addition to the SM fields. Here $\chi$ is boson with charge +2 and $\Psi_\pm$ is fermion with charge +1 under the $U(1)_X$ symmetry. Notice that the SM fields are neutral under this symmetry. We expect the lighter one of $\Psi_\pm$ to be a DM candidate. {\it Scalar sector}: The new Lagrangian for the scalar sector is typically given by \begin{equation} {\cal L}= \partial_\mu\chi^*\partial^\mu\chi+\mu_\chi^2\chi^*\chi-\frac{\lambda_\chi}{2}(\chi^*\chi)^2-\lambda_{H\chi}(\Phi^\dag\Phi)(\chi^*\chi) +{\cal L}_{\rm SM}, \label{lag-higgs} \end{equation} where $\mu_\chi^2$, $\lambda_\chi$, and $\lambda_{\chi\Phi}$ are real. Here we define the scalar fields in the unitary gauge as follows: \begin{equation} \Phi(x) = \frac{1}{\sqrt{2}} \left[ \begin{array}{c} 0\\ v_H+\phi(x) \end{array}\right],\ \chi(x) = \frac{1}{\sqrt{2}} \left[ \begin{array}{c} v_\chi+r(x) \end{array}\right]e^{2i\alpha(x)}, \end{equation} where $ v_H^2=(246\ {\rm GeV})^2$ is the vacuum expectation value (vev) of the SM, and $v_\chi$ is vev of the hidden sector, which can be determined by the analysis of DM data. Then Eq. (\ref{lag-higgs}) can be rewritten as \begin{eqnarray} {\cal L}&=& \frac12\partial_\mu r\partial^\mu r + 2(v_\chi+r)^2\partial_\mu\alpha\partial^\mu\alpha +\frac12 \mu_\chi^2(v_\chi+r)^2 \nonumber\\ &-&\frac{\lambda_\chi}{8}(v_\chi+r)^4 -\frac12\lambda_{H\chi}(\Phi^\dag\Phi)(v_\chi+r)^2 +{\cal L}_{\rm SM}. \label{lag-higgs-expand} \end{eqnarray} The CP-even scalr mass-squared matrix in the basis of $(\phi,r)^t$ can be diagonalized by the following mixing matrix \begin{eqnarray} M^2_{\rm Higgs} &\equiv& \left(\begin{array}{cc} \lambda_Hv_H^2 & \lambda_{H\chi}v_Hv_\chi \\ \lambda_{H\chi}v_Hv_\chi & \lambda_\chi v_\chi^2 \end{array}\right)\\ \!\!&=&\!\! \left(\begin{array}{cc} \cos\theta & \sin\theta \\ -\sin\theta & \cos\theta \end{array}\right) \!\!\left(\begin{array}{cc} m^2_{{1}} & 0 \\ 0 & m^2_{{2}} \end{array}\right)\!\! \left(\begin{array}{cc} \cos\theta & -\sin\theta \\ \sin\theta & \cos\theta \end{array}\right),\nonumber \label{eq:higgs} \end{eqnarray} where $\tan2\theta=2 \lambda_{H\chi}v_Hv_\chi/(\lambda_\chi v_\chi^2-\lambda_Hv_H^2)$. The gauge eigenstate $(\phi,r)^t$ can be rewritten in terms of the mass eigenstate $(H_1,H_2)^t$ as \begin{eqnarray} \phi &=& H_{1} \cos\theta + H_{2}\sin\theta, \nonumber\\ r &=&- H_{1} \sin\theta + H_{2} \cos\theta. \label{eq:mass_weak} \end{eqnarray} Hereafter we regard $H_1$ ($m_{1}$=125 GeV) as the SM Higgs boson, and $H_2$ as a lighter scalar boson whose mass is expected to be small of the order 500 MeV to accommodate a significant deviation in the effective neutrino number $\Delta N_{\rm eff} =0.36$~\cite{Weinberg:2013kea}. {\it Dark sector}: The new Lagrangian for the DM sector is given by \begin{eqnarray} {\cal L}&=& \frac{i}{2}\left(\bar\Psi_+\gamma^\mu\partial_\mu\Psi_+ + \bar\Psi_-\gamma^\mu\partial_\mu\Psi_-\right)\nonumber\\ &-& \frac{i}{4v_\chi}\partial_\mu\alpha'\left(\bar\Psi_+\gamma^\mu\Psi_- - \bar\Psi_-\gamma^\mu\Psi_+\right)\nonumber\\ &-& \frac{f}{2}(- H_{1} \sin\theta + H_{2} \cos\theta)\left(\bar\Psi_+\Psi_+ - \bar\Psi_-\Psi_-\right)\nonumber\\ &-& \frac{1}{2}\left(m_+\bar\Psi_+\Psi_+ +m_- \bar\Psi_-\Psi_-\right), \label{int-dm} \end{eqnarray} where we redefined $\alpha=\alpha'/(2v_\chi)$. Here we can take $f>0$ without loss of generality. Then $\Psi_-$ is a DM candidate with mass $M_{\rm DM} = m_-$. We also obtain the mass difference, $\Delta m\equiv m_+ - m_- = 2 f v_\chi$. It turns out that $\Delta m$ is very large in our scenario as we will see later. To get the large mass difference we need some degree of fine-tuning to get $M_{\rm DM}$ at electroweak scale. \section{ Dark Matter } In our DM analysis, we focus on explaining $\gamma$-ray excess at 130 GeV reported by the Fermi-LAT experiment. It is however worth mentioning the constraints from the other experiments before we go to the main part. {\it Invisible decay of SM Higgs}: The current experiment at LHC tells us that the invisible branching ratio of the SM Higgs (${\rm B_{inv}}$) is conservatively estimated to be less than 20\%~\cite{Belanger:2013kya}. There are two invisible modes: $H_1\to 2\alpha'$ and $H_1\to 2 {\rm DM}$, and their decay rates ($\Gamma_{\rm inv}$) are given by \cite{Garcia-Cely:2013nin} \begin{eqnarray} &&\Gamma_{\rm inv}\equiv \Gamma(H_1\to2\alpha')+\Gamma(H_1\to2 {\rm DM}),\\ &&\Gamma(H_1\to2\alpha')=\frac{m_{1}^3}{32\pi v_H^2}\sin^2\theta,\\ &&\Gamma(H_1\to2{\rm DM})=\frac{f^2 \sin^2\theta}{16\pi m_1^2}(m_1^2-4 M_{\rm DM})^{3/2}. \end{eqnarray} However, we consider $M_{\rm DM} > m_1/2$ and the latter mode is forbidden kinematically in our scenario. One obtains the following relation~\cite{Garcia-Cely:2013nin} \begin{equation} \Gamma_{\rm inv}< \frac{{\rm B_{inv}}\cos^2\theta}{1-{\rm B_{inv}}}\Gamma^{\rm SM}_{\rm Higgs}, \end{equation} where $\Gamma^{\rm SM}_{\rm Higgs}$ is the total decay width of the SM Higgs boson and estimated as $4.1\times 10^{-3}$ GeV at $m_{1}$=125 GeV. The upper bound on the invisible decay of Higgs restricts $\theta \lesssim 0.06$. This constraint is much weaker than that from the direct detection of DM. Another decay mode is $H_1 \to 2 H_2$ whose rate is given by \begin{equation} \Gamma(H_1\to2H_2)\simeq \frac{m_{1}^3(v_H\cos\theta-v_\chi\sin\theta)^2}{128\pi v_H^2v_\chi^2}\sin^22\theta. \label{eq:H22} \end{equation} We assume this rate is about 10\% of that of $H_1 \to 2 \alpha^\prime$ to explain the Fermi-LAT gamma-ray line. {\it Direct detection}: \begin{figure}[t] \centering \includegraphics[width=0.35\textwidth]{Weinberg-DD.eps} \caption{The Feynman diagram for the elastic scattering of DM off the matter. } \label{fig:DD} \end{figure} The relevant process contributing to the spin independent scattering cross section is the $t-$channel diagram mediated by the lighter scalar as shown in Fig.~\ref{fig:DD}. The corresponding elastic cross section is estimated as \begin{equation} \sigma_p \approx 0.27^2 \frac{f^2 m_p^2 M_{\rm DM}^2}{4 \pi v_H^2 (M_{\rm DM} + m_p)^2 } \left({1 \over m_1^2} -{1 \over m_2^2}\right)^2 \sin^2 2\theta, \end{equation} where $m_p \approx 1$ GeV is the proton mass. It suggests the following constraint \cite{Garcia-Cely:2013nin}, which is derived from the current upper bound reported by XENON100 and LUX~\cite{xenon100}: \begin{equation} |f\sin2\theta| \le {\cal O}(10^{-5}). \label{eq:DD} \end{equation} {\it Fermi-LAT and Relic density}: \begin{figure}[htb] \centering \includegraphics[width=0.15\textwidth,height=3cm]{Weinberg-H2-pp.eps} \hspace{1cm} \includegraphics[width=0.15\textwidth,height=3cm]{Weinberg-ann-22.eps} \caption{The dominant Feynman diagrams for the relic density, where the black circle in the right panel represents all the possible channels at tree level. } \label{fig:ann} \end{figure} The possible dominant annihilation channels to obtain the current relic density are shown in Fig.~\ref{fig:ann}. They are i) 2 DM $\to 2 \pi$, ii) 2 DM $\to 2 H_2$, iii) 2 DM $\to 2 \alpha^\prime$ and iv) DM coannihilation channels. But the channel i) is strongly suppressed because its amplitude has the same parametric combination with Eq.~(\ref{eq:DD}). Therefore the relic density is achieved either by one of ii), iii), iv) or combination of them. The Ref.~\cite{Garcia-Cely:2013nin} shows that iv) is dominant when it is allowed, which is not allowed in our case. The annihilation cross sections for the processes ii) and iii) are estimated to be \begin{eqnarray} \langle \sigma v \rangle_{2 H_2} \approx \frac{3 f^4 v_{\rm rel}^2}{128 \pi M_{\rm DM}^2}, \quad \langle \sigma v \rangle_{2 \alpha^\prime} \approx \frac{ f^4 v_{\rm rel}^2}{32 \pi m_{+}^2}. \end{eqnarray} In our case it turns out the mode ii) is dominant because $M_{\rm DM} \ll m_{+}$. The photon line observed by Fermi-LAT comes from ii) when $H_2 \to 2 \pi$. Then the pion decays into two photons. It implies that the mass of DM is fixed to be 214 GeV~\footnote{If the gamma-ray line were emitted from $H_2$, its maximum energy would be just $M_{\rm DM}$ In our case the peak energy $E_\gamma=130$ GeV is obtained for $M_{\rm DM} = 214$ GeV, since it comes from the decay of $H_2$ into two pions.}. For either case, the condition Eq.~(\ref{eq:sigmav_Fermi}) can be satisfied if \begin{eqnarray} {\rm B}(H_2 \to 2 \pi^0)\approx 1.1~\%, \label{eq:R} \end{eqnarray} because ${\rm B}(\pi^0\to2\gamma)\simeq 99\%$~\cite{Beringer:1900zz}. This can be understood as follows: if we set $x_0 = {\rm B}(H_2 \to \pi^0 \pi^0)$, $x_+ = {\rm B}(H_2 \to \pi^+ \pi^-)$, $x_\alpha = {\rm B}(H_2 \to \alpha' \alpha')$, and considering $\psi_- \psi_- \to H_2 H_2$ dominates, Eq.~(\ref{eq:sigmav_Fermi}) requires \begin{eqnarray} 4 x_0^2 + 2 (2 x_\alpha x_0) + 2 (2 x_0 x_+) \approx 4.2~\%. \end{eqnarray} Using $x_0 + x_+ + x_\alpha \approx 1$, we get $x_0 \approx 1.1~\%$. We define ratio $R$ as \begin{eqnarray} R \equiv \frac{ {\rm B}(H_2 \to 2 \pi^0)}{ {\rm B}(H_2 \to \alpha' \alpha')} \approx 1.1~\%. \end{eqnarray} The ratio $R$ is given in the Ref.~\cite{Cheung:2013oya}, \begin{eqnarray} R &=& \theta^2 \frac{v_\chi^2}{v_H^2} \left(1-\frac{4 m_\pi^2}{m_2^2}\right)^{1/2} \left(1+\frac{2 m_\pi^2}{m_2^2}\right)^{2}\nonumber\\ & \approx& 0.011 \left(\theta \over 10^{-5} \right)^2 \left(v_\chi \over 2.5 \times 10^6 {\rm GeV} \right)^2. \end{eqnarray} We can see $f \approx 0.9$, $\theta \approx 10^{-5}$, and $v_\chi \approx 2.5 \times 10^6$ GeV satisfies both Eq.~(\ref{eq:DD}) and Eq.~(\ref{eq:R}), using micromegas \cite{Belanger:2008sj}. As a result we can obtain the correct annihilation cross section necessary to explain the Fermi-LAT gamma-ray line. It is worth mentioning that the shape line at 130 GeV is rather wide, if the photons are produced via neutral pions \cite{ non-SUSY-2, pion}, although the fall-off of the peak can be explained and the data still show fluctuation in the spectrum. We also note that in our scenario there is no associated $Z$-boson or Higgs boson production contrary to most other works where the signal comes from the loops of charged particles. Future experiments with more data will give more clues on the possible scenarios. \section{Conclusions} \label{sec:Conclusions} We considered a dark global $U(1)_X$ model with a Goldstone boson and a dark matter~\cite{Weinberg:2013kea}. The Goldstone boson can contribute to the effective neutrino number $\Delta N_{\rm eff} =0.36$ if the dark scalar mass is about 500 MeV. We showed that this light dark scalar produced by the dark matter annihilation can mix with the SM Higgs boson and about 4\% of them can decay into two neutral pions. These pions finally decay into two photons with energy 130 GeV if the dark matter mass is 214 GeV. Our benchmark parameters for the Fermi-LAT gamma-ray line are dark scalar coupling with the dark matter $\sim 1$, the mixing angle of the dark scalar with the SM Higgs $\sim 10^{-5}$, and the vev of the dark scalar $\sim 3.5 \ \times 10^{6}$ GeV. The dark matter relic density can possibly be obtained by four channels: i) 2 DM $\to 2 \pi$, ii) 2 DM $\to 2 H_2$, iii) 2 DM $\to 2 \alpha^\prime$ and iv) DM coannihilation. But the strong constraint from the direct detection makes the channel i) always negligible. The parameter space explaining the Fermi-LAT gamma-ray line makes the iii) and iv) suppressed. Therefore only ii) is dominant contribution to the cross section for the relic density. The obtained gamma-ray spectrum is broad box shape and does not fit to the data perfectly, but the data show there may be fluctuation in the spectrum. There is no associated $Z\gamma$ or $h \gamma$ production signal contrary to most other works where the signal comes from the loops of charged particles. The annihilation into the other SM particles are highly suppressed due to the small mixing from the direct detection, so we can avoid the constraints from the indirect detection easily. Future experiments with more data will give more clues on the possible scenarios. The generic signature of the model at the collider is the production of $H_1$ via $g g \to H_1$ and its subsequent decay $H_1 \to H_2 H_2 \to (\pi\pi) (\alpha^\prime \alpha^\prime)$~\cite{Cheung:2013oya}. However, the branching ratio $\Gamma(H_1 \to 2 H_2)$ is very small in our scenario as can be seen in Eq.~(\ref{eq:H22}). \acknowledgments We are grateful to P. Ko and T. Toma for fruitful discussions and KIAS Workshop in Jeju island for providing us with the nice environment for discussion. This work is partly supported by NRF Research Grant 2012R1A2A1A01006053 (SB).
\section{Introduction} \setcounter{equation}{0} \label{s1} One of the main tasks of nondifferentiable optimization is to extend some optimality conditions to more general classes of nondifferentiable functions. There are necessary and sufficient conditions in unconstrained optimization in terms of various generalized derivatives. The most of them are of first- and second-order. The higher-order conditions are rather limited. Such results were obtained in \cite{aub90,gin02,ban01,hof78,jim08,lin82,pal91,stu86,stu99,war94}. Even the conditions of first- and second-order are satisfied for restricted classes of functions when we apply the known directional derivatives: locally Lipschitz, continuously differentiable, lower semicontinuous, the class C$^{1,1}$, and so on functions. Consider, for example, the lower Dini directional derivative, which is one of the most simple and popular ones. The higher-order necessary conditions hold for an arbitrary nondifferentiable function, but the sufficient ones do not. Recently, Bedna\v r\'ik and Pastor \cite{pas08} introduced a new class of $\ell$--stable functions and generalized some second-order sufficient criteria applying these objects. The class of $\ell$-stable functions includes the functions with locally Lipschitz gradient. These facts motivate us to search for a new directional derivative such that higher-order optimality conditions in terms of it are satisfied for an arbitrary function. In our opinion, the following question is important for nondifferentiable optimization: Are there any directional derivatives such that both the necessary conditions for a local minimum and the sufficient ones are satisfied for any function, not necessarily differentiable. In particular, these derivatives should extend the classical Fr\'echet derivatives. In this paper, we introduce a new generalized directional derivative of order $n$ ($n$ is a positive integer) such that the necessary conditions for optimality and the sufficient ones hold for arbitrary not necessarily differentiable function. We obtain necessary conditions for a local minimum, sufficient ones for a strict local minimum, and complete characterizations of isolated local minimizers of order $n$ ($n$ is a positive integer) in terms of this derivative. We derive our criteria for arbitrary proper extended real functions. The convergence in the definition of the derivatives is of Hadamard type. We introduce a subdifferential of order $n$ and apply it in the optimality criteria. We additionally prove conditions of order $n$, which are both necessary and sufficient for a given point to be a global minimizer. They concern a new class of invex functions of order $n$. We introduce another new class of functions that we denote by $\mathcal F_n$, where $n$ is a positive integer such that $n\ge 2$. We obtain necessary and sufficient conditions for a point $\bar x$ to be an isolated minimizer of order $n$ of a function $f\in\mathcal F_n$, which are quite different from the case of an arbitrary function. The derivative and the subdifferential of order $n$ of the given function do not appear in our criteria for an isolated local minimizer of order $n$. At last, we compare our optimality conditions with some known results. We show that some of the theorems by Huang and Ng \cite{hua94} and Chaney \cite{cha87}, which concern the second-order derivative of Chaney follow from our optimality conditions. The main result of Ben-Tal and Zowe \cite{bt85} also is a consequence of our results. Higher-order necessary and higher-order sufficient conditions for an isolated minimum were obtained by Studniarski \cite{stu86}. In his paper, Studniarski applied some derivatives of Hadamard type, which were introduced by him. Other higher-order derivatives of Hadamard type are studied in \cite{aub90,gin02,ban01}. In contrast of our derivatives, the derivatives in \cite{aub90,gin02,stu86} are not consistent with the classical Fr\'echet derivatives. Some of their conditions for optimality do not concern arbitrary functions like our ones. \section{Higher-order directional derivatives and subdifferentials of Ha\-da\-mard type} \label{s2} \setcounter{equation}{0} In this paper, we suppose that $\mathbb E$ is a real finite-dimensional Euclidean space. Denote by $\mathbb R$ the set of reals and $\overline{\mathbb R}=\mathbb R\cup\{-\infty\}\cup\{+\infty\}$. Let $X$ and $Y$ be two linear spaces and $L(X,Y)$ be the space of all continuous linear operators from $X$ to $Y$. Then denote by $L^1(\mathbb E)$ the space $L(\mathbb E,\mathbb R)$, by $L^2(\mathbb E)$ the space $L(\mathbb E,L^1(\mathbb E))$ and so on. If $n$ is an arbitrary positive integer such that $n>1$, let $L^n(\mathbb E)$ be the linear space $L(\mathbb E,L^{n-1}(\mathbb E))$. Consider a proper extended real function $f : \E\to\R\cup\{+\infty\}$, that is a function, which never takes the value $-\infty$. The domain of a proper extended real function is the set: \[ {\rm dom}\; f:=\{x\in\mathbb E\mid f(x)<+\infty\}. \] \begin{definition}\label{def1} The lower Hadamard directional derivative of a function $f : \E\to\R\cup\{+\infty\}$ at a point $x\in\dom f$ in direction $u\in\mathbb E$ is defined as follows: \[ \ld 1 f(x;u)=\liminf_{t\downarrow 0,u^\prime\to u}\, t^{-1}[f(x+t u^\prime)-f(x)]. \] Here $t$ tends to 0 with positive values, and $u^\prime\to u$ implies that the norm $\norm{u^\prime-u}$ approaches $0$. \end{definition} \begin{definition} Recall that the lower Hadamard subdifferential of a function $f : \E\to\R\cup\{+\infty\}$ at some point $x\in\dom f$ is defined by the following relation: \[ \lsubd 1 f(x)=\{ x^*\in L^1(\mathbb E)\mid x^*(u)\le\ld 1 f (x;u)\quad\textrm{for all directions}\quad u\in\mathbb E\}. \] \end{definition} We introduce the following definitions: \begin{definition} Let $f : \E\to\R\cup\{+\infty\}$ be an arbitrary proper extended real function. Suppose that $x^*_1$ is a fixed element from the lower Hadamard subdifferential $\lsubd 1 f (x)$ at the point $x\in\dom f$. Then the lower second-order derivative of Hadamard type of $f$ at $x\in\dom f$ in direction $u\in\mathbb E$ is defined as follows: \[ \ld 2 f(x;x^*_1;u)=\liminf_{t\downarrow 0,u^\prime\to u}\, 2t^{-2}[f(x+t u^\prime)-f(x)-tx^*_1(u^\prime)]. \] \end{definition} \begin{definition} Let $f : \E\to\R\cup\{+\infty\}$ be an arbitrary proper extended real function. Suppose that $x\in\dom f$, $x^*_1\in\lsubd 1 f(x)$. The lower second-order Hadamard subdifferential of the function $f : \E\to\R\cup\{+\infty\}$ at the point $x\in\dom f$ is defined by the following relation: \[ \lsubd 2 f(x;x^*_1)=\{ x^*\in L^2(\mathbb E)\mid x^*(u)(u)\le\ld 2 f (x;x^*_1;u)\quad\textrm{for all directions}\quad u\in\mathbb E\}. \] \end{definition} \begin{definition}\label{def3} Let $f : \E\to\R\cup\{+\infty\}$ be an arbitrary proper extended real function, and $n$ be any positive integer. Suppose that the lower Hadamard subdifferential \[ \lsubd i f (x;x_1^*,x_2^*,\dots,x_{i-1}^*),\quad i=1,2,\dots, n-1 \] of order $i$ at the point $x\in\dom f$ is nonempty and $x^*_i$ is a fixed point from it. Then the lower derivative of Hadamard type of order $n$ of $f$ at $x\in\dom f$ in direction $u\in\mathbb E$ is defined as follows: \[ \ld n f(x;x^*_1,x^*_2,\dots,x^*_{n-1};u)=\liminf_{t\downarrow 0,u^\prime\to u}\,\Delta_n, \] where \[ \Delta_n= n!\, t^{-n}\, [f(x+t u^\prime)-f(x)-\sum_{i=1}^{n-1}\frac{t^i}{i!}\, x^*_i\underbrace{(u^\prime)(u^\prime)\dots (u^\prime)}_{i-\text{times}}]. \] This derivative is well defined as element of $\bar\mathbb R$, because only the term $f(x+t u^\prime)$ can be infinite in the expression for $\Delta_n$. \end{definition} \begin{definition}\label{def4} Suppose that $f : \E\to\R\cup\{+\infty\}$ is an arbitrary proper extended function, and $n$ is any positive integer. Let $x^*_i$ be a fixed point from the lower Hadamard subdifferential $\lsubd i f (x;x_1^*,x_2^*,\dots,x_{i-1}^*)$, $i=1,2,\dots, n-1$ of order $i$ at the point $x\in\dom f$. Then the lower subdifferential of Hadamard type of order $n$ of $f$ at $x\in\dom f$ is defined as follows: \[ \lsubd n f(x;x^*_1,x^*_2,\dots,x^*_{n-1})=\{ x^*\in L^n(\mathbb E)\mid x^*\underbrace{(u)(u)\dots (u)}_{n-\text{times}} \] \[ \le\ld n f (x;x^*_1,x^*_2,\dots,x^*_{n-1};u),\;\forall u\in\mathbb E\}. \] \end{definition} The essence of the next result is that the derivatives, defined in Definition \ref{def3} generalize the usual classical ones in contrast of the derivative in \cite{stu86} and a lot of other derivatives. \begin{proposition}\label{pr3} Let the function $f : \E\to\R\cup\{+\infty\}$ have Fr\'echet derivatives \[ \nabla f(y),\nabla^2 f(y),\dots,\nabla^{n-1} f(y) \] at each point $y\in\mathbb E$ from some neighborhood of the point $x\in\mathbb E$, and let there exists the n-th order Fr\'echet derivative $\nabla^n f(x).$ Then the lower derivatives of order $m$ exist for every integer $m$ such that $1\le m\le n$ and we have the following relations: \[ \ld 1 f(x;u)=\nabla f(x)(u);\quad \lsubd 1 f(x)=\{\nabla f(x)\}; \] \[ \ld m f(x;\nabla f(x),\nabla^2 f(x),\dots,\nabla^{m-1} f(x);u)=\nabla^m f(x)\underbrace{(u)(u)\dots (u)}_{m-\text{times}},\quad m=2,3,\dots, n; \] \begin{equation}\label{28} \nabla^m f(x)\in\lsubd m f(x;\nabla f(x),\nabla^2 f(x),\dots,\nabla^{m-1} f(x)) ,\quad m=2,3,\dots, n. \end{equation} \end{proposition} \begin{proof} The first-order relations are well known, because they concern the Hadamard directional derivative. We prove by induction the relations of order $m>1$. Suppose that they are satisfied for every positive integer $k<m$. It follows from here that \[ \ld m f(x;\nabla f(x),\nabla^2 f(x),\dots,\nabla^{m-1} f(x);u) \] is well defined. By Taylor's expansion formula with a reminder in the form of Peano \cite{il82} we have \[ f(x+t u^\prime)=f(x)+\sum_{i=1}^m\,\frac{1}{i!}\, \nabla^i f(x)\underbrace{(tu^\prime)(tu^\prime)\dots (tu^\prime)}_{i-\text{times}}]+o(t^m), \] where $o(h)$ is a function such that $\lim_{h\to 0}\, o(h)/h=0$. Then we conclude from Definition \ref{def3} that \[ \ld m f(x;\nabla f(x),\nabla^2 f(x),\dots,\nabla^{m-1}f(x);u) \] \[ =\liminf_{t\downarrow 0,u^\prime\to u}\,[\nabla^m f(x)\underbrace{(u^\prime)(u^\prime)\dots (u^\prime)}_{m-\text{times}}+o(t^m)/t^m]=\nabla^m f(x)\underbrace{(u)(u)\dots (u)}_{m-\text{times}}. \] By Definition \ref{def4} we obtain that Inclusions (\ref{28}) are satisfied. \end{proof} \section{Conditions for a local minimum} \label{s3} \setcounter{equation}{0} \begin{theorem}\label{th1} Let $\bar x\in\dom f $ be a local minimizer of the proper extended real function $f : \E\to\R\cup\{+\infty\}$. Then \begin{equation}\label{14} 0\in\lsubd 1 f(\bar x),\quad 0\in\lsubd n f(\bar x;\underbrace{0,0,\dots ,0)}_{(n-1)-times}\quad\text{for all}\quad n=2,3,4,\dots \end{equation} \end{theorem} \begin{proof} Since $\bar x$ is a local minimizer, then there exists a neighbourhood $N\ni\bar x$ with $f(x)\ge f(\bar x)$ for all $x\in N$. Let $u\in\mathbb E$ be an arbitrary chosen direction. Then $f(\bar x+t u^\prime)\ge f(\bar x)$ for all sufficiently small positive numbers $t$ and for all directions $u^\prime$, which are sufficiently close to $u$. It follows from Definition \ref{def1} that $\ld 1 f (\bar x;u)\ge 0$. Therefore $0\in\lsubd 1 f(\bar x)$, because $u\in\mathbb E$ is an arbitrary direction. By the definition of the second-order lower derivative, using that $0\in\lsubd 1 f(\bar x)$ we obtain that $\ld 2 f(\bar x;u;0)$ is well defined and \[ \ld 2 f (\bar x;0;u)=\liminf_{t\downarrow 0,u^\prime\to u,}\, 2\, t^{-2} [f(\bar x+t u^\prime)-f(\bar x)]\ge 0 \] for all directions $u\in\mathbb E$. Therefore $0\in\lsubd 2 f(\bar x;0)$. Let $n$ be an arbitrary positive integer and \[ 0\in\lsubd i f(\bar x;\underbrace{0,0,\dots ,0)}_{(i-2)-times},\quad i=1,2,\dots,n-1 \] Hence $\ld n f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)$ has sense and \[ \ld n f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)=\liminf_{t\downarrow 0,u^\prime\to u,}\, n!\, t^{-n} [f(\bar x+t u^\prime)-f(\bar x)]\ge 0,\quad\forall u\in\mathbb E. \] It follows from the definition of the lower subdifferential of order $n$ that $0\in\lsubd n f(\bar x;\underbrace{0,0,\dots ,0)}_{(n-1)-times}$. \end{proof} \begin{remark} Condition (\ref{14}) is equivalent to the following one: \begin{equation}\label{15} \ld n f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)\ge 0,\quad\textrm{ for all }u\in\mathbb E,\;\textrm{ for all positive integers } n \end{equation} \end{remark} \begin{corollary} If $\bar x$ is a local minimizer, then for every positive integer $n$ there exist $x_1^*$, $x_2^*$,\dots,$x_{n-1}^*$, which do not depend on $u$ such that \begin{equation}\label{30} x_1^*\in\lsubd 1 f(\bar x),\; x_i^*\in\lsubd i f(\bar x;x_1^*,x_2^*,\dots, x_{i-1}^*),\; i=2,3,\dots, n \end{equation} and \begin{equation}\label{26} \ld 1 f(\bar x;u)\ge 0,\quad \ld i f(\bar x;x^*_1,x^*_2,\dots,x^*_{i-1};u)\ge 0,\; i=2,3,\dots,n. \end{equation} \end{corollary} \begin{proof} We choose $x_1^*=0$, $x_2^*=0$,\dots,$x_{n-1}^*=0$. \end{proof} The following example shows that Condition (\ref{15}) is not sufficient for $\bar x$ to be a local minimizer: \begin{example}\label{ex2} Consider the function of one variable $f:\mathbb R\to\mathbb R$ defined by: \[ f(x)=\left\{ \begin{array}{ll} -\exp\,(-1/x^2), & \textrm{if}\quad x\ne 0, \\ 0, & \textrm{if}\quad x=0. \end{array}\right. \] Let us take $\bar x=0$. Then we have \[ \ld n f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)=0\quad\textrm{ for all }u\in\mathbb R,\;\textrm{ for all positive integers } n, \] \[ \lsubd n f(\bar x;\underbrace{0,0,\dots ,0)}_{(n-1)-times}=\{0\}\;\textrm{if }n\textrm{ is odd,}\quad \lsubd n f(\bar x;\underbrace{0,0,\dots ,0)}_{(n-1)-times}=(-\infty,0]\;\textrm{if }n\textrm{ is even.} \] Hence Condition {\rm (\ref{15})} is satisfied, but $\bar x$ is not a local minimizer. Really, it is a global maximizer. \end{example} On the other hand the following sufficient conditions hold: \begin{theorem}\label{th4} Let be given a proper extended real function $f : \E\to\R\cup\{+\infty\}$ and a point $\bar x\in\dom f$. Suppose that for every direction $u\in\mathbb E$, $u\ne 0$ we have $\ld 1 f(\bar x;u)>0$, or there exists a positive integer $n=n(u)$, $n\ge 2$, which depend on $u$, and such that the following conditions hold: \begin{equation}\label{4} 0\in\lsubd 1 f(\bar x),\quad 0\in\lsubd i f(\bar x;\underbrace{0,0,\dots ,0)}_{(i-1)-times}\quad\text{for all}\quad i=1,2,\dots, n-1 \end{equation} and \begin{equation}\label{5} \ld {n(u)} f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)>0. \end{equation} Then $\bar x$ is a strict local minimizer. \end{theorem} \begin{proof} Let $u\ne 0$ be an arbitrary direction. It follows from (\ref{4}) and (\ref{5}) that there exists $\alpha>0$ with \[ \liminf_{t\downarrow 0,u^\prime\to u}\,n!\,t^{-n}[f(\bar x+t u^\prime)-f(\bar x)]>2\alpha>0. \] Therefore, there exist $\delta>0$ and $\varepsilon>0$ such that \begin{equation}\label{29} f(\bar x+t u^\prime)\ge f(\bar x)+\alpha\, t^n/n!>f(\bar x) \end{equation} for every $t\in (0,\delta)$ and arbitrary $u^\prime$ with $\norm{u^\prime-u}<\varepsilon$. Without loss of generality we may suppose that $u$ belongs to the unit sphere $S:=\{u\in\mathbb E\mid\norm{u}=1\}$. Since $u$ is arbitrary chosen, then we can cover $S$ by neighbourhoods $N(u;\varepsilon):=\{u^\prime\in S\mid\norm{u^\prime-u}<\varepsilon\}$ such that (\ref{29}) is satisfied. Taking into account that the unit sphere is compact, then we can choose a finite number of neighbourhoods $N(u_1;\varepsilon_1)$, $N(u_2,\varepsilon_2)$,\dots $N(u_s;\varepsilon_s)$ that cover $S$. Let the respective values of $\delta$ are $\delta_1$, $\delta_2$,\dots $\delta_s$ and $\bar \delta=\min\{\delta_i\mid 1\le i\le s\}$. Then we have \[ f(\bar x+t u^\prime)>f(\bar x),\quad\forall u^\prime\in S,\;\forall t\in(0,\bar\delta). \] Hence, $f(x)>f(\bar x)$ for all $x\in\mathbb E$ such that $\norm{x-\bar x}<\bar\delta$, which implies that $\bar x$ is a strict local minimizer. \end{proof} \section{Isolated minimizers and optimality conditions} \label{s4} \setcounter{equation}{0} The following definition was introduced by Studniarski \cite{stu86} as a generalization of the respective notion of order $1$ and $2$ in \cite{aus84}. \begin{definition} Let $n$ be a positive integer. A point $\bar x\in\dom f$ is called an isolated local minimizer of order $n$ for the function $f : \E\to\R\cup\{+\infty\}$ iff there exist a neighbourhood $N$ of $\bar x$ and a constant $C>0$ with \begin{equation}\label{1} f(x)\ge f(\bar x)+C\norm{x-\bar x}^n,\quad\forall x\in N. \end{equation} \end{definition} \begin{theorem}\label{th2} Let be given a proper extended real function $f : \E\to\R\cup\{+\infty\}$ and $\bar x\in\dom f$. Then $\bar x$ is an isolated local minimizer of order $n$, where $n$ is a positive integer such that $n\ge 2$, if and only if (\ref{4}) is satisfied and \begin{equation}\label{35} \ld {n} f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)>0,\quad\forall\; u\in\mathbb E\setminus\{0\}. \end{equation} \end{theorem} \begin{proof} Let $\bar x$ be an isolated local minimizer of order $n$. We prove that Conditions (\ref{4}) and (\ref{35}) hold. Suppose that $u\in\mathbb E$ is arbitrary chosen. It follows from Inequality (\ref{1}) that there exist numbers $\delta>0$, $\varepsilon>0$ and $C>0$ with \begin{equation}\label{3} f(\bar x+tu^\prime)\ge f(\bar x)+Ct^n\norm{u^\prime}^n \end{equation} for all $t\in (0,\delta)$ and every $u^\prime$ such that $\norm{u^\prime-u}<\varepsilon$. Therefore \begin{equation}\label{2} \ld 1 f (\bar x;u)=\liminf_{t\downarrow 0,u^\prime\to u,}\,t^{-1}[f(\bar x+t u^\prime)-f(\bar x)]\ge\liminf_{t\downarrow 0,u^\prime\to u,}\, Ct^{n-1}\norm{u^\prime}^{n}=0 \end{equation} if $m>1$, and $0\in\ld 1 f(\bar x)$ if $m=1$. Therefore $0\in\lsubd 1 f(\bar x)$. Suppose that $m$ is an arbitrary positive integer such that $1\le m<n$ and we have \[ 0\in\lsubd i f(\bar x;\underbrace{0,0,\dots ,0)}_{(i-1)-times}\quad\text{for all}\quad i<m. \] It follows from (\ref{3}) that \[ \ld m f(\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};u)=\liminf_{t\downarrow 0,u^\prime\to u,}\,m!\, t^{-m}[f(\bar x+t u^\prime)-f(\bar x)] \ge\liminf_{t\downarrow 0,u^\prime\to u,}\, Ct^{n-m}\norm{u^\prime}^{n}\ge 0,\;\forall u\in\mathbb E. \] and $0\in\lsubd m f(\bar x;\underbrace{0,0,\dots ,0)}_{(m-1)-times}$. Then it follows from (\ref{3}) that \[ \ld n f(\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};u)\ge\liminf_{t\downarrow 0,u^\prime\to u,}\, C\norm{u^\prime}^{n} >0,\quad\forall u\in\mathbb E\setminus\{0\}. \] Conversely, suppose that Conditions {\rm (\ref{4})} and {\rm (\ref{35})} hold. We prove that $\bar x$ is an isolated local minimizer of order $n$. Assume the contrary that $\bar x$ is not an isolated minimizer of order $n$. Therefore, for every sequence $\{\varepsilon_k\}_{k=1}^\infty$ of positive numbers converging to zero, there exists a sequence $\{x_k\}$ with $x_k\in\dom f$ such that \begin{equation}\label{6} \norm{x_k-\bar x}\le \varepsilon_k,\quad f(x_k)< f(\bar x)+\varepsilon_k\norm{x_k-\bar x}^n, \end{equation} It follows from (\ref{6}) that $x_k\to\bar x$. Denote $t_k=\norm{x_k-\bar x}$, $d_k=(x_k-\bar x)/t_k$. Passing to a subsequence, we may suppose that $d_k\to d$ where $\norm{d}=1$. It follows from here that \[ \ld 1 f (\bar x;d)\le\liminf_{k\to\infty}\, t^{-1}_k[f(\bar x+t_k d_k)-f(\bar x)] =\liminf_{k\to\infty}\, t^{-1}_k[f(x_k)-f(\bar x)] \le\liminf_{k\to\infty}\,\varepsilon_k t_k^{n-1}=0. \] It follows from $0\in\subd f(\bar x)$ that $\ld 1 f (\bar x;d)=0$. Let $m$ be any integer with $1<m\le n$ such that $\ld i f(\bar x;\underbrace{0,0,\dots ,0}_{(i-1)-times};d)=0$ for $i<m$. Therefore \begin{equation}\label{16} \ld {m} f(\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};d)=\liminf_{t\downarrow 0,d^\prime\to d,}\,m!\,t^{-m}[f(\bar x+t d^\prime)-f(\bar x)] \le\liminf_{k\to\infty}\, m!\,\varepsilon_k\, t_k^{n-m}=0, \end{equation} because $n-m\ge 0$ and $\varepsilon_k\to 0$. Then it follows from (\ref{4}) that \[ \ld m f(\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};d)=0\quad {\rm if}\quad m<n. \] We conclude from the case $m=n$ that Inequality (\ref{16}) contradicts Condition (\ref{35}). \end{proof} \begin{example} Consider the function $f:\mathbb R^2\to\mathbb R$ defined by \[ f(x_1,x_2)=\left\{ \begin{array}{ll} \exp\,(-1/(x_1^2+x_2^2)), & \textrm{if}\quad (x_1,x_2)\ne (0,0), \\ 0, & \textrm{if}\quad (x_1,x_2)=(0,0). \end{array}\right. \] The point $\bar x=(0,0)$ is a strict global minimizer, but there is no a positive integer $n$ such that $\bar x$ is an isolated minimizer of order $n$. \end{example} \section{Global optimality conditions with a higher-order invex function} \label{s5} \setcounter{equation}{0} Example \ref{ex2} shows that the necessary conditions for a local minimum are not sufficient for a global one. Then the following question arises: Which is the largest class of functions such that the necessary optimality conditions from Theorem \ref{th1} become sufficient for a global minimum. Recently, Ivanov \cite{optimization} introduced a new class of Fr\'echet differentiable functions called second-order invex ones in terms of the usual second-order directional derivative. They extend the so called invex ones and obey the following property: A Fr\'echet differentiable function is second-order invex if and only if each second-order stationary point is a global minimizer. We extend the notions invexity and second-order invexity to nondifferentiable functions in terms of the lower Hadamard directional derivatives of order $n$. Some more developments to inequality constrained problems in terms of the usual second-order directional derivative are recently obtained by Ivanov \cite{jogo2011}. First, we recall the definition of an invex function \cite{han81} in terms of the lower Hadamard directional derivative. \begin{definition} A proper extended real function $f : \E\to\R\cup\{+\infty\}$ is called invex in terms of the lower Hadamard directional derivative iff there exists a map $\eta_1: \mathbb E\times \mathbb E\to\mathbb E$ such that the following inequality holds for all $x\in\mathbb E $, $y\in\mathbb E$: \begin{equation}\label{13} f(y)-f(x)\ge\ld 1 f(x;\eta_1(x,y)). \end{equation} \end{definition} We introduce the following two definitions: \begin{definition} We call a proper extended function $f : \E\to\R\cup\{+\infty\}$ invex of order $n$ in terms of the lower Hadamard derivatives iff for every $\bar x\in\dom f$, $x\in\mathbb E$ such that there exist at least one $(i-1)$-ple $(x_1^*,x_2^*,\dots,x_{i-1}^*)$ with \begin{equation}\label{7} x_1^*\in\lsubd 1 f(\bar x),\; x_i^*\in\lsubd i f(\bar x;x_1^*,x_2^*,\dots, x_{i-1}^*),\; i=2,3,\dots, n \end{equation} there are $\eta_1$, $\eta_2,\dots$, $\eta_n$, which depend on $\bar x$ and $x$ such that the following inequality holds \begin{equation}\label{8} f(x)-f(\bar x)\ge\ld 1 f(\bar x;\eta_1(\bar x,x))+\sum_{i=2}^n\ld i f(\bar x; x^*_1,x_2^*,\dots,x_{i-1}^* ;\eta_i(\bar x,x)) \end{equation} for all $x_i^*$, $i=1,2,\dots,n$ satisfying Conditions (\ref{7}). If there exist $\eta_1(\bar x,x)$, $\eta_2(\bar x,x)$, $\eta_3(\bar x,x),\dots$ such that (\ref{7}) and (\ref{8}) are satisfied with $n=+\infty$, then we call $f$ invex in generalized sense (or invex of order $+\infty$). \end{definition} \begin{definition} Let $f : \E\to\R\cup\{+\infty\}$ be a given proper extended real function and $n$ be a positive integer. We call a stationary point of order $n$ every point $\bar x\in\dom f$ which satisfies the necessary optimality conditions (\ref{30}) and (\ref{26}). The notion of a 1-stationary point coincides with the notion of a stationary point. If (\ref{30}) and (\ref{26}) are satisfied for every $n=1,2,3,\dots$, then we call $\bar x$ stationary point in generalized sense (or order $+\infty)$. \end{definition} \begin{theorem}\label{npi} Let $n$ be a positive integer or $+\infty$ and $f : \E\to\R\cup\{+\infty\}$ a proper extended function. Then $f$ is invex of order $n$ if and only if each stationary point $\bar x\in\dom f$ of order $n$ $(n<+\infty$ or $n=+\infty)$ is a global minimizer of $f$. \end{theorem} \begin{proof} We prove the case $n<+\infty$. The other case is similar. Suppose that $f$ is invex of order $n$. If the function has no stationary points, then obviously every stationary point is a global minimizer. Suppose that the function has at least one stationary point, that is a point satisfying the necessary optimality conditions (\ref{30}) and (\ref{26}). Suppose that $\bar x\in\dom f $ is a given stationary point of order $n$. We prove that it is a global minimizer of $f$. Suppose that $x$ is an arbitrary point from $\mathbb E$. It follows from invexity of order $n$ that there exist $\eta_i(\bar x,x)$, $i=1,2,\dots, n$ such that \begin{equation}\label{9} f(x)-f(\bar x)\ge\ld 1 f(\bar x;\eta(\bar x,x))+\sum_{i=2}^n\ld i f(\bar x; x^*_1,x_2^*,\dots,x_{i-1}^*;\eta_i(\bar x,x) ) \end{equation} for all $x_i^*$, $i=1,2,\dots,n-1$ Since $\bar x$ is a stationary point of order $n$, then there exist $x_1^*$, $x_2^*$,\dots, $x_{n-1}^*$ with \[ \ld 1 f(\bar x;u)\ge 0,\;\ld i f(\bar x; x_1^*,x_2^*,\dots ,x_{i-1}^*;u)\ge 0\quad\text{for all}\quad i=2,3,\dots,n,\;\forall u\in\mathbb E. \] Hence \[ \ld 1 f(\bar x;\eta_1(\bar x,x))\ge 0,\;\ld i f(\bar x; x_1^*,x_2^*,\dots ,x_{i-1}^*;\eta_i(\bar x,x) )\ge 0\quad\text{for all}\quad i=2,3,\dots,n. \] It follows from (\ref{9}) that $f(x)\ge f(\bar x)$. Therefore $\bar x$ is a global minimizer. Conversely, suppose that every stationary point of order $n$ is a global minimizer. We prove that $f$ is invex of order $n$. Assume the contrary. Hence, there exists a pair $(\bar x,x)\in\dom f\times \mathbb E$ such that for every $\eta_i\in\mathbb E$, $i=1,2,\dots,n$ there are $x_1^*$, $x_2^*$ , $\dots$, $x_{n-1}^*$ satisfying Conditions (\ref{7}) and \begin{equation}\label{10} f(x)-f(\bar x)<\ld 1 f(\bar x;\eta_1)+\sum_{i=2}^n\ld i f(\bar x; x^*_1,x_2^*,\dots,x_{i-1}^*;\eta_i ). \end{equation} First, we prove that $f(x)<f(\bar x)$. Let us choose in (\ref{10}) $\eta_i=0$, $i=1,2,\dots,n$. We have \[ \ld 1 f(\bar x;0)\le\liminf_{t\downarrow 0}\,t^{-1}(f(\bar x+t.0)-f(\bar x))=0. \] Let $i$ be an arbitrary integer such that $1<i\le n$. Then \[ \ld i f(\bar x;x^*_1,x^*_2,\dots,x^*_{i-1};0)\le \liminf_{t\downarrow 0,u^\prime\to 0}\,\frac{i!}{t^i}\,[f(\bar x+t.u^\prime)-f(\bar x)-\sum_{k=1}^{i-1}\frac{t^k}{k!}x^*_k\underbrace{(u^\prime)(u^\prime)\dots (u^\prime)}_{k-\text{times}}] \] \[ \le\liminf_{t\downarrow 0}\,\frac{i!}{t^i}\,[f(\bar x+t.0)-f(\bar x)- \sum_{k=1}^{i-1}\frac{t^k}{k!}x^*_k\underbrace{(0)(0)\dots (0)}_{k-\text{times}}]=0,\quad i=2,3,\dots, n. \] It follows from (\ref{10}) that $f(x)<f(\bar x)$. Second, we prove that \begin{equation}\label{11} \ld 1 f(\bar x;u)\ge 0,\quad\forall u\in\mathbb E. \end{equation} Suppose the contrary that there exists at least one point $v\in\mathbb E$ with $\ld 1 f(x;v)<0$. The lower Hadamard directional derivative is positively homogeneous with respect to the direction, that is \[ \ld 1 f(\bar x;\tau u)=\tau\ld 1 f(\bar x;u),\quad\forall \bar x\in\dom f,\;\forall u\in\mathbb E,\;\forall \tau\in(0,+\infty). \] Then inequality (\ref{10}) is satisfied when $\eta_1=tv$, $t>0$, $\eta_i=0$, $i\ne 1$, that is \[ f(x)-f(\bar x)<t\ld 1 f(\bar x;v),\quad\forall t>0, \] which is impossible, because $f(x)-f(\bar x)$ is finite and $\ld 1 f(x;v)<0$. Therefore, $\ld 1 f(\bar x;u)\ge 0,\quad\forall u\in\mathbb E$. Third, we prove that for all $u\in\mathbb E$ there are $x_1^*$, $x_2^*$,\dots,$x_{n-1}^*$ with \begin{equation}\label{12} \ld i f(\bar x;x^*_1,x^*_2,\dots,x^*_{i-1};u)\ge 0\quad\textrm{ for } i=2,3,\dots, n. \end{equation} Suppose the contrary that there exists $v\in\mathbb E$ with $\ld i f(\bar x;x^*_1,x^*_2,\dots,x^*_{i-1};v)<0$ for all $x^*_1$, $x^*_2$,...,$x^*_{i-1}$ satisfying Conditions (\ref{7}). The lower Hadamard directional derivative of order $i$ is positively homogeneous of degree $i$ with respect to the direction, that is \[ \ld i f(\bar x;x_1^*,x_2^*,\dots,x_{i-1}^*;tv)=t^i\ld i f(\bar x;x_1^*,x_2^*,\dots,x_{i-1}^*;v),\quad\forall t>0. \] Then it follows from (\ref{10}) with $\eta_i=tv$, $t>0$, $\eta_k=0$ when $k\ne i$ that \[ f(x)-f(\bar x)<t^i\ld i f(\bar x;x_1^*,x_2^*,\dots,x_{i-1}^*;v ),\quad\forall t>0, \] which is impossible when $t$ is sufficiently large positive number. The following is the last part of the proof. It follows from (\ref{11}) and (\ref{12}) that $\bar x$ is a stationary point of order $n$. According to the hypothesis $\bar x$ is a global minimizer, which contradicts the inequality $f(x)<f(\bar x)$. \end{proof} In the next claim we show that the class of invex functions of order $(n+1)$ contains all invex functions of order $n$ in terms of the lower Hadamard directional derivative. \begin{proposition} Let $f : \E\to\R\cup\{+\infty\}$ be an invex function of order $n$. Then $f$ is invex of order $(n+1)$. Every invex function of order $n$ is invex of order $+\infty$. \end{proposition} \begin{proof} It follows from Equation (\ref{8}) that $f$ is invex of order $(n+1)$ keeping the same maps $\eta_1$, $\eta_2$,..., $\eta_n$ and taking $\eta_{n+1}=0$, because $\ld {n+1} f(\bar x;x^*_1,x^*_2,\dots,x^*_n;0)\le 0$ for all elements of the respective subdifferentials $x^*_1$, $x^*_2$,...,$x^*_n$. \end{proof} The converse claim is not satisfied. There are a lot of second-order invex functions, which are not invex. The following example is extremely simple. \begin{example} Consider the function $f:\mathbb R^2\to\mathbb R$ defined by \[ f(x_1,x_2)=-x_1^2-x_2^2. \] We have $\ld 1 f(x;u)=-2x_1 u_1-2x_2 u_2$ where $u=(u_1,u_2)$ is a direction. Its only stationary point is $\bar x=(0,0)$. This point is not a global minimizer. Therefore, the function is not invex. We have $\lsubd 1 f(\bar x)\equiv \{(0,0)\}$ and $\ld 2 f(\bar x;0;u)=-2u_1^2-2u_2^2$. It follows from here that $f$ has no second-order stationary points. Hence, every second-order stationary point is a global minimizer, and the function is second-order invex. \end{example} \begin{example}\label{npc} Consider the function $f_n:\mathbb R\to\mathbb R$, where $n\ge 2$ is a positive integer: \[ f_n(x)=\left\{ \begin{array}{rr} x^n\,, & x\ge 0\,, \\ (-1)^{n-1}x^n\,, & x<0\,. \end{array}\right. \] If $n$ is an odd number, then $f_n=x^n$. For $n$ even $f_n$ is a function from the class C$^{n-1}$ but not from the class C$^n$. It has no stationary points of order $n$. Therefore, every stationary point of order $n$ is a global minimizer. According to Theorem \ref{npi}, it is invex of order $n$. On the other hand, the point $x=0$ is stationary of order $(n-1)$. Taking into account that $x=0$ is not a global minimizer, we conclude from the same theorem that the function is not invex of order $(n-1)$. \end{example} \section{Second-order conditions for a special class of functions in terms of the lower Dini derivatives} \label{s6} \setcounter{equation}{0} In this section, we derive optimality conditions for an isolated minimum of order two for a special class of functions. Strongly pseudoconvex functions were introduced by Diewert, Avriel and Zang \cite{d-a-z81}. Their definition assumes additionally strict pseudoconvexity. It was proved by Hadjisavvas and Schaible \cite{had93} that in the differentiable case, strict pseudoconvexity of the function is superfluous; in other words each function, which satisfies the next definition is strictly pseudoconvex. \begin{definition}[\cite{had93}]\label{def2} Let $S$ be an open convex subset of $\mathbb E$. A Fr\'echet differentiable function $f:S\to\mathbb R$ is said to be strongly pseudoconvex iff, for all $x\in S$, $u\in\mathbb E$ such that $\norm{u}=1$ and $\nabla f(x)(u)=0$, there exist positive numbers $\delta$ and $\alpha$ with $x+\delta u\in S$ and \[ f(x+t u)\ge f(x)+\alpha t^2,\quad 0\le t<\delta. \] \end{definition} Recall the definition of a strongly pseudoconvex function in terms of lower Dini directional derivative, which were introduced by Biancki \cite{bia96}. \begin{definition} The lower Dini directional derivative of a function $f : \E\to\R\cup\{+\infty\}$ at the point $x\in\dom f$ in direction $u\in\mathbb E$ is defined as follows: \[ \dini f(x;u)=\liminf_{t\downarrow 0}\, t^{-1}[f(x+tu)-f(x)]. \] \end{definition} We adopt the next two definitions to proper extended real functions: \begin{definition}[\cite{bia96}]\label{DefBia96} A proper extended function $f : \E\to\R\cup\{+\infty\}$ is said to be strongly pseudoconvex iff, \begin{enumerate} \item[{\rm (i)}] $f$ is strictly pseudoconvex, that is for all $x\in\dom f $, $u\in\mathbb E$ such that $\norm{u}=1$ the condition \[ \dini f(x;u)\ge 0\quad {\rm implies}\quad f(x+tu)>f(x),\;\forall\; t>0; \] \item[{\rm (ii)}] if $\dini f(x;u)=0$, then there exist positive numbers $\delta$ and $\alpha$ with \[ f(x+t u)\ge f(x)+\alpha t^2,\quad\forall t\in[0,\delta). \] \end{enumerate} \end{definition} It was shown by example in \cite{bia96} that Condition (i) cannot be omitted in Definition \ref{DefBia96}, that is Condition (ii) does not imply Condition (i). The following notion extends the Lipschitz continuity of the gradient, which were applied in the sufficient conditions due to Ben-Tal and Zowe \cite{bt85}: \begin{definition}[\cite{pas08}] A proper extended function $f : \E\to\R\cup\{+\infty\}$ is called $\ell$-stable at the point $x\in\dom f$ iff there exist a neigbourhood $U$ of $x$ and a constant $K>0$ such that \[ |\dini f(y;u)-\dini f(x;u)|\le K\,\norm{y-x}\,\norm{u},\quad\forall y\in U\cap\dom f,\;\forall u\in\mathbb E. \] \end{definition} \begin{proposition}[\cite{pas08}]\label{pr1} Let the proper extended function $f : \E\to\R\cup\{+\infty\}$ be continuous near $x\in\dom f$ and $\ell$-stable at $x$. Then $f$ is strictly differentiable at $x$, hence Fr\'echet differentiable at $x$. \end{proposition} The following mean-value theorem is due to Diewert \cite{die81}. \begin{lemma}\label{mv} Let $\varphi:[a,\,b]\to\mathbb R\cup\{+\infty\}$ be a lower semicontinuous function of one real variable with $a\in\dom \varphi$. Then there exists $\xi$, $a<\xi\le b$, such that \[ \varphi(a)-\varphi(b)\le \dini \varphi(\xi; a-b). \] \end{lemma} In the next theorem, we derive necessary and sufficient conditions for an isolated local minimum of second-order of a function, which satisfies Condition (ii) from Definition \ref{DefBia96} at the same given point $\bar x$: \begin{theorem}\label{th3} Let the proper extended function $f : \E\to\R\cup\{+\infty\}$ be continuous near $\bar x\in\dom f$ and $\ell$-stable at $\bar x$. Suppose that $f$ satisfies Condition {\rm (ii)} from Definition \ref{DefBia96} only at the point $\bar x$. Then $\bar x$ is an isolated local minimizer of second-order if and only if $\nabla f(\bar x)=0$. \end{theorem} \begin{proof} Let $\bar x$ be an isolated local minimizer of second-order. We conclude from Proposition \ref{pr1} that $\nabla f(\bar x)$ exists. Then it is obvious that $\nabla f(\bar x)=0$. We prove the converse claim. Suppose that $\nabla f(\bar x)=0$, but $\bar x$ is not an isolated local minimizer of second-order. Therefore, for every sequence $\{\varepsilon_k\}_{k=1}^\infty$ of positive numbers converging to zero, there exists a sequence $\{x_k\}$, $x_k\in\dom f$ such that \begin{equation}\label{27} \norm{x_k-\bar x}\le \varepsilon_k,\quad f(x_k)< f(\bar x)+\varepsilon_k\norm{x_k-\bar x}^2. \end{equation} It follows from here that $x_k\to \bar x$. Denote $t_k=\norm{x_k-\bar x}$, $d_k=(x_k-\bar x)/t_k$. Passing to a subsequence, we may suppose that the sequence $\{d_k\}_{k=1}^\infty$ is convergent and $d_k\to d$, where $\norm{d}=1$. Therefore \begin{equation}\label{25} \liminf_{k\to\infty}\, t^{-2}_k [f(\bar x+t_k d_k)-f(\bar x)]=\liminf_{k\to \infty}\, t^{-2}_k[f(x_k)-f(\bar x)]\le\lim_{k\to \infty}\, \varepsilon_k=0. \end{equation} We have \[ f(\bar x+t_k d)-f(\bar x)=[f(\bar x+t_k d)-f(\bar x+t_kd_k)]+[f(\bar x+t_kd_k)-f(\bar x)], \] since $f(x_k)$ is finite by (\ref{27}). It follows from Diewert's mean-value theorem that there exists $\theta_k\in[0,1)$ with \[ f(\bar x+t_k d)-f(\bar x+t_k d_k)\le D_{-} f(y_k;t_k(d-d_k))=t_k\, \dini f(y_k;d-d_k) \] where $y_k=\bar x+t_k d_k+t_k\theta_k(d-d_k)$. Since $\nabla f(\bar x)=0$ and $f$ is $\ell$-stable at $\bar x$, then there exists $K>0$ such that \[ |\dini f(y_k;d-d_k)|\le K\,\norm{y_k-\bar x}\,\norm{d-d_k} \] for all sufficiently large integers $k$. Taking into account that $d_k\to d$ when $k\to\infty$, we conclude that \[ \liminf_{k\to\infty}\, t^{-2}_k [f(\bar x+t_k d)-f(\bar x)]\le 0. \] On the other hand, according to the hypothesis $f$ satisfies Condition (ii) from Definition \ref{DefBia96}. Therefore \[ f(\bar x+t_k d)\ge f(\bar x)+\alpha t_k^2 \] for all sufficiently large $k$. Hence, \[ \liminf_{k\to\infty}\, t^{-2}_k [f(\bar x+t_k d)-f(\bar x)]\ge\liminf_{k\to\infty}\,\alpha=\alpha>0, \] which is a contradiction. \end{proof} The following example shows that Theorem \ref{th3} is not true for functions, which are not $\ell$-stable: \begin{example}\label{ex1} Consider the function \[ f=|\,x_2-\sqrt[3]{x_1^4}\,|^{\, 3/2}. \] Of course, the point $\bar x=(0,0)$ is a local and global minimizer, but it is not an isolated local minimizer of order two. Even it is not a strict local minimizer, because $f(x)=0$ for all $x=(x_1,x_2)$ over the curve $x_2=x_1^{4/3}$. We have $\nabla f(\bar x)=(0,0)$. Simple calculations show that this function satisfy Condition {\rm (ii)} from Definition \ref{DefBia96} at $\bar x$. Let $v=(v_1,v_2)$ be an arbitrary vector, whose norm is $1$. If $v_2>0$ or $v_2<0$, then \[ \lim_{t\downarrow 0}\,[f(\bar x+tv)-f(\bar x)]/t^2=\lim_{t\downarrow 0}\, f(tv)/t^2=+\infty. \] If $v_2=0$, then $v_1=\pm 1$ and \[ \lim_{t\downarrow 0}\,[f(\bar x+tv)-f(\bar x)]/t^2=\lim_{t\downarrow 0}\, f(tv)/t^2=1. \] Therefore, for every $v\in\mathbb R^2$ there exists $\delta>0$ and $C>0$ such that \[ f(tv)>Ct^2 \quad\textrm{for all}\quad t\in(0,\delta). \] The sufficient conditions of Theorem \ref{th3} are not satisfied, because $f$ is not $\ell$-stable at $\bar x$. Indeed, if we take $x^\prime_k=(0,k^{-1})$, then $\nabla f(x_k)=(0,3/2k^{-1/2})$ and there do not exist $K>0$ such that \[ \norm{\nabla f(x^\prime_k)-\nabla f(\bar x)}\le K\norm{x^{\prime}_k-\bar x} \] for all sufficiently large integers $k$. We have \[ \lim_{k\to+\infty}\norm{\nabla f(x^{\prime}_k)}\, / \, \norm{x^{\prime}_k}=+\infty. \] \end{example} \section{Higher-order conditions for a special class of functions in terms of the lower Hadamard derivatives} \label{s7} \setcounter{equation}{0} We introduce the following notion: \begin{definition} We say that a proper extended real function $f : \E\to\R\cup\{+\infty\}$ belongs to the class $\mathcal F_2(x)$ iff, $x\in\dom f$ and for every $u\in\mathbb E$ such that $\norm{u}=1$, $\ld 1 f(x;u)=0$, there exist positive numbers $\varepsilon$, $\delta$, and $\alpha$ which satisfy the inequality \begin{equation}\label{19} f(x+tu^\prime)\ge f(x)+\alpha t^2, \end{equation} for all $t\in\mathbb R$ and $u^\prime\in\mathbb E$ with $0\le t<\delta$, $\norm{u^\prime-u}<\varepsilon$. \end{definition} The class containing all functions $f$ such that $f\in\mathcal F_2(x)$ for every $x\in\mathbb E$, and which are additionally strictly pseudoconvex coincides with the set of all strongly pseudoconvex functions with respect to the lower Hadamard directional derivative. \begin{definition} Let $x\in\mathbb E$. For any positive integer $n\ge 3$, we say that a proper extended real function $f : \E\to\R\cup\{+\infty\}$ belongs to the class $\mathcal F_n(x)$ iff, $x\in\dom f$ and for every $u\in\mathbb E$ such that $\norm{u}=1$, \begin{equation}\label{22} 0\in\lsubd 1 f(x),\quad 0\in\lsubd i f (x;x_1^*,x_2^*,\dots,x_{i-1}^*),\; i=2,3,\dots, n-2, \end{equation} \begin{equation}\label{23} \ld 1 f(x;u)=0,\quad \ld i f(x;\underbrace{0,0,\dots ,0}_{(i-1)-times};u)=0,\; i=2,3,\dots,n-1, \end{equation} there exist positive numbers $\varepsilon$, $\delta$, and $\alpha$ which satisfy the inequality \begin{equation}\label{24} f(x+tu^\prime)\ge f(x)+\alpha t^n, \end{equation} for all $t\in\mathbb R$ and $u^\prime\in\mathbb E$ with $0\le t<\delta$, $\norm{u^\prime-u}<\varepsilon$. \end{definition} \begin{theorem}\label{th5} Let $n\ge 2$ be a positive integer, $\bar x\in\mathbb E$, and $f\in\mathcal F_n(\bar x)$. Then $\bar x$ is an isolated local minimizer of order $n$ if and only if \begin{equation}\label{20} 0\in\lsubd 1 f(\bar x),\quad 0\in\lsubd i f (\bar x;\underbrace{0,0,\dots ,0}_{(i-1)-times}),\; i=2,3,\dots, n-1. \end{equation} \end{theorem} \begin{proof} It follows from Theorem \ref{th1} that the condition $\bar x$ is an isolated local minimizer of order $n$ implies that Inclusions (\ref{20}) are satisfied. Conversely, suppose that Inclusions (\ref{20}) are satisfied. We prove that $\bar x$ is an isolated local minimizer of order $n$. Assume the contrary. Therefore, for every sequence $\{\varepsilon_k\}_{k=1}^\infty$ of positive numbers converging to zero, there exists a sequence $\{x_k\}$, $x_k\in\dom f$ such that \begin{equation}\label{21} \norm{x_k-\bar x}\le \varepsilon_k,\quad f(x_k)< f(\bar x)+\varepsilon_k\norm{x_k-\bar x}^n, \end{equation} It follows from (\ref{21}) that $x_k\to\bar x$. Denote $t_k=\norm{x_k-\bar x}$, $d_k=(x_k-\bar x)/t_k$. Passing to a subsequence, we may suppose without loss of generality that $d_k\to d$, where $\norm{d}=1$. Let $m$ be an arbitrary positive integer with $1\le m\le n$. It follows from (\ref{20}) that \[ \ld m f (\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};d)=\liminf_{k\to\infty}\,m!\, t^{-m}_k[f(\bar x+t_k d_k)-f(\bar x)]\le\liminf_{k\to 0}\, m!\, t^{n-m}\varepsilon_k=0 \] if $m>1$ or $\ld 1 f(\bar x;d)\le 0$ if $m=1$. If $1<m<n$, then we conclude from \[ 0\in\lsubd m f (\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times}) \] that $\ld m f (\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};v)\ge0$ for all $v\in\mathbb E$. Therefore $\ld m f (\bar x;\underbrace{0,0,\dots ,0}_{(m-1)-times};d)=0$. Using similar arguments, we can prove that $\ld 1 f(\bar x;d)=0$. On the other hand, by Inequality (\ref{24}) we obtain that \[ \ld n f (\bar x;\underbrace{0,0,\dots ,0}_{(n-1)-times};d)=\liminf_{t\downarrow 0,d^\prime\to d} n!\,t^{-n} [f(\bar x+t d^\prime)-f(\bar x)] \ge\liminf_ {t\downarrow 0,d^\prime\to d}\,n!\,\alpha=n!\,\alpha>0, \] which is a contradiction. \end{proof} \begin{proposition}\label{pr2} Let $x\in\mathbb E$. Then $\mathcal F_{n-1}(x)\subset\mathcal F_n(x)$ for every positive integer $n$, $n\ge 3$. \end{proposition} \begin{proof} Suppose that there exists a function $f\in\mathcal F_{n-1}(x)$ with $f\notin\mathcal F_n(x)$. Therefore, $f$ satisfies Conditions (\ref{22}) and (\ref{23}), but it does not fulfil (\ref{24}). Then it follows from $f\in\mathcal F_{n-1}(x)$ that there exist positive numbers $\varepsilon$, $\delta$ and $\alpha$ such that the following inequality holds \[ f(x+tu^\prime)\ge f(x)+\alpha t^{n-1}, \] for all $t\in\mathbb R$ and $u^\prime\in\mathbb E$ with $0\le t<\delta$, $\norm{u^\prime-u}<\varepsilon$. Therefore \[ \ld {n-1} f (x;\underbrace{0,0,\dots ,0}_{(n-2)-times};u)=\liminf_{t\downarrow 0,u^\prime\to u}\, (n-1)!\, t^{-(n-1)}\, [f(x+t u^\prime)-f(x)] \] \[ \ge\liminf_ {t\downarrow 0,u^\prime\to u}\,(n-1)!\,\alpha=(n-1)!\,\alpha>0, \] which contradicts the assumption $\ld {n-1} f (x;\underbrace{0,0,\dots ,0}_{(n-2)-times};u)=0$. Hence, in the case when $f\in\mathcal F_{n-1}(x)$, (\ref{22}) and (\ref{23}) cannot be satisfied together. Therefore, the implication $\{(\ref{22}), (\ref{23})\Rightarrow (\ref{24})\}$ is fulfiled. Consequently $f\in\mathcal F_n(x)$. \end{proof} The following example shows that the inclusion in Proposition \ref{pr2} is strict: \begin{example} Consider the function $f:\mathbb R^2\to\mathbb R$ defined by \[ f(x)=|x_1|^n+|x_2|^n. \] We prove that $f\in\mathcal F_n(\bar x)$, where $\bar x=(0,0)$. Let $u=(u_1,u_2)$ be a direction. We have \[ \ld 1 f(\bar x;u)=0\quad {\rm and}\quad (0,0)\in\lsubd 1 f(\bar x). \] Therefore Conditions (\ref{22}) and (\ref{23}) hold. We prove that for every $\delta>0$ and each $\varepsilon\in(0,1)$ there exists $\alpha>0$ which satisfies (\ref{24}); in other words there exists $\alpha>0$ such that $\alpha<|u_1^\prime|^n+|u_2^\prime|^n$, when \[ (u_1^\prime-u_1)^2+(u_2^\prime-u_2)^2<\varepsilon^2\quad{\rm and }\quad u_1^2+u_2^2=1. \] Assume the contrary that such positive number $\alpha$ does not exist. Therefore \[ \inf\,\{|u_1^\prime|^n+|u_2^\prime|^n\mid(u_1^\prime-u_1)^2+(u_2^\prime-u_2)^2<\varepsilon^2\}=0. \] Therefore, there exist infinite sequences of positive numbers $\{y_k\}$ and $\{z_k\}$ converging to 0 such that \[ (y_k-u_1)^2+(z_k-u_2)^2<\varepsilon^2, \] which is impossible, because $\norm{u}=1$. Therefore $f\in\mathcal F_n(\bar x)$. We prove that $f\notin\mathcal F_{n-1}(\bar x)$. Suppose the contrary that $f\in\mathcal F_{n-1}(\bar x)$. Conditions (\ref{22}) and (\ref{23}) are satisfied when $x=(0,0)$. Suppose that (\ref{24}) is also fulfiled. Therefore, there exist $\delta>0$ and $\alpha>0$ such that $f(\bar x+tu)\ge f(\bar x)+\alpha t^{n-1}$, $0<t<\delta$. Hence \[ t(|u_1|^n+|u_2|^n)>\alpha,\quad\forall t\in(0,\delta), \] which is obviously impossible. \end{example} \section{Comparison with some previous results} \label{s8} \setcounter{equation}{0} In several papers Chaney introduced and studied a second-order directional derivative (see, for example, \cite{cha87}) which is called the derivative of Chaney. We recall its definition. It is called that a sequence $\{x_k\}$, $x_k\in\mathbb R^n$, $x_k\ne x$ converges to a point $x\in\mathbb R^n$ in direction $u\in\mathbb R^n$, $u\ne 0$ iff the sequence $\{(x_k-x)/\norm{x_k-x}\}$ converges to $u$. Let $f:\mathbb R^n\to\mathbb R$ be a locally Lipschitz function. Denote its Clarke generalized gradient at the point $x$ by $\partial f(x)$. Suppose that $u$ is a nonzero vector in $\mathbb R^n$. Denote by $\partial_u f(x)$ the set of all vectors $x^*$ such that there exist sequences $\{x_k\}$ and $\{x_k^*\}$ with $x_k^*\in\partial f(x_k)$, $\{x_k\}$ converges to $x$ in direction $u$, and $\{x_k^*\}$ converges to $x^*$. Really $\partial_u f(x)\subset\partial f(x)$. \begin{definition}[\cite{cha87}]\label{def5} Let $f:\mathbb R^n\to\mathbb R$ be a locally Lipschitz function. Suppose that $x\in\mathbb R^n$, $u\in\mathbb R^n$, and $x^*\in\partial_u f(x)$. Then the second-order lower derivative of Chaney $\sld f(x;x^*;u)$ at $(x,x^*)$ in direction $u$ is defined to be the infimum of all numbers \[ \liminf\,[f(x_k)-f(x)-x^*(x_k-x)]/t^2_k, \] taken over all triples of sequences $\{t_k\}$, $\{x_k\}$, and $\{x_k^*\}$ for which \begin{enumerate} \item[{\rm (a)}] $t_k>0$ for each $k$ and $\{x_k\}$ converges to $x$, \item[{\rm (b)}] $\{t_k\}$ converges to $0$ and $\{(x_k-x)/t_k\}$ converges to $u$, \item[{\rm (c)}] $\{x_k^*\}$ converges to $x^*$ with $x_k^*\in\partial f(x_k)$ for each $k$. \end{enumerate} \end{definition} The following claims due to Huang and Ng \cite[Theorems 2.2, 2.7 and 2.9]{hua94} are very important necessary and sufficient conditions for optimality in unconstrained optimization. The necessary conditions are generalizations of the respective results due to Chaney \cite[Theorem 1]{cha87} where the function is semismooth. \begin{proposition}\label{pr4} Let $\bar x$ be a local minimum point of the locally Lipschitz function $f$ and $u\in\mathbb R^n$ with norm $1$ such that $\dini f(\bar x;u)=0$. Then $0\in\partial_u f(\bar x)$, and $\sld f(\bar x;0;u)\ge 0$. \end{proposition} \begin{proposition}\label{pr7} Let $f:\mathbb R^n\to\mathbb R$ be a locally Lipschitz function. Suppose that $\dini f(x;v)\ge 0$, for all $v\in\mathbb R^n$. For $u\in\mathbb R^n$ with norm $1$, if $\dini f(x;u)=0$, then $0\in\partial_u f(x)$. \end{proposition} \begin{proposition}\label{pr5} Let $f:\mathbb R^n\to\mathbb R$ be a locally Lipschitz function. Suppose that \[ \dini f(\bar x;v)\ge 0,\quad\forall v\in\mathbb R^n,\; v\ne 0. \] If $\sld f(\bar x;0;u)>0$ for all unit vectors $u\in\mathbb R^n$ for which $\dini f(\bar x;u)=0$, then $\bar x$ is a strict local minimizer. \end{proposition} \begin{lemma}\label{lema1} Let $f:\mathbb R^n\to\mathbb R$ be a locally Lipschitz function. Suppose that $x\in\mathbb R^n$ and $u\in\mathbb R^n$. If $0\in\partial f_u(x)$ and $0\in\lsubd 1 f(x)$, then $\sld f(x;0;u)=\ld 2 f(x;0;u)$. \end{lemma} \begin{proof} Denote $u_k=(x_k-x)/t_k$. Then \[ \sld f(x;0;u)=\liminf\,[f(x+t_k u_k)-f(x)]/t^2_k, \] where the limes infimum is taken over all pairs of sequences $\{t_k\}$, $\{u_k\}$ which satisfy Conditions (a) and (b) from Definition \ref{def5}. It follows from here that \[ \sld f(x;0;u)=\liminf_{t\downarrow 0,u^\prime\to u}\,[f(x+t u^\prime)-f(x)]/t^2=0.5\,\ld 2 f(x;0;u), \qedhere \] \end{proof} \bigskip {\it Proof of Proposition \ref{pr5} as corollary of Theorem \ref{th4}.} Since $f$ is locally Lipschitz, then $\ld 1 f(\bar x;v)=\dini f(\bar x;v)\ge 0$ for all directions $v\in\mathbb R^n$. Therefore $0\in\lsubd 1 f(\bar x)$. Suppose that $\dini f(\bar x;u)=0$ for some unit direction $u$. It follows from Lemma \ref{lema1} that \[ \ld 2 f(\bar x;0;u)=\sld f(\bar x;0;u)>0, \] because by Proposition \ref{pr7} we have $0\in\partial_u f(\bar x)$. Then, according to Theorem \ref{th4} the point $\bar x$ is a strict local minimizer. \qed \bigskip It is seen that our proof is shorter than the proof of Huang and Ng \cite{hua94}. \bigskip {\it Proof of Proposition \ref{pr4} as corollary of Theorem \ref{th1}.}\; Let $\dini f(\bar x;u)=0$ for some unit direction $u$. By Proposition \ref{pr7}, we have $0\in\partial_u f(\bar x)$. Then, by Lemma \ref{lema1}, $\sld f(\bar x;0;u)=\ld 2 f(\bar x;0;u)$. Thus the claim follows from Theorem \ref{th1}. \qed \bigskip Ben-Tal and Zowe introduced the following second-order derivative of a function $f:\mathbb E\to\mathbb R$ at the point $x\in\mathbb E$ in directions $u\in\mathbb E$ and $z\in\mathbb E$: \[ f^{\prime\pr}_{BZ}(x;u,z):=\lim_{t\downarrow 0}\, t^{-2}[f(x+tu+t^2 z)-f(x)-t f^{\prime}(x;u)], \] where $f^{\prime}(x;u):=\lim_{t\downarrow 0}\, t^{-1}[f(x+tu)-f(x)]$ is the usual directional derivative of first-order. The following conditions are necessary for a local minimum in terms of the derivative of Ben-Tal and Zowe \cite{bt85}: \begin{proposition} Let $\bar x$ be a local minimizer of $f:\mathbb E\to\mathbb R$. Then $$ f^\prime(\bar x;u)\ge 0,\quad\forall u\in\mathbb E,\eqno{\rm (BZ_1)} $$ $$ f^\prime(\bar x;u)=0\quad\Rightarrow\quad f^{\prime\pr}_{BZ}(\bar x;u,z)\ge 0,\;\forall z\in\mathbb E.\eqno{\rm (BZ_2)} $$ \end{proposition} In the next result, we prove that Conditions (BZ$_1$) and (BZ$_2$) are consequence of (\ref{14}): \begin{proposition} Let $f:\mathbb E\to\mathbb R$ and $\bar x\in\mathbb E$ be a given function and a point respectively, such that the derivatives $ f^\prime(\bar x;u)$ and $f^{\prime\pr}_{BZ}(\bar x;u,z)$ exist for all directions $u\in\mathbb E$ and $z\in\mathbb E$. Then Conditions (\ref{14}) imply that $({\rm BZ_1})$ and $({\rm BZ_2})$ are satisfied at $\bar x$. \end{proposition} \begin{proof} Suppose that (\ref{14}) holds. Then the inequality $f^\prime(\bar x;u)\ge\ld 1 f(\bar x;u)\ge 0$ implies that (BZ$_1$) is satisfied. Let $f^\prime(\bar x;u)=0$. Then the chain of relations \[ f^{\prime\pr}_{BZ}(\bar x;u,z)=\lim_{t\downarrow 0} t^{-2}[f(\bar x+t(u+tz))-f(\bar x)] \ge\liminf_{t\downarrow 0,u^\prime\to u} t^{-2}[f(\bar x+tu^\prime)-f(\bar x)]= 0.5\,\ld 2 f(\bar x;0;u)\ge 0 \] show that (BZ$_2$) is also satisfied for arbitrary $z\in\mathbb E$. \end{proof} Examples 2.1 and 2.2 in \cite{stu86} show that the optimality conditions given there cannot solve arbitrary set constrained problem. The next example was given in \cite{stu86}: \begin{example} Consider the problem \[ \inf\{f(x)\mid x\in C\}, \] where the function $f:\mathbb R^2\to\mathbb R$ is defined by \[ f(x_1,x_2)=\left\{ \begin{array}{ll} |x_1|^n, & \textrm{if}\quad x_2\ne 0;\\ 0, & \textrm{if}\quad x_2=0, \end{array}\right. \] and $C=\mathbb R\times\{0\}$. Here $n$ is a positive integer. It is shown in \cite[Example 2.2]{stu86} that $\bar x=(0,0)$ is an isolated local minimizer of order $n$ over the set $C$, but the sufficient conditions in this paper cannot establish this fact, because the required derivatives are identical to $0$. The derivatives that we study in the present work can solve the problem. Consider the function $g$ such that \[ g(x)=\left\{ \begin{array}{ll} f(x), & \textrm{if}\quad x\in C;\\ +\infty, & \textrm{if}\quad x\notin C. \end{array}\right. \] We obviously have \[ \bar x\in {\rm Argmin}\,\{f(x)\mid x\in C\}\quad\Longleftrightarrow\quad \bar x\in {\rm Argmin}\,\{g(x)\mid x\in\mathbb E\}. \] We prove that $\bar x$ is an isolated minimizer of order $n$. Let $u\in\mathbb R^2$ be an arbitrary direction such that $\norm{u}=1$. Then $\ld 1 g(\bar x;u)=0$ and $\ld i g(\bar x;0,\dots,0;u)=0$ if $u=(\pm 1,0)$ for $i<n$. We have $\ld 1 g(\bar x;u)=+\infty$ and $\ld i g(\bar x;0,\dots,0;u)=+\infty$ if $u\ne (\pm 1,0)$ for $i<n$. Moreover, $0\in\lsubd 1 g(\bar x)$, $0\in\lsubd i g(\bar x,0,\dots,0)$ for $i<n$. At last, we obtain that $\ld n g(\bar x;0,\dots,0;u)=n!$ if $u=(\pm 1,0)$ and $\ld n g(\bar x;0,\dots,0;u)=+\infty$ if $u\ne (\pm 1,0)$. It follows from the sufficient conditions in Theorem \ref{th2} that $\bar x$ is an isolated local minimizer of order $n$. \end{example} Ginchev \cite{gin02} introduced the following directional derivatives of Hadamard type. Let be given an proper extended real function $f : \E\to\R\cup\{+\infty\}$. The derivatives begin with the derivative of order $0$: \[ \gin 0 f(x;u):=\liminf_{t\downarrow 0,u^\prime\to u}\, f(x+t u^\prime). \] Let $n$ be a positive integer. Then the derivative of order $n$ is defined as follows: \[ \gin n f(x;u):=\liminf_{t\downarrow 0,u^\prime\to u}\,\frac{n!}{t^n}\,[f(x+t u^\prime)-\sum_{i=0}^{n-1}\frac{t^i}{i!}\gin i f(x;u)]. \] In \cite[Theorem 1]{gin02} the author derived necessary optimality conditions of order $n$ for a local minimum under the assumption that the required derivatives exist. It is not discussed when these derivatives exist. The derivatives in \cite{aub90,gin02,stu86} are not consistent with the classical Fr\'echet derivatives. Really, for every twice Fr\'echet differentiable convex function $\gin 2 f(x;u)$ does not coincide with the second-order Fr\'echet directional derivative. On the other hand, by Proposition \ref{pr3}, the derivatives introduced in Definition \ref{def3} are consistent with the classical ones. We can find only higher-order necessary conditions for a local minimum and second-order sufficient conditions for a weak local minimum in the book \cite[Section 6.6]{aub90}. \begin{example} Consider the function $f:\mathbb R^2\to\mathbb R$ such that \[ f(x_1,x_2)=x_1^2+x_2^2. \] Let $u=(u_1,u_2)$ be a direction. We have \[ \gin 0 f(x;u)=x_1^2+x_2^2,\quad \gin 1 f(x;u)=2x_1 u_1+2x_2 u_2. \] Therefore, \[ \gin 2 f(x;u)=2(u_1^2+u_2^2)+\liminf_{t\downarrow 0,u^\prime\to u}\,4 t^{-1}[x_1(u_1^\prime-u_1)+x_2(u_2^\prime-u_2)]. \] If we take $t=t_k=1/k^2$, $u_2^\prime=u_2=0$, $u_1=1$, $u_1^\prime=1-1/k$, $x_1>0$, where $k$ is a positive integer with $k\to +\infty$, then we see that $\gin 2 f(x;u)=-\infty\ne\nabla^2 f(x)(u)(u)$. \end{example} This example also shows that the second-order contingent epi-derivative in \cite{aub90} is not consistent with the classical second-order derivative, and it equals $-\infty$ if we take the same values of the variables as in the example.
\section{Motivation} \label{novak_rudolf sec: motivation} Consider the following example. We want to compute \[ \mathbb{E}_G(f) = \frac{1}{{\rm vol}_d(G)} \int_G f(x)\, \text{\rm d} x, \] where $f$ belongs to some class of functions and $G$ belongs to some class of sets. We assume that $G\subset \mathbb{R}^d$ is measurable with $0<{\rm vol}_d(G)<\infty$, where ${\rm vol}_d$ denotes the Lebesgue measure. Thus, we want to compute the expected value of $f$ with respect to the uniform distribution on $G$. The input $(f, G)$ is given by an oracle: For $x \in G$ we can compute $f(x)$ and $G$ is given by a membership oracle, i.e. we are able to check whether any $x\in \mathbb{R}^d$ is in $G$ or not. We always assume that $G$ is convex and will work with the class \begin{equation} \label{novak_rudolf classGrd} \mathcal{G}_{r,d} = \{ G \subset \mathbb{R}^d \colon G \text{\,is convex},\; B_d \subset G \subset rB_d\}, \end{equation} where $r\geq 1$ and $rB_d=\{x\in \mathbb{R}^d \colon \vert x \vert \leq r \}$ is the Euclidean ball with radius $r$. A first approach might be a simple acceptance/rejection method. The idea is to generate a point in $rB_d$ according to the uniform distribution and if it is in $G$ it is accepted, otherwise it is rejected. If $x_1,\dots,x_n \in G$ are the accepted points then we output the mean value of the $f(x_i)$. However, this method does not work reasonably since the acceptance probability can be extremely small, it can be $r^{-d}$. It seems that all known efficient algorithms for this problem use Markov chains. The idea is to find a sampling procedure that approximates a sample with respect to the uniform distribution in $G$. More precisely, we run a Markov chain to approximate the uniform distribution for any $G\in\mathcal{G}_{r,d}$. Let $X_1,X_2,\dots,X_{n+n_0}$ be the first $n+n_0$ steps of such a Markov chain. Then \[ S_{n,n_0}(f,G)= \frac{1}{n} \sum_{j=1}^n f(X_{j+n_0}) \] is an approximation of $\mathbb{E}_G(f)$. The additional parameter $n_0$ is called burn-in and, roughly spoken, is the number of steps of the Markov chain to get close to the uniform distribution. \section{Approximation of expectations by MCMC} \subsection{Preliminaries} We provide the basics of Markov chains. For further reading we refer to the paper \cite{novak rudolf:RoRo04} of Roberts and Rosenthal which surveys various results about Markov chains on general state spaces. A Markov chain is a sequence of random variables $(X_n)_{n\in\mathbb{N}}$ which satisfies the Markov property. For $i\in \mathbb{N}$, the conditional distribution of $X_{i+1}$ depends only on $X_i$ and not on $(X_1,\dots,X_{i-1})$, \[ \mathbb{P}(X_{i+1}\in A \mid X_1,\dots,X_i) = \mathbb{P}(X_{i+1}\in A\mid X_i). \] By $\mathcal{B}(G)$ we denote the Borel $\sigma$-algebra of $G$. Let $\nu$ be a distribution on $(G,\mathcal{B}(G))$ and let $K \colon G \times \mathcal{B}(G) \to [0,1]$ be a \emph{transition kernel}\index{transition kernel}, i.e. $K(x,\cdot)$ is a probability measure for each $x \in G$ and $K(\cdot,A)$ is a $\mathcal{B}(G)$-measurable real-valued function for each $A \in \mathcal{B}(G)$. A transition kernel and a distribution $\nu$ give rise to a Markov chain $(X_n)_{n\in\mathbb{N}}$ in the following way. Assume that the distribution of $X_1$ is given by $\nu$. Then, for $i\geq 2$ and a given $X_{i-1}=x_{i-1}$, we have $X_i$ with distribution $K(x_{i-1},\cdot)$, that is, for all $A\in\mathcal{B}(G)$, the conditional probability that $X_i \in A$ is given by $K(x_{i-1},A)$. We call such a sequence of random variables a Markov chain with transition kernel $K$ and initial distribution $\nu$. In the whole paper we only consider Markov chains with \emph{reversible}\index{reversible} transition kernel, we assume that there exists a probability measure $\pi$ on $\mathcal{B}(G)$ such that \[ \int_A K(x,B)\, \pi({\rm d}x) = \int_B K(x,A)\, \pi({\rm d}x), \quad A,B \in \mathcal{B}(G). \] In particular any such $\pi$ is a stationary distribution of $K$, i.e., \[ \pi(A) = \int_G K(x,A) \, \pi({\rm d} x), \quad A\in\mathcal{B}(G). \] Further, the transition kernel induces an operator on functions and an operator on measures given by \[ P f(x) = \int_G f(y)\, K(x,{\rm d}x), \quad \text{and} \quad \nu P(A) = \int_G K(x,A)\, \nu({\rm d}x), \] where $f$ is $\pi$-integrable and $\nu$ is absolutely continuous with respect to $\pi$. One has \[ \mathbb{E}[f(X_n) \mid X_1 = x] = P^{n-1} f(x) \quad \text{and} \quad \mathbb{P}_\nu(X_n\in A) = \nu P^{n-1} (A), \] for $x\in G$, $A \in \mathcal{B}(G)$ and $n\in \mathbb{N}$, where $\nu$ in $\mathbb{P}_\nu$ indicates that $X_1$ has distribution $\nu$. By the reversibility with respect to $\pi$ we have $ \frac{d(\nu P)}{d\pi}(x) = P (\frac{d \nu}{d \pi})(x), $ where $\frac{d \nu}{d \pi}$ denotes the density of $\nu$ with respect to $\pi$. Further, for $p\in[1,\infty)$ let $L_p=L_p(\pi)$ be the space of measurable functions $f\colon G \to \mathbb{R}$ which satisfy \[ \Vert f \Vert _{p} = \left( \int_G \vert f(x) \vert^p \pi({\rm d} x) \right)^{1/p} < \infty. \] The operator $P \colon L_p \to L_p$ is linear and bounded and by the reversibility $P\colon L_2 \to L_2$ is self-adjoint. The goal is to quantify the speed of convergence, if it converges at all, of $\nu P^{n}$ to $\pi$ for increasing $n\in\mathbb{N}$. For this we use the \emph{total variation distance}\index{total variation distance} between two probability measures $\nu,\mu$ on $(G,\mathcal{B}(G))$ given by \[ \Vert \nu-\mu \Vert_{\text{\rm tv}} =\sup_{A\in\mathcal{B}(G)}\vert \nu(A)-\mu(A) \vert. \] It is helpful to consider the total variation distance as an $L_1$-norm, see for example \cite[Proposition~3, p.~28]{novak rudolf:RoRo04}. \begin{lemma} \label{novak_rudolf tv_present} Assume the probability measures $\nu,\mu$ have densities $\frac{d \nu}{d \pi}, \frac{d \mu}{d \pi} \in L_1$, then $ \Vert \nu-\mu \Vert_{\text{\rm tv}} =\frac{1}{2}\left \Vert \frac{d\nu}{d\pi}-\frac{d\mu}{d\pi}\right \Vert_{1}. $ \end{lemma} Now we ask for an upper bound of $\left \Vert \nu P^n - \pi\right \Vert_{\text{\rm tv}}$. \begin{lemma} \label{novak_rudolf lem: est_tv} Let $\nu$ be a probability measure on $(G,\mathcal{B}(G))$ with $\frac{d\nu}{d\pi}\in L_1$ and let $ S(f) = \int_G f(x)\,\pi(\text{\rm d} x). $ Then, for any $n\in\mathbb{N}$ holds \[ \left \Vert \nu P^n - \pi \right \Vert_{\text{\rm tv}} \leq \left \Vert P^n-S \right \Vert_{L_1\to L_1} \frac{1}{2} \left \Vert \frac{d\nu}{d\pi}-1\right \Vert_{1} \leq \left \Vert P^n-S \right \Vert_{L_1\to L_1} \] and \[ \left \Vert \nu P^n - \pi \right \Vert _{\text{\rm tv}} \leq \left \Vert P^n-S \right \Vert_{L_2\to L_2} \frac{1}{2} \left \Vert \frac{d\nu}{d\pi}-1 \right \Vert_{2}. \] \end{lemma} \begin{proof} By Lemma~\ref{novak_rudolf tv_present}, by $P^n 1 = 1$ and by the reversibility, in particular $ \frac{d(\nu P^n)}{d\pi}(x) = P^n (\frac{d \nu}{d \pi})(x), $ we have \begin{align*} 2 \left \Vert \nu P^n - \pi \right \Vert _{\text{\rm tv}} & = \left \Vert \frac{d(\nu P^n)}{d\pi} - 1 \right \Vert_{1} = \left \Vert P^n\left(\frac{d\nu}{d\pi}-1\right)\right \Vert_{1} = \left \Vert (P^n-S)\left(\frac{d\nu}{d\pi}-1\right)\right \Vert_{1}. \end{align*} Note that the last equality comes from $S(\frac{d\nu}{d\pi}-1)=0$. \end{proof} Observe that for $\nu = \pi$ the left-hand side and also the right-hand side of the estimates are zero. Let us consider $\left \Vert P^n-S \right \Vert_{L_2\to L_2}$. Because of the reversibility with respect to $\pi$ we obtain the following, see for example \cite[Lemma~3.16, p.~45]{novak rudolf:Ru12}. \begin{lemma} \label{novak_rudolf lem: norm_spec_gap} For $n\in \mathbb{N}$ we have \[ \left \Vert P^n-S \right \Vert_{L_2 \to L_2 } = \left \Vert (P-S)^n \right \Vert_{L_2 \to L_2} = \left \Vert P-S \right \Vert_{L_2 \to L_2}^n. \] \end{lemma} The last two lemmata motivate the following two convergence properties of transition kernels. \begin{definition} [\index{$L_1$-exponential convergence}$L_1$-exponential convergence] Let $\alpha\in[0,1)$ and $M\in(0,\infty)$. Then the transition kernel $K$ is $L_1$-exponentially convergent with $(\alpha,M)$ if \begin{equation} \label{novak_rudolf eq: L_1_exp_con} \left \Vert P^n-S \right \Vert_{L_1 \to L_1} \leq \alpha^n M,\quad n\in\mathbb{N}. \end{equation} A Markov chain with transition kernel $K$ is called $L_1$-exponentially convergent if there exist an $\alpha\in[0,1)$ and $M\in(0,\infty)$ such that \eqref{novak_rudolf eq: L_1_exp_con} holds. \end{definition} \begin{definition}[\index{$L_2$-spectral gap}$L_2$-spectral gap] We say that a transition kernel $K$ and its corresponding Markov operator $P$ have an $L_2$-spectral gap if \[ {\rm gap}(P)=1-\left \Vert P-S\right \Vert_{L_2 \to L_2}>0. \] \end{definition} If the transition kernel has an $L_2$-spectral gap, then by Lemma~\ref{novak_rudolf lem: est_tv} and Lemma~\ref{novak_rudolf lem: norm_spec_gap} we have that \[ \left \Vert \nu P^n - \pi \right \Vert_{\text{\rm tv}} \leq (1-{\rm gap}(P))^n \left \Vert \frac{d\nu}{d\pi}-1\right \Vert_{2}. \] Next, we define other convergence properties which are based on the total variation distance. \begin{definition}[\index{uniform ergodicity}uniform ergodicity and \index{geometric ergodicity}geometric ergodicity] Let $\alpha\in [0,1)$ and $M\colon G \to (0,\infty)$. Then the transition kernel $K$ is called \emph{geometrically ergodic with $(\alpha,M(x))$} if one has for $\pi$-almost all $x\in G$ that \begin{equation} \label{novak_rudolf eq: pi_erg} \left \Vert K^n(x,\cdot)-\pi\right \Vert_{\text{\rm tv}}\leq M(x)\, \alpha^n ,\quad n\in\mathbb{N}. \end{equation} If the inequality $\eqref{novak_rudolf eq: pi_erg}$ holds with a bounded function $M(x)$, i.e. \[ \sup_{x\in G} M(x) \leq M' <\infty, \] then $K$ is called \emph{uniformly ergodic with $(\alpha,M')$}. \end{definition} Now we state several relations between the different properties. Since we assume that the transition kernel is reversible with respect to $\pi$ we have the following: \begin{equation} \label{novak_rudolf eq: diagramm} \begin{array}{ccc} \mbox{uniformly ergodic} & \Longleftrightarrow & \mbox{$L_1$-exponentially convergent } \\ \mbox{with } (\alpha,M) & & \mbox{with } (\alpha,2M) \\[1ex] \mathbin{\text{\rotatebox[origin=c]{90}{$\Longleftarrow$}}} & & \mathbin{\text{\rotatebox[origin=c]{90}{$\Longleftarrow$}}} \\[1ex] \mbox{geometrically ergodic} & & \mbox{$L_2$-spectral gap $\geq$}\\ \mbox{with } (\alpha,M(x)) & & 1-\alpha. \end{array} \end{equation} The fact that uniform ergodicity implies geometric ergodicity is obvious. For the proofs of the other relations and further details we refer to \cite[Proposition~3.23, Proposition~3.24]{novak rudolf:Ru12}. Further, if the transition kernel is $\varphi$-irreducible, for details we refer to \cite{novak rudolf:RoRo97} and \cite{novak rudolf:RoTw01}, then \begin{equation} \begin{array}{ccc} \mbox{geometrically ergodic} &\quad \Longleftrightarrow \quad & \mbox{$L_2$-spectral gap $\geq$}\\ \mbox{with } (\alpha,M(x)) & & 1-\alpha. \end{array} \end{equation} \subsection{Mean square error bounds of MCMC} The goal is to compute \[ S(f) = \int_G f(x)\, \pi(\text{\rm d} x). \] We use an average of a finite Markov chain sample as approximation of the mean, i.e. we approximate $S(f)$ by \[ S_{n,n_0}(f) = \frac{1}{n} \sum_{j=1}^n f(X_{j+n_0}). \] The number $n$ determines the number of function evaluations of $f$. The number $n_0$ is the \emph{burn-in}\index{burn-in} or \emph{warm up} time. Intuitively, it is the number of steps of the Markov chain to get close to the stationary distribution $\pi$. We study the mean square error of $S_{n,n_0}$, given by \[ e_\nu(S_{n,n_0},f) = \left( \mathbb{E}_{\nu,K} \vert S_{n,n_0}(f)-S(f) \vert \right)^{1/2}, \] where $\nu$ and $K$ indicate the initial distribution and transition kernel. We start with the case $\nu=\pi$, where the initial distribution is the stationary distribution. \begin{lemma} \label{novak_rudolf thm: err_bound_stat} Let $(X_n)_{n\in\mathbb{N}}$ be a Markov chain with transition kernel $K$ and initial distribution $\pi$. We define \[ \Lambda = \sup\{ \alpha \colon \alpha \in {\rm {s}pec}(P-S) \}, \] where ${\rm {s}pec}(P-S)$ denotes the spectrum of the operator $P-S\colon L_2 \to L_2$, and assume that $\Lambda<1$. Then \[ \sup_{\left \Vert f\right \Vert_{2}\leq 1} e_\pi(S_{n,n_0},f)^2 \leq \frac{2}{n(1-\Lambda)}. \] \end{lemma} For a proof of this result we refer to \cite[Corollary~3.27]{novak rudolf:Ru12}. Let us discuss the assumptions and implications of Lemma~\ref{novak_rudolf thm: err_bound_stat}. First, note that for the simple Monte Carlo method we have $\Lambda=0$. In this case we get (up to a constant of 2) what we would expect. Further, note that ${\rm gap}(P) =1-\left \Vert P-S\right \Vert_{L_2 \to L_2}$ and \[ \left \Vert P-S\right \Vert_{L_2 \to L_2} = \sup \{\vert \alpha \vert \colon \alpha \in {\rm {s}pec}(P-S)\}, \] so that ${\rm gap}(P)\leq 1-\Lambda$. This also implies that if $P\colon L_2 \to L_2$ is positive semidefinite we obtain ${\rm gap}(P) = 1-\Lambda$. Thus, whenever we have a lower bound for the spectral gap we can apply Lemma~\ref{novak_rudolf thm: err_bound_stat} and can replace $1-\Lambda$ by ${\rm gap}(P)$. Further note if $\gamma\in[0,1)$, $M\in(0,\infty)$ and the transition kernel is $L_1$-exponentially convergent with $(\gamma,M)$ then we have, using \eqref{novak_rudolf eq: diagramm}, that ${\rm gap}(P)\geq 1-\gamma$. Now we ask how $e_\nu(S_{n,n_0},f)$ behaves depending on the initial distribution. The idea is to decompose the error in a suitable way. For example in a bias and variance term. However, we want to have an estimate with respect to $\left \Vert f \right \Vert_{2}$ and in this setting the following decomposition is more convenient: \[ e_\nu(S_{n,n_0},f)^2 = e_\pi(S_{n,n_0},f)^2 + \mbox{rest}, \] where rest denotes an additional term such that equality holds. Then, we estimate the remainder term and use Lemma~\ref{novak_rudolf thm: err_bound_stat} to obtain an error bound. For further details of the proof of the following error bound we refer to \cite[Theorem~3.34 and Theorem~3.41]{novak rudolf:Ru12}. \begin{theorem} \label{novak_rudolf thm: err_bound} Let $(X_n)_{n\in\mathbb{N}}$ be a Markov chain with reversible transition kernel $K$ and initial distribution $\nu$. Further, let \begin{equation*} \Lambda = \sup\{ \alpha \colon \alpha \in {\rm {s}pec}(P-S) \}, \end{equation*} where ${\rm {s}pec}(P-S)$ denotes the spectrum of the operator $P-S\colon L_2 \to L_2$, and assume that $\Lambda<1$. Then \begin{equation} \label{novak_rudolf eq: expl_err_bound} \sup_{\left \Vert f \right \Vert_{p}\leq 1 } e_\nu(S_{n,n_0},f)^2 \leq \frac{2}{n(1-\Lambda)} + \frac{2\, C_\nu \gamma^{n_0}}{n^2(1-\gamma)^2} \end{equation} holds for $p=2$ and for $p=4$ under the following conditions \begin{enumerate \item for $p=2$, $\frac{d\nu}{d\pi}\in L_\infty$ and a transition kernel $K$ which is $L_1$-exponentially convergent with $(\gamma,M)$ where $ C_\nu = M \left \Vert \frac{d\nu}{d\pi}-1\right \Vert_{\infty}; $ \item \label{novak_rudolf en: p_4} for $p=4$, $\frac{d\nu}{d\pi}\in L_2$ and $1-\gamma={\rm gap}(P)>0$ where $ C_\nu = 64 \left \Vert \frac{d\nu}{d\pi}-1\right \Vert_{2}. $ \end{enumerate} \end{theorem} Let us discuss the results. If the transition kernel is $L_1$-exponentially ergodic, then we have an explicit error bound for integrands $f\in L_2$ whenever the initial distribution has a density $\frac{d \nu}{ d \pi} \in L_\infty$. However, in general it is difficult to provide explicit values $\gamma$ and $M$ such that the transition kernel is $L_1$-exponentially convergent with $(\gamma,M)$. This motivates to consider transition kernel which satisfy a weaker convergence property, such as the existence of an $L_2$-spectral gap. In this case we have an explicit error bound for integrands $f\in L_4$ whenever the initial distribution has a density $\frac{d \nu}{ d \pi} \in L_2$. Thus, by assuming a weaker convergence property of the transition kernel we obtain a weaker result in the sense that $f$ must be in $L_4$ rather than $L_2$. However, with respect to $\frac{d \nu}{d \pi}$ we do not need boundedness anymore, it is enough that $\frac{d \nu}{d \pi} \in L_2$. In Theorem~\ref{novak_rudolf thm: err_bound} we provided explicit error bounds and we add in passing that also other error bounds are known, see \cite{novak rudolf:BeCh09,novak rudolf:JoOl10,novak rudolf:LaMiNi11,novak rudolf:Ru12}. If we want to have an error of $\varepsilon \in(0,1)$ it is still not clear how to choose $n$ and $n_0$ to minimize the total amount of steps $n+n_0$. How should we choose the burn-in $n_0$? Let $e(n,n_0)$ be the right hand side of \eqref{novak_rudolf eq: expl_err_bound} and assume that $\Lambda= \gamma$. Further, assume that we have computational resources for $N=n+n_0$ steps of the Markov chain. We want to get an $n_{\text{opt}}$ which minimizes $e(N-n_0,n_0)$. In \cite[Lemma~2.26]{novak rudolf:Ru12} the following is proven: For all $\delta>0$ and large enough $N$ and $C_\nu$ the number $n_{\text{opt}}$ satisfies \[ n_{\text{opt}} \in \left[ \frac{\log C_\nu}{\log\gamma^{-1}}, (1+\delta)\frac{\log C_\nu}{\log\gamma^{-1}} \right]. \] Further note that $\log \gamma^{-1} \geq 1-\gamma$. Thus, in this setting $ n_{\text{opt}} = \lceil\frac{\log C_\nu}{1-\gamma} \rceil$ is a reasonable and almost optimal choice for the burn-in. \section{Application of the error bound and limitations of MCMC} First, we briefly introduce a technique to prove a lower bound of the spectral gap if the Markov operator of a transition kernel is positive semidefinite on $L_2$. The following result, known as \emph{Cheeger's inequality}\index{Cheeger's inequality}, is in this form due to Lawler and Sokal \cite{novak rudolf:LaSo88}. \begin{proposition} \label{novak_rudolf prop: cheeger} Let $K$ be a reversible transition kernel, which induces a Markov operator $P\colon L_2 \to L_2$. Then \[ \frac{\varphi^2}{2} \leq1-\Lambda \leq 2 \varphi, \] where $ \Lambda = \sup\{ \alpha \colon \alpha \in {\rm {s}pec}(P-S) \} $ and \[ \varphi = \inf_{0<\pi(A)\leq 1/2} \frac{\int_A K(x,A^c) \, \pi({\rm d} x)}{\pi(A)} \] is the conductance of $K$. \end{proposition} Now we state different applications of Theorem~\ref{novak_rudolf thm: err_bound}. \subsection{Hit-and-run algorithm\index{Hit-and-run algorithm}} \label{novak_rudolf sec: har} We consider the example of Section~\ref{novak_rudolf sec: motivation}. Let $G \in \mathcal{G}_{r,d}$, see \eqref{novak_rudolf classGrd}, and let $\mu_G$ be the uniform distribution in $G$. We define \begin{equation} \label{novak_rudolf Frd} \mathcal{F}_{r,d} = \{ (f,G)\colon G \in \mathcal{G}_{r,d}, \, f\in L_4(\mu_G),\, \left \Vert f \right \Vert _{4}\leq 1\}. \end{equation} The goal is to approximate \[ S(f,\mathbf{1}_G) = \frac{1}{{\rm vol}_d(G)} \int_G f(x)\, \text{\rm d} x, \] where $(f,G) \in \mathcal{F}_{r,d}$. The hit-and-run algorithm defines a Markov chain which satisfies the assumptions of Theorem~\ref{novak_rudolf thm: err_bound}. A step from $x\in G$ of the hit-and-run algorithm works as follows \begin{enumerate} \item Choose a direction, say $\theta$, uniformly distributed on the sphere $\partial B_d$. \item Choose the next state, say $y \in G$, uniformly distributed in $G\cap \{ x+\theta r\colon r\in \mathbb{R} \}$. \end{enumerate} After choosing a direction $\theta$ one samples the next state $y\in G$ with respect to the uniform distribution in the line determined by the current state $x$ and the direction $\theta$ restricted to $G$. The random number, say $u \in [0,1]$, for the second part is chosen independently of the first part and also all steps are independent. Lova\'sz and Vempala prove in \cite[Theorem 4.2, p. 993]{novak rudolf:LoVe06} a lower bound of the conductance $\varphi$, see Proposition~\ref{novak_rudolf prop: cheeger} for the definition of the conductance. \begin{proposition} Let $G\in \mathcal{G}_{r,d}$. Then, the conductance of the hit-and-run algorithm is bounded from below by $2^{-25}(dr)^{-1}$. \end{proposition} It is known that the hit-and-run algorithm induces a positive semidefinite Markov operator, say $H$, see \cite{novak rudolf:RuUl12}. By Proposition~\ref{novak_rudolf prop: cheeger} we obtain \[ {\rm gap}(H) \geq \frac{2^{-51}}{(dr)^2} \] and Theorem~\ref{novak_rudolf thm: err_bound} implies the following error bound for the class $\mathcal{F}_{r,d}$, see~\eqref{novak_rudolf classGrd} and~\eqref{novak_rudolf Frd}. \begin{theorem} \label{novak_rudolf thm: har} Let $\nu$ be the uniform distribution on $B_d$. Let $(X_n)_{n\in\mathbb{N}}$ be a Markov chain with transition kernel, given by the hit-and-run algorithm, and initial distribution $\nu$. Let \[ n_0 = \lceil 4.51\cdot 10^{15} d^2 r^2 ( d \log r + 4.16) \rceil. \] Then \[ \sup_{(f,G) \in \mathcal{F}_{r,d}} e_\nu(S_{n,n_0},(f,\mathbf{1}_G)) \leq 9.5\cdot 10^7 \frac{d r}{\sqrt{n}} + 6.4\cdot 10^{15} \frac{d^2 r^2}{n}. \] \end{theorem} This result states that the number of oracle calls for $f$ and $G$ to obtain an error $\varepsilon>0$ is bounded by $ \kappa\, d^2 r^2 (\varepsilon ^{-2} + d\log r), $ for an explicit constant $\kappa>0$. Hence the computation of $S(f,\mathbf{1}_G)$ on the class $\mathcal{F}_{r,d}$ is polynomially tractable, see \cite{novak rudolf:NoWo08,novak rudolf:NoWo10,novak rudolf:NoWo12}. The tractability result can be extended also to other classes of functions, see \cite{novak rudolf:Ru13}. Note that we applied the second statement of Theorem~\ref{novak_rudolf thm: err_bound}. It is known that the hit-and-run algorithm is $L_1$-exponentially ergodic with $(\gamma,M)$, for some $\gamma\in(0,1)$ and $M\in(0,\infty)$. But the best known numbers $\gamma$ and $M$ are exponentially bad in terms of the dimension, see \cite{novak rudolf:Sm84}. \subsection{Metropolis-Hastings algorithm\index{Metropolis-Hastings algorithm}} Let $G \subset \mathbb{R}^d$ and $\rho \colon G \to (0,\infty)$, where $\rho$ is integrable with respect to the Lebesgue measure. We define the distribution $\pi_\rho$ on $(G,\mathcal{B}(G))$ by \[ \pi_\rho(A) = \frac{\int_A \rho(x)\, {\rm d} x}{\int_G \rho(x)\, {\rm d} x}, \qquad A\in \mathcal{B}(G). \] The goal is to compute \[ S(f,\rho) = \int_G f(x)\, \pi_\rho({\rm d}x) = \frac{\int_G f(x) \rho(x)\, {\rm d}x }{ \int_G \rho(x)\,{\rm d}x} \] for functions $f\colon G \to \mathbb{R} $ which are integrable with respect to $\pi_\rho$. The \emph{Metropolis-Hastings algorithm} defines a Markov chain which approximates $\pi_\rho$. We need some further notations. Let $q\colon G\times G \to [0,\infty]$ be a function such that $q(x,\cdot)$ is Lebesgue integrable for all $x\in G$ with $\int_G q(x,y)\,{\rm d} y\leq 1$. Then \[ Q(x,A)= \int_A q(x,y)\,{\rm d} y + \mathbf{1}_A(x)\left(1-\int_G q(x,y)\, {\rm d} y\right), \quad x\in G,\; A\in\mathcal{B}(G), \] is a transition kernel and we call $q(\cdot,\cdot)$ \emph{transition density}. The idea is to modify $Q$, such that $\pi_\rho$ gets a stationary distribution of the modification. We propose a state with $Q$ and with a certain probability, which depends on $\rho$, the state is accepted. Let $\alpha(x,y)$ be the acceptance probability \[ \alpha(x,y) = \begin{cases} 1 & \mbox{if } q(x,y)\rho(x)=0,\\ \min\{ 1 , \frac{q(y,x)\rho(y)}{q(x,y)\rho(x)}\} & \mbox{otherwise}. \end{cases} \] The transition kernel of the Metropolis-Hastings algorithm is \begin{align*} K_\rho(x,A) & = \int_A \alpha(x,y)\, q(x,y) {\rm d} y + \mathbf{1}_A(x) \left[ 1-\int_G \alpha(x,y)\,q(x, y) {\rm d} y\right] \end{align*} for $x\in G$ and $A\in \mathcal{B}(G)$. The transition kernel $K_\rho$ is reversible with respect to $\pi_\rho$. {}From the current state $x\in G$ a single transition of the algorithm works as follows: \begin{enumerate} \item Sample a proposal state $y\in G$ with respect to $Q(x,\cdot)$. \item With probability $\alpha(x,y)$ return $y$, otherwise reject $y$ and return $x$. \end{enumerate} Again, all steps are done independently of each other. If $q(x,y)=q(y,x)$, i.e. $q$ is symmetric, then $K_\rho$ is called \emph{Metropolis algorithm} and if $q(x,y)=\eta(y)$ for a function $\eta\colon G\to (0,\infty)$ for all $x,y\in G$, then $K_\rho$ is called \emph{independent Metropolis algorithm}.\\ Let $G\subset \mathbb{R}^d$ be bounded and for $C\geq 1$ let \begin{equation} \label{novak_rudolf R_C} \mathcal{R}_C = \{ \rho\colon G \to (0,\infty) \mid 1\leq \rho(x) \leq C \}. \end{equation} Thus, for any $\rho \in \mathcal{R}_C$ holds $\sup \rho/ \inf \rho \leq C$. If $\rho\colon G \to (0,\infty)$ satisfies $\sup \rho/ \inf \rho \leq C$, then \[ \frac{\left \Vert \rho \right \Vert_{\infty}}{C} \leq \rho(x) \leq C \inf \rho. \] Thus, $ C\cdot\rho /\left \Vert \rho\right \Vert_{\infty} \in \mathcal{R}_C$. We consider an independent Metropolis algorithm. The proposal transition kernel is \[ Q(x,A)= \mu_G (A) = \frac{{\rm vol}_d(A)}{{\rm vol}_d(G)}, \quad A\in\mathcal{B}(G), \] i.e. a state is proposed with the uniform distribution in $G$. Then \[ K_\rho(x,A) = \int_A \alpha(x,y) \frac{{\rm d} y}{{\rm vol}_d(G)} + \mathbf{1}_A(x) \left(1-\int_G \alpha(x,y)\, \frac{{\rm d} y}{{\rm vol}_d(G)}\right), \] where $\alpha(x,y)= \min\{ 1,\rho(y)/\rho(x) \}$. The transition operator $P_\rho\colon L_2(\pi_\rho) \to L_2(\pi_\rho)$, induced by $K_\rho$, is positive semidefinite. For details we refer to \cite{novak rudolf:RuUl12}. Thus, ${\rm gap}(P_\rho)=1-\Lambda_\rho$, with $\Lambda_{\rho} = \Lambda$. Further, for $\rho \in \mathcal{R}_C$ Theorem~2.1 of \cite{novak rudolf:MeTw96} provides a criterion for uniform ergodicity of the independent Metropolis algorithm. Namely, $K_\rho$ is uniformly ergodic with $(\gamma,1)$ for $\gamma = 1-C^{-1}/{\rm vol}_d (G)$. Thus, by \eqref{novak_rudolf eq: diagramm} we have that it is $L_1$-exponentially ergodic with $(\gamma,2)$. Further, by \eqref{novak_rudolf eq: diagramm} we obtain \[ 1-\Lambda_\rho={\rm gap}(P_\rho) \geq \frac{C^{-1}}{{\rm vol}_d (G) }. \] Let \begin{equation} \label{novak_rudolf F_C,d} \mathcal{F}_{C,d} = \{ (f,\rho) \colon \rho \in \mathcal{R}_{C},\, f\in L_2(\pi_\rho),\, \left \Vert f\right \Vert_{2}\leq 1 \}. \end{equation} We apply Theorem~\ref{novak_rudolf thm: err_bound} and obtain for the class $\mathcal{F}_{C,d}$ (see \eqref{novak_rudolf R_C} and \eqref{novak_rudolf F_C,d}) \begin{theorem} \label{novak_rudolf TH3} Let $(X_n)_{n\in\mathbb{N}}$ be a Markov chain with transition kernel, given by the Metropolis algorithm with proposal $\mu_G$, and initial distribution $\mu_G$. Let \[ n_0 = \left\lceil C {\rm vol}_d(G)\log(2C) \right \rceil. \] Then \[ \sup_{(f,\rho) \in \mathcal{F}_{C,d}} e_\nu(S_{n,n_0},(f,\rho))^2 \leq \frac{2\, C\, {\rm vol}_d(G) }{n} + \frac{4\, C^2\, {\rm vol}_d(G)^2 }{n^2}. \] \end{theorem} The upper bound in Theorem~\ref{novak_rudolf TH3} does not depend on the dimension $d$, as long as ${\rm vol}_d(G)$ and $C$ do not depend on $d$. In some applications, however, the upper bound is rather useless since $C=C_d$ is exponentially large in $d$. Assume, for example, that \begin{equation} \label{novak_rudolf eq: dens_normal} \rho(x)= \exp(-\alpha \vert x \vert ^2), \end{equation} i.e. $\rho$ is the non-normalized density of a $N(0, \sqrt{2\alpha^{-1}})$ random variable. We consider scaled versions of $\rho$. If $G=B_d $, then $\exp(\alpha)\rho \in \mathcal{R}_{\exp(\alpha)}$ and if $G=[-1,1]^d$, then $\exp(\alpha d)\rho \in \mathcal{R}_{\exp(\alpha d)}$. This is bad, since $C$, for example $\exp(\alpha)$ or $\exp(\alpha d)$, might depend exponentially on $\alpha$ and $d$. This example shows that we would greatly prefer an upper bound where $C$ is replaced by a power of $\log C$. However, on the class $\mathcal{F}_{C,d}$ this is not possible. The same proof as in \cite[Theorem~1]{novak rudolf:MaNo07} leads to the following lower bound for \emph{all} randomized algorithms. \begin{theorem} Any randomized algorithm $S_n$ that uses $n$ values of $f$ and $\rho$ satisfies the lower bound \[ \sup_{(f,\rho) \in \mathcal{F}_{C,d}} e(S_n,(f,\rho)) \geq \frac{\sqrt{2}}{6} \begin{cases} \sqrt{\frac{C}{2n}} & 2n \geq C-1,\\ \frac{3C}{C+2n-1} & 2n < C-1. \end{cases} \] \end{theorem} The class $\mathcal{F}_{C,d}$ is too large. Thus the error bound is not satisfying. In the following we prove a much better upper bound for a smaller class of densities. Let $G=B_d$ and let $\rho$ be log-concave, i.e. for all $\lambda \in (0,1)$ and for all $x,y\in B_d$ we have \begin{equation} \label{novak_rudolf eq: log_conc} \rho(\lambda x + (1-\lambda) y) \geq \rho(x)^\lambda \rho(y)^{1-\lambda}. \end{equation} Then let \begin{equation} \label{novak_rudolf Rad} \mathcal{R}_{\alpha,d} = \{ \rho \colon B_d \to (0,\infty)\mid \rho\; \mbox{is log-concave},\, \vert \log\rho(x)-\log\rho(y) \vert \leq \alpha \vert x-y \vert \}. \end{equation} We consider log-concave densities where $\log \rho$ is Lipschitz continuous with constant $\alpha$. Note that the setting is more restrictive compared to the previous one. The goal is to get an upper error bound which is polynomially in $\alpha$ and $d$. We consider a \emph{Metropolis algorithm based on a ball walk}. For $\delta >0$ the transition kernel of the $\delta$ ball walk is \[ B_\delta(x,A) = \frac{{\rm vol}_d(A\cap B_\delta(x))}{{\rm vol}_d(B_\delta(0))} + \mathbf{1}_A(x) \left( 1- \frac{{\rm vol}_d(G\cap B_\delta(x))}{{\rm vol}_d(B_\delta(0))} \right), \quad x\in G,\,A\in \mathcal{B}(G), \] where $B_\delta(x)$ denotes the Euclidean ball with radius $\delta$ around $x$. Let $K_{\rho, \delta}$ be the transition kernel of the Metropolis algorithm with ball walk proposal $B_\delta$, let $P_{\rho,\delta}$ be the corresponding transition operator and let $\Lambda_{\rho,\delta}$ be the largest element of the spectrum of $P_{\rho,\delta}-S\colon L_{2}(\pi_\rho) \to L_{2}(\pi_\rho)$. In \cite[Corollary~1]{novak rudolf:MaNo07} the following result is proven. \begin{proposition} \label{novak_rudolf prop: metro_bw_conduct} Let $\rho\in \mathcal{R}_{\alpha,d}$ and let $\delta=\min\{1/\sqrt{d+1},{\alpha^{-1}}\}$. Then, the conductance of $K_{\rho, \delta}$ is bounded from below by \[ \frac{0.0025}{\sqrt{d+1}} \min\left\{\frac{1}{\sqrt{d+1}},\frac{1}{\alpha}\right\}. \] \end{proposition} By Proposition~\ref{novak_rudolf prop: cheeger} and Proposition~\ref{novak_rudolf prop: metro_bw_conduct} we have a lower bound of $1-\Lambda_{\rho,\delta}$. However, to apply Theorem~\ref{novak_rudolf thm: err_bound} we need a lower bound on ${\rm gap}(P_{\rho,\delta})$. Let $\widetilde{K}_{\rho,\delta}$ be the transition kernel of the lazy version of $K_{\rho, \delta}$, i.e. for $x\in G$ and $A\in \mathcal{B}(G)$ holds $ \widetilde{K}_{\rho,\delta}(x,A) = (K_{\rho, \delta}(x,A) + \mathbf{1}_{A}(x) )/2. $ In words, $\widetilde{K}_{\rho,\delta}$ can be described as follows: With probability $1/2$ stay at the current state and with with probability $1/2$ do one step with $K_{\rho, \delta}$. This transition kernel induces a positive semidefinite operator $\widetilde{P}_{\rho,\delta} \colon L_{2}(\pi_\rho) \to L_{2}(\pi_\rho)$ with \[ {\rm gap}(\widetilde{P}_{\rho,\delta}) = \frac{1}{2}(1 + \Lambda_{\rho,\delta}). \] Let \begin{equation} \label{novak_rudolf Falphad} \mathcal{F}_{\alpha,d} = \{ (f,\rho) \colon \rho \in \mathcal{R}_{\alpha,d},\, f\in L_4(\pi_\rho),\, \left \Vert f \right \Vert_{4}\leq 1 \}, \end{equation}\ and recall that $\mathcal{R}_{\alpha,d}$ is defined in \eqref{novak_rudolf Rad}. Note that we assumed $G=B_d$. Now we can apply Theorem~\ref{novak_rudolf thm: err_bound} for the lazy Metropolis algorithm with ball walk proposal $\widetilde{K}_{\rho,\delta}$. \begin{theorem} \label{novak_rudolf thm: metro_bw_mse} Let $\nu$ be the uniform distribution on $B_d$ and let us assmue that $\delta=\min\{1/\sqrt{d+1},\alpha^{-1}\}$. Let $(X_n)_{n\in\mathbb{N}}$ be a Markov chain with transition kernel $\widetilde{K}_{\rho,\delta}$, i.e. the lazy version of the Metropolis algorithm with ball walk proposal $B_\delta$, and initial distribution $\nu$. Let \[ n_0 = \lceil 5.92\cdot 10^{6} (d+1) \max\{ \alpha^2,d+1 \} ( 2\alpha + 4.16) \rceil. \] Then \begin{align*} \sup_{(f,G) \in \mathcal{F}_{\alpha,d}} e_\nu(S_{n,n_0},(f,\rho)) & \leq 1089 \frac{\sqrt{d+1} \max\{ \alpha, \sqrt{d+1}\} }{\sqrt{n}} \\ & \qquad+ 8.38 \cdot 10^5 \frac{(d+1) \max\{ \alpha^2, d+1\} }{n} . \end{align*} \end{theorem} The last theorem states that the number of oracle calls of $f$ and $\rho$ to obtain an error $\varepsilon>0$ is bounded by $\kappa\, d \max\{ \alpha^2,d \}(\varepsilon^2+\alpha)$. Hence the computation of $S(f,\rho)$ is polynomially tractable. Note that $\mathcal{R}_{\alpha,d}$ might be interpreted as a subclass of $\mathcal{R}_C$ with $C=\exp (2\alpha)$ and $G=B_d$, since $ \rho\in \mathcal{R}_{\alpha,d}$ implies $ \exp(2\alpha) \rho/\left \Vert \rho \right \Vert_{\infty} \in \mathcal{R}_{\exp (2\alpha)}. $ Thus, by Theorem~\ref{novak_rudolf thm: metro_bw_mse} we obtain that the number of oracle calls to get an error $\varepsilon$ also depends polynomially on $\log C$, since $C=\exp(2\alpha)$. \section{Open problems and related comments} \begin{itemize} \item We do not know whether an error bound as in Theorem~\ref{novak_rudolf thm: err_bound} holds for $f \in L_2$ if ${\rm gap} (P)>0$. \item In \cite{novak rudolf:RuSc13} error bounds of $S_{n,n_0}$ for $f\in L_p$ with $1 < p \leq 2$ are proven. Then one needs a new error criterion, here the absolute mean error \[ \mathbb{E}_{\nu,K} \vert S_{n,n_0}(f)-S(f) \vert \] is used. If the Markov chain is $L_1$-exponentially convergent, then the error bound decreases with $n^{1/p-1}$. For a Markov chain with $L_2$-spectral gap a similar error bound is shown. \item The tractability results in Theorem~\ref{novak_rudolf thm: har} and Theorem~\ref{novak_rudolf thm: metro_bw_mse} are nice since the degree of the polynomial is small. Nevertheless, the upper bound is not really useful because of the huge constants. Is it possible to prove these or similar results with much smaller constants? \item A related question would be the construction of Markov chain quasi-Monte Carlo methods, see \cite{novak rudolf:ChDiOw11,novak rudolf:DiRuZh13}. Here the idea is to derandomize the Markov chain by using a carefully constructed deterministic sequence of numbers to obtain a sample $x_1,\dots,x_{n+n_0}$. However, explicit constructions with small error bounds are not known. \end{itemize} \input{reference_novak_rudolf} \end{document}
\section{INTRODUCTION} \noindent Along computer science history, interfaces and interactions have been getting more complex. Nowadays computers are everywhere, used by everyone. It is necessary to make them comply with human capabilities, practical to use. This is mostly done by evaluating HCI prior to their public availability. Yet traditional evaluation methods could either be ambiguous, lack real-time recordings, or disrupt the interaction. On the other hand, new technologies arise. Physiological sensors help to improve the ergonomics of human-computer interaction (HCI) \cite{Fairclough2009a}. Systems could be tuned to users by monitoring their mental workload in real-time \cite{Kohlmorgen2007}. Physiological sensors add an insightful information channel. However sensors may be intrusive or require a calibration to record a proper signal, and some are hardly available to consumers. These issues could be resolved by using physiological sensors in HCI evaluation. While designing a user interface (UI) it should be acceptable to add sensors' hindrance to specially enrolled users. Those testers will then help to improve beforehand the UI. Laboratory conditions permit a controlled setup for repeatable measures. Neuroimaging rely on demanding but sensitive sensors. We consider them as an innovative supplement to conventional evaluation methods. Measuring neural activity during HCI can help us to better understand what occurs in the brain when users are interacting \cite{Parasuraman2013}. We highlight in this paper which neuroimaging techniques could be used conveniently within laboratories to overcome the difficulties encountered by traditional evaluation methods alone. We review a repertoire of patterns of users' state which could be used to characterize HCI, and evaluate how neuroimaging objectively measures them. We call those patterns ``constructs'', a term which refers to notions as different as workload and the state of ``flow''. Other papers already began to sense how neurotechnologies benefit HCI, but they do not cover evaluation \cite{George2010}, or if so they do not study many constructs. \cite{Parasuraman2013} only discuss workload, vigilance and error recognition. In the present review we gathered from the HCI literature every major construct which could potentially be evaluated with brain activity. In this review, we first briefly describe the different families of evaluation methods aimed at assessing HCI and UI quality, along with their advantages and drawbacks. We divided them in four categories: behavioral studies (observations of users actions in real-time), inquiries (e.g.~questionnaires, interviews, think aloud), physiological sensors (e.g.~heart rate, galvanic skin response) and neuroimaging (a subset of physiological sensors which records brain activity). We also formalize a new scale (whenever the measure is ``exocentic'' or ``egocentric'') which could help to choose the right combination of methods for evaluations. We show that electroencephalography (EEG) is the neuroimaging technique which offers the best trade-off between spatial and temporal resolution, practical use and cost. Therefore we focus on EEG during the second part. We review there constructs related to the quality of HCI. We identified that workload, attention, vigilance, fatigue, error recognition, emotions, engagement, flow and immersion are useful for evaluation and can be measured with EEG. Finally we outline the challenges and limitations which arise from this encounter between HCI evaluation and neurotechnologies, as well as constructs that could benefit from being measurable with EEG. \section{EVALUATION METHODS} \subsection{Behavioral Studies} Recording users interactions, such as mouse speed, is one standard way to evaluate a UI. ``Behavioral studies'' refers to this method: behavior and actions of users inside a software. Behavioral studies are close to performance measures, as seen in human factors. The easiest way to sense if a UI is well designed is to watch users. How fast do they complete the task? Are they more accurate with a bigger mouse cursor? Such methods helped to formulate a preeminent law in HCI, Fitts's law, which is all about time to reach a target depending on its distance and size \cite{Fitts1954}. Although behavioral studies are able to account in real-time for users' interactions, they can be hard to interpret: measures may not be specific to one construct. E.g. a high reaction time can be caused either by a low concentration level or a high workload \cite{Berka2007}, \cite{Hart1988}. On top of that, behavioral studies may not provide much information on the users' state. With simple tasks in particular, little can be computed beside reaction times and a performance metric. \subsection{Inquiries} \label{inquiries} While it is possible to infer users' thoughts through a behavioral study, it may be simpler to record their opinion. We call this ``inquiries''. In HCI we are interested in questionnaires related to the use of a UI. Standardized questionnaires have been validated across several studies for various measures: e.g.~NASA-TLX for workload \cite{Hart1988}. Unfortunately those ``pen and paper'' tests are discrete and are not good for real-time assessments. The ``think aloud'' protocol \cite{Weber2007} is a way to circumvent this, yet it could influence the interaction as users still have two different things to do: interact and report their experience. It is an example of double task and divided attention \cite{Ogolla2011}. ``Focus groups'' \cite{Bruseberg2002} is the third form of inquiry. It involves experts and advanced users, who exchange about their findings under the control of the designer. Questionnaires, think aloud an focus group are three different forms of inquiry fraught with the same hazards. Resulting measures are prone to be contaminated by ambiguities \cite{Nisbett1977}, social pressure \cite{Picard1995} or participants' memory limitations \cite{Kivikangas2010} -- when answers are not oriented toward experimenters' expectations if subjects figure out what is at stake. \subsection{Physiological Sensors} When humans interact with computers bodily changes co-occurs with mental changes. E.g. pupils dilate while experiencing strong emotions \cite{Partala2003}. Physiological sensors can be used in order to account for such body changes in HCI \cite{Fairclough2009a}, \cite{Dirican2011} or game \cite{Ravaja2009}, \cite{Nacke2009a} research. Galvanic skin response (GSR, also called ``electrodermal activity'') is among those sensors, as well as electrocardiography (ECG, the signal modality heart rate is derived from) and electromyography (EMG, caused by muscular activity, including facial expressions). Even if someone trained could control his heartbeat, physiological cues are great for the ``objectivity'' they bring into HCI (see section \ref{exoego}). Body reactions are sometimes misleading though: you may record ECG to study attention, whereas an increase in heartbeat can also be caused by strong feelings. Muscles and organs are controlled by the peripheral nervous system. Physiological sensors are a second-order inference about the processing which occurs in the central nervous system. \subsection{Neuroimaging} Neuroimaging is a currently rising field used in brain-computer interfaces (BCI) settings \cite{Blankertz2010}, \cite{Hamadicharef2010}. Neuroimaging techniques allows the assessment of brain activity; we classify them apart even if strictly speaking they do belong to physiological sensors. Non-invasive neuroimaging techniques, which do not require surgery, are divided into two main families \cite{Zander2011}. Functional magnetic resonance imaging (fMRI) and functional near-infrared spectroscopy (fNIRS) record brain activity through blood flow variations. fMRI has a very good spatial resolution but is a large device which completely surrounds subjects and costs about one million dollars. fNIRS is a much more lightweight and affordable device. Instead of magnetic fields, it uses direct light for recordings. Sensors are fixed on a cap, hence subjects are free to interact with a computer while wearing it. Compared to fMRI, the spatial resolution of fNIRS is less detailed. It records only the outer region of the brain -- light is absorbed by tissues. fMRI and fNIRS share a poor temporal resolution. With a latency reaching up to several seconds it is difficult to observe fast and short responses. The second family of neuroimaging uses electrical currents generated by neural activity. Magnetoencephalography (MEG) records magnetic fields. It is less heavy and expensive than fMRI, but still hardly manageable for uses in HCI contexts. MEG has a high temporal resolution, down to the millisecond. Electroencephalography (EEG) also has a high temporal resolution. It is comparable in size to fNIRS. EEG measures electrical current onto the scalp. Electrodes are ``dry'' -- no electrolyte solution -- or, more frequently, ``wet'' -- solvent is either water or gel. Despite its poor spatial resolution it is a relatively cheap equipment for a laboratory. Because it is portable and non invasive, it interferes little with HCI setting. Experimenters must be cautious with the limitations of the device they choose. Is the signal-to-noise ratio sufficient for what they intend to measure? What artifacts could pollute their data? Are they in control of the algorithms producing measures from raw signals? That said, EEG is the most promising candidate to assist inquiries and other physiological sensors in a wide range of evaluation measures. Compared to others neuroimaging devices, EEG offers the best compromise between spatial and temporal resolution, practical use and cost. Therefore we focus mostly on this type of brain activity recordings in this paper. \subsection{A New Continuum for HCI Evaluation Methods} \label{exoego} We have previously mentioned how the evaluation methods do bring different levels of ``objectivity'' in their measures. Unfortunately, in such context ``objective'' and ``subjective'' are scarcely defined in the literature. According to \cite{Laar2013}, ``the objective methods are based on overt and covert user responses during interaction while the subjective methods rely on user expressions after the interaction''. From that perspective, inquiries are ``subjective'' while behavioral studies, physiological sensors and neuroimaging are ``objective''. While we agree such a distinction is required, a more rigorous vocabulary is needed. We also doubt the ``time'' variable should be involved in the definition. As stated in section \ref{inquiries}, results of inquiries are prone to social pressure and other self-interpretations, and this is also true for the real-time think aloud. Moreover, when studying emotions, it could be argued that only ``subjective'' feelings are recorded, as the evaluation is centered on the user. Hence, without a complex phrasing (i.e. ``objective measure of subjective feelings''), employing such words is open to criticisms. As an alternative ``direct'' and ``indirect'' could be considered. But then those concepts are more likely to refer to how measures are reported, not where they originate from (e.g.~EMG vs an external observer annotating facial expressions). As such, we would like to introduce a new nomenclature to name those two aspects and avoid ambiguities: exocentric and egocentric. Those terms are borrowed from spacial navigation research \cite{Brandt1973} and bring the notion of the self. Exocentric measures are here close to the stimuli, to the source, while egocentric measures are close to the conscious thoughts of the user, to the outcome. \begin{figure}[htbp] \centering \includegraphics{exo_ego_scale.png} \caption{Proposal of an ``exocentric / egocentric'' scale aimed at classifying evaluation methods for HCI. \label{exo_ego_fig}} \end{figure} We therefore create a continuous space between two extremes (see Figure \ref{exo_ego_fig}). We illustrate this scale with the measurement of pain. The pressure of a needle on a finger would represent a perfect exocentric measure: the stimulus' strength, a value disconnected from human body and perceptions. When the pressure is transmitted to nociceptors in the skin, the measure shifts a little from exocentric to egocentric. As nerves are transmitting signals from the peripheral nervous system to the brain, we go further to the right of the axis. Since we may not be interested in skin's thickness, this neural activity represents the first interesting value from this side of the exo/egocentric scale. Neuroimaging techniques record such activity, hence it is the most exocentric evaluation method. When the signal reaches the central nervous system, autonomic responses are triggered -- increase in heart rate, galvanic skin response \cite{Loggia2011}. Those reactions could be recorded through physiological sensors, a step further from the exocentric extreme. As the pain grows, it will alter behaviors and thoughts. A runner may slow down when experiencing pain in a foot, no matter his willingness. Behavioral studies are able to sense modifications occurring against the will of the subject; that could be placed somewhere in the middle of our scale. Concurrently, most of the time, the person is being aware of the pain and could phrase it if asked to. Many other cognitive processes are involved in such a high level of consciousness (e.g.~planning, awareness), thus measures recorded by inquiries are close to the far-end of the scale and are indeed egocentric. This scale can be used for various evaluations. Eventually, it is possible to add ``objective/subjective'' and ``direct/indirect'' to describe a whole framework. A construct could be objective (usability) or subjective (emotions). A tool could be either direct (sensor) or indirect (observer). A method is more exocentric (neuroimaging) or egocentric (inquiries). E.g. the work of an experimenter assessing workload with ECG can be described as objective/exocentric/direct. \section{CONSTRUCTS} \noindent ``Constructs'' designate the patterns of users' state which could be used to characterize interactions. This part reviews relevant constructs from an HCI evaluation perspective that can be assessed using neuroimaging techniques. We grouped similar measurements. \subsection{Workload} \subsubsection{Definition} Humans have a limited set of resources to process information \cite{Just2003}. The ratio between processing power and data coming from the environment determines mental workload. Workload increases as cognitive resources lessen or as the quantity of demands grows. If the workload is too high subject's performance decreases, sometimes dramatically. \subsubsection{Neuroimaging} Using a device with 9 channels \cite{Berka2007} correlated EEG with workload. With a better equipment \cite{Mathan2007} showed how EEG measures more subtle changes compared to ECG. fNIRS is another well-tried technology: neurons require more energy, hence more oxygen, as the load increases. fNIRS showed better results compared to EEG, with 82\% of correct classifications between 2 classes (low vs high workload) and 50\% with 3 classes (low, medium, high) \cite{Hirshfield2009}. In \cite{Blankertz2010} EEG online analyses (i.e.~in real-time) discriminate 2 classes with a 70\% accuracy. A 2 minutes time window enables scores from 80\% to 90\% \cite{Brouwer2012}. With 2 classes still, reviews report scores close to 100\% if EEG is combined with other physiological sensors \cite{Erp2010}. \cite{Grimes2008} claim 99\% success in distinguishing 2 memory load levels, 88\% with 4. \subsection{Attention -- Vigilance -- Fatigue} \subsubsection{Definition} Attention, vigilance and fatigue are closely related and regularly measured altogether \cite{Oken2006}. ``Attention'' refers to the ability to focus cognitive resources on a particular stimulus \cite{Kivikangas2010}. A correct selective attention allows to ignore distractors. An insufficient attention level results in a difficulty or an inability to complete the task, whereas too high or narrow attention resources may prevent someone to disengage from a sub-task. While in the literature, ``attention'' designates more frequently the ability to perceive changes from the environment, the term ``vigilance'' then often refers to a broader resource, dependent of both cognitive performance and the arousal level on the sleep--wake spectrum \cite{Oken2006}. In that sense it refers to a state of sustained attention. One needs to maintain a high degree of vigilance over time in order to focus his attention on something. Hereby ``alertness'' will be considered as a synonym of ``vigilance''. ``Fatigue'' is a state in which cognitive resources are exhausted. If the required level of vigilance or attention causes a strain too important on the organism, fatigue arises and performances decrease \cite{Boksem2005}. Then the task cannot be performed correctly and errors appear \cite{Erp2010}. \subsubsection{Neuroimaging} The alpha band is associated with attention. When eyes are closed, or when fatigue occurs, alpha waves amplitude increases \cite{Shaw2003}. This frequency band in the range 8-12Hz is mostly generated by the occipital lobe. It is easily recorded with EEG, even with a single electrode \cite{George2011}. Alpha band analysis discriminates different attention levels \cite{Klimesch1998}. Even more, it enables to detect which side of his visual field a subject is paying attention to while his eyes stare in front of him with 70\% accuracy \cite{Trachel2013}. Other types of brain activity are used, such as delays in event-related potentials (ERP) -- e.g.~visual selective attention in \cite{Saavedra2012}. \cite{Berka2007} suggested that EEG is the only sensor which can accurately report attention and vigilance shifts on a second-by-second timeframe. Works investigating vigilance measures are reviewed in \cite{Parasuraman2013}. Regarding fatigue, if EEG signals are not more accurate than physiological sensors to detect microsleeps, they offer the possibility to detect preceding inattentive states \cite[ sec.~3.1]{Blankertz2010}. Mental fatigue has been detected on 4 seconds time windows with 80\% accuracy, or 94\% over 30 seconds \cite{Laurent2013}. In order to improve reliability, additional frequency ranges were recorded in this study. For instance alpha, theta (4-8Hz) and beta (13-18Hz) bands have been combined. ERP on the other hand have been used to study how fatigue impairs differently cognitive processes \cite{Lorist2000}. \subsection{Error Recognition} \subsubsection{Definition} We call ``error recognition'' the situation that occurs when users detect by themselves an outcome different from what is expected \cite{Nieuwenhuis2001}. It can be something users genuinely trigger but then they realize they did a mistake. Or it can happen due to commands erroneously interpreted by the machine. Error recognition does not occur when a negative feedback is given per se \cite{Ferrez2008}. It is a matter of recognition by the user of a faulty event. In UI evaluation, error recognition could be an objective measure of subjective (mis)representations, an objective assessment of how intuitive an HCI is. \subsubsection{Neuroimaging} ERP are ``peaks'' and ``valleys'' in averaged EEG recordings associated with an external event. ERP differ in their ``shapes'', place on the scalp and latency depending on the source of the stimuli or on the underlying cognitive mechanism. One particular kind of ERP has been discovered: error-related potentials (ErrP) \cite{Schalk2000}. They are triggered when an ``error'' occurs. It can be caused by something users themselves did (response ErrP), by an incorrect response from the command they used (interaction ErrP), by something they witnessed from another user (observation ErrP), and also when an explicit negative feedback is given (feedback ErrP). All of which have distinguishable features \cite{Ferrez2008}. Response ErrP and interaction ErrP suit perfectly our definition of ``error recognition''. Brain signals are elicited even when users are not consciously aware of errors \cite{Nieuwenhuis2001}. ErrP have been used to discriminate between incorrect and correct users decisions. In \cite{Chavarriaga2010a} respectively 76\% and 63\% accuracy were obtained to detect observation ErrP in ``single trial'', i.e.~in detecting ErrP for each user's action. These scores are common in the literature: 79\% and 84\% in a task involving interaction ErrP \cite{Ferrez2008}. Accuracy relates to EEG devices' quality. From 70\% with an entry-level headset and non gel-based electrodes \cite{Vi2012} up to 90\% with a more expansive device \cite{Schmidt2012}. While ErrP detection does not reach 100\% (chance is 50\%), those scores are sufficient to improve HCI reliability \cite{Vi2012}. \cite{Sobolewski2013} recorded EEG while subjects use a mouse and have to reach different targets. In one-fourth of the trials the hand-to-cursor mapping is randomly off-set by several degrees. Users do not expect these shifts and the analysis gives first insights that the amplitudes of elicited ErrP could relate to the degree of error. If this result is confirmed we may link error recognition to ``intuitivity'' evaluation. \subsection{Emotions} \subsubsection{Definition} Psychology and neuroscience showed that emotions are connected to high-level reasoning; they are tightly linked to decision-making processes \cite{Damasio1994}. The valence/arousal model is the most commonly used paradigm to categorize emotions \cite{Picard1995}. In this two-dimensional representation, valence is related to hedonic tone and varies from ``negative'' to ``positive'' (e.g.~frustrated vs pleasant); arousal is related to bodily and mental activation and varies from ``calm'' to ``excited'' (e.g.~satisfied vs happy). This model must be applied with caution with some populations. Children hardly make distinction between different arousal levels \cite{Posner2005}. \subsubsection{Neuroimaging} Technologies with the highest temporal resolution, such as MEG or EEG, are more indicated when a dynamic content is involved \cite{Vecchiato2011}. An asymmetry within frequency bands (e.g.~alpha and theta) in the frontal brain could be related to different emotions (valence), such as pleasantness/unpleasantness \cite{Vecchiato2011}. Still, EEG is not yet a reliable sensor to assess emotions. In \cite{Chanel2011} even if EEG was better than the other studied physiological sensors on short period of times, a 56\% accuracy barely suffices for the differentiation of three emotions (chance level is 33\%). Some Papers report high classifications rates. In \cite{Liu2011} 7 emotions are categorized. Authors state a 85\% accuracy for arousal and 90\% for valence. This using only three channels of an EEG headset which is known to be sensitive to EMG artifacts. In pure EEG studies it is important to control for facial expressions (i.e.~EMG signals), because they can be easily recorder by electrodes. This is even more problematic when emotions are involved. Although we have to be cautious when assessing EEG reliability, there is nothing wrong in combining EEG and EMG (or other sensors) to improve overall performance. Despite the lack of clear indicators of affect in EEG, neuroimaging is nevertheless a good lead for novel research in this topic. For example different patterns of EEG signals have been observed depending on the sense (sight or hearing) which induces an emotion \cite{Muhl2011}. It could then be speculated that neuroimaging one day will be able to discriminate which emotion is elicited by which input modality, or which information channel leads to positive and which to negative user experience. \subsection{Engagement -- Flow -- Immersion} \subsubsection{Definition} Definitions of ``engagement'', ``immersion'' and ``flow'' overlap. From \cite{Matthews2002}, task engagement is defined as an ``effortful striving towards task goals''. Authors add that task engagement increases during a demanding cognitive task and decreases when participants perform a sustained and monotonous vigilance task, see also \cite{Fairclough2009a}. In \cite{Chanel2011} ``engagement'' is treated as one particular emotion, expressed as ``positive excited'' in the valence/arousal model. Engagement is at a crossroads between several concepts studied in this paper: workload, attention and emotions. ``Flow'' originates from psychological studies involving challenge and/or creativity. It is a state in which someone is totally involved in what he is doing. Flow happens when the skills of the person meet a sufficient amount of challenge. A too important challenge brings anxiety, for too much skills it is boredom, and too few of both results in apathy \cite{Nacke2009}. Here again, several measures are involved. Challenge relates to workload and the resulting state to emotions. By definition, flow implies engagement. ``Immersion'' is studied mainly in virtual reality (VR) litterature. In \cite{Slater2009} immersion stands for the modalities hardware gives to users, how well devices can preserve fidelity in VR compared to reality. Then the subjective feeling of being in the VR is called ``presence''. Unfortunately this distinction between ``immersion'' and ``presence'' is less clear-cut in other papers, see \cite{Nacke2009}. \subsubsection{Neuroimaging} In neuroimaging literature \cite{Fairclough2009a}, \cite{George2010} engagement assessment studies are mentioned, but they often relate only to sub-components such as workload or attention. \cite{Berka2007} see engagement as a process related to information gathering, visual scanning, and sustained attention. This study managed to discriminate workload and engagement by using EEG but the tasks involved (mental additions, recalls) are close to what is seen elsewhere in attention/vigilance protocols. Engagement is often left entangled with other states in a ``performance'' measure, see \cite[ sec.~3.2]{Blankertz2010}. Experiments conducted during the FUGA project showed that flow could be related to fMRI measures \cite{Ravaja2009}. The analysis with EEG of frequency bands shows different pattern across three conditions of interaction: boredom (i.e.~not engaged), flow and immersion in a pilot study \cite{Nacke2010}. \cite{Berta2013} improved on this work and achieved a 66\% classification accuracy. \section{CHALLENGES} \noindent We saw how constructs relevant to HCI can be investigated with neuroimaging techniques. In this section we will argue that two of them could benefit drastically from neurotechnologies: error recognition and attention. Besides accuracy, both could reach a new level of description. Furthermore we will emphasize the need for the evaluation of a whole HCI to account for constructs of higher level, to study usability and user experience. Finally we have to take care of EEG devices and reliability in order to make it casual for experimenters to use neuroimaging techniques. \subsection{Improving on Constructs} Measuring of two constructs would particularly benefit from improvements in neuroimaging. First, as it may enable a real-time measure of how intuitive a UI is, we would benefit from a continuous and modulated measure of error recognition. We saw how error recognition can be indicated through ErrP \cite{Schalk2000}. This means that it is possible to detect when an interaction runs against users' expectations \cite{Ferrez2008}, i.e.~when it is not intuitive. At the moment only a binary measure and poorly detailed data -- ``an ErrP is detected or not'' -- is reliably obtained. Fortunately it seems possible to measure a modulated ErrP \cite{Sobolewski2013}, thus sensing by how much an operation in the UI has perturbed users. If it is to be confirmed, this would enable a quantitative and qualitative data assessment. We saw how single trial detection can be achieved with EEG. Promising work reported ErrP detection as the movement is occurring, within a 400ms timeframe \cite{Milekovic2013}. At the moment this near continuous detection uses an invasive technique. The construct evolving around attention would be the second one to profit from neuroimaging. To distinguish clearly in their measurements vigilance and fatigue would be one point. On the other hand EEG studies showed that visual artifacts in images or videos are detected by subjects beyond consciousness \cite{Scholler2012}, whenever it is conscious perception or attention \cite{Mustafa2012}. This would suggest that ERP could be used to anticipate how much information users are able to process, before even considering their attention level. A (highly) speculative experimental design where various cues are hidden within sensory modalities in order to elicit evoked potentials would create a ``human bandwidth'' assessment, upstream from vigilance and attention. \subsection{Assessing New Constructs} Three constructs sit apart in our nomenclature. Both usability and comfort are more closely related to UI properties than to users' state, and user experience is entirely based on previously seen measures. Since they are the subject of many HCI papers, it is worth to shape their meaning in this review in order to encourage their assessment with neuroimaging techniques. \subsubsection{Usability -- Comfort} ``Usability'' groups together the notions of ``ease of use'' and ``usefulness'' \cite{Bowman2002}. It relates to speed, accuracy and error rates in task completion. The learnability of UI is also a key point of usability. As such a good affordance of UI elements -- how perceptions of objects induce a proper use -- will improve overall usability. Usability is impacted by UI nature and constrains. E.g. an input device based on body gestures is likely to be more tiring than a joystick, given that it requires more energy from the user. Usability is inextricably bound to users' comfort. Although usability could be investigated through behavioral studies or inquiries \cite{Jankowski2013}, to our knowledge there is no neuroimaging study which accounts solely for this construct. Neuroimaging has been used instead as an indicator, for example workload through fNIRS \cite{Hirshfield2009}. In conjunction with other evaluation methods, real-time recordings from physiological sensors and neuroimaging give additional insights and help to contextualize data \cite{Pikea2012}. \subsubsection{User Experience} For \cite{Mandryk2006} user experience (UX) is a shift from usability analysis by bringing emotions and entertainment into the equation. UX embeds ``usability/comfort'', ``emotions'' and ``engagement/flow/immersion''. UX is a higher comprehension level of what users experience during interactions. \cite{Ravaja2009} compiled various methods to measure media enjoyment. It is possible to refer to UX when studying the social aspect of interactions -- e.g.~GSR is different if the opponent in a sport game is played by a friend or a computer \cite{Mandryk2006}. Assessing UX every time new technologies are used could guide the HCI community in its choices, e.g.~with BCI \cite{Laar2013}. \subsection{Hardware -- Signal Processing} Some limitations observed in EEG research are yet to be resolved to make EEG-based evaluation of HCI more operable. EEG devices, while practical compared to other neuroimaging techniques, take long to set up. Hence experiments can be tedious both for the experimenter and for the subject. This is why there are often only few subjects during EEG or BCI experiments, which is a problem for the reliability of the results. EEG signals contain many potential artifacts (e.g.~muscular activity and electrical parasites); the quality of the device is essential. EEG signals must be calibrated, processed and interpreted carefully. Since a few years new EEG devices have appeared, oriented toward a larger public. Their electrodes use no conductive solution, or water as solvent. These electrodes are faster to set-up -- no more gel to be put on each one after the device has been installed -- but may be less sensitive, see \cite[ sec.~2.1]{Blankertz2010}. Hence some companies, while transforming EEG into a mass-product, bring less reliable technology to the market. Those devices often possess fewer electrodes. Without a cap the electrodes are difficult to place in a standardized position on the scalp. Finally they are often packaged with software development kits which hide the signal processing from the users. Constructs like attention or emotions are then claimed to be directly measured, without further justification or muscular artifact control, see \cite{Heingartner2009}. Nevertheless, while experimenters must be aware of such limits if their intent is to rely solely on brain activity, this increasing appeal in favor of cheap EEG devices is a great opportunity to push forward the use of neuroimaging in HCI. Improvements in signal processing, either in features extraction or classification, could benefit every technology. Constructs, such as emotions, are not yet accurately assessed with pure EEG signals. When too many classes (e.g.~emotions and workload levels) are assessed altogether, the classifier performance drops -- e.g.~see how the ``curse-of-dimensionality'' relates to classifiers' complexity \cite{Friedman1997}. Improvements in mathematical analysis and machine learning algorithms, as well as a better understanding of brain activity, would increase the reliability of the whole system by a great amount and favour every construct. Finally, no matter how lightweight they are, EEG and physiological sensors change the way users interact. Movements could be restrained by the devices (less immersion) and users could perceive a more stressful context, potentially biasing their experience. As a result, a framework integrating physiological sensors and traditional evaluation methods has to be conceived to profit from the potential of these novel methods, while avoiding their limitations and pitfalls. \section{CONCLUSION} \noindent We reviewed how neuroimaging techniques could assess constructs relevant for HCI evaluation. Between the four categories of evaluation methods, inquiries could deliver more qualitative data, while physiological sensors and neuroimaging are exocentric measures (the most ``objective'' measures of subjectively perceived stimuli). It is particularly interesting to combine those methods for constructs otherwise difficult to assess with exactitude, as investigated in many studies \cite{Ravaja2009}, \cite{Nacke2009}, \cite{Erp2010}, \cite{Chanel2011}. Our analysis of neuroimaging techniques focused on EEG as it promises a good trade-off between cost, time resolution and ease of installation. We advocate that neurotechnologies can bring useful insights to HCI evaluation. EEG devices are not yet perfectly reliable and practical to use; hardware and software processing are still evolving. However their cumbersomeness is partially avoided if they are used during a dedicated evaluation phase in the HCI development process, with specially enrolled users (testers). \begin{figure}[htbp] \centering \includegraphics{constructs_simple.png} \caption{One possible view of a simplified characterization of the constructs. In the middle circles are the constructs (dotted = not yet measurable with EEG). The inner circles represent the HCI components the most closely related to the constructs, or on which it would be easier to leverage. The outer circles give a hint about what an evaluation would be useful for. \label{constructs_simple}} \end{figure} We studied workload, attention, vigilance, fatigue, error recognition, emotions, engagement, flow and immersion. Figure \ref{constructs_simple} stimulates thoughts about their relationships with HCI components. Some constructs should benefit more than the others from EEG measures: 1) workload, EEG being more sensible to changes compared to other methods \cite{Mathan2007}; 2) attention, because event related potentials could help to anticipate how many details users register \cite{Mustafa2012}; 3) emotions, with an arousal/valence state measured over a short time-frame \cite{Chanel2011}. Error recognition could hardly be assessed precisely with anything but neuroimaging. Such construct highlights how innovative this evaluation method is. Among the outlined challenges, a continuous and modulated error recognition would greatly help to assess usability and comfort. Next studies should start to combine the various constructs, along with a comprehensive framework which gathers every evaluation method, one's advantages preventing others' drawbacks. This should lead to an increase of the overall user experience. \bibliographystyle{apalike} {\small
\section{Introduction} The Coulomb gas is the infinite system of point particles which carry positive or negative unit electric charges, interact via the electrostatic potential and are subject to thermal disorder. In this paper we consider the {\it neutral} case, in which the total charge of the particle system is zero. (This is the case of major importance in physics; a non-neutral Coulomb gas could also be defined, see Section II.B.2 of \citep{Min87}, and has a different phenomenology.) The mathematical difficulty of the model, as well as the reason of physical interest, stem from the fact that the electrostatic potential in dimension two is very long range: for large $|x|$ it is \begin{equation}} \def\ee{\end{equation}\lb{0} V(x)=-\frac{1}{2\p} \ln |x| + c_E + o(1), \ee where the constant $c_E$ depends on the microscopic regularization. The study of the two-dimensional Coulomb gas began in theoretical physics with the suggestion of \cite{Ber71} and of \cite{KoTh73} that this model, as well as the related classical XY model, undergo a new kind of phase transition, named after them. Shortly after, the Berezinskii-Kosterlitz-Thouless (BKT) transition became one of the fundamental paradigms of the theory of critical phenomena: on the one hand, BKT transitions were predicted for several other two-dimensional toy models, including solid-on-solid models, vertex models, interacting dimers and other lattice systems that can be described in terms of a ``height function'' (see \citep{JKKN77,Kad78,dN83,Nie84,AJMPMT05}); on the other hand, the BKT transition turned out to explain the outcomes of several experiments on real-world systems, such as trapped atomic gases, liquid helium films and arrays of Josephson junctions (see \citep{NeKo77,RGSN81,Min87,HKCBD06,HZGC11} and references therein). A precise description of the phase diagram of the Coulomb gas was elaborated by \cite{Kos74,JKKN77,GiSc89}. The properties of the gas are determined by two parameters: the activity $z$ (large $z$ corresponds to high density of particles) and the inverse temperature $\b$ (large $\b$ corresponds to small thermal disorder). With a non-rigorous renormalization group (RG) argument, Kosterlitz found the picture given in Fig.\ref{f0}, which has the following interpretation. \insertplot{420}{100} {}% {f0}{Diagram of phases: the thicker curve is the BKT transition line.\lb{f0}}{0} The thicker line, $\b=\b_{\rm BKT}(z)$, called BKT transition line, divides the $\b$--$z$ plane into two regions, the {\it dipole phase} on the right and the {\it plasma phase} on the left, which are characterized by a different behavior of the correlations. Let us call {\it charge--$\h$ correlation}, $\r_\h(x-y)$, the system response to a probe of charge $\h\in (0,1]$ in position $x$ and a probe of charge $-\h$ in position $y$; and let us call {\it charge--$\h$ density}, $\r_{1,\h}$, the system response to a probe of charge $\h$ at a point $x$. (A more precise definitions of the former quantity will be given below. The latter quantity is by definition non-zero only if $\h=1$). The truncated charge correlation is $\r^T_\h(x-y)=\r_\h(x-y)-\r_{1,\h}\r_{1,-\h}$. It is expected that: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item For $\b>\b_{\rm BKT}(z)$, the truncated charge correlation display a power law decay \begin{equation}} \def\ee{\end{equation}\lb{1} \r^T_\h(x-y)\sim \frac{C}{|x-y|^{2\kappa}}\;, \ee where $C\=C(z,\b)$ is a prefactor and $\kappa$ is the {\it correlation critical exponent}. Each thinner line is Fig.\ref{f0} is the locus of $(\b,z)$ corresponding to a constant value of the critical exponent \begin{equation}} \def\ee{\end{equation}\lb{2cr} \kappa=\frac{\b_{\rm eff}}{4\p}\h^2 \ee with a $\b_{\rm eff}\=\b_{\rm eff}(z,\b)>8\p$. \item Along the BKT line $\b=\b_{BKT}(z)$ the truncated charge correlations decay as a power law, but with a multiplicative logarithmic correction \begin{equation}} \def\ee{\end{equation}\lb{1m} \r^T_\h(x-y)\sim \begin{cases} \frac{C}{|x-y|^{2\kappa}}(\ln|x-y|)^{\frac12}\qquad &\text{for $\h=\frac12$}, \\ \frac{C}{|x-y|^{2\kappa}\; }(\ln|x-y|)^{-\kappa}\qquad&\text{otherwise}. \end{cases}\; \ee The critical exponent $\kappa$ is constant and given by \pref{2cr} for $\b_{\rm eff}=8\p$. \item For $\b<\b_{BKT}(z)$, truncated charge correlations decay exponentially (but only if specific boundary conditions are imposed). \ed The curves in Fig.\ref{f0} were obtained by Kosterlitz as orbits of the ODE \bal\lb{ds} &\dot{s}(\ell)=-8\p^2 e^{8\p c_E} z(\ell)^2 \notag\\ &\dot{z}(\ell)=-2 s(\ell) z(\ell) \eal where $\ell$ is a length parameter and $s(\ell)=1-\frac{8\p}{\b(\ell)}$. $\b(\ell)$ and $z(\ell)$ are effective parameters, obtained by averaging fluctuations over $\ell$-size subparts of the systems: hence $\b(0)=\b$ and $z(0)=z$ are the true parameters of the model; while $\b(\io)$ and $z(\io)$ are the parameters that determine the long-distance asymptotic behavior of the correlations. The orbits of \pref{ds} are hyperbolas in the $s,z$ variables; a sketch of them is in Fig.\ref{fig1}. \insertplot{600}{150} {}% {f1}{Diagram of phases. The BKT line is the separatrix of the dynamical system; the asterisks denote the semi-line of fixed points. \lb{fig1}}{0} Only the initial data $(s(0), z(0))$ on the right of a separatrix asymptotically evolve to one of the fixed points of the horizontal axis. The separatrix is then identified as the BKT line. The speed of convergence towards the fixed point turns out to be exponential, except when the initial data are along the separatrix: in this case the convergence is much slower \begin{equation}} \def\ee{\end{equation}\lb{negotiate} s(\ell)=2\p e^{4\p c_E} z(\ell)= \frac{s(0)}{1+2s(0) \ell} \ee and this explains the appearance of a logarithmic correction in the truncated charge correlations along the BKT line. This description of the phases diagram was a breakthrough discovery in physics for the theoretical and the experimental implications mentioned at the beginning of this Introduction; however, it has eluded a mathematical validation for a long time. Indeed physicists' results relied on an RG computation at second order in $z$ only; higher orders are difficult to be taken into account for it is not known whether the perturbation theory is ultimately convergent even for small $z$ (see \citep{GN85}). Besides, in the plasma region, the second order approximation of the RG flow is divergent and so scarcely reliable. Remarkably, the exponential decay of the charge correlations in the plasma phase was proven to hold by \cite{Yan87}, although only in a region of the $\b$--$z$ plane that is far from the BKT line and only for $\h=1$. His approach was not based on an RG argument, but rather on an expansion about mean field theory which was used by \cite{BryF80} to prove the Debye screening in dimension three. That said, from now on we will focus on the dipole phase and the BKT line. The fundamental step towards the mathematical understanding of the dipole phase was made by \cite{FrSp81b}: first, by Jensen's inequality, they obtained a power law lower bound for $\r_\h(x-y)$; second, they developed a sophisticated {\it multi-scale} decomposition of $\r_\h(x-y)$ that provides an upper bound that is also power law. Their result, among the most celebrated ones in rigorous statistical mechanics, had however three substantial limitations: 1) the multi-scale method applied only to {\it fractional charge} correlation, i.e. for $\h\in (0,1)$, and 2) only in a region of the dipole phase that is far from the BKT line; 3) the upper and lower bounds, being power-laws with different exponents, cannot rule out the presence of multiplicative logarithmic corrections. Their multi-scale method was later improved in a series of papers \citep{MKP90,Ma90,Br91,MK91} so to make it applicable in a region of dipole phase that touches the BKT line at $z=0$; but the other limitations remained. Noteworthily, Fr\"ohlich-Spencer's calculations suggested an important refinement of the conjectures: for $\b\ge \b_{\rm BKT}(z)$, the correct formula for the critical exponent $\kappa$ cannot be \pref{2cr}, but one should rather expect that \begin{equation}} \def\ee{\end{equation}\lb{2crc} \kappa= \begin{cases} \frac{\b_{\rm eff}}{4\p}\h^2 & {\rm if\ } \h\in (0,\frac12]\\ \frac{\b_{\rm eff}}{4\p}(1-\h)^2 & {\rm if\ } \h\in [\frac12,1)\\ 4& {\rm if\ } \h=1. \end{cases} \ee To our understanding, the second and third of \pref{2crc} were overlooked by physicists, who mostly had in mind applications with $\h\in (0,\frac12]$. Several authors advocated the use of a rigorous RG approach to have a more direct access to the conjectures. This direction was followed by \cite{DH00}, who used the general RG approach of \cite{BY90} and some new bounds for the charged clusters of particles to obtain a convergent series representation of the {\it free energy} of the Coulomb gas. This was an important work because it provides a method to obtain, in the RG scheme, some of the ``power counting estimates'' which are implicit in Kosterlitz's analysis. However, it is based on some technical ideas that appear to be applicable neither to the study of charge correlations anywhere in the dipole phase, nor to the evaluation of the free energy at the BKT transition line. These technical problems have prevented further mathematical progress in the study the two dimensional Coulomb gas for the last ten years. The aim of this and of a previous paper, \citep{Fa12}, is to show that the Brydges-Yau's technique is truly an effective method to deal with the BKT line of the Coulomb gas. In \citep{Fa12}, building on a technical suggestion due to D. Brydges and on the general scheme of \citep{Br07} (see also \citep{Dim09,BrSl10}), we already showed that some difficulties of \citep{DH00} can be avoided; and that a convergent series representation for the free energy {\it along } the BKT line, for $z$ small enough, can be provided. In this paper we take up the mathematically more sophisticated and physically more interesting objective of studying the long-distance decay of fractional charge correlations \pref{1m}, again along the BKT curve and for $z$ small enough. Sharp upper bounds for correlations had already been obtained in the general Brydges-Yau's scheme in the case of a different model, the Dipole gas, \citep{DH92,BrKe94}; however, those approaches do not appear to be directly applicable to correlations displaying an {\it anomalous} decay, such as the power law with logarithmic factors that is expected along the BKT line. Besides, our interest here is the critical exponents, therefore we rather need exact long-distance asymptotic formulas. For these reasons we introduce in this paper a new method to deal with correlations, which is inspired, partially, on the study of \cite{BGPS94} of fermion systems with anomalous critical exponents. For clarity's sake in this paper we only consider the most interesting aspect of the dipole phase: the correlation of two fractional charges for $(\b,z)$ along the BKT line. However, our approach is also applicable to the case of integer charges and everywhere in the dipole phase, at least if $z$ is small enough. Furthermore, we believe that results on the $n$-points correlations and their scaling limits can also be obtained building on a method which was introduced in \citep{Fa06,BFM07} to deal with fermion systems $n$-points correlations. \section{Definition and Results} The electrostatic interaction is usually defined as the inverse Laplacian; in dimension two, however, the subtraction of a divergent term is needed to make sense of it. For $L$ an odd integer and $R$ another integer, consider the finite square lattice $$\L=\Big\{(x_0,x_1)\in \mathbb{Z}^2: \max\{|x_0|, |x_1|\}< \frac{L^{R}}{2}\Big\} $$ endowed with periodic boundary condition. Define the Yukawa interaction on $\L$ with inverse Debye screening length $m>0$ as \begin{equation}} \def\ee{\end{equation}\lb{yuk} W_\L(x;m):={1\over |\L|}\sum_{k\in \L^*}{e^{ikx}\over m^2-\hat\Delta} \def\L{\Lambda} \def\X{\Xi(k)}, \ee where: $\L^*=\{\frac{2\p}{L^{R}}(n_0,n_1): (n_0,n_1)\in \L\}$ is the reciprocal lattice of $\L$; $|\L|=L^{2R}$ is the volume of $\L$; $\hat\Delta} \def\L{\Lambda} \def\X{\Xi(k)=-2\sum_{j=0,1}(1-\cos k_j)$ is the Fourier transform of the discrete Laplacian on $\L$. The {\it two dimensional electrostatic potential} is \begin{equation}} \def\ee{\end{equation}\lb{elp} W_\L(x|0):=\lim_{m\to 0} \left} \def\rgt{\right[W_\L(x;m)-W_\L(0;m)\rgt]= {1\over |\L|}\sum_{k\in \L^*\bs \{0\}}{e^{ikx}-1\over -\hat\Delta} \def\L{\Lambda} \def\X{\Xi(k)}. \ee It is a classical result, \citep{Sto50}, that the large $|x|$ asymptotic formula for the infinite volume limit of $W_\L(x|0)$ is \pref{0}, for the $o(1)$ term that is actually $O(\frac{1}{|x|^2})$ and for $c_E=-\frac{2\g_E +\ln 8}{4\p}$, where $\g_E$ is the Euler's constant. We can now define the probabilistic model. Consider a system of point particles labeled with numbers $j=1, 2, 3,\ldots,n$; a configuration $\o$ is the assignment to each particle $j$ of a charge $\s_j=\pm1$ and of a position $x_j\in \L$. Let $\O^0_n$ be the set of the {\it neutral} configurations of $n$ particles, i.e. the configurations of $n$ particles such that $\s_1+\cdots+\s_n=0$. The total energy of $\o\in\O^0_n$ is \begin{equation}} \def\ee{\end{equation}\lb{ene} H_\L(\o):=\sum_{i<j=1}^n\s_i\s_j W_\L(x_i-x_j|0). \ee We consider $\O^0_0$ as made of one configuration, the ``no particle'' one, with zero total energy. For activity $z\ge 0$ and inverse temperature $\b>0$, the Grand Canonical partition function of the two dimensional Coulomb gas is \begin{equation}} \def\ee{\end{equation}\lb{gf} Z_\L(\b,z):=\sum_{n\ge 0} {z^n\over n!} \sum_{\o\in \O^0_n} e^{-\b H_\L(\o)}. \ee In the previous paper, \citep{Fa12}, we studied the free energy, \begin{equation}} \def\ee{\end{equation}\lb{freeen} p(\b,z):=-\lim_{\L\to \io} \frac1{\b |\L|} \ln Z_\L(\b,z). \ee In this paper we focus on the fractional charge correlation, which is defined as a ratio of partition functions. Consider two probes: $p_1$, which is a particle of charge $\h\in (0,1)$ at the lattice site $x$; and $p_2$, which is a particle of charge $-\h$ at the lattice site $y$. Let $\o\wedge \{p_1,p_2\}$ be the configuration $\o$ augmented of the two probes. Set \begin{equation}} \def\ee{\end{equation}\lb{gfprobe} Z^{p_1,p_2}_\L(\b,z):=\sum_{n\ge 0} {z^n\over n!} \sum_{\o\in \O^0_n} e^{-\b H_\L(\o\wedge \{p_1,p_2\})} \ee (namely the probes contribute to the energy but not to the entropy of the system). The precise definition of $\r_\h(x-y)$ in the Introduction is then \begin{equation}} \def\ee{\end{equation}\lb{infvollim} \r_\h(x-y):=\lim_{\L\to \io} \frac{Z^{p_1,p_2}_\L(\b,z)}{Z_\L(\b,z)}. \ee The invariance of \pref{infvollim} under translations of the probes is a consequence of the definition. The existence of the infinite volume limits will be proved in the theorem below. When $z=0$, the BKT point is at $\b=8\p$; at these values of the parameters and for every $\h$, a simple computation gives \begin{equation}} \def\ee{\end{equation}\lb{0dec} \r_\h(x)= \frac{e^{8\p \h^2 c_E}}{|x|^{4\h^2} }(1+o(1)), \ee where $o(1)$ is vanishing in the limit of $|x|\to \io$. When $z\neq 0$ the situation is more complicated. \begin{theorem}\lb{t1} Fixed $\h\in(0,1)$, there exist an $L_0\=L_0(\h)>1$, a $z_0\=z_0(\h)>0$ and an inverse temperature $\b_{BKT}(z)\ge 8\p$ such that if $L\ge L_0$, $0<z\le z_0$ and $\b=\b_{BKT}(z)$, the limit \pref{infvollim} exists and: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item If $\h\neq \frac12$, then \begin{equation}} \def\ee{\end{equation}\lb{zdec} \r_\h(x)=\r^{(a)}_{\h}(x) + \r^{(b)}_{\h}(x), \ee where, for $x$-independent $f_a,f_b,f$, \bal\lb{zdec1} &\r^{(a)}_{\h}(x)=\frac{e^{8\p \h^2 c_E} +f_a}{|x|^{4\h^2} \left} \def\rgt{\right(1+f\ln|x|\rgt)^{2\h^2} }\left} \def\rgt{\right(1+o(1)\rgt), \notag\\ &\r^{(b)}_{\h}(x)=\frac{f_b}{|x|^{4(1-\h)^2} \left} \def\rgt{\right(1+f\ln|x|\rgt)^{2(1-\h)^2} }\left} \def\rgt{\right(1+o(1)\rgt). \eal \item If $\h=\frac12$, then, for $x$-independent $f_a,f$, \begin{equation}} \def\ee{\end{equation}\lb{zdec2} \r_{\frac12}(x)=\frac12\frac{e^{2\p c_E}+f_a}{|x| }\left} \def\rgt{\right(1+f\ln|x|\rgt)^{\frac12} \left} \def\rgt{\right(1+o(1)\rgt). \ee \ed In the above formulas, $o(1)$ are vanishing terms for $|x|\to\io$; $f=c z$ for $c>0$, $f_b=c(\h)^2 z^2(1+\tilde f_b)$ for $c(\h)>0$; $f_a$, $\tilde f_b$ are vanishing in the limit $z\to 0$. Besides $z_0(\h)$ is such that, for every $[a,b]\subset(0,1)$, one has $\inf\{z_0(\h):\h\in [a,b]\}>0$. \end{theorem} This is the main result of the paper. \brrm \item In the limit $|x|\to \io$, \pref{zdec} and \pref{zdec2} coincide with \pref{1m} for exponent \pref{2crc} and $\b_{\rm eff}=8\p$. For this reason, we identify the curve $\b=\b_{BKT}(z)$ with the Berezinskii-Kosterlitz-Thouless transition line. Whether $\b<\b_{BKT}(z)$ implies an exponential decay of truncated correlations is an open problem; the only available rigorous result, \citep{Yan87}, is for $\b\ll \b_{BKT}(z)$. \item A heuristic interpretation of the result, neglecting for a moment the logarithmic corrections, is the following. A probe charge $\h$ at an inverse temperature $\b=\b_{BKT}(z)$ placed inside the interacting system is equivalent to a ``virtual'' point charge $\h+m$ at inverse temperature $8\p$ placed inside a free system. Here $m$ represents a local fluctuation of unit charges and can be any positive or negative integer value. By choosing the two smallest values of the virtual charge critical exponent, $2(\h+m)^2$, one obtains the leading parts of $\r^{(a)}$ and $\r^{(b)}$ in \pref{zdec}. \item A justification of the logarithmic factor in \pref{zdec} is more subtle and will emerge from the multi-scale approach used in the proof. The different formula for the case $\h=\frac12$ is related to the fact that, continuing with the argument in the previous point, only at this value of $\h$ there are two different values of $m$ that minimize the virtual charge correlation exponent. \item If $\h\in(\frac12, 1)$, the critical exponent of the free case, $2\h^2$, differs from the one of the interacting case, $2(1-\h)^2$; despite that, there is no discontinuity in the behavior of the correlation at $z=0$. Indeed, note that $\r_\h^{(a)}$ has a prefactor $O(1)$, whereas $\r_\h^{(b)}$ has a prefactor $O(z)$. Therefore, the smaller $z$, the larger the threshold distance passed which $\r_\h^{(b)}$ dominates over $\r_\h^{(a)}$; for $z\to 0$ such threshold distance is infinite, and the free case critical exponent is recovered. \item Since the logarithmic corrections have $O(z)$ prefactors, by the same argument of the previous point, in the limit $z\to 0$ the purely power law decay of the free case is recovered. \item The prefactor $\frac12$ in \pref{zdec} is absent in formula for the correlation in the case $z=0$. Again, this is not a sign of discontinuity: as it can be traced in the proof of the Theorem, among the $o(1)$ terms in \pref{zdec} there is one that in the limit $z\to0$ does not vanish, ceases to be subleading and, with its contribution, restores the prefactor $1$ in the leading term. \errm In the next section we provide the detailed renormalization group construction that directly implies Theorem \ref{t1}. The reader with some familiarity with physicists' renormalization group jargon will recognize in the right hand side of \pref{lk} the {\it beta function} of the model; and in the right hand side of \pref{sk} the {\it gamma function}. The major technical novelty of \citep{Fa12}, with respect to \citep{DH00}, was the derivation, in the setting of the Brydges-Yau's expansion, of the dynamical system \pref{lk} and of new bounds to control it. That allowed us to obtain a convergent series representation of the free energy at the BKT transition. The most important contribution of this paper is the introduction, again in the framework of the Brydges-Yau's technique, of {\it renormalization constants} for the observables --namely for the {\it fractional charges}-- which are described by the dynamical system \pref{lk}. That allows us to obtain \pref{zdec} and \pref{zdec2}, partly by bounds and partly by an explicit computation of the leading term of the solution of \pref{lk}. In the forthcoming analysis, we will work with five parameters: given a charge $\h\in (0,1)$ and $0<\t\le \t_0$, we will need $L\ge L_0(\h,\t)$, $A\ge A_0(\h,\t, L)$ and $0<z\le z_0(\h,\t, L, A)$ in order for the results to be valid. We will also have other two parameters, $\a$ and $h\=h(\a)$, which however are eventually fixed by the condition $\a^2=8\p$. Finally, in our notation, $C$, $C_0$, $C_1$ or $c_0$ might represent different prefactors when they appear in different bounds. \section{Strategy of the Proof}\lb{strategy} \subsection{Multiscale approach} Since $W_\L(x-y; m)$ has strictly positive Fourier transform, a Gaussian field $\{\f_x:x\in \L\}$ is defined by assigning zero mean and covariance \begin{equation}} \def\ee{\end{equation}\lb{wruble} \mathbb{E}_{m,\b}\left} \def\rgt{\right[\f_x \f_y\rgt]=\b W_\L(x-y; m). \ee By means of the {\it sine-Gordon transformation}, such a finite-dimensional measure provides a functional integral representation for the partition function \begin{equation}} \def\ee{\end{equation}\lb{whitman0} Z_\L(\b,z)=\lim_{m\to 0}\mathbb{E}_{m,\b}\left} \def\rgt{\right[e^{2z\sum_{x\in\L}\cos \f_x}\rgt], \ee as well as for the correlation \begin{equation}} \def\ee{\end{equation}\lb{whitman} \r_\h(x-y)=\lim_{\L\to\io}{\langle}} \def\ra{{\rangle} e^{i\h(\f_{x}-\f_{y})}\ra_{\L}, \ee where $$ {\langle}} \def\ra{{\rangle}\cdot\ra_\L:= \frac{\lim_{m\to 0}\mathbb{E}_{m,\b}\left} \def\rgt{\right[e^{2z\sum_{x\in\L}\cos \f_x}\;\cdot\;\rgt]} {\lim_{m\to 0}\mathbb{E}_{m,\b}\left} \def\rgt{\right[e^{2z\sum_{x\in\L}\cos \f_x}\rgt]}. $$ The proof of \pref{whitman0} and \pref{whitman} is in Appendix \ref{appA}. In the RG approach it is natural to study \pref{whitman0} and \pref{whitman} through the generating functional of the correlations of $e^{i\h \f_{x}}$: define \begin{equation}} \def\ee{\end{equation}\lb{skf} \O(J,\L):=\lim_{m\to 0}\mathbb{E}_{m,\b}\Big[ e^{2z\sum_{x\in\L} \cos\f_x+ \sum_{x\in\L}\left} \def\rgt{\right(J_{x,+}e^{i\h\f_x}+ J_{x,-} e^{-i\h\f_x}\rgt)}\Big] \ee where $\{J_{x,\s}: x\in \L, \s=\pm1\}$ are real variables; then \bal &p(\b,z)=-\lim_{R\to \io} \frac{1}{\b |\L|}\ln \O(J,\L)\Big|_{J\=0}, \lb{pres0} \\ \lb{correlazione0} &\r_\h(x-y) = \lim_{R\to \io}\frac{1}{Q(J,\L)} \frac{\dpr^2\O(J,\L)}{\dpr J_{x,+}\dpr J_{y,-}}\Big|_{J\=0}. \eal The point of departure of the RG analysis is a multi-scale representation of $\O(J,\L)$. We need some further notations. The two independent unit vector of the lattice are $e_0=(1,0)$ and $e_1=(0,1)$. Consider the set of unit vectors $\hat u=\{\pm e_0, \pm e_1\}$: for any $\m\in \hat u$ define the discrete partial derivative as $\dpr^\m \f_x:=\f_{x+\m}- \f_{x}$ if $\m=e_0, e_1$, or as $\dpr^\m \f_x:=\f_{x}- \f_{x+\m}$ if $\m=-e_0, -e_1$. Correspondingly define the vector component $x^\m:=x\cdot \m$ if $\m=e_0, e_1$, and $x^\m:=-x\cdot \m$ if $\m=-e_0, -e_1$. In our notation, every sum $\sum_{\m\in \hat u}$ will also imply a factor $\frac12$ that we do not write explicitly. This means, for example, that the Fourier transform of $\sum_{\m\in \hat u} \dpr^{-\m} \dpr^\m$ coincides with $\hat \Delta} \def\L{\Lambda} \def\X{\Xi(k)$ defined after \pref{yuk}; and that the discrete form of the first order Taylor expansion of a lattice function $f_y-f_x$ is $\sum_{\m\in \hat u} (\dpr^\m f_x)(y^\m-x^\m)$% \footnote {in the sense that, for any lattice path $p_{x,x}$ that joins $x=(x_0,x_1)$ with $y=(y_0, y_1)$ and has length $|y_0-x_0| +|y_1-x_1|$, $$ \Big|f_y-f_x-\sum_{\m\in \hat u} (\dpr^\m f_x)(y^\m-x^\m)\Big| \le 4\max_{j=0,1} |y_j-x_j|^2 \max_{z\in p_{x,y}}\max_{\m_1,\m_2\in \hat u}|\dpr^{\m_1}\dpr^{\m_2} f_z| $$ }. In Appendix \ref{appA} we prove the multiscale functional integral representation \begin{equation}} \def\ee{\end{equation}\lb{pot0} \O(J,\L)= e^{E|\L|}\lim_{m\to 0} \mathbb{E}_R \cdots \mathbb{E}_0 \left} \def\rgt{\right[e^{\VV(J,\z^{(0)}+\z^{(1)}+ \cdots+ \z^{(R)} )}\rgt], \ee where, fixed any $s\in (0,\frac12)$ and for $\a^2:=\b(1-s)$: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item $E=\frac12\ln (1-s)$ and the interaction $\VV(J,\f)$ is \bal\lb{pot} \VV(J,\f):={s\over 2}\sum_{x\in \L\atop \m\in \hat u}(\dpr^\m\f_x)^2+ z\sum_{x\in\L\atop\s=\pm1}e^{i\s\a\f_x}+\sum_{x\in \L\atop\s=\pm1} J_{x,\s} e^{i\h\a\s\f_x}. \eal \item $\z^{(0)}, \ldots,\z^{(R)}$ are two-by-two independent Gaussian fields, each of which has zero mean and covariance \begin{equation}} \def\ee{\end{equation}\lb{expectations} \mathbb{E}_j[ \z^{(j)}_x\z^{(j)}_y]= \begin{cases} \G_j(x-y)\qquad &\text{for } j=0,1,\ldots, R-1\\ \G'_R(x-y) \qquad & \text{for } j=R. \end{cases} \ee Each $\G_{j}$ is independent of $m$ and $\L$ and, for positive $C_q$ and $c$, \bal \lb{p1}&\G_j(x)=0 &{\rm for\ }|x|\ge L^{j+1}/2, \\ \lb{p2}&|\dpr^{\m_1}\cdots\dpr^{\m_q} \G_j(x)|\le C_q L^{-jq} &\text{for any $\m_j\in \hat u$ and any $q\ge1$}, \\ \lb{p3}&\G_j(0)={1\over 2\p}\ln L + c_j(L) & {\rm for\ } |c_j(L)|\le c L^{-\frac{j}{4}}, \eal The covariance $\G'_R$, instead, depends upon $m$ and $\L$. One has \begin{equation}} \def\ee{\end{equation}\lb{p4a} \lim_{m\to 0} \G'_R(0)=+\io, \ee while, if $\G'_R(x|0):=\G'_R(x)-\G'(0)$, \begin{equation}} \def\ee{\end{equation}\lb{p4b} \lim_{R\to \io}\lim_{m\to 0} \G'_R(x|0)=0. \ee The limit \pref{p4b} implies \begin{equation}} \def\ee{\end{equation}\lb{p4} \G_{\io,0}(x|0):=\sum_{j=0}^\io \left} \def\rgt{\right[\G_{j}(x)-\G_{j}(0)\rgt]= -\frac{1}{2\p} \log |x| +c_E+o(1). \ee \ed The meaning of \pref{p2} and \pref{p1} is that $\G_j$ carries a typical momentum $O(L^{-j})$ and has a compact support of side length $O(L^{j+1})$. The precise construction of $\G'_{R}$ and of $\G_0, \ldots, \G_{R-1}$ was given in \citep{Fa12} building on \citep{BrGuMi04}; a review is in Appendix \ref{appA}. Note that the expectations in \pref{expectations} are independent of $\b$, while the interaction in \pref{pot} is dependent on the new parameters $\a$ and $s$. The relationship among $\a$, $s$ and $\b$ and their role in the forthcoming analysis is the following. The parameter $s\=s(z)$ is introduced so that the curve in Fig.\ref{f0} that corresponds to a system with effective inverse temperature $\a^2$ has graph $\b=\b_{\a}(z)$, where \begin{equation}} \def\ee{\end{equation} \lb{p4} \b_{\a}(z)=\frac{\a^2}{1-s(z)}. \ee Although in many sub-results we will leave an explicit dependence on $\a$, for Theorem \ref{t1} we will eventually set $\a^2=8\p$, which means that in the statement of that Theorem $\b_{ BKT}(z)\=\b_{\sqrt{8\p}}(z)$. The RG approach consists in computing the integrals in \pref{pot0} progressively from the random variable with highest momentum to the one with lowest. First, set \bal\lb{itr} &\O_{1}(J,\f,\L):=e^{E|\L|}\mathbb{E}_0\left} \def\rgt{\right[e^{\VV(J,\f+\z^{(0)})}\rgt]; \eal then, inductively for $j=2,\ldots, R$, set \bal\lb{itrj} &\O_{j}(J,\f,\L) :=\mathbb{E}_{j-1}\left} \def\rgt{\right[\O_{j-1}(\L;J,\f+\z^{(j-1)})\rgt]; \eal at last, one finds \bal\lb{0itr} &\O(J,\L)=\lim_{m\to 0}\mathbb{E}_{R}\left} \def\rgt{\right[\O_{R}(\L;J,\z^{(R)})\rgt]. \eal In this way the evaluation of the partition function is transformed into the evaluation of a sequence of {\it effective generating functionals} $\O_1,\ldots, \O_{R}, \O$. \subsection{Polymer gas representation} Following \citep{BY90,Br07,BrSl10}, each $\O_j$ can be efficiently represented as a {\it polymer gas}. Before describing this formulation, we have to introduce a multiscale decomposition of the lattice and, correspondingly, special types of lattice domains. \\ {\it\0a) Blocks.} Set $|x|:=\max\{|x_0|,|x_1|\}$. Recall that each side of the square lattice $\L$ is made of $L^{R}$ sites, where $L$ is odd; for $j=0,1,\ldots,R$, pave the periodic lattice $\L$ with $L^{2(R-j)}$ disjoint squares of $L^{2j}$ sites, in such a way that there is a central square, $$ \big\{x\in \L:|x|\le L^{j}/2 \big\} $$ and all the other squares are translations of this one by vectors in $L^{j}\mathbb{Z}$. An example is in Fig. \ref{fig2}. \insertplot{420}{220} {}% {f2}{Lattice paving with blocks of different sizes in the case $L=3$ and $R=3$. \lb{fig2} }{0} We call such squares $j$--blocks, and we denote the set of all $j$--blocks by $\BB_j\=\BB_j(\L)$. $0$--blocks are made of single points: $\BB_0=\L$. \\ {\it\0b) Polymers.} A union of two-by-two different $j$--blocks is called $j$--polymer, and the set of all $j$--polymers in $\L$ is denoted ${\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\={\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j(\L)$. Suppose $X$ is a $j$--polymer: $\dpr X$ is the set of sites in $X$ with a nearest neighbor outside $X$; $\dpr_{ext} X$ is the set of sites outside $X$ with a nearest neighbor inside $X$; $\BB_j(X)$ is the set of the $j-$blocks in $X$; $|X|_j$ is the cardinality of $\BB_j(X)$; the {\it closure} $\bar X$ is the smallest polymer in ${\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}(\L)$ that contains $X$. \\ {\it\0c) Connectivity.} A polymer made of two different blocks, $B,B'\in \BB_j$, is connected if there exist $x\in B$ and $x'\in B'$ s.t. $|x-x'|=1$; the definition extends to connected polymers of more blocks in the usual way. For example, in Fig. \ref{fig2} there is one connected 2-polymer, which is the closure of three connected $1$--polymers, which in turn are the closure of ten connected $0$--polymers. ${\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c\={\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_j(\L)$ is the set of the connected $j$--polymers; the collection of the maximal connected parts of a $j$--polymer $X$ (each of which is a $j-$polymer by construction) is called ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)$. \\ {\it\0d) Small polymers.} The polymer $X$ is {\it small} if it is connected and $|X|_j\le 4$. The set of the small $j$--polymers will be called $\SS_j\=\SS_j(\L)$; the set of the connected $j$--polymers that are not small will be called ${\SS\hskip-.6em /}_j\={\SS\hskip-.6em /}_j(\L)$; the number of small $j$--polymers that contain a given $j$--block is independent of $j$ and will be called $S$. The {\it small set neighborhood} of a $j$--polymer $X$ is the set $X^*:=\cup\{Y\in \SS_j:Y\cap X\neq \emptyset\}$. \\ {\it\0d) Empty set.} The empty set is considered as an element of ${\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$, but not of ${\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_j$. We will assume that $L\ge 16$ so that, if $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c$, then the set $X^*\bs X$ is a ``small margin'' around $X$, in the following sense: if $X, Y\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c$ and $\bar X$, $\bar Y$ are separated by at least one $j+1$ block, then \begin{equation}} \def\ee{\end{equation}\lb{separation} \min\{|x-y|: x\in X^*, y\in Y^*\}\ge L^{j+1}- 8L^{j}\ge \frac12 L^{j+1} \ee which, by \pref{p1}, is larger than the range of $\G_j$. Now we pass to the polymer gas representation of the generating functional. Set $\F=(J,\f)$. For each scale $j=1, \ldots, R$, assume that five real parameters, $E_j$ and $t_j:=(s_j, z_j, Z_j, \bar Z_j)$ are given; and assume that $\O_j(\F,\L)$ has the form \bal\lb{pfr} \O_j(\F,\L)=e^{|\L|E_j} \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j} e^{U_j(\F,\L\bs X)} \prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} K_j(\F,Y), \eal where the definitions of the interaction $U_j$ and of the polymer activity $K_j$ follow. Given a $j$--block $B$, the interaction is \begin{equation}} \def\ee{\end{equation}\lb{uu} U_j(\F,B)=V_j(\F,B)+W_j(\F,B). \ee The first term, $V_j$, is similar to the initial interaction \pref{pot} and is the sum of $V_{0,j}$ and $V_{1,j}$, for \bal \lb{vv} &V_{0,j}(\f,B)={s_j\over 2}\sum_{x\in B\atop \m\in \hat u} (\dpr^\m\f_x)^2+z_jL^{-2j}\sum_{x\in B\atop \s=\pm1}e^{i\a\s\f_x}, \notag\\ &V_{1,j}(\F,B)= Z_j L^{-2j}\sum_{x\in B\atop \s=\pm1}J_{x,\s}e^{i\h\a\s\f_x}+ \bar Z_j L^{-2j}\sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\bar\h\a\s\f_x}. \eal Here, $\bar \h:=\h-1$; therefore, as $\h\in (0,1)$, also $-\bar\h\in (0,1)$. The factors $L^{-2j}$ make $V_j$ explicitly dependent on the scale $j$; besides, $V_j$ depends upon the fields $\{\f_x: x\in B\cup \dpr_{ext} B\}$ and $\{J_{x,\s}: x\in B, \s=\pm1\}$ and upon the parameters $t_j$. Note that $z_j$ and $s_j$ play the role of the the effective parameters discussed in the Introduction; whereas $Z_j$ and $\bar Z_j$ are the ``fractional charge renormalization constants''. The second term in \pref{uu}, $W_j(\F,B)$, is generated by the multi-scale integration: $W_0(\F,B)=0$; while, for $j\ge 1$, inductively assume that $W_j(\F,B)$ depends upon the scale $j$, upon the fields $\{\f_x, J_{x,\s}:x\in B^*, \s=\pm1\}$, and upon the parameters $t_j$. We give now a partially explicit formula for $W_j$; the $w$'s functions that appear in \pref{wj0}, \pref{wj1} and \pref{wj2} will be defined in Section \ref{sc20}. $W_j$ is the sum of three terms: $W_{0,j}(\f,B)$, $W_{1,j}(\F,B)$ and $W_{2,j}(\F,B)$, where the enumeration corresponds to the powers of $J$ as we now explain. $W_{0,j}$ contains terms that are independent of $J$ and quadratic in $s_j$, $z_j$: \bal\lb{wj0} &W_{0,j}(\f,B) = -s^2_j \sum_{y\in \mathbb{Z}^2\atop \m, \nu} \def\p{\pi} \def\r{\rho\in \hat u} w_{0,a,j}^{\m\nu} \def\p{\pi} \def\r{\rho}(y) \sum_{x\in B}(\dpr^\m\f_x) \Big[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f_x)\Big] \notag\\ &+z^2_j\sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y) \sum_{x\in B\atop\s=\pm} \Bigg[ e^{i\s\a(\f_x-\f_{x+y})} -1 +|y|^2\frac{\a^2}4\sum_{\m\hat u}(\dpr^{\m}\f_x)^2 \Bigg] \notag\\ &+z^2_j\sum_{y\in \mathbb{Z}^2} w_{0,c,j}(y) \sum_{x\in B\atop\s=\pm} e^{i\s\a(\f_x+\f_{x+y})} \notag\\ &+z_j s_j \sum_{y\in \mathbb{Z}^2\atop \m\in \hat u} w^\m_{0,d,j}(y) \sum_{x\in B\atop\s=\pm} i\s\left} \def\rgt{\right[ e^{i\s\a\f_x}(\dpr^\m\f_{x+y})- e^{i\s\a\f_{x+y}}(\dpr^{-\m}\f_x)\rgt] \notag\\ &-z_j s_j \sum_{y\in \mathbb{Z}^2} w_{0,e,j}(y)\sum_{x\in B\atop\s=\pm} \left} \def\rgt{\right(e^{i\s\a\f_{x+y}}-e^{i\s\a\f_x}\rgt). \eal $W_{1,j}$ contains terms linear in $J$, and linear in $s_j$ or $z_j$: \bal\lb{wj1} &W_{1,j}(\F, B)= z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} w_{1,b,j}(y)\sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s(\h\f_x+\f_{x+y})} \notag\\ &+z_j \bar Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} \bar w_{1,b,j}(y)\sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s(\bar \h\f_x-\f_{x+y})} \notag\\ &+z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} w_{1,c,j}(y)\sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s\bar \h\f_x} \left} \def\rgt{\right[e^{-i\a\s(\f_{x+y}-\f_x)}-1+i\a\s y^\m\sum_{\m\in \hat u}(\dpr^\m \f_{x})\rgt] \notag\\ & +z_j \bar Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} \bar w_{1,c,j}(y)\sum_{x\in B\atop \s=\pm} J_{x,\s}e^{i\a\s\h\f_x}\left} \def\rgt{\right[ e^{i\a\s(\f_{x+y}-\f_x)} - 1-i\a\s y^\m\sum_{\m\in \hat u}(\dpr^\m \f_{x})\rgt] \notag\\ & + s_j Z_j L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in \hat u}w^\nu} \def\p{\pi} \def\r{\rho_{1,d,j} (y) \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\h\a\s\f_x} \s \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x})\rgt] \notag\\ & + s_j \bar Z_j L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in \hat u}\bar w^\nu} \def\p{\pi} \def\r{\rho _{1,d,j} (y) \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\bar\h\a\s\f_x} \s \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x})\rgt]. \eal Finally, $W_{2,j}$ contains the terms quadratic in $J$, and independent of $s$ or $z$: \bal\lb{wj2} &W_{2,j}(\F, B)= \sum_{y\in\mathbb{Z}^2\atop\e=\pm}w^\e_{2,a,j}(y) \sum_{x\in B\atop \s=\pm}J_{x,\s}J_{x+y,\s\e} e^{i\h\a\s(\f_x+\e\f_{x+y})} \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop\e=\pm}\bar w^\e_{2,a,j}(y) \sum_{x\in B\atop \s=\pm}J_{x,\s}J_{x+y,\s\e} e^{i\bar\h\a\s(\f_x+\e \f_{x+y})} \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop\e=\pm}w^\e_{2,b,j}(y) \sum_{x\in B\atop \s=\pm} J_{x,\s}J_{x+y,\s\e} \left} \def\rgt{\right[e^{i\a\s(\h\f_x+\e\bar\h \f_{x+y})}+ e^{i\a\s(\bar\h\f_x+\e\h \f_{x+y})}\rgt] \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop \e=\pm}w^\e_{2,c,j}(y) \sum_{x\in B\atop \s=\pm} J_{x,\s}J_{x+y,\e\s} e^{i\a\s (1+\e)(\h-\frac12)\f_x}. \eal We extend these definitions from $j$--blocks to $j$--polymers additively: for $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$: \begin{equation}} \def\ee{\end{equation}\lb{nt} U_j(\F,X):= \sum_{B\in \BB_j(X)}U_j(\F,B); \ee $V_j(\F,X)$ and $W_j(\F,X)$ are defined in the same way. Returning to the explanation of \pref{pfr}, the polymer activity, $K_j(\F,X)$, is also generate by the multi-scale integration: $K_0(\F,X)=0$; while, for $j\ge 1$, $K_j(\F,X)$ depends upon $\{\f_x, J_{x,\s}:x\in X^*, \s=\pm1\}$ and is the sum of four terms, \begin{equation}} \def\ee{\end{equation}\lb{pinning0} K_j(\F,X)=K_{0,j}(\f,X)+K_{1,j}(\F,X)+K_{2,j}(\F,X) +K_{\ge3,j}(\F,X) \ee where, again, the enumeration refers to the powers of $J$. The last term is proportional to the third power or an higher power of $J$: it will not play any role in the analysis of this paper, since eventually we are only interested in up to two derivatives in $J$ at $J=0$. The second and third terms can be further decomposed: \bal\lb{pinning} K_{1,j}(\F,X)= &L^{-2j}\sum_{x\in X\atop \s=\pm1} J_{x,\s} \left} \def\rgt{\right[Z_j K_{1,j}(\f,X, x,\s) +\bar Z_j K^\dagger_{1,j}(\f,X, x,\s)\rgt], \notag\\ K_{2,j}(\F,X)= &\sum_{x_1\in X, x_2\in X^*\atop \s_1,\s_2=\pm1} J_{\s_1, x_1}J_{\s_2, x_2}K_{2,j}(\f,X,x_1,\s_1,x_2,\s_2). \eal Note that $K_{1,j}(\f,X, x,\s)$ and $K^\dagger_{1,j}(\f,X, x,\s)$ are ``pinned'' at $x$ in the sense that they are defined by \pref{pinning} only for $x\in X$; we extend their definitions by setting $K_{1,j}(\f,X, x,\s)=K^\dagger_{1,j}(\f,X, x,\s)=0$ whenever $x\not\in X$. In the same way, $K_{2,j}(\f,X,x_1,\s_1,x_2,\s_2)$ is pinned at $x_1$ and $x_2$ and we set $K_{2,j}(\f,X,x_1,\s_1,x_2,\s_2)=0$ if $x_1\not\in X$ or $x_2\not\in X^*$. Besides note that at least one power of $J$ is assumed to be restricted to the set $X$ (indeed, the same sort of dependence in $J$ is assumed in \pref{wj1} and \pref{wj2}). This completes the explanation of the inductive assumption \pref{pfr}. As we read from \pref{pot0} and \pref{pot}, \pref{pfr} holds at $j=0$, for $$ E_0\=E=\frac12\ln(1-s), \qquad (s_0,z_0, Z_0, \bar Z_0) =(s,z,1,0),\qquad W_0\=K_0\=0. $$ We shall see that it also holds by induction for any $j=1,2, \ldots, R$, with: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item Effective couplings $(s_j, z_j)$ and effective polymer activity $K_{0,j}$ given by \bal\lb{lk} &s_{j+1}=s_j-a_j z_j^2+\FF_j \notag\\ &z_{j+1}=L^2 e^{-\frac{\a^2}2 \G_j(0)}\left} \def\rgt{\right[z_j-b_j s_jz_j+\MM_j\rgt] \notag\\ &K_{0,j+1}=\LL_{0,j} + {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}, \eal for coefficients $a_j$,$b_j$, and functionals $\FF_j\=\FF_j(K_{0,j})$, $\MM_j\=\MM_j(K_{0,j})$, ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}\={\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}(z_j, s_j, K_{0,j})$ and $\LL_j\=\LL_j(K_{0,j})$. The functionals $\FF_j$, $\MM_j$ and ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}$ will play the role of ``remainder parts'' with respect to the other terms in the equation. The functional $\LL_{0,j}$ will be a contraction with respect to suitable norms. \item Effective free energy $E_j$ given by \begin{equation}} \def\ee{\end{equation}\lb{elk} E_{j+1}=E_j + L^{-2j} \left} \def\rgt{\right[\EE_{1,j}+ s_j \EE_{2,j}+ s^2_j \EE_{3,j}+ z^2_j \EE_{4,j}\rgt], \ee for coefficients $\EE_{2,j}$, $\EE_{3,j}$, $\EE_{3,j}$ and for a functional $\EE_{1,j}\=\EE_{1,j}(K_{0,j})$. \item Fractional charge renormalization constants $Z_j$ and $\bar Z_j$, \bal\lb{sk} &Z_{j+1}= L^2 e^{-\h^2\frac{\a^2}2\G_j(0)} \left} \def\rgt{\right[(1 -s_j m_{1,1,j}+\MM_{1,1,j}) Z_j+\left} \def\rgt{\right(z_j m_{1,2,j} +\MM_{1,2,j}\rgt)\bar Z_j\rgt], \notag\\ &\bar Z_{j+1}= L^2 e^{-\bar\h \frac{\a^2}2\G_j(0)} \left} \def\rgt{\right[(1- s_j m_{2,2,j}+\MM_{2,2,j}) \bar Z_j +\left} \def\rgt{\right(z_j m_{2,1,j} +\MM_{2,1,j}\rgt)Z_j\rgt], \notag\\ &K_{1,j+1}=\LL_{1,j} + {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}, \eal for coefficients $\{m_{p,q,j}:p,q=1,2\}$ and functionals $\{\MM_{p,q,j}\=\MM_{p,q,j}(K_{1,j}):p,q=1,2\}$, $\LL_{1,j}\=\LL_{1,j}(K_{1,j})$ and ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}\={\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}(s_j,z_j,K_{0,j}, K_{1,j})$. The functional $\LL_{1,j}$ will be a contraction with respect to suitable norms. \ed For every $j=0,1,\ldots, R$, all the coefficients and functionals appearing in \pref{lk}, \pref{elk} and \pref{sk} are independent of $\L$: this will simplify the discussion of the calculation of the limit $\L\to \io$. Note that at $\a^2=8\p$, because of \pref{p3}, $L^2 e^{-\frac{8\p}2 \G_j(0)}\sim 1$ and the map \pref{lk} is our rigorous counterpart of Kadanoff's ODE for the effective coupling constants, \pref{ds}. Note also that \pref{elk} and \pref{sk} depend on the flow \pref{lk}, but do not affect it; therefore the study of \pref{lk} done in \citep{Fa12} remains valid for the developments of this paper. The last step of the RG is \bal\lb{0pfr} \O(J,\L) &=e^{E_R|\L|} \lim_{m\to 0}\mathbb{E}_{R}\left} \def\rgt{\right[e^{U_{R}(J,\z^{(R)},\L)}+K_R(J,\z^{(R)}, \L)\rgt]. \eal Suppressing the dependence in the set $\L$ of interactions and polymer activities, and setting $\d E_R:=E_{R+1}-E_R$, $\x_x=\x_x^{(R)}$, $\F_x=(J_x, \x_x^{(R)})$, we have: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item For the free energy, \bal\lb{finalen} &\frac1{|\L|}\ln \O(0,\L):=E_{R+1} \notag\\ &=E_R +L^{-2R}\lim_{m\to 0}\ln\mathbb{E}_R\left} \def\rgt{\right[1+\left} \def\rgt{\right(e^{V_{0,R}(\z) + W_{0,R}(\z)}-1\rgt) + K_{0,R}(\z)\rgt]. \eal \item For the fractional charge correlation \bal\lb{finalcorr} &\frac{\dpr^2 \O}{\dpr J_{x,+}\dpr J_{0,-}}(0,\L) =e^{-\d E_R|\L|} \notag\\ &\qquad\times \lim_{m\to 0}\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} \left} \def\rgt{\right(\frac{\dpr V_{1,R}(\F)}{\dpr J_{x,+}} +\frac{\dpr W_{1,R}(\F)}{\dpr J_{x,+}}\rgt)\left} \def\rgt{\right(\frac{\dpr V_{1,R}(\F)}{\dpr J_{0,-}} +\frac{\dpr W_{1,R}(\F)}{\dpr J_{0,-}}\rgt)\rgt]_{J=0} \notag\\ & +e^{-\d E_R|\L|} \lim_{m\to 0}\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} \frac{\dpr^2 W_{2,R}(\F)}{\dpr J_{x,+}\dpr J_{0,-}} +\frac{\dpr^2 K_{2,R}(\F)}{\dpr J_{x,+}\dpr J_{0,-}}\rgt]_{J=0}. \eal \ed \subsection{Bounds on the RG map} To control the limit $R\to \io$ of \pref{finalen} and \pref{finalcorr}, we need bounds for all the intermediate steps of the RG map. In the previous paper, \citep{Fa12}, we dealt with \pref{lk}, \pref{elk} and the formula for the free energy \pref{pres0}. We showed that there exists a unique choice of the initial value $s$ as function of $z$ such that the limit for $j\to \io$ of $s_j$ $z_j$ and $K_j$ is vanishing. More precisely, we found the following results. \begin{lemma}[\citep{Fa12}] \lb{scott} Consider the coefficients $a_j$ and $b_j$ in \pref{lk}. For $\a^2=8\p$, there exists a $j$-independent $C\=C(L)$ and a number $\tilde c_E$ such that \bal\lb{scott1} &|a_j-a|\le C L^{-\frac14 j}, &&|b_j-b|\le C L^{-\frac14 j}, \eal where $a=8\p^2 e^{-8\p \tilde c_E} \ln L$ and $b=2\ln L$. \end{lemma} The constant $\tilde c_E$ in this Lemma is not the same as $c_E$ in \pref{0} --although it has a similar origin; note, however, that $\tilde c_E$ will not explicitly appear in the final results \pref{zdec} and \pref{zdec2}. For stating the next results, set, for any $j\ge 1$, \begin{equation}} \def\ee{\end{equation}\lb{negotiate0} q_j:=\frac {q_1}{1+|q_1| (j-1)}, \qquad q_1:= \sqrt{ab} z_1. \ee Hence $q_1=z_1 4\p e^{4\p \tilde c_E} \ln L $ and $q_j$ is almost a discrete version of $2 s(\ell)$ in \pref{negotiate}. Given two parameters, $h>0$ and $A>1$, in Section \ref{s5} we will introduce the norm $\|\cdot\|_{h,T_j}\=\|\cdot\|_{h,T_j}(A)$, that will measure the size of polymer activities. \begin{theorem}[\citep{Fa12}]\lb{t.sm} Given a $\t>0$ small enough, for $L$ and $A$ large enough, there exists an $\e\=\e(A,L,\t)$ such that the following statement holds. If $0<z\le \e$, there exists a unique $s\=s(z)$ such that the solution of \pref{lk} with initial data $(z_0,s_0)=(z, s(z))$ satisfies \bal\lb{slk} &\left} \def\rgt{\right|s_j-\frac{|q_j|}{b}\rgt|\le \frac{\t}{b}\frac{|q_1|}{[1+|q_1|(j-1)]^{\frac32}}, \notag\\ &\left} \def\rgt{\right|z_j-\frac{q_j}{\sqrt {ab}}\rgt|\le \frac{\t}{\sqrt {ab}}\frac{|q_1|}{[1+|q_1|(j-1)]^{\frac32}}, \notag\\ &\|K_{0,j}\|_{h,T_j}\le \frac{\t^2 |q_1|^2}{[1+|q_1|(j-1)]^3}, \eal for all $j=1, \ldots, R$. Besides, the choice of the parameters $L$, $A$, $\e$ and the function $s(z)$ are independent of $|\L|$. \end{theorem} As anticipated, the $s(z)$ found in this Theorem determines the graph of the BKT transition line, $\b=\b_{\rm BKT}(z)$, via \pref{p4}. This result was instrumental to control \pref{elk} and to prove the convergence of \pref{pres0}. \begin{theorem}[\citep{Fa12}]\lb{t3.5} There exists $C\=C(\a,L)$ such that, given any $j=0,1,\ldots, R$, if $|s_j|, |z_j|,\|K_{0,j}\|_{h,T_j}\le c_0|q_j|$, then \begin{equation}} \def\ee{\end{equation}\lb{r2} |E_{j+1}-E_{j}|\le C L^{-2j}|q_j|. \ee Besides, $E_0, \ldots, E_{R}$ (but not $E_{R+1}$) are independent of $|\L|$. \end{theorem} The consequence of this result is a convergent series representation of the free energy $$ p(\b,z)=-\frac1{2\b} \log(1-s(z))-\frac1\b\sum_{j\ge 0} (E_{j+1}-E_j), $$ which was the main result of \citep{Fa12}. In this paper we study \pref{sk} and \pref{finalcorr}. For this task, we need to introduce a norm for activities with one pinning point, $\|\cdot\|_{1,h,T}$, and a norm for activities with two pinning points, $\|\cdot\|_{2,h,T}$, see discussion after \pref{pinning}; such norms will be defined in Section \ref{s5}. In the following result, we control the activities $K_{1,j}$ and $K^\dagger_{1,j}$. \begin{theorem}\lb{leibler} There exists a $C\=C(\a)>0$ such that, under the same hypothesis of Theorem \ref{t.sm}, \begin{equation}} \def\ee{\end{equation}\lb{leibler0} \|K_{1,j}\|_{1,h,T_j}\le C |q_j|^2,\qquad \|K^\dagger_{1,j}\|_{1,h,T_j}\le C |q_j|^2. \ee \end{theorem} The proof is in Section \ref{7.1}. Next, we study the coefficients in the flow \pref{sk}. \begin{lemma}\lb{t3.1bb} There exists a $j$-independent $C\=C(\a,L)$ such that, for any $p=1,2$, \begin{equation}} \def\ee{\end{equation}\lb{scott2ab} |\MM_{p,1,j}|\le C A^{-1} \|K_{1,j}\|_{1,h,T_j},\qquad |\MM_{p,2,j}|\le C A^{-1} \|K^\dagger_{1,j}\|_{1,h,T_j}. \ee \end{lemma} \begin{lemma}\lb{t3.1b} Consider $a$ and $b$ given in Lemma \ref{scott}. There exists a $j$-independent $C\=C(\a,L)$ such that: if $\a^2\ge 8\p$ and $p,q=1,2$, \begin{equation}} \def\ee{\end{equation}\lb{scott2a} |m_{q,p,j}|\le C; \ee besides, if $\a^2=8\p$, \bal |m_{1,1,j}-\h^2 b|\le C L^{-\frac j4}, \qquad |m_{2,2,j}-\bar \h^2 b|\le C L^{-\frac j4}; \lb{scott2} \eal finally, if $\a^2=8\p$ and $\h=-\bar \h= \frac12$, then $\MM_{1,1,j}=\MM_{2,2,j}$, $\MM_{1,2,j}=\MM_{1,2,j}$ and \bal |m_{2,1,j}-\frac{\sqrt{ab}}2|\le C L^{-\frac j4}, \qquad |m_{1,2,j}-\frac{\sqrt{ab}}2|\le C L^{-\frac j4}. \lb{scott3}\eal \end{lemma} This Lemma does not provide the exact asymptotic values of $m_{2,1,j}$ and $m_{1,2,j}$ if $\h\neq \frac12$; however, they will not be necessary for studying \pref{sk}. To formulate the next result, set $Z^+_j:=Z_j+\bar Z_j$, $Z^-_j:=Z_j-\bar Z_j$ and $$ g_j:=-\p \sum_{k=1}^j[\G_k(0)-\frac 1{2\p}\log L], $$ which is a bounded sequence because of \pref{p3}. \begin{theorem}\lb{leiblerbis} In the same hypothesis of Theorem \ref{t.sm}, for $j=1,\ldots,R$: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item If $\h=-\bar \h=\frac12$, there exist two coefficients $\{c_\s: \s=\pm\}$ that are vanishing for $z\to 0$ and are such that \bal\lb{leibler1} &Z^+_{j+1}=Z_{1}^+ L^{\frac32 j}(1+|q_1| j)^{\frac14}e^{g_j+c_++r_{1,j}}, \notag\\ &Z^-_{j+1}=Z^-_1 L^{\frac32 j}(1+|q_1| j)^{-\frac34}e^{g_j+c_{-}+r_{2,j}}; \eal in the above formulas, for a constant $C$ and for $m=1,2$, $$ |r_{m,j}|\le C \frac{\t}{\sqrt{1+|q_1|j}}. $$ \item If $0\le \h<\frac12$, there exist two coefficients, $c_1$, $c_2$, which are vanishing in the limit $z\to 0$ and are such that \bal\lb{leibler2} &Z_{j+1}=L^{2j(1-\h^2)}(1+|q_1| j)^{-\h^2} e^{4\h^2 g_j+c_{1}} \left} \def\rgt{\right[e^{r_{1,j}}Z_1+c_{2}e^{s_{1,j}}\bar Z_1 \rgt], \notag\\ &\bar Z_{j+1}=L^{2j(1-\h^2)}(1+|q_1| j)^{-\h^2} e^{4\h^2 g_j}\left} \def\rgt{\right[r_{2,j}Z_1+s_{2,j}\bar Z_1 \rgt], \eal where, for a $C_0\=C_0(\h)$ and any $m=1,2$, $$ |r_{m,j}|\le C_0 \frac{\t}{\sqrt{1+|q_1|j}}, \qquad |s_{m,j}|\le C_0 \frac{1}{\sqrt{1+|q_1|j}}+C_0 L^{-2(\bar \h^2-\h^2)j}. $$ A formula for $c_2$ is, for a $c(\h)>0$, $$ c_2=z e^{4\p (\bar\h^2-\h^2)\G_{0}(0)} \left} \def\rgt{\right[c(\h) - m_{1,2,0}\rgt]+ O(z^{\frac32}). $$ \item If $\frac12 <\h <1$, \pref{leibler2} holds after interchanging $Z_j$ with $\bar Z_j$ and $\h$ with $-\bar \h$ (hence the formula for $c_2$ becomes $z e^{4\p (\h^2-\bar \h^2)\G_{0}(0)}[c(-\bar \h)-m_{2,1,0}] + O(z^{\frac32})$). \ed Finally, for every $\h\in (0,1)$, \bal\lb{12:57am} &Z_1=L^{2}e^{-\h^2 \frac{\a^2}{2}\G_0(0)}(1+O(z)), \notag\\ &\bar Z_1=L^{2}e^{-\bar \h^2 \frac{\a^2}{2}\G_0(0)} m_{2,1,0}z. \eal \end{theorem} \begin{theorem}\lb{pherson} Under the same hypothesis of Theorem \ref{t.sm} and if $A\ge e^2$, there exists a $C>0$ such that, for any $j=1,2,\ldots, R$ (suppressing the dependence in the variables $\f,X,x_1,\s_1,x_2,\s_2$), \bal\lb{maldacena} K_{2,j}=&\sum_{k=0}^j 2^{-(j-k)} e^{-L^{-k}|x_1-x_2|} L^{-4k} \notag\\ &\times\left} \def\rgt{\right[Z_k^2 K^{(a,k)}_{2,j}+\bar Z_k^2 K^{(\bar a,k)}_{2,j}+ Z_k\bar Z_k K^{(b,k)}_{2,j}\rgt], \eal where, for any $\d=a,\bar a, b$, \begin{equation}} \def\ee{\end{equation}\lb{maldacena1} \|K^{(\d,k)}_{2,j}\|_{2,h,T_j} \le C|q_k|. \ee \end{theorem} As a consequence of the above Theorems we can finally turn to the calculation of the fractional charge correlation. Consider the $w$'s function in \pref{wj2}. \begin{theorem} \lb{finale} The limits \bal\lb{canaletto0} &w^-_{2,a}(x):=\lim_{R\to \io}w^-_{2,a, R}(x)\qquad w^-_{2,\bar a}(x):=\lim_{R\to \io}\bar w^-_{2,a, R}(x) \notag\\ &w^-_{2,b}(x):=\lim_{R\to \io}w^-_{2,b, R}(x)\qquad w^-_{2,c}(x):=\lim_{R\to \io} w^-_{2,c, R}(x) \eal exist and, under the same hypothesis of Theorem \ref{t.sm}, \begin{equation}} \def\ee{\end{equation}\lb{canaletto} \lim_{R\to \io}\frac{\dpr^2 \O}{\dpr J_{x,+}\dpr J_{0,-}}(0,\L) =2w_{2,a}^-(x)+ 2 w_{2,\bar a}^-(x) +2w_{2,c}^-(x). \ee (While $w^-_{2,b}(x)$ does not contribute to the correlation.) \end{theorem} The last ingredient for the proof of the main Theorem is then an exact evaluation of the long $|x|$ asymptotic formulas for the functions in \pref{canaletto0}. \begin{theorem}\lb{finalevero} For coefficients $f, f_a, f_{\bar a}, \tilde f_{b}$ that are vanishing for $z\to 0$, and for a constant $C$: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item If $\h=-\bar\h=\frac12$, then, for $\d=a,\bar a$, \bal &w^-_{2,\d}(x)= \frac{e^{2\p c_E} + f_{\d}}{8|x|} \left} \def\rgt{\right(1+f \ln |x|\rgt)^{\frac12}(1+o(1)), \\ &|w^-_{2,c}(x)|\le \frac{C}{|x|}\left} \def\rgt{\right(1+f\ln |x|\rgt)^{-\frac12}. \eal \item If $\h\neq \frac12$, then, for the same $c(\h)$ of Theorem \ref{leiblerbis}, \bal\lb{finale1} &w^-_{2,a}(x)+w^-_{2,\bar a}(x) =\frac{e^{8\p \h^2 c_E}+f_a}{2|x|^{4\h^2}}\left} \def\rgt{\right(1+f\ln |x|\rgt)^{-2\h^2}(1+o(1)) \notag\\ &\qquad\qquad\qquad+\frac{c(\h)^2 z^2(1+\tilde f_b)}{2|x|^{4\bar \h^2}} \left} \def\rgt{\right(1+f\ln |x|\rgt)^{-2\bar \h^2}(1+o(1)), \\ &|w^-_{2,c}(x)|\le \frac{C}{|x|^{4\h^2}}\left} \def\rgt{\right(1+f\ln |x|\rgt)^{-2\h^2-1} +\frac{C}{|x|^{4\bar \h^2}}\left} \def\rgt{\right(1+f\ln |x|\rgt)^{-2\bar \h^2-1}. \eal \ed (While, for every $\h\in (0,1)$, $w^-_{2,b}(x)=0$.) Besides, $f=4\p e^{4\p \tilde c_E}L^2 e^{-4\p\G_0(0)} z$. \end{theorem} Our main result, Theorem \ref{t1}, is then a direct consequence of Theorem \ref{finale} and Theorem \ref{finalevero}. \section{Dimensional bounds}\lb{s5} Here we set up scale dependent norms that we will use to control the size of the polymer activities. We will also show how norms encode the {\it dimensional analysis} used in physics to adapt renormalization group ideas to this model. \subsection{Norms and regulators: definitions}\lb{nr} We mainly follow \citep{Br07}. Let $j\in \mathbb{N}$. For $n=0,1,2$ and for $\dpr^\m$ the discrete derivative introduced before \pref{pot0}, define \begin{equation}} \def\ee{\end{equation}\lb{ln} \|\nabla^n_j\f\|_{L^\io (X^*)}:=\max_{\m_1,\ldots,\m_n\atop \m_j\in \hat u}\max_{x\in X^*} L^{nj}\big|\dpr^{\m_1}\cdots \dpr^{\m_n} \f_x\big|. \ee For $X$ a connected $j$--polymer, let ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)$ be the linear space of the functions $\f:X^*\to \mathbb{C}$ with norm $$ \|\f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}:=\max_{n=0,1,2}\|\nabla_j^n\f\|_{L^\io(X^*)}. $$ Observe that $\nabla_j$ is $L^j \dpr$, which makes the norm explicitly scale dependent; besides, we are using the notation ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)$ even though the domain involved in the definition of the norm is the set $X^*$. Let $\NN_j(X)$ be the space of the smooth complex activities of the polymer $X^*$, i.e. the set of $C^\io$ functions $F(\f,X):{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)\rightarrow \mathbb{C}$. The $n$-order derivative of $F$ along the directions $f_1,\ldots,f_n\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)$ is \begin{equation}} \def\ee{\end{equation}\lb{fdiff} D^n F(\f,X)\cdot (f_1,\ldots,f_n) :=\sum_{x_1,\ldots,x_n\in X^*}(f_1)_{x_1}\cdots (f_n)_{x_n} {\dpr^n F\over \dpr \f_{x_1}\cdots\dpr \f_{x_n}}(\f,X). \ee Again, despite the notation $\NN_j(X)$, the relevant set here is the bigger set $X^*$. The size of the differential of order $n$ is given by \begin{equation}} \def\ee{\end{equation}\lb{tjn} \|D^n F(\f,X)\|_{T^n_j(\f,X)}:=\sup_{\|f_i\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}=1} \big|D^n F(\f,X)\cdot (f_1,\ldots,f_n)\big|. \ee Then, given any $h>1$, define the norm \bal\lb{sn} &\|F(\f,X)\|_{h,T_j(\f,X)} :=\sum_{n\ge 0}{h^n\over n!}\|D^n F(\f,X)\|_{T^n_j(\f,X)}. \eal In order to control the norm of the activities as function of the field $\f$, for any scale $j$ and any $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_j$ introduce the {\it field regulators}, $G_j(\f,X)\ge 1$, that depends upon derivatives of $\f$ only. An explicit choice will be provided below. Then, define \bal\lb{rn} \|F(X)\|_{h,T_j(X)} &:=\sup_{\f\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}{\|F(\f,X)\|_{h,T_j(\f,X)}\over G_j(\f,X)}. \eal Finally, we have to weight the polymer activity w.r.t. the size of the set. Given a parameter $A>1$, define \bal\lb{an} \|F\|_{h,T_j}\=\|F\|_{h,T_j}(A):=\sup_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c} A^{|X|_j}\|F(X) \|_{h,T_j(X)}. \eal Inspired by the discussion after \pref{pinning}, given a charge $\s=\pm$ and a lattice point $x$, we will call an activity of the form $F(\f,X, x, \s)$ pinned at the lattice point $x$ if $F(\f,X, x, \s)=0$ whenever $x\not\in X$. For such activities, we define $$ \|F\|_{1,h,T_j}:=\sup_{x\in \mathbb{Z}^2,\s=\pm}\|F(\cdot,\cdot, x, \s)\|_{h,T_j}(A). $$ Likewise, given two charges $\s, \s'=\pm1$ and two lattice points $x$ and $x'$, an activity $F(\f,X, x, \s, x', \s')$ is pinned at $x$ and $x'$ if $F(\f,X, x, \s, x', \s')=0$ whenever $x\not\in X$ or $x'\not\in X^*$. For such activities we set $$ \|F\|_{2,h,T_j}:=\sup_{x,x'\in \mathbb{Z}^2\atop \s,\s'=\pm} \|F(\cdot,\cdot, x, \s, x', \s')\|_{h,T_j} (A^{\frac12}). $$ In the last definition, note that the weight in the size of the polymer has been reduced to $A^{\frac12}$. This concludes the set up of the norms, except for the choice of some parameters and functions that were involved in the definition. The parameter $h\=h(\a)$ is chosen to be $h:=\max\{1, 2 \mathfrak{h}_j(\a):j\ge 0\}$, where \begin{equation}} \def\ee{\end{equation}\lb{hut} \mathfrak{h}_j(\a):=\max\{\|h_j\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}: 0\in X\in \SS_j\} \ee and $h_j(x)$ is the function $\a[\G_j(x)-\G_j(0)]$. The usefulness of this choice will become clear in Appendix \ref{ptdh}. It is not difficult to see that, by \pref{p2}, $\mathfrak{h}_j$ is bounded in $j$ and so the definition of the constant $h$ makes sense. The parameter $A$ will be chosen large enough in various points below. Next, we have to choose $G_j$. Here we follow \citep{Fa12}. Given two positive constants $c_1$, $c_3$, and a positive function of $L$, $\kappa_L$, if $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c$, the function $G_j$ is such that \begin{equation}} \def\ee{\end{equation}\lb{6.69} \ln G_j(\f,X)=c_1\kappa_L \|\nabla_j\f\|^2_{L^2_j(X)}+c_3\kappa_L\|\nabla_j\f\|^2_{L^2_j(\dpr X)} +c_1\kappa_L W_j(\nabla^2_j\f,X)^2, \ee where we have used $L^2$-type norms \bal\lb{ln2} &\|\nabla^n_j\f\|^2_{L^2_j(X)}:= L^{-2j}\sum_{x\in X}\sum_{\m_1,\ldots,\m_n} L^{2nj}\big|\dpr^{\m_1}\cdots \dpr^{\m_n} \f_x\big|^2, \notag\\ &\|\nabla^n_j\f\|^2_{L^2_j(\dpr X)}:= L^{-j}\sum_{x\in \dpr X}\sum_{\m_1,\ldots,\m_n} L^{2nj}\big|\dpr^{\m_1}\cdots \dpr^{\m_n} \f_x\big|^2, \eal \begin{equation}} \def\ee{\end{equation}\lb{6.109} W_j(\f, X)^2:=\sum_{B\in \BB_j(X)} \|\f\|^2_{L^{\io}(B^*)}. \ee To control the field dependence of $U_j$ we shall occasionally use an auxiliary field regulator, called {\it strong field regulator}, $G^{\rm str}_j$: for $B\in \BB_j$ and $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$, \begin{equation}} \def\ee{\end{equation}\lb{6.66} \ln G^{\rm str}_j(\f,B):=\kappa_L \max_{n=1,2}\|\nabla^n_j\f\|^2_{L^\io(B^*)}, \quad G_j^{\rm str}(\f,X):=\prod_{B\in \BB_{j}(X)}G_j^{\rm str}(\f,B). \ee \subsection{Norms and regulators: properties} First, it is important to observe that $\NN_j(X)$ with the norm $\|\cdot\|_{h,T_j(X)}$ is a Banach space. We now list some useful features of the field regulators. As apparent from the definition, if $X\in P_{j+1}$, \begin{equation}} \def\ee{\end{equation}\lb{nw} G^{\rm str}_j(\f',X)\le G^{\rm str}_{j+1}(\f',X). \ee Consider a polymer $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$. From the definitions, we have $L^2_j(X)=\sum_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} L^2(Y)$. Besides, since two $Y$'s in ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)$ have disjoint boundaries, we also have $L^2_j(\dpr X)=\sum_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} L^2(\dpr Y)$. Therefore \begin{equation}} \def\ee{\end{equation}\lb{6.51} \prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)}G_j(\f,Y)=G_j(\f,X). \ee For the following results to hold, $c_3$ and $c_1$ must be large enough, but independently of the scale $j$ and the size $L$. Unless otherwise stated, $j=0,1,\ldots, R-1$. \begin{lemma}\lb{l6.100b} For any polymer $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$, \bal\lb{6.100b} G^{\rm str}_j(\f,X)\le G_j(\f,X). \eal For any polymer $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$ and any block $B\in \BB_j$, but $B$ not inside $X$, \bal\lb{6.52} G^{\rm str}_j(\f,B)G_j(\f,X)\le G_j(\f,B\cup X). \eal \end{lemma} This Lemma corresponds to formula (6.52) of \citep{Br07}: the proof can be found in that paper after Lemma 6.21. The role of the field regulators in the forthcoming analysis is to have a standard function to integrate with respect to the Gaussian measures. \begin{lemma}\lb{l6.53} Let $\kappa_L=c(\log L)^{-1}$ with $c>0$ and small enough. \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item For $j=0, 1, \ldots, R-1$ and any connected polymer $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c$, \begin{equation}} \def\ee{\end{equation}\lb{6.54} \mathbb{E}_j \left} \def\rgt{\right[G_j(\f,X)\rgt]\le 2^{|X|_j} G_{j+1}(\f',\bar X); \ee if instead $j=R$, \begin{equation}} \def\ee{\end{equation}\lb{ex6.53bis} \mathbb{E}_R \left} \def\rgt{\right[G_R(\f,\L)\rgt]\le 2. \ee \item For $j=0, 1, \ldots, R-1$, $m=1,2,3$ and any small polymer $X\in \SS_j$, there exists a $C_m>1$ such that \bal\lb{6.58} \left} \def\rgt{\right(1+\max_{n=1,2}\|\nabla^n_{j+1}\f'\|_{L^\io(X^*)}\rgt)^{m} \mathbb{E}_j \left} \def\rgt{\right[G_{j}(\f,X)\rgt]\le \frac{C_m}{\kappa_L^{m/2}} 2^{|X|_j} G_{j+1}(\f',\bar X); \eal besides the last formula holds even if $G_{j}(\f,X)$ on the left hand side member is replaced by $\sup_{t\in[0,1]} G_{j}(t\f'+\z,X)$. \ed \end{lemma} The proof is in Section D of \citep{Fa12}. From the definitions set up so far, we can derive some simple bounds that will be needed in the next section. For any $\f\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_{j+1}(X)$, we have $\|\f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}\le \|\f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_{j+1}(X)}$, so that, for any $F\in \NN_j(X)$ \begin{equation}} \def\ee{\end{equation}\lb{6.8} \|F(\f,X)\|_{h,T_{j+1}(\f,X)} \le\|F(\f,X)\|_{h,T_{j}(\f,X)} . \ee If $Y\subset X$, for any $\f\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)$ we have $\|\f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(Y)}\le \|\f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}$, so that ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)\subset {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(Y)$ and \begin{equation}} \def\ee{\end{equation}\lb{6.36} \|F(\f,X)\|_{h,T_{j}(\f,X)} \le\|F(\f,X)\|_{h,T_{j}(\f,Y)}. \ee For any two polymers $Y_1, Y_2$ not necessarily disjoint and such that $Y_1\cup Y_2\subset X$, and any two polymer activities, $F_{1}\in\NN_j(Y_1)$ and $F_{2}\in\NN_j(Y_2)$, we have: a generalized {\it triangular inequality} \begin{equation}} \def\ee{\end{equation}\lb{6.5} \|F_{1}(\f,Y_1)+F_{2}(\f,Y_2)\|_{h,T_{j}(\f,X)} \le\|F_{1}(\f,Y_1)\|_{h,T_{j}(\f, Y_1)} + \|F_{2}(\f, Y_2)\|_{h,T_{j}(\f, Y_2)} , \ee (which is stronger than the usual triangular inequality because different norms appear in the two members); and the {\it factorization property} \begin{equation}} \def\ee{\end{equation}\lb{6.37} \|F_{1}(\f,Y_1)F_{2}(\f,Y_2)\|_{h,T_{j}(\f,X)} \le\|F_{1}(\f,Y_1)\|_{h,T_{j}(\f, Y_1)} \|F_{2}(\f, Y_2)\|_{h,T_{j}(\f, Y_2)}. \ee Details of the proofs of these inequalities are in \citep{Br07}. Finally, given $\|F(X)\|_{h,T_{j}(X)}$, in order to have an estimate of the size of $\|F\|_{h,T_{j+1}}$ one needs to sum over the position of the polymer. Let us consider separately the case of configurations on small sets and on large sets. For $\l\in (0,1)$ and $\r=s,l$, set \bal\lb{gld} &k_\r(A,\l):=\sup_{V\in P^c_{j+1}}A^{|V|_{j+1}} \sum_{Y\in O_\r}^{\bar Y=V}(\l A)^{-|Y|_j}, \eal where $O_s=\SS_j$ and $O_l= \not\!\SS_j$. Besides, consider also the case of a pinning point in the sum and set \bal\lb{gld2} k^*_s(A,\l):=\sup_{V\in P^c_{j+1}}A^{|V|_{j+1}}\sup_{x\in V} \sum_{Y\in \SS_j\atop Y\ni x}^{\bar Y=V}(\l A)^{-|Y|_j}. \eal Note that $k_s(A,\l)$, $k_l(A,\l)$ and $k^*_s(A,\l)$ are $j$-independent, and so the notation is consistent. \begin{lemma}\lb{l6.90} There exist $c >0$ and $\vartheta} \def\O{\Omega} \def\Y{\Upsilon>0$ such that, for $A$ large enough \begin{equation}} \def\ee{\end{equation}\lb{6.90} k_s(A,\l)\le c L^2,\qquad k^*_s(A,\l)\le c,\qquad k_l(A,\l)\le A^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon}. \ee \end{lemma} For the proof see Lemma 6.19 and Lemma 6.18 in \citep{Br07}. In brief, when the sum is over small sets and there is no pinning point, the bound is proportional to a volume factor $L^2$; when the sum is over large sets, the bound is finite in $L$ and vanishing for large $A$. \subsection{Dimensional analysis} We now return to the actual polymer activities of our RG treatment of the Coulomb Gas. To reproduce the physicists' analysis we first need to decompose the polymer activity into terms which represents clusters of particles with given total charge. To do so, note that $K_{0,j}$ contains terms that, as functions of the fields, either are periodic of period $2\p/\a$ or are derivative terms; therefore $K_{0,j}$ is invariant under $\f_\xx\to \f_\xx+ \frac{2\p}{\a}t$ for any constant, integer field $t$. As explained Appendix \ref{abCG}, such invariance provides via a Fourier analysis the following decomposition into {\it charged components} for $K_{0,j}$ as well as for $K_{1,j}$, $K^\dagger_{1,j}$ and $K^{(\d,k)}_{2,j}$. \begin{lemma}\lb{ldec3} For $j=0, 1, \ldots, R$ and for any $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_j$, \bal\lb{dec3} &K_{0,j}(\f,X)=\sum_{q\in \mathbb{Z}} \hK_{0,j}(q,\f,X), \notag\\ &K_{1,j}(\f,X,x,\s)= \sum_{q\in \mathbb{Z}} \hK_{1,j}(q,\f,X,x,\s), \notag\\ &K^\dagger_{1,j}(\f,X,x,\s)= \sum_{q\in \mathbb{Z}} \hK^\dagger_{1,j}(q,\f,X,x,\s), \eal and, for $\d=a,\bar a, b$ \begin{equation}} \def\ee{\end{equation}\lb{dec3.0} K^{(\d,k)}_{2,j}(\f,X,x,\s,x',\s')= \sum_{q\in \mathbb{Z}} \hK^{(\d,k)}_{2,j}(q,\f,X,x,\s,x',\s'). \ee The above series are absolutely convergent and, if $\vartheta} \def\O{\Omega} \def\Y{\Upsilon$ is a constant field, \bal\lb{dec4} &\hK_{0,j}(q,\f,X) = e^{i q\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK_{0,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X), \notag\\ &\hK_{1,j}(q,\f,X,x,\s)= e^{i (q+\h\s)\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK_{1,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X,x,\s), \notag\\ &\hK^\dagger_{1,j}(q,\f,X,x,\s)= e^{i (q+\bar \h \s)\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK^\dagger_{1,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X,x,\s), \notag\\ &\hK^{(a,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\h \s+\h \s')\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK^{(a,k)}_{2,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X,x,\s,x',\s'), \notag\\ &\hK^{(\bar a,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\bar \h \s+\bar \h\s')\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK^{(a,k)}_{2,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X,x,\s,x',\s'), \notag\\ &\hK^{(b,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\h \s+\bar \h\s')\a\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\hK^{(a,k)}_{2,j}(q,\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X,x,\s,x',\s'). \eal Besides, \begin{equation}} \def\ee{\end{equation}\lb{cin0} \|\hK_{0,j}\|_{h,T_j}\le \|K_{0,j}\|_{h,T_j}, \ee \begin{equation}} \def\ee{\end{equation}\lb{cin1} \|\hK_{1,j}\|_{1,h,T_j}\le \|K_{1,j}\|_{1,h,T_j},\qquad \|\hK^\dagger_{1,j}\|_{1,h,T_j}\le \|K^\dagger_{1,j}\|_{1,h,T_j}, \ee and for $\d=a,\bar a, b$ \begin{equation}} \def\ee{\end{equation}\lb{cin2} \|\hK^{(\d,k)}_{2,j}\|_{2,h,T_j}\le \|K^{(\d,k)}_{2,j}\|_{2,h,T_j}. \ee \end{lemma} The meaning of \pref{dec4} is: $\hK_{0,j}(q,\f,X)$ represents clusters of particles with a total charge $q$; $\hK_{1,j}(q,\f,X,x,\s)$ and $\hK^\dagger_{1,j}(q,\f,X,x,\s)$ represent clusters of particle with total charge $q+\h \s$ and $q+\bar \h \s$ respectively; similarly for $\hK^{(\d,k)}_{2,j}(q,\f,X,x,\s,x',\s')$. Now we can discuss the typical bound we need in the rest of the paper. By \pref{6.8} and \pref{6.54}, for any connected polymer $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j^c$ \bal\lb{6.4} \|\mathbb{E}_j\left} \def\rgt{\right[K_{0,j}(\f,X)\rgt]\|_{h,T_{j+1}(\f',X)} \le \|K_{0,j}\|_{h,T_j}\left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X); \eal and, by \pref{cin0}, for the charged component $\hK_{0,j}(q,\f,X)$, \bal\lb{6.4bis} \|\mathbb{E}_j\left} \def\rgt{\right[\hK_{0,j}(q,\f,X)\rgt]\|_{h,T_{j+1}(\f',X)} \le \|K_{0,j}\|_{h,T_j}\left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X); \eal similar bounds can be derived for $\hK_{1,j}(q,\f,X,x,\s)$ and $\hK^\dagger_{1,j}(q,\f,X,x,\s)$; and also for $\hK^{(\d,k)}_{2,j}(q,\f,X,x,\s,x',\s')$. Then we could use \pref{6.90} to sum over the polymer $X$. However, following this procedure, the sum over the small polymers $X$ will generate a bound proportional to the {\it volume factor} $L^2$, which would exponentially increase the size of the bound for $\|K_{0,j}\|_{h,T_j}$ at each step. To avoid that, we need to improve \pref{6.4} and \pref{6.4bis} whenever $X$ is a small set to beat such an $L^2$. Observe that we passed from scale $j+1$ to scale $j$ by the bound \pref{6.8} which is of general validity. Under special circumstances, this step can be done in a more efficient way. To formulate the next results in a simplified notation, in general we will say that $F(\f,X)$ is a {\it charge $p$ activity} if, for any constant complex field $\vartheta} \def\O{\Omega} \def\Y{\Upsilon$, one has $$ F(\f,X)=e^{i\a p \vartheta} \def\O{\Omega} \def\Y{\Upsilon} F(\f-\vartheta} \def\O{\Omega} \def\Y{\Upsilon,X). $$ \begin{theorem}\lb{dh} Consider a charge $p$ activity $F(\f,X)$, with $X\in \SS_j$. There exists a $C\=C(\a)$ such that, \begin{equation}} \def\ee{\end{equation}\lb{eqdh} \|\mathbb{E}_j\left} \def\rgt{\right[F(\f,X)\rgt]\|_{h,T_{j+1}(\f',X)}\le\r(p,\a) \|F\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X) \ee for a ``dimensional factor'' $$ \r(p,\a)= C^{1+|p|} L^{-d(p)\frac{\a^2}{4\p}}, $$ where $d(p)=p^2$ if $|p|\le 1$ and $d(p)=2|p|-1$ otherwise. \end{theorem} \pref{eqdh} differs from \pref{6.4} by the prefactor $\r(p,\a)$. The proof, mostly borrowed from \citep{DH00}, is in Appendix \ref{ptdh}. As an application consider the charged components of $K_{0,j}$ and of $\hat K_{1,j}$ with total charge $p:|p|> 1$. Setting $F(\f):=\hat K_{0,j}(q, \f,X)$, the hypothesis of the theorem is satisfied for $p=q$; therefore \bal &\|\mathbb{E}_j\left} \def\rgt{\right[\hat K_{0,j}(q, \f,X)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le C^{1+|q|} L^{-(2|q|-1)\frac{\a^2}{4\p}}\|K_{0,j}\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal Considering that $\a^2\ge 8\p$, if $|q|\neq 0,1$ and $L$ is large enough, the prefactor $C^{1+|q|} L^{-(2|q|-1)\frac{\a^2}{4\p}}\le (C^2L^{-3})^{|q|}$ beats the volume factor $L^2$ that will be generated by \pref{6.90} once we sum the above bound over $X\in \SS_j$. The same conclusion holds for $\hat K_{1,j}(q,\f,X,x,\s)$. Indeed, the theorem applies with $p=q+\h \s$ and we have \bal\lb{eqdh1} &\|\mathbb{E}_j\left} \def\rgt{\right[\hat K_{1,j}(q,\f,X,x,\s)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le C^{2|q+\h \s|} L^{-d(q+\h \s)\frac{\a^2}{4\p}}\|K_{1,j}\|_{1,h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal Therefore, if $|q+\h\s|>1$, the prefactor is $C^{2|q+\h \s|} L^{-(2|q+\h \s|-1)\frac{\a^2}{4\p}}\le (C^2L^{-2})^{|q+\h\s|}$ and beats the volume factor $L^2$. For completeness, we also state that \bal\lb{eqdh1bar} &\|\mathbb{E}_j\left} \def\rgt{\right[\hat K^\dagger_{1,j}(q,\f,X,x,\s)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le C^{2|q+\bar \h \s|} L^{-d(q+\bar \h \s)\frac{\a^2}{4\p}}\|K^\dagger_{1,j}\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal Finally, for $\d=a,\bar a, b$ and $0\le k\le j$, \bal\lb{lili} &\|\mathbb{E}_j\left} \def\rgt{\right[\hat K^{(\d,k)}_{2,j}(q,\f,X,x,\s,x',\s')\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le C^{1+|p|} L^{-d(p)\frac{\a^2}{4\p}}\|K^{(\d,k)}_{2,j}\|_{2,h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X), \eal where $$ p= \begin{cases} q+\h(\s+\s')\qquad & \text{ if $\d=a$}\\ q+\bar \h(\s+\s')\qquad & \text{ if $\d=\bar a$}\\ q+(\h\s+\bar \h \s')\qquad & \text{ if $\d=b$}\\ \end{cases} . $$ For other terms for which the above power counting improvement is not sufficient we need to extract some finite order of the Taylor expansion, which we now define. Let $F(\x,X)$ be a smooth function of the field $\{\x_x:x\in X^*\}$; the $n$-order Taylor expansion of $F(\x,X)$ at $\x=0$ is \bal\lb{defTay} &(\Tay_{n,\x} F)(\x,X) :=\sum_{m=0}^n {1\over m!} \sum_{x_1\ldots, x_m\in X^*}\x_{x_1}\cdots\x_{x_m} {\dpr^m F\over \dpr \x_{x_1}\cdots \dpr \x_{x_m}}(0,X); \eal the $n$-order remainder is \begin{equation}} \def\ee{\end{equation}\lb{defRem} (\Rem_{n,\x} F)(\x,X):= F(\x,X)-(\Tay_{n,\x} F)(\x,X). \ee The next theorem provides the power counting improvement in such cases. \begin{theorem}\lb{by} Consider a charge $p$ activity $F(\f,X)$ with support $X\in \SS_j$ and fix any point $x_0\in X$. For any $m\in \mathbb{N}$, there exist $C\=C(\a)$ and $C_m$ such that, if $(\d\f)_x:=\f_x-\f_{x_0}$ \begin{equation}} \def\ee{\end{equation} \|\Rem_{m,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[F(\f,X)\rgt]\|_{h,T_{j+1}(\f',X)}\le \r_m(p,\a) \|F\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X) \ee for a ``dimensional factor'' $$ \r_m(p,\a):= C^{1+|p|} C_m L^{-d(p)\frac{\a^2}{4\p}}(\sqrt{\kappa_L} L)^{-(m+1)} $$ where, again, $d(p)=p^2$ if $|p|\le 1$ and $d(p)=2|p|-1$ otherwise. \end{theorem} The proof of this theorem, mostly borrowed from \citep{Fa12}, is in Appendix \ref{callan}. $\kappa_L= c(\log L)^{-1}$ as stated in Lemma \ref{l6.53}. There are various consequences of this Theorem that interest us. First, it applies to the neutral components of $K_{0,j}$. Setting $F(\f,X):= \hat K_{0,j}(0,\f,X)$, \bal\lb{6.62} &\|\Rem_{2,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hat K_{0,j}(0,\f,X)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le\r_2(0,\a)\|K_{0,j}\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal For $L$ large enough, the dimensional factor $\r_2(0,\a)= C C_2 (\sqrt{\kappa_L} L)^{-3}$ beats the volume factor $L^2$. Second, this theorem applies to the components of $\hat K_{0,j} $ with charges $q=\pm1$. Indeed, for $F(\f,X):=\hat K_{0,j}(q,\f, X)$ the hypothesis holds for $p=q$ and then \bal &\|\Rem_{0,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hat K_{0,j}(q,\f,X)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le \r_0(q,\a)\|K_{0,j}\|_{h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal For $q=\pm1$ and $L$ large enough, the dimensional factor $\r_0(1,\a)=C_0 C^2 (\sqrt{\kappa_L} L)^{-1} L^{-\frac{\a^2}{4\p}}$ is smaller than the volume factor $L^2$. The third application is the charged components of $\hat K_{1,j}$. We find \bal\lb{6.62b} &\|\Rem_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hat K_{1,j}(q,\f,X,x,\s)\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le \r_1(q+\h\s,\a)\|K_{1,j}\|_{1,h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X). \eal Finally, for $\d=a,\bar a, b$ and $0\le k\le j$, \bal\lb{lili2} &\|\Rem_{0,\d\f'}\mathbb{E}_j\left} \def\rgt{\right[\hat K^{(\d,k)}_{2,j}(q,\f,X,x,\s,x',\s')\rgt]\|_{h,T_{j+1}(\f',X)} \notag\\ &\le C^{1+|p|} C_0 L^{-2d(p)}\left} \def\rgt{\right(\sqrt{k_L} L\rgt)^{-1}\|K^{(\d,k)}_{2,j}\|_{2,h,T_j} \left} \def\rgt{\right({A\over 2}\rgt)^{-|X|_j} G_{j+1}(\f',\bar X), \eal where $$ p= \begin{cases} q+\h(\s+\s')\qquad & \text{ if $\d=a$}\\ q+\bar \h(\s+\s')\qquad & \text{ if $\d=\bar a$}\\ q+(\h\s+\bar \h \s')\qquad & \text{ if $\d=\bar a$}.\\ \end{cases} $$ We can now describe the ``power counting'' argument that will drive our analysis in the rest of the paper: a) by Theorem \ref{dh}, terms with charge $q$ contract by a factor $L^{-\frac{\a^2}{4\p}d(q)}$; b) by Theorem \ref{by}, terms proportional to $(\dpr \f')^n$ contract by a factor $L^{-n}$; c) as a consequence of Lemma \ref{l6.90}, all terms are increased by a volume factor $L^2$. Therefore, at $\a^2=8\p$, the action of the RG to contract the size of: i) the terms of total integer charge $p$, with $|p|\ge 2$; ii) the terms of total charge $p$, $|p|=1$, after that the 0-th order Taylor expansion in $\dpr \f'$ has been extracted; iii) the terms of total charge $p$, with $|p|=\h$ or $\bar \h$, after that the 1-th order Taylor expansion in $\dpr \f'$ has been extracted; iv) neutral terms, after that the 2-th order Taylor expansion in $\dpr \f'$ has been extracted. The terms that are extracted at points ii), iii) and iv) are absorbed into $E_j, t_j$ (see definitions before \pref{pfr}) to generate $E_{j+1}, t_{j+1}$. These ideas will be made precise in the next sections. \section{Renormalization Group Map}\lb{RGM} In the present and in the following section we adopt an abridged notation for the fields. In general, we remove the labels $j$ because they will be clear from the context, and we label the sum of the fields on higher scales with a prime, so that $\z_x:=\z^{(j)}_x$ and $\f'_x:=\z^{(R)}_x+\z^{(R-1)}_x+\cdots+\z^{(j+1)}_x$; besides, $\f_x:=\f'_x+\z_x$. We also set $\F=(J,\f)$ and $\F'=(J,\f')$. We indicate with $O(F_1, \ldots F_n)$ a term that is proportional to the fist power, at least, of each of $F_j$'s. Besides in the context of the inductive hypothesis described in Section \ref{strategy}, we will also assume the following symmetry properties. Define the $\p/2$ rotation $R(x_0,x_1):=(-x_1,x_0)$ and the translation $T_y x:=x+y$; and extend these transformations in a natural way to lattices subsets; besides, let $(R\f)_x:=\f_{R x}$ and $(T_y\f)_x:=\f_{x+y}$. We inductively assume that, for $\SS=R, T_y$, \begin{equation}} \def\ee{\end{equation}\lb{sym1} \hK_{0,j}(q,\SS\f,\SS Y)=\hK_{0,j}(q,\f,Y), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym2} \hK_{1,j}(q,\SS\f,\SS Y, \SS x, \s)=\hK_{1,j}(q,\f,Y, x, \s), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym3} \hK^\dagger_{1,j}(q,\SS\f,\SS Y, \SS x, \s)=\hK^\dagger_{1,j}(q,\f,Y, x, \s), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym04} \hK^{(\d,k)}_{2,j}(q,\SS\f,\SS Y,\SS x,\s,\SS x',\s')= \hK^{(\d,k)}_{2,j}(q,\f,Y,x,\s,x',\s'). \ee Besides, \begin{equation}} \def\ee{\end{equation}\lb{sym4} \hK_{0,j}(-q,-\f,Y)=\hK_{0,j}(q,\f,Y), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym5} \hK_{1,j}(-q,-\f,Y, x, -\s)=\hK_{1,j}(q,\f,Y, x, \s), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym6} \hK^\dagger_{1,j}(-q,-\f,Y, x, -\s)=\hK^\dagger_{1,j}(q,\f,Y, x, \s), \ee \begin{equation}} \def\ee{\end{equation}\lb{sym7} \hK^{(\d,k)}_{2,j}(-q,-\f,Y,x,-\s,x',-\s')= \hK^{(\d,k)}_{2,j}(q,\f,Y,x,\s,x',\s'). \ee We now discuss the RG procedure at a generic scale $j=1, \ldots, R-1$; subsequently we will discuss the slightly different procedure at scales $j=0$. \subsection{General RG step}\lb{s4.2} Assume by induction that at a given scale $j=1,2,\ldots,R-1$ the formula \pref{pfr} holds. We want to provide a useful way to recast $\O_{j+1}=\mathbb{E}_j\left} \def\rgt{\right[\O_j(J,\f'+\z)\rgt]$ into the same form of \pref{pfr}: \begin{equation}} \def\ee{\end{equation}\lb{pfr2} \O_{j+1}(\F')=e^{E_{j+1}|\L|} \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}}e^{U_{j+1}(\F',\L\bs X)} \prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X)} K_{j+1}(\F',Y). \ee We have the freedom to decide what to include in $K_{j+1}$ and what in $U_{j+1}$. Our aim will be to have a formula for $K_{j+1}$ of the form $\LL_j + {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$ where $\LL_j$ contains the linear order in $K_j$ and the linear and quadratic orders in $s_j$ and $z_j$; besides, we want $\LL_j$ to be a contraction. To obtain that, as explained in the end of the previous section, we need to implement the extraction based on the power counting argument. The next Lemma can be read in this way: there is a natural tentative choice for $K_{j+1}$, which at lowest orders contains the terms $\mathbb{E}_j [K_j]$ and $\mathbb{E}^T_j[V_j;V_j]$; from such a choice, a term $Q_j= O(K_{j})$ is extracted from $\mathbb{E}_j [K_j]$ and a term $Q^*_j=O(V^2_{j})$ is extracted from $\mathbb{E}^T_j[V_j;V_j]$; next, $Q_j$ and $Q^*_j$ are stored into $U_{j+1}$ and generate the new--scale parameters, $E_{j+1}, t_{j+1}$, from the old ones, $E_j, t_{j}$. Before stating the Lemma, we need some definitions. Introduce the two ``extraction activities'': \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item The activity $Q_j(\F',B,X)$, which is nonzero only for $X\in \SS_j$ and $B\in \BB_j(X)$. It is assumed to depend upon the fields $\{\f'_x, J_{x,\s}:x\in X^*, \s=\pm1\}$; however, it is also assumed that the dependence in at least one power of $J$ is restricted to the block $B$ (as opposed to the larger $X^*$). \item The activity $Q^*_j(\F',D,Y)$, which is nonzero only for $|Y|_{j+1}\le 2$ and $D\in \BB_{j+1}(Y)$. It is assumed to depend upon the fields $\{\f'_x, J_{x,\s}:x\in D^*, \s=\pm1\}$; but, again, one power of $J$ is in fact restricted to the set $D$ (as opposed to $D^*$). \ed Then define a new polymer activity $J_j$, which contains the extraction activities: \bal\lb{dfj} &J_j(\F',D,Y):=Q^*_{j}(\F',D, Y) +\sum_{B\in \BB_j(D)} \sum_{X\in \SS_j\atop X\supset B}^{\bar X=Y} Q_{j}(\F',B,X) \cr &\qquad-\d_{D,Y}\sum_{Y'\in \SS_{j+1}}^{Y'\supset D} \left} \def\rgt{\right[Q^*_{j}(\F',D,Y')+\sum_{B\in \BB_j(D)} \sum_{X\in \SS_j\atop X\supset B}^{\bar X=Y'} Q_j(\F',B,X)\rgt]. \eal Hence $J_j(\F',D,Y)$ is zero unless $Y\in \SS_{j+1}$ and $D\in \BB_{j+1}(Y)$. As the conditions $X\in \SS_j$ and $X\supset B$ together imply $X^*\subset D^*$ for $D=\bar B$, then $J_j(\F',D,Y)$ depends upon $\{\f'_x, J_{x,\s}: x\in D^*, \s=\pm1\}$; however, one power of $J$ is actually restricted to $D$. The second line of \pref{dfj} (with $\d_{D,Y}=1$ if $Y=D$ and $\d_{D,Y}=0$ otherwise) has been included so to obtain the crucial property of zero average: \bal\lb{5.23} \sum_{Y\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_{j+1}} J_j(\F',D,Y)=0. \eal For $Y\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_{j+1}$, define \bal\lb{5.23ancora} \tilde K_j(\F, Y):=\sum_{X'\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j}(Y)}^{\overline {X'}=Y} e^{U_{j}(\F,Y\bs X')} \prod_{Y'\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j}(X')} K_j(\F, Y'), \eal which depends on $\{\f_x, J_{x,\s}: x\in Y^*, \s=\pm1\}$, Now we are ready for the extractions. For every block $D\in \BB_{j+1}$, define \bal\lb{4.19bis} &P_j(\F',\z,D):=e^{U_j(\F,D)}-e^{U_{j+1}(\F',D)+(E_{j+1}-E_j)|D|}, \eal which depends on $\{\z_x:x\in \cup_{B\in \BB_j(D)} B^*\}$ and on $\{\f'_x,J_{x,\s}:x\in D^*,\s=\pm1\}$. For every connected polymer $Y\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}^c$, define \bal \lb{5.9bis} &R_j(\F',\z,Y):=\tilde K_j(\F,Y)-\sum_{D\in \BB_{j+1}(Y)}J_j(\F',D,Y), \eal which depends on $\{\z_x:x\in Y^*\}$ and on $\{\f'_x, J_{x,\s}:x\in Y^*, \s=\pm1 \}$. Note that in \pref{5.23ancora} and \pref{5.9bis} one power of $J$ is restricted to $Y$; likewise, in \pref{4.19bis} one power of $J$ is restricted to $D$. \begin{lemma} \lb{l5.3} Given formula \pref{pfr} with certain $t_j$, $E_j$ and $K_j$; given any two extraction activities as defined above and such that \begin{equation}} \def\ee{\end{equation}\lb{cond0} Q_{j}(\F',B,X)=O(K_j), \qquad Q^*_{j}(\F',D,Y)=O(V_j^2); \ee and given parameters $E_{j+1},t_{j+1}$ that satisfy \bal \lb{cond1} (E_{j+1}-E_{j})|D|+V_{j+1}(\F',D)-\mathbb{E}_j\left} \def\rgt{\right[V_{j}(\F,D)\rgt] &=O(K_{j}, V_{j}^2), \eal the following holds. A possible choice for $K_{j+1}$ in \pref{pfr2} is \bal\lb{j+1} K_{j+1}(\F', Y')=&\sum_{X_0,X_1\atop Z, (D) }^{\to Y'} e^{-(E_{j+1}-E_j)|W|+U_{j+1}(\F',Y'\bs W)} \notag\\ &\qquad\times \mathbb{E}_j\left} \def\rgt{\right[P_j(\F',\z)^ZR_j(\F',\z)^{X_1}\rgt] J_j(\F')^{X_0,(D)}, \eal where the notation is: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item The sum with label $\to Y'$ indicates the sum over three $j+1$--polymers $X_0$, $X_1$, $Z$, contained in $Y'$, and over one $j+1$--block, $D_Y\in \BB_{j+1}(Y)$, per each polymer $Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)$, such that: a) $X_0$ and $X_1$ are separated by at least by one $j+1$--block, namely ${\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)= {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)+ {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)$; b) $Z\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}(Y'\bs (X_0\cup X_1))$; c) each connected component of $X_0$ is $j+1$-small; d) $\cup_{Y}D^*_Y\cup Z\cup X_1=Y'$. Besides, $W\=X_0\cup X_1\cup Z$. \item For polymers $Z, X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}$, we set \begin{equation}} \def\ee{\end{equation}\lb{mixed} P_j(\F',\z)^Z:=\prod_{D\in\BB_{j+1}(Z)}P_j(\F',\z,D),\qquad R_j(\F',\z)^X:=\prod_{Y\in{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X)}R_j(\F',\z,Y). \ee \item Given $X_0\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}$ and one $D_Y\in \BB_{j+1}(Y)$ for each $Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)$, we set \begin{equation}} \def\ee{\end{equation}\lb{mixed2} J_j(\F')^{X_0,(D)}:=\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)} J_j(\F', D_Y, Y). \ee \ed Such choice of $K_{j+1}(\F',Y')$ can be decomposed in the sum of two parts, the leading one, $\LL_j (\F',Y')$ and the remainder one ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j(\F',Y')$ in a way that: the latter is an higher order correction in the sense that if $V_j$, $V_{j+1}$ are scaled by $t$ and $W_j$, $W_{j+1}$, $K_j$ are scaled by $t^2$, for small parameter $t$, then ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j(\F',Y')=O(t^3)$; while the former has an explicit formula $$ \LL_j(\F',Y')=\LL^{(a)}_j(\F',Y')+\LL^{(b)}_j(\F',Y')+\LL^{(c)}_j(\F',Y'), $$ where, for $\d E_j:=E_{j+1}-E_j$, \bal\lb{5.25} &\LL^{(a)}_j(\F',Y') =\sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}^c_j(Y')}^{\overline X=Y'} \left} \def\rgt{\right[\mathbb{E}_j[K_j(\F,X)]-\sum_{B\in \BB_j(X)} Q_{j}(\F',B,X)\rgt], \notag\\ &\LL^{(b)}_j(\F',Y') = \frac12\sum_{B_0, B_1\in \BB_{j}(Y')}^{\overline{B_0\cup B_1}=Y'} \mathbb{E}^T_j\left} \def\rgt{\right[V_j(\tilde t_j, \F,B_0);V_j(\tilde t_j,\F,B_1)\rgt] -\sum_{D\in \BB_{j+1}(Y')}Q^*_{j}(\F',D, Y'), \notag\\ &\LL^{(c)}_j(\F',Y') =-\sum_{B\in \BB_{j}}^{\bar B=Y'} \Bigg[\d E_j|B|+V_{j+1}(\F',B) -\mathbb{E}_j\left} \def\rgt{\right[V_j(\F,B)\rgt] -\sum_{X\in \SS_j}^{X\supset B} Q_{j}(\F',B,X)\Bigg] \notag\\ &-\sum_{D\in \BB_{j+1}}^{D=Y'} \Bigg[W_{j+1}(\F',D)- \mathbb{E}_j\left} \def\rgt{\right[W_j(\tilde t_{j},\F,D)\rgt] -\sum_{Y\in \SS_{j+1}}^{Y\supset D} Q^*_{j}(\F',D,Y)\Bigg] \eal for any $\tilde t_j$ such that $\tilde t_j-t_j= (O(z^2), O(z^2), O(z), O(z))$. Besides the scale $j+1$ activity, $K_{j+1}$, can be decomposed into charged terms as stated in \pref{pinning0}, \pref{pinning}, \pref{dec3} and \pref{dec4} for the scale $j$ activity. \end{lemma} {\it\0Proof.}\ Starting from \pref{pfr} and re-blocking the polymers on scale $j+1$, we obtain an equivalent formulation for $\O_j$: \bal\lb{pfj+1} \O_j(\F)=e^{E_j|\L|}\sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}} \Big[\prod_{D\in \BB_{j+1}(\L\bs X)} e^{U_{j}(\F,D)}\Big] \prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X)} \tilde K_j(\F, Y) \eal for $\tilde K_j$ given by \pref{5.23ancora}. Plugging \pref{4.19bis} and \pref{5.9bis} in \pref{pfj+1} and expanding, we find \pref{pfr2}, for $K_{j+1}$ given by \pref{j+1}. Observe that, to derive it, we also used the factorization of $\mathbb{E}_j$ over sets that are in two different connected components of a $j+1$--polymer as explained in \pref{separation}. Besides, in some terms we have the parameters $\tilde t_j$ instead of the more natural $t_j$ because the difference can be left inside ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$. Finally, by construction, $W\subset Y'$ so that that $K_{j+1}(\F',Y)$ depends on the fields $\{\f'_x,J_{x,\s}: x\in Y^*,\s=\pm1\}$; and, in particular, one power of $J$ is restricted to $Y$, as required. We have to prove that the linear part in $K_j$, $Q_j$, $Q^*_j$ and second order part in $V_j$ of this choice of $K_{j+1}$ is \pref{5.25}: expanding formula \pref{j+1}, using \pref{cond0} and \pref{cond1}, we obtain \pref{5.25} via two simple identities, \bal\lb{*id1} &\sum_{D\in \BB_{j+1}(Y')}J_j(\F',D,Y')= \sum_{D\in \BB_{j+1}(Y')}Q^*_{j}(\F',D,Y') +\sum_{X\in\SS_j}^{\bar X=Y'} \sum_{B\in \BB_{j}(X)}Q_{j}(\F',B,X) \cr &\qquad\qquad- \sum_{D\in \BB_{j+1}}^{D=Y'}\sum_{Y\in \SS_{j+1}}^{Y\supset D} Q^*_{j}(\F',D,Y) -\sum_{B\in \BB_j}^{\bar B=Y'} \sum_{X\in \SS_j}^{X\supset B} Q_{j}(\F',B,X); \eal and, by \pref{5.23}, \bal\lb{*id2} &\sum_{Y\in \SS_{j+1}} \sum_{D\in \BB_{j+1}(Y)}^{D^*=Y'}\; J_j(\F',D,Y) =\sum_{D\in \BB_{j+1}}^{D^*=Y'} \sum_{Y\in \SS_{j+1}}^{Y\supset D}\; J_j(\F',D,Y)=0. \eal This completes the proof of the Lemma.\quad\qed\vskip.8em The usefulness of \pref{5.25} is that, as planned before, in $\LL^{(a)}_j(\F',Y')$ and $\LL^{(b)}_j(\F',Y')$ we read the extraction of $Q_{j}$ and $Q^*_{j}$ from $\mathbb{E}_j[K_j]$ and $\mathbb{E}^T_j[V_j;V_j]$ respectively; in $\LL^{(c)}_j(\F',Y')$ the same terms are re-absorbed into $E_{j},t_{j}$ so generating $E_{j+1},t_{j+1}$. Note that by construction $\LL_j$ depends on $t_j$, $K_j$, $Q_{j}$, $Q^*_{j}$, $\tilde t_j$, $\d E_j$ and $t_{j+1}$; however, in Section \ref{s6}, we will determine the last five of them as function of $t_j$ and $K_j$, so that also $\LL_j$ is ultimately only a function on $t_j$ and $K_j$. In fact, as stated in the next Theorem, also the dependence on $t_j$ disappears from $\LL_j$. Decompose $$ \LL_j(\F',Y)= \LL_{0,j}(\f',Y) + \LL_{1,j}(\F',Y)+ \LL_{2,j}(\F',Y) + \LL_{\ge 3,j}(\F',Y) $$ where the enumeration refers to the powers of $J$. The term that in linear in $J$ is \bal \LL_{1,j}(\F',Y) &=L^{-2(j+1)} Z_{j+1} \sum_{x\in Y\atop\s=\pm} J_{x,\s} \LL_{1,j}(\f',Y,x,\s) \notag\\ &+L^{-2(j+1)} \bar Z_{j+1} \sum_{x\in Y\atop\s=\pm} J_{x,\s} \LL^{\dagger}_{1,j}(\f',Y,x,\s). \eal The term that is quadratic in $J$ is \bal \LL_{2,j}(\F',Y)&=\sum_{x_1\in Y, x_2\in Y^*\atop \s_1,\s_2=\pm1} J_{\s_1, x_1}J_{\s_2, x_2}\LL_{2,j}(\f',Y,x_1,\s_1,x_2,\s_2) \eal and can be further decomposed into (suppressing the dependence in $\f',Y,x_1,\s_1,x_2,\s_2$) $$ \LL_{2,j}=\sum_{k=0}^{j} 2^{-(j-k)}L^{-4k } e^{-L^{-k}|x_1-x_2|} \left} \def\rgt{\right[Z^2_k\LL^{(a,k)}_{2,j}+ \bar Z^2_k\LL^{(\bar a,k)}_{2,j} +Z_k \bar Z_k\LL^{(b,k)}_{2,j} \rgt]. $$ \begin{theorem}\lb{faculty2} For a suitable choice of $Q_{j}$, $Q^*_{j}$, $\tilde t_j$, $\d E_j$ and $t_{j+1}$ as functions of $t_j$, $K_j$, the leading part $\LL_j$ is independent of $t_j$ and is linear in $K_j$. Besides, under the inductive assumption that $$ \left} \def\rgt{\right|\frac{Z_j}{Z_{j+1}}\rgt|\le 1, \qquad \left} \def\rgt{\right|\frac{\bar Z_j}{\bar Z_{j+1}}\rgt|\le 1, $$ $\LL_j$ satisfies the following bounds: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item for the term of $\LL_j$ that is independent of $J$, \begin{equation}} \def\ee{\end{equation}\lb{allen2} \|\LL_{0,j}\|_{h, T_{j+1}}\le \r(L,A)\|K_{0,j}\|_{h, T_j}; \ee where $\r(L,A)$ is arbitrarily small for $L$ and $A$ large enough; \item for the terms of $\LL_j$ that are linear or quadratic in $J$, \bal\lb{arkani2} &\|\LL_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K_{1,j}\|_{1,h,T_{j}} \notag\\ &\|\LL^{\dagger}_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K^\dagger_{1,j}\|_{1,h,T_{j}} \eal \begin{equation}} \def\ee{\end{equation}\lb{goddar2} \|\LL^{(\d, k)}_{2,j}\|_{2,h,T_{j+1}}\le \r(L,A,\h) \|K^{(\d, k)}_{2,j}\|_{2,h,T_{j}} \ee for a $\r(L,A,\h)$ that is arbitrarily small for any $\h\in (0,1)$ if $L$ and $A$ are large enough. \ed \end{theorem} The first point of the result was already proven in \cite{Fa12}. The proof of the second point is a direct consequence of Lemma \ref{faculty}, Lemma \ref{bois}, Lemma \ref{bourgain} and Lemma \ref{bois2} in Section \ref{s6}. There we will also explain how to obtain \pref{lk}, \pref{elk}, \pref{sk} and the following formula for $\tilde t_j$, \begin{equation}} \def\ee{\end{equation}\lb{5.25def} \tilde t_j=(s_{j+1}, z_{j+1} L^{-2} e^{\frac{\a^2}2 \G_j(0)}, Z_{j+1} L^{-2} e^{\h^2\frac{\a^2}2 \G_j(0)}, \bar Z_{j+1} L^{-2} e^{\bar \h^2\frac{\a^2}2 \G_j(0)}). \ee Consider now the remainder part. Using \pref{j+1} for $K_{j+1}$ and formula \pref{5.25} for its leading part, we obtain the following formula for ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$: \begin{equation}} \def\ee{\end{equation}\lb{exprem} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j(\F', Y'): \sum_{n=1}^9 {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(n)}_j(\F', Y') \ee where, suppressing the dependence in the field (again $\d E_{j}:=E_{j+1}-E_j$), \bal {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(1)}_j(Y') &= \sum_{D\in \BB_{j+1}}^{D=Y'}\Big[\mathbb{E}_j\left} \def\rgt{\right[P_j(D)\rgt]+V_{j+1}(D)- \mathbb{E}_j[V_j(D)]-\d E_{j+1}|D| \notag\\ &\qquad\qquad -\frac12 \mathbb{E}^T[V_j(D);V_j(D)]+W_{j+1}(D)-\mathbb{E}_j[W_j(D)]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(2)}_j(Y')&= \frac12\sum_{D_1, D_2\in \BB_{j+1}\atop D_1\neq D_2}^{D_1\cup D_2=Y'} \Big[\mathbb{E}_j\left} \def\rgt{\right[P_j(D_1) P_j(D_2)\rgt]- \mathbb{E}^T_j\left} \def\rgt{\right[V_j(D_1); V_j(D_2)\rgt]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(3)}_j(Y') &= \sum_{D\in \BB_{j+1}}^{D=Y'}\Big[\mathbb{E}_j[W_j(D)]-\mathbb{E}_j[W_j(\tilde t_j, D)]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(4)}_j(Y') &= \frac12\sum_{D_1, D_2\in \BB_{j+1}}^{D_1\cup D_2=Y'} \Big[\mathbb{E}^T_j\left} \def\rgt{\right[V_j(D_1); V_j(D_2)\rgt]- \mathbb{E}^T_j\left} \def\rgt{\right[V_j(\tilde t_j, D_1); V_j(\tilde t_j, D_2)\rgt]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(5)}_j(Y')&= \sum_{|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)|\ge 1\atop |Z|_{j+1}+|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)|\ge 2}^{\to Y'} \mathbb{E}_j\left} \def\rgt{\right[P_j^Z R_j^{X_1}\rgt] J_j^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(6)}_j(Y')&= \sum_{|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)|\ge 1}^{\to Y'} \left} \def\rgt{\right(e^{-\d E_j|Y'|+U_{j+1}(Y'\bs W)}-1\rgt)\mathbb{E}_j\left} \def\rgt{\right[P_j^Z R_j^{X_1}\rgt] J_j^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(7)}_j(Y')&= \sum_{Z\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}\atop |Z|_{j+1}\ge 3}^{Z=Y'}\mathbb{E}_j\left} \def\rgt{\right[P_j^Z\rgt]+ \left} \def\rgt{\right(e^{-\d E_j|Y'|}-1\rgt) \mathbb{E}_j\left} \def\rgt{\right[P_j^{Y'}\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(8)}_j(Y')&= \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop |{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|\ge 2}^{\bar X=Y'}\mathbb{E}_j\left} \def\rgt{\right[\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} K_j(Y)\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(9)}_j(Y')&=\sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j}^{\bar X=Y'} \mathbb{E}_j\left} \def\rgt{\right[\left} \def\rgt{\right(e^{U_j(Y'\bs X)}-1\rgt)\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} K_j(Y)\rgt]. \eal ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$, as well as each ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(n)}_j$, can be decomposed in terms with increasing powers of $J$, \begin{equation}} \def\ee{\end{equation}\lb{rdec1} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j(\F',Y)={\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}(\f',Y)+{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}(\F',Y)+{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{2,j}(\F',Y) +{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{\ge 3,j}(\F',Y). \ee The term that is linear in $J$ is \bal\lb{rdec2} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}(\F',Y) &= L^{-2(j+1)}Z_{j+1}\sum_{x\in Y\atop \s=\pm1} J_{x,\s}{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}(\f',Y, x,\s) \notag\\ &+ L^{-2(j+1)}\bar Z_{j+1} \sum_{x\in Y\atop \s=\pm1} J_{x,\s}{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{\dagger}_{1,j}(\f',Y, x,\s). \eal The term that is quadratic in $J$ is \bal\lb{rdec3} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{2,j}(\F',Y)&=\sum_{x_1\in Y, x_2\in Y^*\atop \s_1,\s_2=\pm1} J_{\s_1, x_1}J_{\s_2, x_2}{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{2,j}(\f',Y,x_1,\s_1,x_2,\s_2) \eal and can be further decomposed into (suppressing the dependence in $\f',Y,x_1,\s_1,x_2,\s_2$) \begin{equation}} \def\ee{\end{equation}\lb{rdec4} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{2,j}=\sum_{k=0}^{j} 2^{-(j-k)}L^{-4k } e^{-L^{-k}|x_1-x_2|} \left} \def\rgt{\right[Z^2_k{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(a,k)}_{2,j}+ \bar Z^2_k{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(\bar a,k)}_{2,j}+Z_k \bar Z_k{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(b,k)}_{2,j} \rgt]. \ee \begin{theorem}\lb{crone00} If $z>0$ is small enough and $|s_j|,|z_j|\le c_0|q_j|$, $\|K_{0,j}\|_{h,T_j}\le c_0|q_j|^2$, there exists $C\=C(A, L, \a)$ such that, \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item for the term of ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$ that is independent of $J$ \bal\lb{old3.22} &\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}- \dot{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,j}\|_{h, T_{j+1}} \notag\\ &\le C \left} \def\rgt{\right[|q_j|^2|s_j-\dot s_j|+ |q_j|^2|z_j-\dot z_j|+ |q_j|\|K_{0,j}-\dot K_{0,j}\|_{h,T_j}\rgt] \eal where $\dot R_{0,j}$ is obtained from $R_{0,j}$ by replacing $s_j,z_j,K_{0,j}$ with any $\dot s_j, \dot z_j, \dot K_{0,j}$ that satisfy $|\dot s_j|,|\dot z_j|\le c_0|q_j|$ and $\|\dot K_{0,j}\|_{h,T_j}\le c_0|q_j|^2$; \item for the terms of ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_j$ that are linear in $J$, \begin{equation}} \def\ee{\end{equation}\lb{cosmo00} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}\|_{1, h, T_{j+1}}\le C \left} \def\rgt{\right[|q_j|^2 +|q_j|\|K_{1,j}\|_{1,h,T_j}+|q_j|\|K^\dagger_{1,j}\|_{1,h,T_j}\rgt], \ee and the same bound is valid for $\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^\dagger_{1,j}\|_{1, h, T_{j+1}}$; \item for the terms of $R_{j}$ that are quadratic in $J$, with the extra assumption that $\|K_{1,j}\|\le c_0 |q_j|^2$ and $\|K^\dagger_{1,j}\|\le c_0 |q_j|^2$, \bal\lb{geary00} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(\d,k)}_{2,j}\|_{2, h, T_{j+1}} \le \begin{cases} C |q_k| &\text{for $k=j$} \\ C |q_j| \|K^{(\d,k)}_{2,j}\|_{2,h,T_j}\qquad&\text{for $0\le k\le j-1$} \end{cases} \eal \ed \end{theorem} The first point was already proven in \citep{Fa12}. The second and third points are consequence of Lemma \ref{crone} and Lemma \ref{fassin}. \subsection{First RG step} The starting point is formula \pref{itr} that in our current notations reads \begin{equation}} \def\ee{\end{equation}\lb{itr0} \O_1(\F)=e^{E_0|\L|} \mathbb{E}_0\left} \def\rgt{\right[e^{V_0(\F,\L)}\rgt]. \ee As already noted, the term in the square brackets of \pref{itr0} has the form \pref{pfr} for $j=0$ for $W_{0}(\F,B)=0$, $K_0(\F,Y)=0$, and for parameters $E_0=E$ and $t_0=(s, z, 1, 0)$. We want to recast $\O_1$ into the form \pref{pfr} for $j=1$. For doing so, we apply Lemma \ref{l5.3} to the scale $j=0$: since $K_{0}=0$, in Section \ref{s6} we will see that there exists a choice of $Q_0^*$, $\tilde t_0$, $\d E_0$ and $t_1$ such that \begin{equation}} \def\ee{\end{equation}\lb{5.25bis} \LL_{0}(\F,Y)\=0. \ee However, since the choice for $Q^*_0$ will differ from the general formula for $Q^*_j$ in the part that does not depend on $J$ (we do this for merging with the treatment of \citep{Fa12}) the remainder part is slightly different from \pref{exprem} at $j=0$, $W_0=0$ and $K_0=0$; indeed we have \begin{equation}} \def\ee{\end{equation}\lb{exprem0} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_0(\F', Y'): \sum_{n=1}^6 {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(n)}_0(\F', Y') \ee where, suppressing the dependence in the field, \bal {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(1)}_0(Y') &= \sum_{D\in \BB_{1}}^{D=Y'}\Big[\mathbb{E}_0\left} \def\rgt{\right[P_0(D)\rgt]+V_{1}(D)- \mathbb{E}_0[V_0(D)]-\d E_0|D| \notag\\ &\qquad\qquad -\frac12 \mathbb{E}^T_0[V_0(D);V_0(D)]+W_{1}(D)+\frac12 \mathbb{E}^T_0[V_{0,0}(D);V_{0,0}(D)]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(2)}_0(Y')&= \frac12\sum_{D_1, D_2\in \BB_{1}\atop D_1\neq D_2}^{D_1\cup D_2=Y'} \Big[\mathbb{E}_0\left} \def\rgt{\right[P_0(D_1) P_0(D_2)\rgt]- \mathbb{E}^T_0\left} \def\rgt{\right[V_0(D_1); V_0(D_2)\rgt] \notag\\ &\qquad\qquad+ \mathbb{E}^T_0\left} \def\rgt{\right[V_{0,0}(D_1); V_{0,0}(D_2)\rgt]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(3)}_0(Y') &= \frac12\sum_{D_1, D_2\in \BB_{1}}^{D_1\cup D_2=Y'} \Big[\mathbb{E}^T_0\left} \def\rgt{\right[V_0(D_1); V_0(D_2)\rgt]- \mathbb{E}^T_0\left} \def\rgt{\right[V_0(\tilde t_0, D_1); V_0(\tilde t_0, D_2)\rgt]\Big] \notag\\ &- \frac12\sum_{D_1, D_2\in \BB_{1}}^{D_1\cup D_2=Y'} \Big[\mathbb{E}^T_0\left} \def\rgt{\right[V_0(D_1); V_0(D_2)\rgt]- \mathbb{E}^T_0\left} \def\rgt{\right[V_{0,0}(\tilde t_0, D_1); V_{0,0}(\tilde t_0, D_2)\rgt]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(4)}_0(Y')&= \sum_{|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{1}(X_0\cup X_1)|\ge 1\atop |Z|_{1}+|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{1}(X_0\cup X_1)|\ge 2}^{\to Y'} \mathbb{E}_0\left} \def\rgt{\right[P_0^Z R_0^{X_1}\rgt] J_0^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(5)}_0(Y')&= \sum_{|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{1}(X_0\cup X_1)|\ge 1}^{\to Y'} \left} \def\rgt{\right(e^{-\d E_0|Y'|+U_{1}(Y'\bs W)}-1\rgt)\mathbb{E}_0\left} \def\rgt{\right[P_0^Z R_0^{X_1}\rgt] J_0^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(6)}_0(Y')&= \sum_{Z\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{1}\atop |Z|_{1}\ge 3}^{Z=Y'}\mathbb{E}_0\left} \def\rgt{\right[P_0^Z\rgt]+ \left} \def\rgt{\right(e^{-\d E_0|Y'|}-1\rgt) \mathbb{E}_0\left} \def\rgt{\right[P_0^{Y'}\rgt]. \eal The decompositions \pref{rdec1}, \pref{rdec2},\pref{rdec3} and \pref{rdec4} are valid also at $j=0$. \begin{theorem}\lb{crone001} Under the same hypothesis of Theorem \ref{crone00}, \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item for the term of ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_0$ that is independent of $J$ \bal\lb{old3.221} &\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,0}-\dot {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{0,0}\|_{h, T_{1}} \le C |q_0|\left} \def\rgt{\right[|s_0-\dot s_0|+ |z_0-\dot z_0|\rgt]; \eal \item for the terms of ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_0$ that are linear in $J$, \begin{equation}} \def\ee{\end{equation}\lb{cosmo001} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,0}\|_{1, h, T_{1}}\le C |q_0|^2, \ee and the same bound is valid for $\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^\dagger_{1,0}\|_{1, h, T_{1}}$; \item for the terms of $R_{0}$ that are quadratic in $J$, \bal\lb{geary001} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(\d,0)}_{2,0}\|_{2, h, T_{1}} \le C|q_0|. \eal \ed \end{theorem} As for Theorem \ref{crone00}, we only need to prove the second and third points, which are consequence of Lemma \ref{crone} and Lemma \ref{fassin}. Note that \pref{cosmo001} and \pref{geary001} coincide with \pref{cosmo00} and \pref{geary00} at $j=0$; while \pref{old3.221} differs from \pref{old3.22} at $j=0$ and is the same as in \citep{Fa12}. \section{Leading part of the RG map}\lb{s6} \subsection{Running coupling constants}\lb{sc20} The choice of $Q_{j}$ requires Taylor expansion in $\nabla \f'$. For any point $x_0\in X$, if $(\d \f')_x:= \f'_x-\f'_{x_0}$ (which is a sum of $\nabla \f'$'s), using \pref{dec4}, we have \bal &\hK_{0,j}(q,\f,X)= e^{i\a q \f'_{x_0}} \hK_{0,j}(q,\d\f'+\z,X), \notag\\ &\hK_{1,j}(q,\f,X, x, \s)= e^{i\a (q+\h \s) \f'_{x_0}} \hK_{1,j}(q,\d\f'+\z,X,x, \s), \notag\\ &\hK_{1,j}^\dagger(q,\f,X, x,\s)= e^{i\a (q+\bar \h \s) \f'_{x_0}}\hK^\dagger_{1,j}(q,\d\f'+\z,X, x,\s), \notag\\ &\hK^{(a,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\h \s+\h \s')\a\f'_{x_0}}\hK^{(a,k)}_{2,j}(q,\d\f'+\z,X,x,\s,x',\s'), \notag\\ &\hK^{(\bar a,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\bar \h \s+\bar \h\s')\a\f'_{x_0}}\hK^{(a,k)}_{2,j}(q,\d\f'+\z,X,x,\s,x',\s'), \notag\\ &\hK^{(b,k)}_{2,j}(q,\f,X,x,\s,x',\s')= e^{i (q+\h \s+\bar \h\s')\a\f'_{x_0}}\hK^{(a,k)}_{2,j}(q,\d\f'+\z,X,x,\s,x',\s'). \eal We now choose $Q_{j}$. Set $Q_{j}(\F',B,X)=0$ if $X\not\in S_j$ or $B\not\in \BB_j(X)$; otherwise $Q_{j}(\F',B,X)$ is the sum of the following four terms. \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item A term proportional to $K_{0,j}$: \bal\lb{qbr} Q_{0,j}(\f',B,X) =&\frac{1}{|X|}\sum_{x_0\in B} \Tay_{2,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{0,j}(0,\d\f'+\z,X)\rgt] \notag\\ &+ \frac{1}{|X|}\sum_{x_0\in B\atop\s=\pm1}e^{i\s\a\f'_{x_0}}\Tay_{0,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{0,j}(\s,\d\f'+\z,X)\rgt]. \eal \item Two terms proportional to $K_{1,j}$: \bal Q_{1,j}(\F',B,X)= & Z_j L^{-2j}\sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\h\a\s \f'_x} \Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\d\f'+\z,X, x,\s)\rgt] \notag\\ +& Z_j L^{-2j}\sum_{x\in B\atop \s=\pm1} J_{x,\s}e^{i\bar\h\a\s \f'_x} \Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(-\s,\d\f'+\z,X, x,\s)\rgt], \notag\\\notag\\ Q^\dagger_{1,j}(\F',B,X)=& \bar Z_j L^{-2j}\sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\bar \h\a\s \f'_x} \Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK^\dagger_{1,j}(0,\d\f'+\z,X, x,\s)\rgt] \notag\\ +& \bar Z_j L^{-2j} \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i \h\a\s \f'_x}\Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK^\dagger_{1,j}(\s,\d\f'+\z,X, x,\s)\rgt], \lb{qbr1} \eal where the special point in $\d \f'$ is $x$. Even though the pinning at $x$ prevents the generation of the volume factor $L^2$ when we sum these terms over $X$, note that here the extraction is guided by the standard power counting. As it will be clear in Section \ref{s6.4}, the reason for doing so is that we want to preserve the prefactor $L^{-2j}$ at each scale, which costs an $L^2$ factor at each step. \item A term proportional to $K_{2,j}$: \bal &Q_{2,j}(\F',B,X) =\sum_{k=0}^j2^{-(j-k)} L^{-4k} \sum_{x_1\in B \atop x_2\in X^*} e^{-L^{-k}|x_1-x_2|} \notag\\ &\times \sum_{\s,\s'=\pm1}J_{x_1,\s}J_{x_2,\s'}e^{i\a(\s+\s')(\h-\frac12)\f'_{x_1}} \notag\\ &\times \left} \def\rgt{\right\{Z_k^2 \Tay_{0,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(a,k)}(-\frac{\s+\s'}2, \d\f'+\z, X, x_1, \s, x_2, \s')\rgt] \rgt. \notag\\ &\qquad+\left} \def\rgt{\right. \bar Z^2_k\Tay_{0,\d\f'}\mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(\bar a,k)}(\frac{\s+\s'}2, \d\f'+\z, X, x_1, \s, x_2, \s') \rgt] \rgt. \notag\\ &\qquad+\left} \def\rgt{\right. Z_k\bar Z_k \Tay_{0,\d\f'}\mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(b,k)}(-\frac{\s-\s'}2, \d\f'+\z, X, x_1, \s, x_2,\s') \rgt]\rgt\}.\lb{qbr2b} \eal As opposed to what we did for \pref{qbr1}, in \pref{qbr2b} the extraction follows a power counting that does {\it not} take account of the volume factor $L^2$. Note that the above term is then irrelevant unless $\h=\frac12$ or $\s=-\s'$; however we extract the same $Q_{2,j}$ for every $\s,s',\h$ not to have an $L$ that be divergent for $\h\to\frac12$. \ed Finally, we have the following result: $\LL^{(a)}_j(\F',Y')$ is of the form $$ \LL^{(a)}_j(\F',Y')= \LL^{(a)}_{0,j}(\f',Y') + \LL^{(a)}_{1,j}(\f',Y') + \LL^{(a)}_{2,j}(\f',Y') $$ where \bal \LL^{(a)}_{1,j}(\f',Y') &=L^{-2(j+1)} Z_{j+1} \sum_{x\in Y'\atop \s=\pm1} J_{x,\s} \LL^{(a)}_{1,j}(\f',Y',x,\s) \notag\\ &+L^{-2(j+1)} \bar Z_{j+1} \sum_{x\in Y'\atop \s=\pm1} J_{x,\s} \LL^{(a)\dagger}_{1,j}(\f',Y',x,\s) \notag \eal and \bal \LL^{(a)}_{2,j}(\f',Y') &=\sum_{x_1\in Y', x_2\in {Y'}^*\atop \s_1,\s_2=\pm1} J_{x_1,\s_1} J_{x_2,\s_2} \LL^{(a)}_{2,j}(\f',Y',x_1,\s_1,x_2,\s_2) \eal for (neglecting the variables that are $\f',Y',x_1,\s_1,x_2,\s_2$ in each term) \bal \LL^{(a)}_{2,j} &=\sum_{k=0}^j 2^{-(j-k)}L^{-4k}e^{-L^{-k}|x_1-x_2|} \left} \def\rgt{\right[Z^2_k \LL^{(a,k)}_{2,j} +\bar Z^2_k \LL^{(\bar a,k)}_{2,j}+Z_k \bar Z_k\LL^{(b,k)}_{2,j}\rgt]. \eal \begin{lemma}\lb{faculty} Assume by induction that $$ \left} \def\rgt{\right|\frac{Z_j}{Z_{j+1}}\rgt|\le 1\qquad \left} \def\rgt{\right|\frac{\bar Z_j}{\bar Z_{j+1}}\rgt|\le 1; $$ then, for large enough $L$, there exist $\r(L,A)$ and $\r(L,A,\h)$ such that \begin{equation}} \def\ee{\end{equation}\lb{allen} \|\LL^{(a)}_{0,j}\|_{h, T_{j+1}}\le \r(L,A)\|K_{0,j}\|_{h, T_j}; \ee \bal\lb{arkani} &\|\LL^{(a)}_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K_{1,j}\|_{1,h,T_{j}}, \notag\\ &\|\LL^{(a)\dagger}_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K^\dagger_{1,j}\|_{1,h,T_{j}}, \eal \begin{equation}} \def\ee{\end{equation}\lb{goddar} \|\LL^{(\d,k)}_{2,j}\|_{2,h,T_{j+1}}\le \r(L,A,\h) \|K^{(\d,k)}_{2,j}\|_{2,h,T_{j}}. \ee Besides, fixed any $\h\in (0,1)$, the prefactors $\r(L,A)$ and $\r(L,A,\h)$ are arbitrarily small for $L$ and $A$ large enough. \end{lemma} The proof is in Section \ref{s6.4}. The next step is to choose $Q^*_{j}$: set $Q^*_{j}(\F',D,Y'):=0$ if $|Y'|_{j+1}\ge 3$ or $D\not\in \BB_{j+1}(Y')$; otherwise, if $j\ge 1$, \bal\lb{qqq} Q^*_{j}(\F',D,Y') &:= {1\over 2}\sum_{B_0\in \BB_j(D)\atop B_1\in \BB_{j}(Y')}^{\overline{B_0\cup B_1}=Y'} \mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\tilde t_j,\f,B_0);V_{0,j}(\f,B_1)\rgt] \notag\\ &+\sum_{B_0\in \BB_j(D)\atop B_1\in \BB_{j}(Y')}^{\overline{B_0\cup B_1}=Y'} \mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B_0);V_{0,j}(\tilde t_j,\f,B_1)\rgt] \notag\\ &+{1\over 2}\sum_{B_0\in \BB_j(D)\atop B_1\in \BB_{j}(Y')}^{\overline{B_0\cup B_1}=Y'}\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B_0);V_{1,j}(\tilde t_j,\F,B_1)\rgt]. \eal If instead $j=0$, we do not include in \pref{qqq} the fist line, i.e. the one proportional to $V_{0,0}^2$. This was also the choice in \citep{Fa12}; and this explains why right hand side of \pref{old3.221} is quadratic (as opposed to cubic) in $s$ and $z$. With this definition of $Q^*_j$, the proof of the following Lemma is a computational verification. \begin{lemma} \lb{bois} $$ \LL^{(b)}_j(\F',Y')=0 $$ \end{lemma} Finally, we have to deal with $\LL^{(c)}_j(\F',Y')$; namely we have to show how $E_j, t_j$ and $Q_j, Q^*_j$ generate $E_{j+1}, t_{j+1}$ so that \pref{cond1} holds and $\LL^{(c)}_j(\F',Y')$ is a contraction. We split this term in two pieces. First define intermediate effective parameters: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item Intermediate effective couplings $s^*_{j}$ and $z^*_{j}$ \bal\lb{lk2} &s^*_j:=s_j +\FF_j, \notag\\ &z^*_j:=L^{2}e^{-\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right[z_j+\MM_j\rgt] \eal where $\FF_j$ and $\MM_j$ are functionals of the fields \bal &\FF_j\=\FF_j(K_j)=\sum_{X\in \SS_j}^{X\ni0} \frac{L^{-2j}}{|X|_{j}|X|}\sum_{x_0\in X\atop x_1,x_2\in X^*} \mathbb{E}_j\left} \def\rgt{\right[\frac{\dpr^2\hK_{0,j}} {\dpr \f_{x_1}\dpr \f_{x_2}}(0,\z,X)\rgt] \sum_{\m\in\hat u}(x_1-x_0)^\m(x_2-x_0)^\m, \notag\\ &\MM_j\=\MM_j(K_j)=\frac{e^{\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1} \sum_{X\in \SS_j}^{X\ni0} \frac{1}{|X|_{j}} \mathbb{E}_j\left} \def\rgt{\right[\hK_{0,j}(\s,\z,X)\rgt]. \eal \item Intermediate effective free energy \begin{equation}} \def\ee{\end{equation}\lb{lk2b0} E^*_j=E_j+ L^{-2j}\left} \def\rgt{\right[\EE_{1,j} + s_j \EE_{2,j}\rgt] \ee for \bal &\EE_{1,j}\=\EE_{1,j}(K_j)=\sum_{X\in \SS_j}^{X\ni0} \frac{1}{|X|_{j}}\mathbb{E}_j\left} \def\rgt{\right[\hK_{0,j}(0,\z, X)\rgt], \notag\\ &\EE_{2,j}=-{L^{2j}\over 2}\sum_{\m\in \hat u} (\dpr^{-\m}\dpr^\m\G_j)(0). \eal \item Intermediate renormalization constants \bal\lb{lk2b} Z^*_j&=L^2e^{-\h^2\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right[\left} \def\rgt{\right(1+\MM_{1,1,j}\rgt)Z_j +\MM_{1,2,j}(K_j)\bar Z_j\rgt], \notag\\ \bar Z^*_j&=L^2e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right[\MM_{2,1,j} Z_j +\left} \def\rgt{\right(1+\MM_{2,2,j}\rgt)\bar Z_j\rgt], \eal where the functionals $\MM_{p,q,j}$ are \bal\lb{dfm1} &\MM_{1,1,j}\=\MM_{1,1,j}(K_j)= \frac{e^{\h^2\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni0} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\z,X, 0,\s)\rgt], \notag\\ &\MM_{1,2,j}\=\MM_{1,2,j}(K_j)= \frac{e^{\h^2\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni0} \mathbb{E}_j\left} \def\rgt{\right[\hK^\dagger_{1,j}(\s,\z,X, 0,\s)\rgt], \notag\\ &\MM_{2,1,j}\=\MM_{2,1,j}(K_j)= \frac{e^{\bar \h^2\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni0} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(-\s,\z,X, 0,\s)\rgt], \notag\\ &\MM_{2,2,j}\=\MM_{2,2,j}(K_j)= \frac{e^{\bar \h^2\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1} \sum_{X\in \SS_j}^{X\ni0} \mathbb{E}_j\left} \def\rgt{\right[\hK^\dagger_{1,j}(0,\z,X, 0,\s)\rgt]. \eal Note that, in the definition of $\MM_{m,n, j}$ we only retained the Tay$_0$ part of \pref{qbr1}; this is because of cancellations due to \pref{sym2} and \pref{sym3}. For example, \pref{sym2} for $\SS=R^2$ gives for any $m=0,1$ \begin{equation}} \def\ee{\end{equation}\lb{conseq1} \sum_{X\in \SS_j}^{X\ni0} \sum_{y\in X^*} \mathbb{E}_j\left} \def\rgt{\right[\frac{\dpr \hK_{1,j}}{\dpr \z_y}(0,\z,X, 0,\s)\rgt]y^\m=0. \ee Besides, we used also the symmetry under charge conjugation \pref{sym5} and \pref{sym6}, which implies, for example, \begin{equation}} \def\ee{\end{equation}\lb{conseq2} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\z,X, 0,1)\rgt]=\frac12\sum_{\s=\pm1} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\z,X, 0,\s)\rgt] \ee These points will be detailed in Section \ref{s6.4b}. \ed Next, split $\LL^{(c)}_j(\F',Y')$ into two terms $$ \LL^{(c)}_j(\F',Y')=\LL^{(c1)}_j(\F',Y')+ \LL^{(c2)}_j(\F',Y') $$ for \bal\lb{irb} \LL^{(c1)}_j(\F',Y'):=\sum_{B\in \BB_{j}}^{\bar B=Y'} \Bigg\{ &(E^*_{j}-E_j)|B|+V_{j+1}(t^*_j,\f',B) -\mathbb{E}_j\left} \def\rgt{\right[V_{j}(\f,B)\rgt] \notag\\ &-\sum_{X\in \SS_j}^{X\supset B} \left} \def\rgt{\right[Q_{j}(\f',B,X)-Q_{2,j}(\f',B,X)\rgt]\Bigg\} \eal By construction, $\LL^{(c1)}_j(\F',Y')$ is made of a part that is $J$-independent, which we call $\LL^{(c1)}_{0,j}(\F',Y')$, and a part that is linear in $J$, which we call $\LL^{(c1)}_{1,j}(\F',Y')$ and which can be further decomposed \bal \LL^{(c1)}_{1,j}(\F',Y') &=L^{-2(j+1)} Z_{j+1} \sum_{x,\s} J_{x,\s} \LL^{(c1)}_{1,j}(\f',Y',x,\s)\notag\\ &+L^{-2(j+1)} \bar Z_{j+1} \sum_{x,\s} J_{x,\s} \LL^{(c1)\dagger}_{1,j}(\f',Y',x,\s)\;. \eal \begin{lemma}\lb{bourgain} For large enough $L$, there exist $\r(L,A)$ and $\r(L,A,\h)$ such that \begin{equation}} \def\ee{\end{equation}\lb{chaniotis} \|\LL^{(c1)}_{0,j}\|_{h, T_{j+1}}\le \r(L,A)\|K_{0,j}\|_{h, T_j}, \ee \bal\lb{chaniotis2} &\|\LL^{(c1)}_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K_{1,j}\|_{1,h,T_j}, \notag\\ &\|\LL^{(c1)\dagger}_{1,j}\|_{1,h,T_{j+1}}\le \r(L,A,\h) \|K^\dagger_{1,j}\|_{1,h,T_j}. \eal Besides, $\r(L,A)$ and $\r(L,A,\h)$ are arbitrarily small for $L$ and $A$ large enough. \end{lemma} The proof is in Section \ref{s6.4b}. By subtraction, the other part of $\LL^{(c)}_j(\F',Y')$ is \bal\lb{eW} &\LL^{(c2)}_j(\F',Y')=\sum_{D\in \BB_{j+1}}^{D=Y'} \Bigg\{ \left} \def\rgt{\right(E_{j+1}-E^*_{j}\rgt)|D| +V_{j+1}(t_{j+1}- t^*_j,\F',D) \notag\\ &+ W_{j+1}(t_{j+1},\F',D)-\mathbb{E}_j\left} \def\rgt{\right[W_j(\tilde t_{j},\F,D)\rgt] \notag\\ &-\sum_{Y\in \SS_{j+1}}^{Y\supset D} Q^*_{j}(\F',D,Y)-\sum_{B\in \BB_j(D)}\sum_{X\in\SS_j}^{X\supset B} Q_{2,j}(\F',B,X) \Bigg\}. \eal We want to choose $E_{j+1}$, $s_{j+1}$ and $z_{j+1}$ so that $\LL^{(c2)}_j(\F',Y')$ vanishes. Because of the identity \bal\lb{srp} \sum_{Y\in \SS_{j+1}}^{Y\supset D} Q^*_j(\F', D,Y) &=\frac12\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\tilde t_j,\f,D);V_{0,j}(\tilde t_j,\f,D^*)\rgt] \notag\\ &+\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,D);V_{0,j}(\tilde t_j,\f,D^*)\rgt] \notag\\ & +\frac12\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,D);V_{1,j}(\tilde t_j,\F,D^*)\rgt], \eal and because of computations in Section \ref{sc2}, we finally set: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item Effective couplings $s_{j+1}$ and $z_{j+1}$, \bal\lb{lk3} &s_{j+1}=s^*_j-a_jz^2_j \notag\\ &z_{j+1}=z^*_j-L^2 e^{-\frac{\a^2}{2}\G_j(0)} b_js_j z_j \eal where, setting $\G_{j,n}(x):=\sum_{m=n}^j\G_{m}(x)$ and $\G_j(0|x):= \G_j(0)-\G_j(x)$, the coefficients in \pref{lk3} are $a_0=0$, $b_0=0$, and, for any $j\ge1$, \bal\lb{dfm3} &a_j:={\a^2\over 2} \sum_{y\in \mathbb{Z}}|y|^2 \left} \def\rgt{\right[ w_{b,j}(y)\left} \def\rgt{\right(e^{-\a^2\G_j(0|y)}-1\rgt) + e^{-\a^2\G_j(0)}\left} \def\rgt{\right(e^{\a^2\G_j(y)}-1\rgt)L^{-4j}\rgt], \notag\\ &b_j:={\a^2 \over 2}\sum_{y\in \mathbb{Z}\atop\m\in\hat u} \left} \def\rgt{\right[\left} \def\rgt{\right(\dpr^\m\G_{j}\rgt)^2(y) + 2\sum_{n=0}^{j-1}\left} \def\rgt{\right(\dpr^\m\G_{n}\rgt)(y)\left} \def\rgt{\right(\dpr^\m\G_{j}\rgt)(y) e^{-{\a^2\over2}\G_{j-1,n}(0)} L^{2(j-n)}\rgt]. \eal \item Effective free energy $E_{j+1}$ \bal \lb{lk3b} &E_{j+1}=E^*_{j}+L^{-2j}\left} \def\rgt{\right[s^2_j\EE_{3,j}+ z^2_j\EE_{4,j}\rgt] \eal where the coefficients in \pref{lk4} are $\EE_{3,0}=\EE_{4,0}=0$ and, for any $j\ge 1$, \bal\lb{dfm4} &\EE_{3,j}:={L^{2j}\over 4}\sum_{y\in \mathbb{Z}^2}\sum_{\m,\nu} \def\p{\pi} \def\r{\rho\in\hat u} \Big[(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho\G_{j})(y) + 2(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho\G_{j-1,1})(y)\Big] (\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho\G_{j})(y), \notag\\ & \EE_{4,j}:=2L^{2j}\sum_{y}w_{0,b,j}(y)\left} \def\rgt{\right[e^{-\a^2\G_j(0|y)}-1-{\a^2\over 2} |y|^2\sum_{\m\in \hat u}(\dpr^{-\m}\dpr^\m \G_j)(0)\rgt] \notag\\ &\qquad\qquad +L^{-2j}\sum_{y}e^{-\a^2\G_j(0)} \left} \def\rgt{\right(e^{\a^2\G_j(y)}-1\rgt). \eal \item Fractional charge renormalization constants $Z_{j+1}$ and $\bar Z_{j+1}$ \bal\lb{lk4} &Z_{j+1}=Z^*_{j}+L^2 e^{-\h^2\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right( -m_{1,1,j} s_j Z_j + m_{1,2,j} z_j \bar Z_j\rgt), \notag\\ &\bar Z_{j+1}=\bar Z^*_j+L^2 e^{-\bar\h^2\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right(-m_{2,2,j} s_j\bar Z_j+m_{2,1,j} z_j Z_j \rgt), \eal where the coefficients are, for any $j\ge 0$, \bal\lb{dfm4} &m_{1,1, j}= \frac{\a^2\h^2}{4} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} \left} \def\rgt{\right[ (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y)+2\sum_{n=0}^{j-1} (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y)\left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y)-(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(0)\rgt] L^{2(j-n)} e^{-\h^2\frac{\a^2}{2} \G_{j-1,n}(0)}\rgt], \notag\\ &m_{2,2, j}= \frac{\a^2\bar \h^2}{4} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y)+2\sum_{n=0}^{j-1} (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y) \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\G_j)(y)-(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(0)\rgt] L^{2(j-n)} e^{-\bar \h^2\frac{\a^2}{2} \G_{j-1,n}(0)} \rgt], \notag\\ &m_{1,2, j}= \sum_{y\in \mathbb{Z}^2} \left} \def\rgt{\right[\bar w_{2,c,j}(y)\left} \def\rgt{\right(e^{-\a^2\bar \h\G_j(y|0)}-1\rgt) +L^{-2j} e^{\bar \h\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\bar \h\a^2\G_j(y)}-1\rgt) \rgt], \notag\\ &m_{2,1, j}= \sum_{y\in \mathbb{Z}^2} \left} \def\rgt{\right[w_{2,c,j}(y)\left} \def\rgt{\right(e^{\a^2\h\G_j(y|0)}-1\rgt) +L^{-2j} e^{-\h\a^2\G_j(0)}\left} \def\rgt{\right(e^{\h\a^2\G_j(y)}-1\rgt) \rgt]. \eal \item The functions $w$'s in \pref{wj0} are all vanishing for $j=0,1$; while, for $j\ge 2$, \bal\lb{lwn} &w_{0,a,j}^{\m\nu} \def\p{\pi} \def\r{\rho}(y)={1\over 2}\sum_{n=1}^{j-1} (\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_{n})(y), \notag\\ &w_{0,b,j}(y)= \frac12\sum_{n=1}^{j-1} e^{-\a^2\G_{j-1,n+1}(0|y)} e^{-\a^2\G_n(0)}\left} \def\rgt{\right(e^{\a^2\G_n(y)}-1\rgt)L^{-4n}, \notag\\ &w_{0,c,j}(y)={1\over 2}\sum_{n=1}^{j-1} e^{-\a^2[\G_{j-1,n+1}(0)+\G_{j-1, n+1}(y)]} e^{-\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\a^2\G_n(y)}-1\rgt)L^{-4n}, \notag\\ &w^{\m}_{0,d,j}(y)={\a\over 2}\sum_{n=1}^{j-1} e^{-{\a^2\over 2}\G_{j-1, n}(0)} (\dpr^\m\G_{n})(y) L^{-2n}, \notag\\ &w_{0,e,j}(y)={\a^2\over 4}\sum_{n=1}^{j-1} e^{-{\a^2\over 2}\G_{j-1,n}(0)} \sum_{\m\in \hat u} \left} \def\rgt{\right[\left} \def\rgt{\right(\dpr^\m\G_{j-1,n}\rgt)^2(y)- \left} \def\rgt{\right(\dpr^\m\G_{j-1,n+1}\rgt)^2(y)\rgt] L^{-2n}. \eal \item The functions $w$'s in \pref{wj1} are \bal\lb{lw1n} w_{1,b,j}(y)&=\sum_{n=0}^{j-1}L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{-\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{-\h\a^2\G_n(y)}-1\rgt), \notag\\ \bar w_{1,b,j}(y)&=\sum_{n=0}^{j-1} L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{\bar \h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{\bar \h\a^2\G_n(y)}-1\rgt), \notag\\ w_{1,c,j}(y)&=\sum_{n=0}^{j-1} L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{\h\a^2\G_n(y)}-1\rgt), \notag\\ \bar w_{1,c,n}(y)&=\sum_{n=0}^{j-1} L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{-\bar\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{-\bar \h\a^2\G_n(y)}-1\rgt), \notag\\ w^{\nu} \def\p{\pi} \def\r{\rho}_{1,d,j}(y)&= i\a\h\sum_{n=0}^{j-1} (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y), \notag\\ \bar w^{\nu} \def\p{\pi} \def\r{\rho}_{1,d,j}(y)&= i\a\bar \h\sum_{n=0}^{j-1} (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y). \eal \item the functions $w$'s in \pref{wj2} are \bal\lb{hofern} w^{\e}_{2,a,j}(y)&=\frac12 \sum_{n=0}^{j-1} Z_n^2L^{-4n} e^{-\h^2(1+\e)\a^2\G_{j-1,n+1}(0)} e^{-\h^2\a^2\e\G_{j-1,n+1}(y|0)} \notag\\ &\qquad\times e^{-\h^2\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\h^2\a^2\e \G_n(y)}-1\rgt), \notag\\ \bar w^{\e}_{2,a,j}(y)&=\frac12 \sum_{n=0}^{j-1} \bar Z_n^2L^{-4n} e^{-\bar \h^2(1+\e)\a^2\G_{j-1,n+1}(0)} e^{-\bar \h^2\a^2\e\G_{j-1,n+1}(y|0)} \notag\\ &\qquad\times e^{-\bar \h^2\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\bar \h^2\a^2\e \G_n(y)}-1\rgt), \notag\\ w^{\e}_{2,b,j}(y)&=\frac12 \sum_{n=0}^{j-1} Z_n\bar Z_n L^{-4n} e^{-(\h+\e\bar\h)^2\frac{\a^2}{2}\G_{j-1,n+1}(0)} e^{-\h\bar \h\a^2\e\G_{j-1,n+1}(y|0)} \notag\\ &\qquad\times e^{-(\h^2+\bar \h^2)\frac{\a^2}{2}\G_n(0)}\left} \def\rgt{\right(e^{-\h\bar \h\a^2\e \G_n(y)}-1\rgt), \notag\\ w^{\e}_{2,c,j}(y)&= \sum_{k=0}^{j-1} L^{-4k} e^{-L^{-k} |y|} \sum_{n=k}^{j-1} e^{-\frac{\a^2}2(1+\e)^2(\h-\frac12)^2 \G_{j-1,n+1}(0)} 2^{-(n-k)} \notag\\ &\times \left} \def\rgt{\right\{Z_k^2 \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(a,k)}\left} \def\rgt{\right(-\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt)\rgt] \rgt. \notag\\ &\quad+\left} \def\rgt{\right. \bar Z^2_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}\mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(\bar a,k)}\left} \def\rgt{\right(\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt) \rgt] \rgt. \notag\\ &\hskip-2em+\left} \def\rgt{\right. Z_k\bar Z_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(b,k)}\left} \def\rgt{\right(-\s\frac{1-\e}2,\z, X, 0, \s, y,\s\e\rgt) \rgt]\rgt\}. \eal Note that, because of the smallness condition on $X$, $w^\e_{2,c,j}(y)=0$ for $|y|\ge 8L^{j-1}$. \ed \begin{lemma} \lb{bois2} $$ \LL^{(c2)}_j(\F',Y')=0. $$ \end{lemma} By \pref{p1}, $W_j(\f, B)$ depends on the fields $\f_x$ and $J_{x,\s}$ for $x$ in a neighborhood of $B$ of diameter $L^j/2$, which is a subset of $B^*$. Finally, joining \pref{lk3} with \pref{lk2} we obtain \pref{lk} and \pref{5.25def}; and condition \pref{cond1} is fulfilled. \subsection{Proof of Lemma \ref{t3.1bb} }\lb{s6.1} The formulas for $\{\MM_{m,n,j}:m,n=1,2\}$ are in \pref{dfm1}. The bounds \pref{scott2ab} directly descend from \pref{eqdh1} and \pref{eqdh1bar} at $\f'=0$. For example, for $\l=\frac12$ and a $C\=C(\a,L)$, \bal\lb{mmj11} &|\MM_{1,1,j}| \le \frac{e^{\h^2\frac{\a^2}{2}\G_j(0)}}{2}\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni0} \|\mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\z,X, 0,\s)\rgt]\|_{h,T_{j+1}(0,X)} \notag\\ &\le C\;\|K_{1,j}\|_{1,h,T_j} \sum_{X\in \SS_j}^{X\ni0} \left} \def\rgt{\right(\frac A2\rgt)^{-|X|_j} \le CS^2\;k^*_s(A,\l) A^{-1} \|K_{1,j}\|_{1,h,T_j}. \eal The other $\MM_{p,q,j}$'s can be studied in a similar way. \subsection{Proof of Lemma \ref{faculty}}\lb{s6.4} In \citep{Fa12} we already proved formula \pref{allen}, for $\r(L,A)= C (L^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon} + A^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon'})$, where $C>1$ and $\vartheta} \def\O{\Omega} \def\Y{\Upsilon,\vartheta} \def\O{\Omega} \def\Y{\Upsilon'>0$. We only need to derive \pref{arkani} and \pref{goddar}. Consider $\LL^{(a)}_{1,j} (\f',V,x,\s)$ and decompose $$ \LL^{(a)}_{1,j} (\f',V,x,\s)=\sum_{n=1}^3\LL^{(n)}_{j} (\f',V,x,\s), $$ where, with Taylor expansions in $\d \f'$, \bal \LL^{(1)}_j (\f',V,x,\s) &:=\frac{Z_j}{Z_{j+1}}L^2 \sum_{Y\in\not\SS_j(V)\atop Y\ni x}^{\bar Y=V} \mathbb{E}_j[K_{1,j}(\f,Y, x,\s)], \lb{el1} \\ \LL^{(2)}_j (\f',V,x,\s) &:=\frac{Z_j}{Z_{j+1}}L^2 \sum_{Y\in \SS_j(V)\atop Y\ni x}^{\bar Y=V} \sum_{q\in \mathbb{Z}\atop |q+\h \s|>1} \mathbb{E}_j\Big[\hK_{1,j}(q,\f,Y, x,\s)\Big], \lb{el2} \\ \LL^{(3)}_j (\f',V,x,\s) &:=\frac{Z_j}{Z_{j+1}}L^2\sum_{Y\in \SS_j(V)\atop Y\ni x}^{\bar Y=V} \sum_{q=0, -\s} \Rem_{1,\d\f'}\;\mathbb{E}_j[\hK_{1,j}(q, \f,Y, x,\s)]. \lb{el3} \eal Let us consider each of the terms, assuming $|Z_j/Z_{j+1}|\le 1$. \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item {\it Norm of $\LL^{(1)}$.} Use \pref{6.5}, as well as a simple extension of \pref{6.4} to activities with a pinning point, to find \bal &\|\LL^{(1)}_{j}(\f',V,x,\s)\|_{h,T_{j+1}(\f',V)} \le L^2 \sum_{Y\in\not\SS_j\atop Y\ni x}^{\bar Y=V} \|\mathbb{E}_j\left} \def\rgt{\right[K_{1,j}(\f,Y, x,\s)\rgt]\|_{h,T_{j+1}(\f',Y)} \notag\\ &\qquad\qquad\qquad\qquad \le G_{j+1}(\f',V)\|K_{1,j}\|_{1,h,T_j} L^2\sum_{Y\in\not\SS_j\atop Y\ni x}^{\bar Y=V} A^{-|Y|_j}2^{|Y|_j} \notag\\ &\qquad\qquad\qquad\qquad \le G_{j+1}(\f',V)A^{-|V|_{j+1}}\;\|K_{1,j}\|_{1,h,T_j}\; L^2 k_l(A,1/2). \eal By \pref{6.90}, we find $$ \|\LL^{(1)}_{j}\|_{1,h,T_{j+1}}\le \d_1(L,A)\|K_{1,j}\|_{1,h,T_j} $$ with $\d_1(A,L)=L^2 A^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon}$: this quantity can be made as small as needed since $A$ is chosen after $L$. \item {\it Norm of $\LL^{(2)}$.} Use \pref{6.5} and \pref{eqdh1} to find, for $C\=C(\a)$ and if $\a^2\ge 8\p$, \bal &\|\LL^{(2)}_j(\f',V, x,\s)\|_{h,T_{j+1}(\f',V)} \le L^2 \sum_{Y\in \SS_j(V)\atop Y\ni x}^{\bar Y=V} \sum_{q\in \mathbb{Z}\atop |q+\h \s|>1} \|\mathbb{E}_{j}\Big[\hK_{1,j}(q,\f,Y, x,\s)\Big]\|_{h,T_{j+1}(\f',Y)} \notag\\ &\qquad \le G_{j+1}(\f',V)A^{-|V|_{j+1}} \;\|K_{1,j}\|_{h,T_j} \; k^*_s(A,1/2)L^2\sum_{q\in \mathbb{Z}\atop |q+\h \s|>1} L^{-4|q+\s\h|+2} C^{2|q+\s\h|}; \eal by \pref{6.90} we obtain $$ \|\LL^{(2)}_{j}\|_{1,h,T_{j+1}}\le \d_2(L,A)\|K_{1,j}\|_{1,h,T_j} $$ with $\d_2(A,L)= C L^{-4\min\{|\h|,|\bar\h|\}}$: this quantity can be made small by taking $L$ large enough, given the choice of $\h$. \item {\it Norm of $\LL^{(3)}$.} By \pref{6.5} and \pref{6.62b} \bal &\|\LL^{(3)}_j (\f',V,x,\s)\|_{h,T_{j+1}(\f',V)} \le L^2 \sum_{Y\in \SS_j(V)\atop Y\ni x}^{\bar Y=V}\sum_{q=0,-\s} \|\Rem_{1,\d\f'}\mathbb{E}_j[\hK_{1,j}(q, \f,Y, x,\s)]\|_{h,T_{j+1}(\f',Y)} \notag\\ &\qquad\qquad \le G_{j+1}(\f',V)A^{-|V|_{j+1}}\;\|K_{1,j}\|_{h,T_j}\; k^*_s(A,1/2)L^2\sum_{q=0,-\s} \r_1(q+\h\s,\a); \eal by \pref{6.90} we obtain $$ \|\LL^{(3)}_{j}\|_{1,h,T_{j+1}}\le \d_3(L,A)\|K_{1,j}\|_{1,h,T_j} $$ with $\d_3(L,A)= \frac{C}{\kappa_L}L^{-\min\{\h^2,\bar\h^2\}}$. \ed This proves the former of \pref{arkani} for $\r(L,A)=\d_1(L,A) +\d_2(L,A) + \d_3(L,A)$. The latter of \pref{arkani} has a similar proof. Let us now consider \pref{goddar}. For $\LL^{(a,k)}_{2,j}$ we have $$ \LL^{(a,k)}_{2,j} (\f',V,x_1,\s_1,x_2,\s_2)=\sum_{p=1}^3\LL^{(a,p,k)}_{2,j} (\f',V,x_1,\s_1,x_2,\s_2) $$ where \bal \LL^{(a,1,k)}_{2,j} (\f',V,x_1,\s_1,x_2,\s_2) &:= \sum_{Y\in\not\SS_j(V)\atop Y\ni x_1}^{\bar Y=V} \mathbb{E}_j\left} \def\rgt{\right[K^{(a,k)}_{2,j}(\f,Y,x_1,\s_1,x_2,\s_2)\rgt], \lb{el1b} \\ \LL^{(a,2,k)}_{2,j} (\f',V,x_1,\s_1,x_2,\s_2) &:= \sum_{Y\in \SS_j(V)\atop Y\ni x_1}^{\bar Y=V} \sum_{q\in \mathbb{Z}\atop q\neq -\frac12(\s_1+\s_2)} \mathbb{E}_j\left} \def\rgt{\right[\hK^{(a,k)}_{2,j}(q,\f,Y,x_1,\s_1,x_2,\s_2)\rgt], \lb{el2b} \\ \LL^{(a,3,k)}_{2,j} (\f',V,x_1,\s_1,x_2,\s_2) &:=\sum_{Y\in \SS_j(V)\atop Y\ni x_1}^{\bar Y=V} \Rem_{0,\d\f'}\;\mathbb{E}_j \left} \def\rgt{\right[\hK^{(a,k)}_{2,j}(-\frac{\s_1+\s_2}{2}, \f,Y,x_1,\s_1,x_2,\s_2)\rgt]. \lb{el3b} \eal With estimates similar to the ones used in the previous discussions, we have a bound \begin{equation}} \def\ee{\end{equation}\lb{goddar3} \|\LL^{(a, p,k)}_{2,j}\|_{2,h,T_{j+1}}\le \d_{a,p,k}(L,A,\h) \|K^{(a,k)}_{2,j}\|_{2,h,T_{j}} \ee where possible choices of the prefactors $\d_{a,p,k}(L,A,\h)$'s are: using the third of \pref{6.90}, $\d_{a,1,k}(L,A,\h)=A^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon}$ for a $\vartheta} \def\O{\Omega} \def\Y{\Upsilon>0$; using the second of \pref{6.90} and \pref{lili}, $\d_{a,2,k}(L,A,\h)= CL^{-d(2\h)}$ for an $\h$ independent $C$ and for $d(2\h)$ defined in Theorem \ref{dh}; using the second of \pref{6.90} and \pref{lili2}, $\d_{a,3,k}(L,A,\h)= C\left} \def\rgt{\right(\sqrt{\kappa_L} L\rgt)^{-1}$. This proves \pref{goddar} for $\d=a$ and $\r(L,A,\h)= \sum_{p=1}^3\d_{a,p,k}(L,A,\h)$. The proof of \pref{goddar} for $\d=\bar a, b$ is similar. \subsection{Proof of Lemma \ref{bourgain}}\lb{s6.4b} \begin{lemma}\lb{sc1} For $V_j$ given in \pref{vv}, \bal\lb{9:01} &\mathbb{E}_j\left} \def\rgt{\right[V_{j}(t_j,\F,B)\rgt]= (\tilde E_j- E_j)+ V_{j+1}(\tilde t_j,\F',B), \eal where \bal \tilde E_j:=E_j-\frac{s_j}{2}|B|& \sum_{\m\in\hat u}(\dpr^{-\m}\dpr^\m\G_j)(0) \eal and $\tilde t_j$ is defined in \pref{5.25}. \end{lemma} {\it\0Proof.}\ From standard results on the correlations of Gaussian measures, \bal\lb{pt1} &\mathbb{E}_j\left} \def\rgt{\right[V_{0,j}(t_j,\F,B)\rgt] = \frac{s_j}{2}\sum_{x\in B\atop \m\in\hat u}(\dpr^\m \f')_x^2 -|B| \frac{s_j}{2}\sum_{\m\in\hat u}(\dpr^{-\m} \dpr^\m\G_j)(0) \notag\\ &\qquad\qquad\qquad\qquad +z_j L^{-2j} e^{-\frac{\a^2}{2}\G_j(0)} \sum_{x\in B\atop \s=\pm1} e^{i\a\s\f'_x} \\\lb{pt1b} &\mathbb{E}_j\left} \def\rgt{\right[V_{1,j}(t_j,\F,B)\rgt] =Z_j L^{-2j} e^{-\frac{\a^2}{2}\h^2 \G_j(0)} \sum_{x\in B\atop \s=\pm}J_{\s,x} e^{i\a\h\s\f'_x} \notag\\ &\qquad\qquad\qquad\qquad +\bar Z_j L^{-2j} e^{-\frac{\a^2}{2}\bar \h^2 \G_j(0)} \sum_{x\in B\atop \s=\pm}J_{\s,x} e^{i\a\bar \h\s\f'_x} \eal These identities give \pref{9:01}. \quad\qed\vskip.8em Via this Lemma, be obtain the following formulas for $\LL^{(c1)}_{1,j}$ and $\LL^{(c1)\dagger}_{1,j}$: \bal &\LL^{(c1)}_{1,j}(\f',V,x,\s) \notag\\ &= \frac{Z_j}{Z_{j+1}}L^2 \sum_{B\in \BB_j(V)\atop B\ni x}^{\bar B=V}\left} \def\rgt{\right[e^{-\h^2\frac{\a^2}{2}\G_j(0)}\MM_{1,1,j} - \sum_{X\in \SS_j}^{X\supset B}\Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0, \d\f'+\z, X, x,\s)\rgt]\rgt]e^{i\h\a\s\f'_x} \notag\\ &+\frac{Z_j}{Z_{j+1}}L^2 \sum_{B\in \BB_j(V)\atop B\ni x}^{\bar B=V}\left} \def\rgt{\right[e^{-\bar\h^2\frac{\a^2}{2}\G_j(0)}\MM_{2,1,j} - \sum_{X\in \SS_j}^{X\supset B}\Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(-\s, \d\f'+\z, X, x,\s)\rgt] \rgt]e^{i\bar\h\a\s\f'_x} \lb{el4} \eal and \bal &\LL^{(c1)\dagger}_{1,j}(\f',V,x,\s) \notag\\ &= \frac{\bar Z_j}{\bar Z_{j+1}}L^2 \sum_{B\in \BB_j(V)\atop B\ni x}^{\bar B=V}\left} \def\rgt{\right[e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)}\MM_{2,2,j} - \sum_{X\in \SS_j}^{X\supset B}\Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}^\dagger(0, \d\f'+\z, X, x,\s)\rgt]\rgt]e^{i\bar \h\a\s\f'_x} \notag\\ &+\frac{\bar Z_j}{\bar Z_{j+1}}L^2 \sum_{B\in \BB_j(V)\atop B\ni x}^{\bar B=V}\left} \def\rgt{\right[e^{-\h^2\frac{\a^2}{2}\G_j(0)}\MM_{1,2,j} - \sum_{X\in \SS_j}^{X\supset B}\Tay_{1,\d\f'} \mathbb{E}_j\left} \def\rgt{\right[\hK^\dagger_{1,j}(\s, \d\f'+\z, X, x,\s)\rgt] \rgt]e^{i\h\a\s\f'_x}. \lb{el4bis} \eal Let us consider the two terms in the square brackets in the first line of \pref{el4}: by \pref{conseq1} and \pref{conseq2} they are equal to \bal\lb{el5} \sum_{X\in \SS_j}^{X\ni x}\sum_{y\in X^*} \mathbb{E}_j\left} \def\rgt{\right[\frac{\dpr\hK_{1,j}}{\dpr \z_y}(0,\z,X,x,\s)\rgt] \left} \def\rgt{\right(\sum_{\m\in\hat u} (y-x)^\m \dpr^\m \f'_x-(\f'_y-\f'_x) \rgt). \eal Note that \pref{el5} depends on $\f'$ only via the factor $u_x(y,\f'):=\sum_{\m\in\hat u} (y-x)^\m \dpr^\m \f'_x-(\f'_y-\f'_x)$ and that, with the notation of \pref{fdiff}, \begin{equation}} \def\ee{\end{equation} D^n u_x(y,\f')\cdot (f_1, \ldots, f_n)= \begin{cases} u_x(y,f_1) \quad &\text{ if $n=1$}\\ 0 \quad &\text{ if $n\ge 2$}\\ \end{cases} .\ee As $X\in S_j$, we have $X^*\subset V^*$ and $|y-x|\le CL^j$, so that \begin{equation}} \def\ee{\end{equation}\lb{sofia} \|u_x(\cdot, \f)\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}\le CL^{-2}\|\nabla_{j+1}^2 \f\|_{L^\io(V^*)}. \ee Finally the $\|\cdot\|_{h,T_{j+1}(\f',X)}$ norm of \pref{el5} is bounded by \bal &\sum_{X\in \SS_j}^{X\ni x}\|\mathbb{E}_j\left} \def\rgt{\right[\hK_{1,j}(0,\z,X,x,\s)\rgt]\|_{h,T_{j+1}(0,X)} \notag\\ &\qquad\times \left} \def\rgt{\right(\| u_x(\cdot,\f)\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}+\sup_{\|f\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_{j+1}}=1}\| u_x(\cdot,f)\|_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}^2_j(X)}\rgt) \notag\\ &\le C L^{-2(1+\h^2)} \|K_{1,j}\|_{1,h,T_j} \left} \def\rgt{\right(1+ \|\nabla_{j+1}^2 \f'\|_{L^\io(X^*)}\rgt) \sum_{X\in \SS_j}^{X\ni x}(A/2)^{-|X|_j} \notag\\ &\le C' \kappa_L^{-1 }L^{-2(1+\h^2)} \|K_{1,j}\|_{1,h,T_j} G^{\rm str}_{j+1}(\f',V) (A/2)^{-1} \eal where, to obtain the second line we used \pref{eqdh} at $\f'=0$ and \pref{sofia}. The other lines in \pref{el4} and \pref{el4bis} can be dealt with exactly the same procedure. Finally, as $|V|_{j+1}=1$, $Z_j/Z_{j+1}\le 1$ and because of \pref{6.100b}, we obtain $$ \|\LL^{(c1)}_{1,j}(\f',V,x,\s)\|_{h,T_{j+1}(\f',V)} \le C \kappa_L^{-1 }L^{-2\h^2} \|K_{1,j}\|_{1,h,T_j} G_{j+1}(\f',V) A^{-|V|_{j+1}} $$ which proves the first of \pref{chaniotis2} for $\r(L,A,\h)=C \kappa_L^{-1 }L^{-2\min\{\h^2, \bar \h^2\}}$. \subsection{Proof of Lemma \ref{bois2}}\lb{sc2} This proof is a detailed calculation of the second order part of the RG map. \begin{lemma} \lb{11:20} If the choice of the $w$'s functions is the one in \pref{lwn}, \pref{lw1n} and \pref{hofern}, and the choice for $E_{j+1}$ and $t_{j+1}$, $\tilde t_j$, $t^*_j$ is the one in Section \ref{sc20}, then, for any $D\in \BB_{j+1}$ and $B\in \BB_j(D)$, \bal\lb{3sc2} &{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B);V_{1,j}(\tilde t_j,\F,D^*)\rgt] +\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B);V_{0,j}(\tilde t_j,\f,D^*)\rgt] \notag\\ &+{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\tilde t_j,\f,B);V_{0,j}(\tilde t_j,\f,D^*)\rgt] +\sum_{X\in \SS_j}^{X\supset B}Q_{2,j}(\F',B,X) \notag\\ &=W_{j+1}(\F',B)-\mathbb{E}_j\left} \def\rgt{\right[W_{j}(\tilde t_j,\F,B)\rgt] +(E_{j+1}-E^*_j)|B|+ V_{j+1}(t_{j+1}-t^*_j,\F',B). \eal \end{lemma} {\it\0Proof.}\ An explicit computation of Gaussian correlations (for $\a$ and $\a'$ any real parameter) yields: \bal\lb{gaussian} &\mathbb{E}^T_j\big[(\dpr^\m\z_x)^2;(\dpr^\nu} \def\p{\pi} \def\r{\rho\z_{x+y})^2\big] =2(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y)^2, \notag\\ &\mathbb{E}^T_j\big[(\dpr^\m\z_x);(\dpr^\nu} \def\p{\pi} \def\r{\rho\z_{x+y})\big] =-(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y), \notag\\ &\mathbb{E}^T_j\big[e^{i\a\z_x};(\dpr^\m\z_{x+y})^2\big] =-\a^2e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\m \G_j)(y)^2, \notag\\ &\mathbb{E}^T_j\big[e^{i\a\z_{x+y}};(\dpr^\m\z_{x})^2\big] =-\a^2e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^{-\m} \G_j)(y)^2, \notag\\ & \mathbb{E}^T_j\big[e^{i\a\z_{x}};(\dpr^\m\z_{x+y})\big] =i\a e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\m \G_j)(y), \notag\\ & \mathbb{E}^T_j\big[e^{i\a\z_{x+y}};(\dpr^\m\z_{x})\big] =-i\a e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^{-\m} \G_j)(y), \notag\\ &\mathbb{E}^T_j\big[e^{i\a\e\z_x};e^{i\a\e \e'\z_{x+y}}\big]= e^{-\frac{\a^2}2\G_j(0)}e^{-\frac{\a'^2}2\G_j(0)} \left} \def\rgt{\right( e^{-\a\a'\G_j(y)}-1\rgt). \eal Let $Y:=D^*\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}$. Let us separate the discussion of \pref{3sc2} into three parts. \* {\it\01. First part.} Our goal is to determine the functions $w_{0,\a, j}(y)$'s and the coefficients $t_{j+1}$ so to satisfy Lemma \pref{bois2} for the part that doesn't depend on $J$: \bal\lb{3sc2first} &\frac12\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\tilde t_j,\F,B);V_{0,j}(\tilde t_j,\F, Y)\rgt] = W_{0,j+1}(\F',B)-\mathbb{E}_j\left} \def\rgt{\right[W_{0,j}(\tilde t_j,\F,B)\rgt] \notag\\ &\qquad+(E_{j+1}-E^*_j)|B| +V_{0,j+1}(t_{j+1}-t^*_j,\F',B). \eal This identity was already verified in \citep{Fa12}. However, here we want to show how to re-derive it by means of an ansatz that can be generalized to the more sophisticated second an third parts. We look for $w_{0,\a, j}(y)$, where $\a$ collects the various labels that appear in \pref{wj0}, into the form of sum of contribution gathered at each scale $n\le j-1$: $$ w_{0,\a, j}(y)=\sum_{n=1}^{j-1} R^{(j-1)}_{0,\a,n}(y). $$ By use of \pref{gaussian}, \bal\lb{tt} &{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\F,B);V_{0,j}(\f, Y)\rgt]= s^2_j|B|\frac{1}{4}\sum_{\m,\nu} \def\p{\pi} \def\r{\rho\in\hat u} \sum_{y\in \mathbb{Z}^2}(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y)^2 \notag\\ &-s^2_j\frac{1}{2} \sum_{\m,\nu} \def\p{\pi} \def\r{\rho\in\hat u} \sum_{y\in \mathbb{Z}^2}(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B}(\dpr^\m\f'_x)(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) \notag\\ &+z^2_j\frac{L^{-4j}}{2} \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)}\left} \def\rgt{\right( e^{{\a^2}\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}e^{i\a\s(\f'_x- \f'_{x+y})} \notag\\ &+z^2_j\frac{ L^{-4j}}{2} \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)}\left} \def\rgt{\right( e^{-{\a^2}\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}e^{i\a\s(\f'_x+\f'_{x+y})} \notag\\ &+z_js_j \frac{i\a L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1}\s\left} \def\rgt{\right[e^{i\a\s\f'_x} (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) -e^{i\a\s\f'_{x+y}} (\dpr^{-\nu} \def\p{\pi} \def\r{\rho}\f'_x)\rgt] \notag\\ &-z_js_j \frac{\a^2L^{-2j}}{4} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}\left} \def\rgt{\right[e^{i\a\s\f'_x} +e^{i\a\s\f'_{x+y}}\rgt]. \eal the above terms have to be re-arranged according to the following rule: each term is to be {\it either power-counting irrelevant} (see discussion after \pref{lili2}) {\it or local} (namely with all the fields $\f'$ dependent by a same point $x$), {\it or constant} (namely independent of $\f'$). Let us discuss each line of the right hand side member. The first line is a constant, which will be absorbed into $E_{j+1}$. The second line appears to be marginal; in fact it is irrelevant, because one can plug in the identity $$ (\dpr^\m\f'_x)(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})=(\dpr^\m\f'_x) \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})-(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x})\rgt]+ (\dpr^\m\f'_x)(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x}) $$ and neglect the last term because of the cancellation \bal &\sum_{\m,\nu} \def\p{\pi} \def\r{\rho\in\hat u}\sum_{y\in \mathbb{Z}^2}(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho\G_j(y))= -\sum_{i,j=0,1} \sin k_i \sin k_j\hat\G_j(k)\Big|_{k=0}=0. \eal The third line is relevant. To write it as the sum of an irrelevant term plus a local one use the identity \bal e^{i\a\s(\f'_x- \f'_{x+y})} &= \left} \def\rgt{\right[e^{i\a\s(\f'_x- \f'_{x+y})}-1+\frac{\a^2}4 |y|^2\sum_{\m\in\hat u} (\dpr^{\m}\f_x)^2\rgt] \notag\\ &+1-\frac{\a^2}4 |y|^2 \sum_{\m\in\hat u} (\dpr^{\m}\f_x)^2. \eal Again, by symmetries, we have neglected a term, the linear order of the Taylor expansion in $y$: when plugged into \pref{tt} this term cancels because it is odd in $\s$. Besides, the term proportional to $|y|^2$ is chosen with a special form thanks to the partial cancellation, for $m,n=0,1$, $$ \sum_{y\in \mathbb{Z}^2} e^{-\a^2 \G_j(0)}\left} \def\rgt{\right(e^{\a^2 \G_j(y)}-1\rgt)y_m y_n= \frac{\d_{m,n}}{2} \sum_{y\in \mathbb{Z}^2} e^{-\a^2 \G_j(0)}\left} \def\rgt{\right(e^{\a^2 \G_j(y)}-1\rgt)|y|^2, $$ which makes irrelevant the sum of the terms in the square brackets. The fourth and the fifth lines of \pref{tt} are irrelevant. The only remaining relevant term is the sixth line. To write it as the sum of an irrelevant term plus a local one use the identity $$ e^{i\a\s\f'_x} +e^{i\a\s\f'_{x+y}}=\left} \def\rgt{\right[e^{i\a\s\f'_{x+y}}-e^{i\a\s\f'_x}\rgt]+2e^{i\a\s\f'_x}. $$ In conclusion, after all such operations, we obtain a new equivalent formula for \pref{tt}: \bal\lb{tt2} &{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\F,B);V_{0,j}(\f, Y)\rgt] \notag\\ &=s^2_j |B|\frac14 \sum_{y\in \mathbb{Z}^2\atop \m,\nu} \def\p{\pi} \def\r{\rho\in\hat u}(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y)+z^2_j |B|L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)} \left} \def\rgt{\right( e^{{\a^2}\G_j(y)}-1\rgt) \notag\\ &-s^2_j\frac12 \sum_{y\in \mathbb{Z}^2\atop \m,\nu} \def\p{\pi} \def\r{\rho\in\hat u}(\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B}(\dpr^\m\f'_x)\left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})-(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x})\rgt] \notag\\ &+z^2_j\frac{L^{-4j}}2 \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)}\left} \def\rgt{\right( e^{{\a^2}\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} \left} \def\rgt{\right[e^{i\a\s(\f'_x- \f'_{x+y})}-1+|y|^2\frac{\a^2}4 \sum_{\m\in\hat u} (\dpr^{\m}\f_x)^2\rgt] \notag\\ &-z^2_j\frac{\a^2L^{-4j} }{2} \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)}\left} \def\rgt{\right( e^{{\a^2}\G_j(y)}-1\rgt)|y|^2 \frac12\sum_{x\in B\atop \m\in\hat u} (\dpr^{\m}\f_x)^2 \notag\\ &+z^2_j\frac{L^{-4j}}{2} \sum_{y\in \mathbb{Z}^2} e^{-{\a^2}\G_j(0)}\left} \def\rgt{\right( e^{-{\a^2}\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}e^{i\a\s(\f'_x+\f'_{x+y})} \notag\\ &+ z_js_j\frac{\a L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1}i\s\left} \def\rgt{\right[e^{i\a\s\f'_x} (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) -e^{i\a\s\f'_{x+y}} (\dpr^{-\nu} \def\p{\pi} \def\r{\rho}\f'_x)\rgt] \notag\\ &-z_js_j\frac{ \a^2L^{-2j}}{4} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}\left} \def\rgt{\right[e^{i\a\s\f'_{x+y}}-e^{i\a\s\f'_x} \rgt] \notag\\ &-z_js_j\frac{ \a^2L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}e^{i\a\s\f'_x} . \eal The first part of our ansatz is that the irrelevant terms which were generated in the above integration provide the $R^{(j)}_{0,\a,j}(y)$'s; more precisely, plugging in \pref{wj0} $\tilde t_j$ instead of $t_j$ (namely replacing $s_j$ and $z_j$ with $s_{j+1}$ and $z_{j+1}$) and then comparing the irrelevant lines with \pref{wj0}, we set \bal\lb{ansatz11} &R^{(j)\m\nu} \def\p{\pi} \def\r{\rho}_{0,a,j}(y)={1\over 2} (\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_{j})(y), \notag\\ &R^{(j)}_{0,b,j}(y)={1\over 2} e^{-\a^2\G_j(0)}\left} \def\rgt{\right(e^{\a^2\G_j(y)}-1\rgt)L^{-4j}, \notag\\ &R^{(j)}_{0,c,j}(y)={1\over 2} e^{-\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\G_j(y)}-1\rgt)L^{-4j}, \notag\\ &R^{(j)\m}_{0,d,j}(y)={\a\over 2} e^{-{\a^2\over 2}\G_{j}(0)} (\dpr^\m\G_{j})(y) L^{-2n}, \notag\\ &R^{(j)}_{0,e,j}(y)={\a^2\over 4} e^{-{\a^2\over 2}\G_{j}(0)} \sum_{\m\in\hat u}(\dpr^\m\G_{j})^2(y) L^{-2j}. \eal Next consider $\mathbb{E}_j\left} \def\rgt{\right[W_{j,0}\rgt]$. As we have done in passing from \pref{tt} to \pref{tt2}, we write the result as a sum of terms each of which is either irrelevant or local, or constant. To do that, we need again some partial cancellations such as, for $m,n=0,1$, $$ \sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y) \left} \def\rgt{\right(e^{\a^2\G_j(y|0)}-1\rgt)y_m y_n=\d_{m,n} \sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y) \left} \def\rgt{\right(e^{\a^2\G_j(y|0)}-1\rgt){|y|^2\over 2}, $$ that is a consequence of the invariance of $w_{0,b,j}(y)$ under the interchange of $y_0$ and $y_1$: this property will be apparent in the final choice of $w_{0,b,j}(y)$ given in \pref{lwn}. The outcome the initial integration and subsequent re-arrangement is \bal\lb{tt3} &\mathbb{E}_j\left} \def\rgt{\right[W_{0,j}(t_j,\f',B)\rgt] \notag\\ &=s_j^2|B| \sum_{y\in \mathbb{Z}^2\atop \m, \nu} \def\p{\pi} \def\r{\rho\in\hat u} w_{0,a,j}^{\m\nu} \def\p{\pi} \def\r{\rho}(y) (\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) +z_j^22 |B|\sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y) \left} \def\rgt{\right(e^{\a^2\G_j(y|0)}-1\rgt) \notag\\ &-z_j^2 |B|\frac{\a^2}{2} \sum_{y\in \mathbb{Z}^2\atop \m\in\hat u}w_{0,b,j}(y) |y|^2 (\dpr^{-\m}\dpr^{\m}\G_j)(0) \notag\\ &-s_j^2 \sum_{y\in \mathbb{Z}^2\atop \m, \nu} \def\p{\pi} \def\r{\rho\in\hat u} w_{0,a,j}^{\m\nu} \def\p{\pi} \def\r{\rho}(y) \sum_{x\in B}(\dpr^\m\f'_x) \Big[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_x)\Big] \notag\\ &+z_j^2\sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y)e^{\a^2\G_j(y|0)} \sum_{x\in B\atop\s=\pm} \Bigg[ e^{i\s\a(\f'_{x}-\f'_{x+y})} -1 +|y|^2\frac{\a^2} 4\sum_{\m\in\hat u}(\dpr^{\m}\f'_x)^2 \Bigg] \notag\\ &-z_j^2 \a^2 \sum_{y\in \mathbb{Z}^2}w_{0,b,j}(y) \left} \def\rgt{\right(e^{\a^2\G_j(y|0)}-1\rgt)|y|^2 \frac12 \sum_{x\in B\atop\m\in\hat u} (\dpr^{\m}\f'_x)^2 \notag\\ &+z_j^2\sum_{y\in \mathbb{Z}^2} w_{0,c,j}(y)e^{-\a^2 (\G_j(0)+\G_j(y))} \sum_{x\in B\atop\s=\pm} e^{i\s\a(\f'_{x}+\f'_{x+y})} \notag\\ &+z_js_j \sum_{y\in \mathbb{Z}^2\atop \m\in\hat u} w^\m_{0,d,j}(y)e^{-\frac{\a^2}{2} \G_j(0)} \sum_{x\in B\atop\s=\pm} i\s\left} \def\rgt{\right[ e^{i\s\a\f'_{x}}(\dpr^\m\f'_{x+y})- e^{i\s\a\f'_{x+y}}(\dpr^{-\m}\f'_x)\rgt] \notag\\ &-z_j s_j \sum_{y\in \mathbb{Z}^2\atop \m\in\hat u} \left} \def\rgt{\right[w_{0,e,j}(y)+\a w^\m_{0,d,j}(y)\dpr^\m \G_j(y)\rgt] e^{-\frac{\a^2}{2} \G_j(0)} \sum_{x\in B\atop\s=\pm} \left} \def\rgt{\right[e^{i\s\a\f'_{x+y}}-e^{i\s\a\f'_{x}}\rgt] \notag\\ &-z_js_j 2\a \sum_{y\in \mathbb{Z}^2\atop \m\in\hat u} w^\m_{0,d,j}(y)e^{-\frac{\a^2}{2} \G_j(0)}\dpr^\m \G_j(y) \sum_{x\in B\atop\s=\pm}e^{i\s\a\f'_{x}}. \eal The second part of our ansatz is that the factors produced in the above integration transform $R^{(j-1)}_{0,\a,n}(y)$ into $R^{(j)}_{0,\a,n}(y)$; more precisely, plugging in \pref{tt3} $\tilde t_j$ instead of $t_j$ and then comparing the irrelevant terms with \pref{wj0}, \bal\lb{ansatz21} &R_{0,a,n}^{(j)\m\nu} \def\p{\pi} \def\r{\rho}(y)=R_{0,a,n}^{(j-1)\m\nu} \def\p{\pi} \def\r{\rho}(y), \notag\\ &R^{(j)}_{0,b,n}(y)= R^{(j-1)}_{0,b,n}(y)e^{\a^2\G_{j}(y|0)}, \notag\\ &R^{(j)}_{0,c,n}(y)= R^{(j-1)}_{0,c,n}(y)e^{-\a^2[\G_{j}(0)+\G_{j}(y)]}, \notag\\ &R^{(j)\m}_{0,d,n}(y)=R^{(j-1)\m}_{0,d,n}(y) e^{-{\a^2\over 2}\G_{j}(0)}, \notag\\ &R^{(j)}_{0,e,n}(y)=R^{(j-1)}_{0,e,n}(y)e^{-{\a^2\over 2}\G_{j}(0)}+R^{(j-1)\m}_{0,d,n}(y) e^{-{\a^2\over 2}\G_{j}(0)}\a \dpr^\m\G_{j}(y). \eal Finally, it is straightforward to solve \pref{ansatz21} with boundary data \pref{ansatz11}; the result is \bal\lb{lw} &R_{0,a,n}^{(j-1)\m\nu} \def\p{\pi} \def\r{\rho}(y)={1\over 2} (\dpr^{-\m}\dpr^\nu} \def\p{\pi} \def\r{\rho \G_{n})(y), \notag\\ &R^{(j-1)}_{0,b,n}(y)= \frac12 e^{-\a^2\G_{j-1,n+1}(0|y)} e^{-\a^2\G_n(0)}\left} \def\rgt{\right(e^{\a^2\G_n(y)}-1\rgt)L^{-4n}, \notag\\ &R^{(j-1)}_{0,c,n}(y)={1\over 2} e^{-\a^2[\G_{j-1,n+1}(0)+\G_{j-1, n+1}(y)]} e^{-\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\a^2\G_n(y)}-1\rgt)L^{-4n}, \notag\\ &R^{(j-1)\m}_{0,d,n}(y)={\a\over 2} e^{-{\a^2\over 2}\G_{j-1, n}(0)} (\dpr^\m\G_{n})(y) L^{-2n}, \notag\\ &R^{(j-1)}_{0,e,n}(y)={\a^2\over 4} e^{-{\a^2\over 2}\G_{j-1,n}(0)} \sum_\m \left} \def\rgt{\right[\left} \def\rgt{\right(\dpr^\m\G_{j-1,n}\rgt)^2(y)- \left} \def\rgt{\right(\dpr^\m\G_{j-1,n+1}\rgt)^2(y)\rgt] L^{-2n}. \eal Besides, collecting the marginal and relevant terms from \pref{tt2} and \pref{tt3} we obtain \bal\lb{dfm2} &a_j:=\a^2\sum_{y\in \mathbb{Z}}|y|^2 \left} \def\rgt{\right[\sum_{n=0}^{j}R^{(j)}_{0,b,n}(y)-\sum_{n=0}^{j-1}R^{(j-1)}_{0,b,n}(y)\rgt], \notag\\ &b_j:=\sum_{y\in \mathbb{Z}\atop\m\in \hat e} \left} \def\rgt{\right[2\a L^{2j} w^\m_{0,d,j}(y)\dpr^\m \G_j(y)+\frac{\a^2}2\left} \def\rgt{\right(\dpr^\m\G_{j}\rgt)^2(y)\rgt]. \eal This proves that \pref{lwn} and \pref{dfm3} yield \pref{3sc2first}. \vskip1em {\it\02. Second term.} This term contains one factor of external field $J_{x,\s}$. We look for $w_{1,\a,j}(y)$ into the form $$ w_{1,\a,j}(y)=\sum_{n=0}^{j-1} R^{(j-1)}_{1,\a,n}(y) $$ where $R^{(j-1)}_{1,\a,n}(y)$ will be determined by means of an ansatz to obtain \bal\lb{3sc2second} \mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B);V_{0,j}(\tilde t_j,\F, Y)\rgt] &=W_{1,j+1}(\F',B)-\mathbb{E}_j\left} \def\rgt{\right[W_{1,j}(\tilde t_j, \F,B)\rgt] \notag\\ &+V_{1,j+1}(t_{j+1}-t^*_j,\F',B). \eal We find \bal\lb{taylor} &\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\F,B);V_{0,j}(\f, Y)\rgt] \notag\\ &=Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\h\f'_x+\f'_{x+y})} \notag\\ &+\bar Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\bar\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{\a^2\bar\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\bar \h\f'_x- \f'_{x+y})} \notag\\ &+Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{\a^2\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\h\f'_x-\f'_{x+y})} \notag\\ &+\bar Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\bar\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\bar \h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\bar \h\f'_x+\f'_{x+y})} \notag\\ & +Z_j s_j i\a \h L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} \s e^{i\h\a\s\f'_x} (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) \notag\\ & +\bar Z_j s_j i\a \bar \h L^{-2j} \sum_{y\in \mathbb{Z}^2\atop \nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} \s e^{i\bar\h\a\s\f'_x} (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) \notag\\ &-Z_js_j \frac{\a^2\h^2L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\h\a\s\f'_x} \notag\\ &-\bar Z_js_j \frac{\a^2\bar \h^2L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\bar \h\a\s\f'_x}. \eal We want to reorganize the summation \pref{taylor} so that every term is either irrelevant or local. The first two lines are irrelevant, because the absolute value of their total charge is $|\h+1|>1$ or $|\bar\h-1|>1 $. The third and fourth lines are relevant; to write it as sum of an irrelevant term and a local one we extract the Taylor expansion in $y^\m$ up to the first order: for example, for the third line this means that we plug in the identity \bal e^{i\a\s(\h\f'_x-\f'_{x+y})} &=e^{i\a\s \bar \h\f'_x} e^{i\a\s(\f'_x-\f'_{x+y})} \notag\\ &=e^{i\a\s \bar \h\f'_x} \left} \def\rgt{\right[e^{i\a\s(\f'_x-\f'_{x+y})}-1-i\a\s\sum_{\m\in\hat u} y^\m(\dpr^\m\f'_x)\rgt] \notag\\ &+e^{i\a\s \bar\h\f'_x} +i\a\s e^{i\a\s \bar \h\f'_x}\sum_{\m\in\hat u} y^\m(\dpr^\m\f'_x). \eal However, since $\sum_{y\in \mathbb{Z}^2}\left} \def\rgt{\right(e^{\a^2\h\G_j(y)}-1\rgt)y^\m=0$, the last term (once replaced into the third line) cancels. The fifth and sixth lines are apparently relevant; in fact, they are irrelevant as one can see by plugging in the identity $$ (\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})=\left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})-(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x})\rgt]+(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x}) $$ and observing that $\sum_{y\in \mathbb{R}^2} (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y)=0$ so that, the last term, which is $y$-independent, give vanishing contribution. Finally, the seventh and eighth lines are relevant; however, they are already local. In conclusion, an equivalent formulation of \pref{taylor} is \bal\lb{taylor2} &\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\F,B);V_{0,j}(\f, Y)\rgt] \notag\\ &=Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\h\f'_x+\f'_{x+y})} \notag\\ &+\bar Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\bar\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{\a^2\bar\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s(\bar \h\f'_x- \f'_{x+y})} \notag\\ &+Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{\a^2\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s \bar \h\f'_x} \notag\\ &\qquad\times \left} \def\rgt{\right[e^{i\a\s(\f'_x-\f'_{x+y})}-1-i\a\s\sum_\m y^\m\dpr^\m\f'_x\rgt] \notag\\ &+Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{\a^2\h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s \bar \h\f'_x} \notag\\ &+\bar Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\bar\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\bar \h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s \h\f'_x} \notag\\ &\qquad\times\left} \def\rgt{\right[e^{-i\a\s(\f'_x-\f'_{x+y})}-1+i\a\s\sum_\m y^\m\dpr^\m\f'_x\rgt] \notag\\ &+\bar Z_jz_j L^{-4j} \sum_{y\in \mathbb{Z}^2} e^{-(1+\bar\h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\a^2\bar \h\G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\a\s \h\f'_x} \notag\\ & +Z_j s_j i\h \a L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\h\a\s\f'_x} \s \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y}) -(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x})\rgt] \notag\\ & +\bar Z_j s_j i \bar \h \a L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \sum_{x\in B\atop \s=\pm1} J_{x,\s} e^{i\bar\h\a\s\f'_x} \s\left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x+y})-(\dpr^\nu} \def\p{\pi} \def\r{\rho\f'_{x})\rgt] \notag\\ &-Z_js_j \frac{\a^2\h^2L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\h\a\s\f'_x} \notag\\ &-\bar Z_js_j \frac{\a^2\bar \h^2L^{-2j}}{2} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} e^{-\bar \h^2\frac{\a^2}{2}\G_j(0)}(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y) \sum_{x\in B\atop \s=\pm1}J_{x,\s} e^{i\bar \h\a\s\f'_x}. \eal Replacing $t_j$ with $\tilde t_j$ and then comparing \pref{taylor2} with \pref{wj1}, we formulate the ansatz \bal\lb{ansatz12} R^{(j)}_{1,b,j}(y)&=L^{-2j} e^{-\frac{\a^2}{2}\G_{j}(0)} \left} \def\rgt{\right(e^{-\a^2\h\G_j(y)}-1\rgt), \notag\\ \bar R^{(j)}_{1,b,j}(y)&= L^{-2j} e^{-\frac{\a^2}{2}\G_{j}(0)} \left} \def\rgt{\right(e^{\a^2\bar \h\G_j(y)}-1\rgt), \notag\\ R^{(j)}_{1,c,j}(y)&= L^{-2j} e^{-\frac{\a^2}{2}\G_{j}(0)} \left} \def\rgt{\right(e^{\a^2\h\G_{j}(y)}-1\rgt), \notag\\ \bar R^{(j)}_{1,c,j}(y)&= L^{-2j} e^{-\frac{\a^2}{2}\G_{j}(0)} \left} \def\rgt{\right(e^{-\a^2\bar \h\G_j(y)}-1\rgt), \notag\\ R^{(j)\nu} \def\p{\pi} \def\r{\rho}_{1,d,j}(y)&= i\a\h (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y) \notag\\ \bar R^{(j)\nu} \def\p{\pi} \def\r{\rho}_{1,d,j}(y)&= i\a\bar \h (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)(y). \eal Next consider $\mathbb{E}_j [W_{1,j}]$: the outcome of the integration and the of the rearrangement into terms that are either irrelevant or local is \bal \lb{taylor3} &\mathbb{E}_j\left} \def\rgt{\right[W_{1,j}(\F, B)\rgt] \notag\\ &= z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} w_{1,b,j}(y) e^{-\frac{\a^2}2 (1+\h^2)\G_j(0)} e^{-\a^2\h\G_j(y)} \sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s(\h\f'_x+\f'_{x+y})} \notag\\ &+z_j \bar Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} \bar w_{1,b,j}(y) e^{-\frac{\a^2}2(1+\bar \h^2)\G_j(0)} e^{\a^2\bar \h\G_j(y)}\sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s(\bar \h\f_x-\f_{x+y})} \notag\\ &+z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} w_{1,c,j}(y) e^{-\frac{\a^2}2\bar\h^2\G_j(0)} e^{\a^2 \h \G_j(y|0)} \sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s\bar \h\f_x} \notag\\ &\qquad\times \left} \def\rgt{\right[e^{-i\a\s(\f_{x+y}-\f_x)}-1+i\a\s y^\m\sum_{\m\in\hat u}(\dpr^\m \f_{x})\rgt] \notag\\ &+z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} w_{1,c,j}(y) e^{-\frac{\a^2}2\bar\h^2\G_j(0)} \left} \def\rgt{\right(e^{\a^2 \h \G_j(y|0)}-1\rgt) \sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s\bar \h\f_x} \notag\\ & +z_j \bar Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} \bar w_{1,c,j}(y) e^{-\frac{\a^2}2 \h^2\G_j(0)} e^{-\a^2 \bar \h \G_j(y|0)} \sum_{x\in B\atop \s=\pm} J_{x,\s}e^{i\a\s\h\f_x} \notag\\ &\qquad\times\left} \def\rgt{\right[ e^{i\a\s(\f_{x+y}-\f_x)} - 1-i\a\s y^\m\sum_{\m\in\hat u}(\dpr^\m \f_{x})\rgt] \notag\\ &+z_j Z_j L^{-2j}\sum_{y\in \mathbb{Z}^2} \bar w_{1,c,j}(y) e^{-\frac{\a^2}2 \h^2\G_j(0)} \left} \def\rgt{\right(e^{-\a^2 \bar \h \G_j(y|0)}-1\rgt) \sum_{x\in B\atop \s=\pm} J_{x,\s} e^{i\a\s \h\f_x} \notag\\ & + s_j Z_j L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u}w^\nu} \def\p{\pi} \def\r{\rho_{1,d,j} (y) e^{-\frac{\a^2}2\h^2\G_j(0)} \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\h\a\s\f_x} \s \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x})\rgt] \notag\\ & + s_j Z_j L^{-2j}i\h\a \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u}w^\nu} \def\p{\pi} \def\r{\rho_{1,d,j} (y) \dpr^\nu} \def\p{\pi} \def\r{\rho \G_j(y) e^{-\frac{\a^2}2\h^2\G_j(0)} \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\h\a\s\f_x} \notag\\ & + s_j \bar Z_j L^{-2j} \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u}\bar w^\nu} \def\p{\pi} \def\r{\rho _{1,d,j} (y) e^{-\frac{\a^2}2\bar\h^2\G_j(0)} \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\bar\h\a\s\f_x} \s \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x+y})- (\dpr^\nu} \def\p{\pi} \def\r{\rho\f_{x})\rgt] \notag\\ & + s_j \bar Z_j L^{-2j}i\bar \h\a \sum_{y\in \mathbb{Z}^2\atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} \bar w^\nu} \def\p{\pi} \def\r{\rho_{1,d,j} (y) \dpr^\nu} \def\p{\pi} \def\r{\rho \G_j(y) e^{-\frac{\a^2}2\bar \h^2\G_j(0)} \sum_{x\in B\atop \s=\pm}J_{x,\s} e^{i\bar \h\a\s\f_x}. \eal Note that, as we did to derive \pref{taylor2}, we used some cancellations, which in this case are consequence of the parity of $w_{1,c,j}(y)$ and $\bar w_{1,c,j}(y)$ in $y$ as seen from \pref{lw1n}: $$ \sum_{y\in \mathbb{Z}^2} w_{1,c,j}(y) e^{-\frac{\a^2}2\bar\h^2\G_j(0)} e^{\a^2 \h \G_j(y|0)} y^\m = \sum_{y\in \mathbb{Z}^2} \bar w_{1,c,j}(y) e^{-\frac{\a^2}2 \h^2\G_j(0)} e^{\a^2 \bar \h \G_j(y|0)} y^\m=0. $$ Therefore, replacing $t_j$ with $\tilde t_j$, we formulate the second part of the ansatz \bal\lb{ansatz22} R^{(j)}_{1,b,n}(y)&=R^{(j-1)}_{1,b,n}(y) e^{-\frac{\a^2}{2}\G_{j}(0)} e^{-\h\a^2\G_{j}(y)}, \notag\\ \bar R^{(j)}_{1,b,n}(y)&=\bar R^{(j-1)}_{1,b,n}(y) e^{-\frac{\a^2}{2}\G_{j}(0)} e^{\bar \h\a^2\G_{j}(y)}, \notag\\ R^{(j)}_{1,c,n}(y)&=R^{(j-1)}_{1,c,n}(y) e^{-\frac{\a^2}{2}\G_{j}(0)} e^{\h\a^2\G_{j}(y)}, \notag\\ \bar R^{(j)}_{1,c,n}(y)&=\bar R^{(j-1)}_{1,c,n}(y) e^{-\frac{\a^2}{2}\G_{j}(0)} e^{-\bar\h\a^2\G_{j}(y)}, \notag\\ R^{(j)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y)&=R^{(j-1)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y), \notag\\ \bar R^{(j)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y)&=\bar R^{(j-1)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y). \eal Finally it is easy to solve \pref{ansatz22} with initial data \pref{ansatz12}: we obtain \bal\lb{lw1} R^{(j-1)}_{1,b,n}(y)&=L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{-\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{-\h\a^2\G_n(y)}-1\rgt), \notag\\ \bar R^{(j-1)}_{1,b,n}(y)&= L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{\bar \h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{\bar \h\a^2\G_n(y)}-1\rgt), \notag\\ R^{(j-1)}_{1,c,n}(y)&= L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{\h\a^2\G_n(y)}-1\rgt), \notag\\ \bar R^{(j-1)}_{1,c,n}(y)&= L^{-2n} e^{-\frac{\a^2}{2}\G_{j-1,n}(0)} e^{-\bar\h\a^2\G_{j-1,n+1}(y)}\left} \def\rgt{\right(e^{-\bar \h\a^2\G_n(y)}-1\rgt), \notag\\ R^{(j-1)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y)&= i\a\h (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y), \notag\\ \bar R^{(j-1)\nu} \def\p{\pi} \def\r{\rho}_{1,d,n}(y)&= i\a\bar \h (\dpr^\nu} \def\p{\pi} \def\r{\rho \G_n)(y). \eal Besides, comparing with \pref{lk4}, the marginal and relevant terms of \pref{taylor2} and \pref{taylor3} give \bal\lb{lw2} m_{2,1, j}&= \sum_{y\in \mathbb{Z}^2} \left} \def\rgt{\right[\sum_{n=0}^{j} R^{(j)}_{1,c,n}(y) e^{-\frac{\a^2}2 (2\h-1) \G_j(0)} -\sum_{n=0}^{j-1} R^{(j-1)}_{1,c,n}(y)\rgt], \notag\\ m_{1,2, j}&= \sum_{y\in \mathbb{Z}^2} \left} \def\rgt{\right[\sum_{n=0}^{j} \bar R^{(j)}_{1,c,n}(y) e^{\frac{\a^2}2 (2\bar\h+1) \G_j(0)} -\sum_{n=0}^{j-1} \bar R^{(j-1)}_{1,c,n}(y)\rgt], \notag\\ m_{1,1, j}&= \frac{\a^2\h^2}2\sum_{y\in \mathbb{Z}^2 \atop \nu} \def\p{\pi} \def\r{\rho\in\hat u} \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y)+ 2(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_{j-1,0})(y) (\dpr^\m \G_j)(y)\rgt], \notag\\ m_{2,2, j}&=\frac{\a^2\bar\h^2}2\sum_{y\in \mathbb{Z}^2 \atop\nu} \def\p{\pi} \def\r{\rho\in\hat u} \left} \def\rgt{\right[(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_j)^2(y)+ 2(\dpr^\nu} \def\p{\pi} \def\r{\rho \G_{j-1,0})(y) (\dpr^\m \G_j)(y)\rgt]. \eal This proves that \pref{lw1n} and \pref{dfm4} yield \pref{3sc2second}. \vskip1em {\it\03. Third term.} This term is quadratic in $J$. We look for $w_{2,\a,j}(y)$ (where again $\a$ is the collections of various labels, compare with \pref{wj2}) into the form $$ w_{2,\a,j}(y)=\sum_{n=1}^{j-1} R^{(j-1)}_{2,\a,n}(y); $$ then $ R^{(j-1)}_{2,\a,n}(y)$ will be determined by an ansatz to obtain \bal\lb{3sc2third} &{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\tilde t_j,\F,B);V_{1,j}(\tilde t_j,\F, Y)\rgt] +\sum_{X\in \SS_j}^{X\supset B}Q_{2,j}(\F',B,X) \notag\\ &=W_{2,j+1}(\F',B)-\mathbb{E}_j\left} \def\rgt{\right[W_{2,j}(\tilde t_j\F,B)\rgt]. \eal The first term in \pref{3sc2third} is \bal\lb{sec1} &{1\over 2}\mathbb{E}^T_j\left} \def\rgt{\right[V_{1,j}(\F,B);V_{1,j}(\F, Y)\rgt] \notag\\ &=Z_j^2\frac{L^{-4j}}{2}\sum_{y\in\mathbb{Z}^2\atop \e=\pm1} e^{-\h^2\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\h^2\a^2\e \G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}J_{x,\s}J_{\s\e,x+y} e^{i\h\a\s(\f'_x+\e \f'_{x+y})} \notag\\ &+Z_j\bar Z_j\frac{L^{-4j}}{2}\sum_{y\in\mathbb{Z}^2\atop \e=\pm1} e^{-(\h^2+\bar \h^2)\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right(e^{-\h\bar \h\a^2\e \G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}\left} \def\rgt{\right[J_{x,\s}J_{\s\e,x+y} e^{i\a\s(\h\f'_x+\e\bar\h \f'_{x+y})}\rgt] \notag\\ &+Z_j\bar Z_j\frac{L^{-4j}}{2}\sum_{y\in\mathbb{Z}^2\atop \e=\pm1} e^{-(\h^2+\bar \h^2)\frac{\a^2}{2}\G_j(0)} \left} \def\rgt{\right(e^{-\h\bar \h\a^2\e \G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}\left} \def\rgt{\right[J_{x,\s}J_{\s\e,x+y} e^{i\a\s(\bar \h\f'_{x}+\e \h \f'_{x+y})}\rgt] \notag\\ &+\bar Z_j^2\frac{L^{-4j}}{2}\sum_{y\in\mathbb{Z}^2\atop \e=\pm1} e^{-\bar \h^2\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\bar\h^2\a^2\e \G_j(y)}-1\rgt) \sum_{x\in B\atop \s=\pm1}J_{x,\s}J_{\s\e,x+y} e^{i\bar\h\a\s(\f'_x+\e \f'_{x+y})}, \eal where the parameters $t_j$ have to be replaced with $\tilde t_j$; taking into account also the second term in \pref{3sc2third}, we set \bal\lb{ansatz13} R^{(j)\e}_{2,a,j}(y)&:=\frac12 Z_j^2L^{-4j} e^{-\h^2\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\h^2\a^2\e \G_j(y)}-1\rgt), \notag\\ \bar R^{(j)\e}_{2,a,j}(y)&:=\frac12 \bar Z_j^2L^{-4j} e^{-\bar \h^2\a^2\G_j(0)}\left} \def\rgt{\right(e^{-\bar \h^2\a^2\e \G_j(y)}-1\rgt), \notag\\ R^{(j)\e}_{2,b,j}(y)&:=\frac12 Z_j\bar Z_j L^{-4j} e^{-(\h^2+\bar \h^2)\frac{\a^2}{2}\G_j(0)}\left} \def\rgt{\right(e^{-\h\bar \h\a^2\e \G_j(y)}-1\rgt), \notag\\ R^{(j)\e}_{2,c,j}(y)&:=\sum_{k=0}^j 2^{-(j-k)} L^{-4k} e^{-L^{-k} |y|} \notag\\ &\times \left} \def\rgt{\right\{Z_k^2 \frac12\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(a,k)}\left} \def\rgt{\right(-\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt)\rgt] \rgt. \notag\\ &+\left} \def\rgt{\right. \bar Z^2_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni 0}\mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(\bar a,k)}\left} \def\rgt{\right(\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt) \rgt] \rgt. \notag\\ &+\left} \def\rgt{\right. Z_k\bar Z_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_j}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,j}^{(b,k)}\left} \def\rgt{\right(-\s\frac{1-\e}2,\z, X, 0, \s, y,\s\e\rgt) \rgt]\rgt\}. \eal Next, we find \bal &\mathbb{E}_j\left} \def\rgt{\right[W_{2,j}(\F,B)\rgt] \notag\\ & =\sum_{y\in\mathbb{Z}^2\atop\e=\pm}w^\e_{2,a,j}(y) e^{-\h^2 \a^2 (1+\e)\G_j(0)}e^{-\h^2 \a^2 \e\G_j(y|0)} \sum_{x\in B\atop \s=\pm}J_{x,\s}J_{\s\e,x+y} e^{i\h\a\s(\f'_x+\e\f'_{x+y})} \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop\e=\pm}\bar w^\e_{2,a,j}(y) e^{-\bar \h^2 \a^2 (1+\e)\G_j(0)}e^{-\bar \h^2 \a^2 \e\G_j(y|0)} \sum_{x\in B\atop \s=\pm}J_{x,\s}J_{\s\e,x+y} e^{i\bar\h\a\s(\f'_x+\e \f'_{x+y})} \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop\e=\pm}w^\e_{2,b,j}(y) e^{-(\h+\e \bar \h)^2 \frac{\a^2}2 \G_j(0)}e^{-\h \bar \h \a^2 \e\G_j(y|0)} \sum_{x\in B\atop \s=\pm} J_{x,\s}J_{\s\e,x+y} \notag\\ &\qquad\qquad\qquad\qquad \times\left} \def\rgt{\right[e^{i\a\s(\h\f'_x+\e\bar\h \f'_{x+y})}+ e^{i\a\s(\bar\h\f'_x+\e\h \f'_{x+y})}\rgt] \notag\\ &+\sum_{y\in\mathbb{Z}^2\atop\e=\pm1}w^\e_{2,c,j}(y) e^{-\frac{\a^2}2(1+\e)^2(\h-\frac12)^2 \G_j(0)} \sum_{x\in B\atop \s=\pm} J_{x,\s}J_{\e\s,x+y} e^{i\a\s(1+\e)(\h-\frac12)\f'_x}. \eal Hence the second part of the ansatz is \bal\lb{ansatz23} R^{(j)\e}_{2,a,n}(y)&= R^{(j-1)\e}_{2,a,n}(y)e^{-\h^2 \a^2 (1+\e)\G_j(0)}e^{-\h^2 \a^2 \e\G_j(y|0)}, \notag\\ \bar R^{(j)\e}_{2,a,n}(y)&=\bar R^{(j-1)\e}_{2,a,n}(y) e^{-\bar \h^2 \a^2 (1+\e)\G_j(0)}e^{-\bar \h^2 \a^2 \e\G_j(y|0)}, \notag\\ R^{(j)\e}_{2,b,n}(y)&=R^{(j-1)\e}_{2,b,n}(y) e^{-(\h+\e \bar \h)^2 \frac{\a^2}2 \G_j(0)}e^{-\h \bar \h \a^2 \e\G_j(y|0)}, \notag\\ R^{(j)\e}_{2,c,n}(y)&=R^{(j-1)\e}_{2,c,n}(y) e^{-\frac{\a^2}2(1+\e)^2(\h-\frac12)^2 \G_j(0)}. \eal Solving \pref{ansatz23} with initial data \pref{ansatz13} we obtain \bal\lb{hofer} R^{(j-1)\e}_{2,a,n}(y)&=\frac12 Z_n^2L^{-4n} e^{-\h^2(1+\e)\a^2\G_{j-1,n+1}(0)} \notag\\ &\qquad\qquad \times e^{-\h^2\a^2\e\G_{j-1,n+1}(y|0)} e^{-\h^2\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\h^2\a^2\e \G_n(y)}-1\rgt) \notag\\ \bar R^{(j-1)\e}_{2,a,n}(y)&=\frac12 \bar Z_n^2L^{-4n} e^{-\bar \h^2(1+\e)\a^2\G_{j-1,n+1}(0)} \notag\\ &\qquad\qquad \times e^{-\bar \h^2\a^2\e\G_{j-1,n+1}(y|0)} e^{-\bar \h^2\a^2\G_n(0)}\left} \def\rgt{\right(e^{-\bar \h^2\a^2\e \G_n(y)}-1\rgt) \notag\\ R^{(j-1)\e}_{2,b,n}(y)&=\frac12 Z_n\bar Z_n L^{-4n} e^{-(\h+\e\bar\h)^2\frac{\a^2}{2}\G_{j-1,n+1}(0)} \notag\\ &\qquad\qquad \times e^{-\h\bar \h\a^2\e\G_{j-1,n+1}(y|0)} e^{-(\h^2+\bar \h^2)\frac{\a^2}{2}\G_n(0)} \left} \def\rgt{\right(e^{-\h\bar \h\a^2\e \G_n(y)}-1\rgt), \notag\\ R^{(j-1)\e}_{2,c,n}(y)&:= e^{-\frac{\a^2}2(1+\e)^2(\h-\frac12)^2 \G_{j-1,n+1}(0)}\sum_{k=0}^n 2^{-(n-k)} L^{-4k} e^{-L^{-k} |y|} \notag\\ &\!\times \left} \def\rgt{\right\{Z_k^2 \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(a,k)}\left} \def\rgt{\right(-\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt)\rgt] \rgt. \notag\\ &+\left} \def\rgt{\right. \bar Z^2_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0}\mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(\bar a,k)}\left} \def\rgt{\right(\s\frac{1+\e}2,\z, X,0, \s, y, \s\e\rgt) \rgt] \rgt. \notag\\ &+\left} \def\rgt{\right. Z_k\bar Z_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ \hat K_{2,n}^{(b,k)}\left} \def\rgt{\right(-\s\frac{1-\e}2,\z, X, 0, \s, y,\s\e\rgt) \rgt]\rgt\}. \eal In conclusion, \pref{hofern} yield \pref{3sc2third}. This concludes the proof of Lemma \ref{11:20}. \quad\qed\vskip.8em \section{Remainder part of the RG map} \begin{lemma}\lb{crone} If $z>0$ is small enough and $|s_j|,|z_j|\le c_0|q_j|$, $\|K_{0,j}\|_{h,T_j}\le c_0|q_j|^2$, there exists $C\=C(A, L, \a)$ such that, \begin{equation}} \def\ee{\end{equation}\lb{cosmo} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}\|_{1, h, T_{j+1}}\le C \left} \def\rgt{\right[|q_j|^2 +|q_j|\|K_{1,j}\|_{1,h,T_j}+|q_j|\|K^\dagger_{1,j}\|_{1,h,T_j}\rgt]; \ee besides the same bound is valid for $\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^\dagger_{1,j}\|_{1, h, T_{j+1}}$. \end{lemma} {\it\0Proof.}\ We begin with an exact formula for ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}$. From \pref{4.19bis} we have $$ P_j(D)= P_{0,j}(D)+ P_{1,j}(D) + P_{2,j}(D)+ P_{\ge 3,j}(D) $$ where, if $\tilde V_{0,j}(D):=V_{0,j}(D)-\mathbb{E}_j\left} \def\rgt{\right[ V_{0,j}(D)\rgt]$ and $\tilde V_{1,j}(D):=V_{1,j}(D)-\mathbb{E}_j\left} \def\rgt{\right[ V_{1,j}(D)\rgt]$, \bal P_{0,j}(D) &= \tilde V_{0,j}(D)-\Big(V_{0,j+1}+\d E_j|D|-\mathbb{E}_j\left} \def\rgt{\right[ V_{0,j}\rgt]\Big) +\left} \def\rgt{\right(e^{U_{0,j}(D)}-1-V_{0,j}(D)\rgt) \notag\\ &- \left} \def\rgt{\right(e^{U_{0,j+1}(D)+\d E_{j}|D|}-1-(V_{0,j+1}(D)+\d E_j|D|)\rgt), \notag\\ P_{1,j}(D) &= \tilde V_{1,j}(D)+ \left} \def\rgt{\right(e^{U_{0,j}(D)}-1\rgt) \tilde V_{1,j}(D) \notag\\ &+\left} \def\rgt{\right(e^{U_{0,j}(D)}-e^{U_{0,j+1}(D)+\d E_{j}|D|}\rgt) \mathbb{E}_j[V_{1,j}(D)] \notag\\ & -e^{U_{0,j+1}(D)+\d E_{j}|D|}\Big(V_{1,j+1}(D)-\mathbb{E}_j[V_{1,j}(D)]\Big) \notag\\ &+e^{U_{0,j}(D)}W_{1,j}(D)-e^{U_{0,j+1}(D)+\d E_{j}|D|} W_{1,j+1}(D), \notag\\ P_{2,j}(D) &=\frac12e^{U_{0,j}(D)}\left} \def\rgt{\right(V_{1,j}(D)^2+W_{1,j}(D)^2+2 W_{2,j}(D)\rgt) \notag\\ &\quad- \frac12e^{U_{0,j+1}(D)+\d E_{j}|D|} \left} \def\rgt{\right(V_{1,j+1}(D)^2+W_{1,j+1}(D)^2+2 W_{2,j+1}(D)\rgt); \eal while $P_{\ge 3,j}(D)$ contains the rest of $P_{j}(D)$. Therefore we find \bal\lb{cosmo:57} {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(1)}_{1,j}(Y')&= \sum_{D\in \BB_{j+1}}^{D=Y'} \mathbb{E}_j\left} \def\rgt{\right[\left} \def\rgt{\right(e^{U_{0,j}(D)}-1-V_{0,j}(D)\rgt)V_{1,j}(D)\rgt] \notag\\ &- \sum_{D\in \BB_{j+1}}^{D=Y'} \left} \def\rgt{\right(e^{U_{0,j+1}(D)+\d E_j|D|}-1-V_{0,j+1}(D)-\d E_j|D|\rgt)\mathbb{E}_j[V_{1,j}(D)] \notag\\ &- \sum_{D\in \BB_{j+1}}^{D=Y'} \left} \def\rgt{\right(e^{U_{0,j+1}(D)+\d E_j|D|}-1\rgt)\Big(V_{1,j+1}(D)-\mathbb{E}_j[V_{1,j}(D)]\Big) \notag\\ &- \sum_{D\in \BB_{j+1}}^{D=Y'} \Big(V_{0,j+1}(D)+\d E_j|D|-\mathbb{E}_j[V_{0,j}(D)]\Big) V_{1,j+1}(D) \notag\\ &- \sum_{D\in \BB_{j+1}}^{D=Y'} \left} \def\rgt{\right(e^{U_{0,j+1}(D)+\d E_j|D|}-1\rgt) W_{1,j+1}(D) \notag\\ &+ \sum_{D\in \BB_{j+1}}^{D=Y'} \mathbb{E}_j\left} \def\rgt{\right[\left} \def\rgt{\right(e^{U_{0,j}(D)}-1\rgt) W_{1,j}(D)\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(2)}_{1,j}(Y')&= \sum_{D_1, D_2\in \BB_{j+1}\atop D_1\neq D_2}^{D_1\cup D_2=Y'} \mathbb{E}_j\left} \def\rgt{\right[\Big(P_{0,j}(D_1)-\tilde V_{0,j}(D_1)\Big)\tilde V_{1,j}(D_2)\rgt] \notag\\ &+ \sum_{D_1, D_2\in \BB_{j+1}\atop D_1\neq D_2}^{D_1\cup D_2=Y'} \mathbb{E}_j\left} \def\rgt{\right[P_{0,j}(D_1)\Big(P_{1,j}(D_2)-\tilde V_{1,j}(D_2)\Big)\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(3)}_{1,j}(Y') &= \sum_{D\in \BB_{j+1}}^{D=Y'}\Big[\mathbb{E}_j[W_{1,j}(D)]-\mathbb{E}_j[W_{1,j}(\tilde t_j, D)]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(4)}_{1,j}(Y') &= \sum_{D_1, D_2\in \BB_{j+1}}^{D_1\cup D_2=Y'} \Big[\mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(D_1); V_{1,j}(D_2)\rgt]- \mathbb{E}^T_j\left} \def\rgt{\right[V_{0,j}(\tilde t_j, D_1); V_{1,j}(\tilde t_j, D_2)\rgt]\Big], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(5)}_{1,j}(Y')&= \sum_{Y_0\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)\ge 1\atop |Z|_{j+1}+{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 2}^{\to Y'} \mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^Z R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0\bs Y_0, (D)}J_{1,j}(D_{Y_0}, Y_0) \notag\\ &+ \sum_{Y_1\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)\ge 1\atop |Z|_{j+1}+{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 2}^{\to Y'} \mathbb{E}_j\left} \def\rgt{\right[P_{j,0}^Z R_{0,j}^{X_1\bs Y_1}R_{1,j}(Y_1) \rgt] J_{0,j}^{X_0, (D)} \notag\\ &+ \sum_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 1\atop B\in \BB_{j+1}(Z)\neq\emptyset}^{\to Y'} \mathbb{E}_j\left} \def\rgt{\right[P_{1,j}(B)P_{0,j}^{Z\bs B} R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(6)}_{1,j}(Y')&= \sum_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 1\atop B\in \BB_{j+1}(Y'\bs W)}^{\to Y'} e^{-\d E_j|Y'|+U_{0,j+1}(Y'\bs W)}\Big(V_{1,j+1}(B)+W_{1,j+1}(B)\Big) \mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^Z R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0, (D)} \notag\\ &+ \sum_{Y_0\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)}^{\to Y'} \left} \def\rgt{\right(e^{-\d E_j|Y'|+U_{0,j+1}(Y'\bs W)}-1\rgt)\mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^Z R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0\bs Y_0, (D)} J_{1,j}(D_{Y_0}, Y_0) \notag\\ &+ \sum_{Y_1\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)}^{\to Y'} \left} \def\rgt{\right(e^{-\d E_j|Y'|+U_{0,j+1}(Y'\bs W)}-1\rgt)\mathbb{E}_j \left} \def\rgt{\right[P_{0,j}^Z R_{0,j}^{X_1\bs Y_1}R_{1,j}(Y_1)\rgt] J_{0,j}^{X_0, (D)} \notag\\ &+ \sum_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 1\atop B\in \BB_{j+1}(Z)}^{\to Y'} \left} \def\rgt{\right(e^{-\d E_j|Y'|+U_{0,j+1}(Y'\bs W)}-1\rgt)\mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^{Z\bs B} P_{1,j}(B) R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0, (D)}, \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(7)}_{1,j}(Y')&= \sum_{Z\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}\atop |Z|_{j+1}\ge 3 }^{Z=Y'}\sum_{B\in \BB_{j+1}(Z)} \mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^{Z\bs B} P_{1,j}(B)\rgt] \notag\\ &+ \left} \def\rgt{\right(e^{-\d E_j|Y'|}-1\rgt) \sum_{Z\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_{j+1}\atop |Z|_{j+1}\ge 2 }^{Z=Y'}\sum_{B\in \BB_{j+1}(Z)} \mathbb{E}_j\left} \def\rgt{\right[P_j^{Y'}\rgt] \notag\\ &+ \left} \def\rgt{\right(e^{-\d E_j|Y'|}-1\rgt) \sum_{B\in \BB_{j+1}}^{B=Y'} \mathbb{E}_j\left} \def\rgt{\right[P_{1,j}(B)- \tilde V_{1,j}(B)\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(8)}_j(Y')&= \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop |{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|\ge 2}^{\bar X=Y'} \sum_{Y_0\in{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X) }\mathbb{E}_{j}\left} \def\rgt{\right[K_{1,j}(Y_0)\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X\bs Y_0)} K_{0,j}(Y)\rgt], \notag\\ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(9)}_j(Y')&=\sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop B\in \BB_{j+1}(Y'\bs X)}^{\bar X=Y'} \mathbb{E}_j\left} \def\rgt{\right[\Big(V_{1,j}(B)+ W_{1,j}(B)\Big) e^{U_{0,j}(Y'\bs X)}\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)} K_{0,j}(Y)\rgt] \notag\\ &+\sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop Y_0\in \mathbb{C}_{j+1}(X)}^{\bar X=Y'} \mathbb{E}_j\left} \def\rgt{\right[\left} \def\rgt{\right(e^{U_{0,j}(Y'\bs X)}-1\rgt)K_{1,j}(Y_0)\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X\bs Y_0)} K_{0,j}(Y)\rgt]. \eal The reason of \pref{cosmo} is that each of the above terms falls into one of two classes: a) those terms which, besides containing a factor of $V_{1,j}$ or $W_{1,j}$, also contain at least two factors of $s_j$, $z_j$, or one factor of $K_{0,j}$; b) those terms which contain one factor of either $K_{1,j}$ or $K^\dagger_{1,j}$ and at least one factor of $s_j,z_j$ or $K_{0,j}$ To proceed more formally, we need formula (6.74) of \citep{Br07} and a simpler version of the bounds in Lemma 14 of \citep{Fa12}. \begin{lemma}[\citep{Br07}] There exists a $\vartheta} \def\O{\Omega} \def\Y{\Upsilon >0$ such that, for any $X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j$, \begin{equation}} \def\ee{\end{equation}\lb{6:25} (1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon) |\bar X|_{j+1}\le |X|_j + 8 (1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon ) |{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|. \ee \end{lemma} \begin{lemma}[\citep{Fa12}] Under the hypothesis of Lemma \ref{crone}, for a $\vartheta} \def\O{\Omega} \def\Y{\Upsilon >0$, there exists a $C\=C(A,L,\a)$ such that \bal \lb{6:27} &\|e^{U_{0,j}(\f,D)}-1\|_{h,T_j(\f,D)}\le C |q_j| G_j^{\rm str}(\f,D), \\ \lb{6:28} &\|e^{U_{0,j+1}(\f',D)+\d E_j|D|}-1\|_{h,T_j(\f',D)}\le C |q_j| G^{\rm str}_{j+1}(\f',D), \\ &\|P_{0,j}(\f',\z,D)\|_{h,T_j(\f',D)} \le C A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)} |q_j|\left} \def\rgt{\right[G^{\rm str}_j(\f,D)+G^{\rm str}_{j+1}(\f',D)\rgt], \lb{6:29} \\ &\|J_{0,j}(\f',D,Y)\|_{h,T_j(\f',Y)} \le C A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|D^*|_{j+1}} |q_j|^2 G^{\rm str}_{j+1}(\f',D), \lb{6:30} \\ \lb{6:31} &\|R_{0,j}(\f',\z,Y)\|_{h,T_j(\f',Y)} \le C A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}}|q_j|^2 \left} \def\rgt{\right[G^{\rm str}_{j+1}(\f',Y)+ G_{j}(\f,Y)\rgt]. \eal \end{lemma} Let us consider, for example, ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(8)}_j(Y)$. Extracting the dependence in $J$ be obtain, as usual, two terms: \bal {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(8)}_j(\f',Y, x,\s)&= \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop |{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|\ge 2}^{\bar X=Y} \sum_{Y_0\in{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)\atop Y_0\ni x } \mathbb{E}_{j}\left} \def\rgt{\right[K_{1,j}(\f,Y_0, x,\s)\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X\bs Y_0)} K_{0,j}(Y)\rgt], \eal and a similar one proportional to $K^\dagger_{1,j}(\f,Y_0, x,\s)$. Using \pref{6.51}, \pref{6.54} and the inequality $A^{-|X|_j}\le A^{-(1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}} A^{8(1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|}$ which is a consequence of \pref{6:25}, a bound for $\|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(8)}_j(\f',Y,x,\s)\|_{h,T_j(\f',Y)}$ is, for a $C\=C(A,L,\a)$, \bal &G_{j+1}(\f',Y)\|K_{1,j}\|_{1,h,T_j} \sum_{X\in {\mathcal P}}\def\EE{{\mathcal E}}\def\MM{{\mathcal M}}\def\VV{{\CMcal V}_j\atop |{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_j(X)|\ge 2}^{\bar X=Y} A^{-|X|_{j}} 2^{|X|_j} (C |q_j|)^{|C_j(X)|-1} \notag\\ &\le G_{j+1}(\f',Y)\|K_{1,j}\|_{1,h,T_j} 4^{L^2|Y|_{j+1}} A^{-(1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}} \sum_{p\ge 2} A^{8(1+2\vartheta} \def\O{\Omega} \def\Y{\Upsilon)p} (C |q_j|)^{p-1} \notag\\ &\le G_{j+1}(\f',Y)\|K_{1,j}\|_{1,h,T_j} A^{-|Y|_{j+1}} C_1 |q_j|, \eal where the last inequality holds if one first chooses $A$ large enough so that $4^{L^2} A^{-2\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\le 1$, and then chooses $|q_1|$ small enough so that the series in $p$ is convergent. To obtain the second line we also used that in the sum in the first line there are no more than $2^{|Y|_j}\le 2^{L^2|Y|_{j+1}}$ terms. As a second sample case, consider one of the terms in ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(6)}_{1,j}(Y)$, which, after the extraction of $Z_j L^{2j}J_{x,\s}$, is \bal\lb{test} \frac{Z_{j+1} }{Z_{j}}L^{-2} e^{i\a\h\s\f'_x} \sum_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 1\atop B\in \BB_{j+1}(Y\bs W), B\ni x}^{\to Y} e^{-\d E_j|Y|+U_{0,j+1}(Y\bs W)} \mathbb{E}_j\left} \def\rgt{\right[P_{0,j}^Z R_{0,j}^{X_1}\rgt] J_{0,j}^{X_0, (D)}. \eal A bound for the norm $\|\cdot\|_{h,T(\f',Y)}$ of this term is made of three kinds of factors: a product of field regulators, a product of factors of $A^{-1}$, and a product of factors of $|q_j|$. Collecting all the factors of field regulators we obtain \bal &G^{\rm str}_{j+1}(\f',Y\bs W)\prod_{D\in \BB_{j+1}(Z)} \Big[G^{\rm str}_j(\f,D)+G^{\rm str}_{j+1}(\f',D)\Big] \notag\\ &\times\prod_{Y\in {\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)}\Big[G_j(\f,Y)+ G^{\rm str}_{j+1}(\f',Y)\Big] G^{\rm str}_{j+1}(\f',X_0) \notag\\ &\le \sum_{W_1\in P_{j+1}(Z)\atop W_2\in ((X_1))_{j+1}} G_j(\f,W_1\cup W_2)G^{\rm str}_{j+1}(\f',Y\bs (W_1\cup W_2)), \eal where $((X_1))_{j+1}$ is the collection of all the possible unions of connected parts of $X_1$. Since the number of terms in the sum over $W_1$ and $W_2$ is not larger than $2^{|Z|_{j+1} +|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)|}$, by \pref{6.54} the expectation of such factors is bounded by $$ 2^{|Z|_{j+1} +|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_1)|}2^{L^2|Z|_{j+1} + L^2|X_1|_{j+1} }G^{\rm str}_{j+1}(\f',Y). $$ Next, collecting the $A^{-1}$ factors coming from \pref{6:29}, \pref{6:30}, \pref{6:31}, we obtain a factor not larger than $A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}}$. In conclusion, a bound for \pref{test} is, for a $C\=C(A,L,\a)$, \bal &G^{\rm str}_{j+1}(\f',Y) 2^{(1+L^2)|Y|_{j+1} }A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}} \notag\\ &\times \sum_{{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)\ge 1}^{\to Y} (C|q_j|)^{|Z|_{j+1}+ 2|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0\cup X_1)|} \notag\\ &\le G^{\rm str}_{j+1}(\f',Y) 2^{(5+L^2)|Y|_{j+1} }A^{-(1+\vartheta} \def\O{\Omega} \def\Y{\Upsilon)|Y|_{j+1}} \sum_{p\ge 1} (C|q_j|)^{2p} \notag\\ & \le C_1 G^{\rm str}_{j+1}(\f',Y) A^{-|Y|_{j+1}} |q_j|^{2}, \eal where we used that the sum in the second line has no more than $4^{|Y|_{j+1}+|{\mathcal C}}\def\FF{{\mathcal F}}\def\HH{{\mathcal H}}\def\WW{{\mathcal W}_{j+1}(X_0)|}\le 2^{4|Y|_{j+1}}$ (indeed each connected component of $X_0$ has to be a small polymer); besides we assumed $A$ large enough so that $2^{(5+L^2)}A^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon}\le 1$, as well as $|q_1|$ small enough so that the series in $p$ is convergent. The other terms of \pref{cosmo:57} can be studied in a similar manner. \quad\qed\vskip.8em \begin{lemma}\lb{fassin} If $z>0$ is small enough and $|s_j|,|z_j|\le c_0|q_j|$, $\|K_{0,j}\|_{h,T_j}\le c_0|q_j|^2$ and $\|K_{1,j}\|_{1,h,T_j}\le c_0|q_j|^2$, there exists $C\=C(A, L, \a)$ such that, \bal\lb{geary} \|{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{\d,k}_{2,j}\|_{2, h, T_{j+1}} \le \begin{cases} C |q_k| &\text{for $k=j$} \\ C |q_j| \|K^{\d,k}_{2,j}\|_{2,h,T_j}\qquad&\text{for $0\le k\le j-1$} \end{cases} \eal \end{lemma} {\it\0Proof.}\ When $k=j$ the term ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{\d,k}_{2,j}$ is generated by a term in $K_{2,j+1}$ which contain at least two factors of $V_{1,j}$, $W_{1,j+1}$ or $K_{1,j}$ and at least one factor of $s_{j}$, $z_j$ or $K_{0,j}$. When $k\le j+1$, term ${\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{\d,k}_{2,j}$ is generated by terms in $K_{2,j+1}$ which contain at least one factor of $K^{\d,k}_{2,j}$ and one factor of $s_{j}$, $z_j$ or $K_{0,j}$. Note that in the case $k=j$ the norms on the right hand side are $\|\cdot\|_{h,T_j}$ or $\|\cdot\|_{1,h,T_j}$, in which the size of the sets are weighed with a factor $A$, whereas on the norm on the left hand side is $\|\cdot\|_{2,h,T_j}$ in which the size of the sets are weighed with a factor $\sqrt A$: this provides the factor $e^{-L^{-k}|x_1-x_2|}$ in \pref{maldacena}. \quad\qed\vskip.8em \subsection{Proof of Theorem \ref{leibler}}\lb{7.1} Let us consider the first of \pref{leibler0}. We have \bal &K_{1,j+1}(\F',Y,x,\s)=\LL_{1,j}(\F',Y,x,\s)+ {\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}_{1,j}(\F',Y,x,\s). \eal From \pref{slk} we see that the assumption of Lemma \ref{crone} is satisfied; therefore, using also \pref{arkani2}, we obtain \bal \|K_{1,j+1}\|_{1,h,T_{j+1}}&\le \r(L,A,\h) \|K_{1,j}\|_{1,h,T_j} \notag\\ &+ C \left} \def\rgt{\right[|q_j|^2 +|q_j|\|K_{1,j}\|_{1,h,T_j}+|q_j|\|K^\dagger_{1,j}\|_{1,h,T_j}\rgt] \eal Assuming by induction that $\|K_{1,j}\|_{1,h,T_j}\le 2C|q_j|^2$ and that $\|K^\dagger_{1,j}\|_{1,h,T_j}\le 2C|q_j|^2$, we obtain that $\|K_{1,j+1}\|_{1,h,T_{j+1}}\le 2C|q_{j+1}|$. (We also used that $\r(L,A,\h)\le 1/4$ for $L$ and $A$ large enough; that $|q_j|/|q_{j+1}|\le 1+ \sqrt{ab}|z|\le 2$ for $|z|$ small enough; and that $|q_j|\le c_0|z|\le \frac18$ for $|z|$ small enough.) This proves the first of \pref{leibler0}. The second can be obtained in a similar way. \subsection{Proof of Theorem \ref{pherson}} Let us consider the bound for $K^{(a,k)}_{2,j+1}$. From the previous definitions, suppressing the dependence in $(\F',Y, x_1, \s_1, x_2, \s_2)$, we have \bal K^{(a,k)}_{2,j+1}=2\LL^{(a,k)}_{2,j} + 2{\mathcal R}}\def\LL{{\mathcal L}}\def\JJ{{\mathcal J}}\def\QQ{{\mathcal Q}^{(a,k)}_{2,j} \eal where the factors $2$ stem from the prefactor $2^{-(j-k)}$ in \pref{maldacena}. Because of \pref{slk} the assumptions in Lemma \ref{fassin} are satisfied; therefore, with the aid of \pref{goddar2}, we have \bal \|K^{(a,k)}_{2,j+1}\|_{2,h,T_{j+1}}&\le 2\r(L,A,\h)\|K^{(a,k)}_{2,j}\|_{2,h,T_{j}} \notag\\ &+ \begin{cases} 2C |q_k| &\text{for $k=j$} \\ 2C |q_j| \|K^{(a,k)}_{2,j}\|_{2,h,T_j}\qquad&\text{for $0\le k\le j-1$} \end{cases} . \eal Therefore, for $L$ and $A$ large enough and $|z|$ small enough it is easy to show inductively a bound such as $\|K^{(a,k)}_{2,j}\|_{2,h,T_{j}}\le 4C|q_k|$. $K^{(\bar a,k)}_{2,j+1}$ and $K^{(b,k)}_{2,j+1}$ can be dealt with in a similar way. \section{Flow of the fractional charge renormalization} Merging \pref{lk4} and \pref{lk2b} we obtain \pref{sk}, namely the equation that describe the flow of the renormalization parameters $Z_j$ and $\bar Z_j$. To study such flows we need an explicit computation of some of the coefficients. In this section we set $\a^2=8\p$. The calculation of \pref{scott1}, was already done in \citep{Fa12}. Note that \pref{scott3} is only valid for $\h=\frac12$; for other values of $\h$ in $(0,1)$ we just need that $|m_{2,1,j}|, |m_{1,2,j}|$ are bounded, see below. Using \pref{slk} and \pref{scott1}, \pref{scott2}, the equation for the fractional charge renormalization constants \pref{sk} becomes \bal\lb{matrix} \begin{pmatrix} Z_{j+1}\\\\\bar Z_{j+1} \end{pmatrix} =&\begin{pmatrix} L^2e^{-4\p\h^2\G_j(0)}& 0\\\\ 0&L^2e^{-4\p\bar \h^2\G_j(0)} \end{pmatrix} \notag\\ &\qquad\times \begin{pmatrix} 1 -\h^2 |q_j|+\tilde \MM_{1,1,j} & m_{1,2,j} z_j+ \MM_{1,2,j} \\\\ m_{2,1,j} z_j+ \MM_{2,1,j}& 1 -\bar \h^2 |q_j|+\tilde \MM_{2,2,j} \end{pmatrix} \begin{pmatrix} Z_j \\\\ \bar Z_j \end{pmatrix} \eal where \bal &\tilde \MM_{1,1,j}=-(m_{1,1,j}s_j-\h^2|q_j|)+ \MM_{1,1,j}, \notag\\ &\tilde \MM_{2,2,j}=-(m_{2,2,j}s_j-\bar \h^2|q_j|)+ \MM_{2,2,j}. \eal Because of \pref{slk}, \pref{leibler0} and \pref{scott2}, for a $C\=C(L)$ and $m=1,2$, \begin{equation}} \def\ee{\end{equation}\lb{pacm} |\tilde \MM_{m,m, j}|\le C \left} \def\rgt{\right[|q_j| L^{-\frac14j} + \frac{\t |q_1|}{[1+|q_1| (j-1)]^{\frac32}}\rgt]. \ee Let us consider two different cases, $|\h|=|\bar \h|$, or $|\h|<|\bar \h|$; the case $|\bar\h|<|\h|$ gives the same formulas after interchanging $Z_j$ and $\h$ with $\bar Z_j$ and $-\bar\h$. \subsection{Case $|\h|=|\bar \h|$} In this case $\h=-\bar \h=\frac12$ and \pref{scott3} holds. Therefore, if we introduce $Z^+_j:= Z_j+\bar Z_j$ and $Z^-_j=Z_j-\bar Z_j$: then \bal\lb{iteration} Z^+_{j+1} &=L^{2}e^{-\p\G_j(0)} \left} \def\rgt{\right(1 -\frac14|q_j|+\frac12 q_j +\MM_{+,j}\rgt) Z^+_{j}, \notag\\ Z^-_{j+1} &=L^{2}e^{-\p\G_j(0)} \left} \def\rgt{\right(1 -\frac14|q_j|-\frac12 q_j +\MM_{-,j}\rgt) Z^-_{j}, \eal where \bal &\MM_{+,j}:= \left} \def\rgt{\right(m_{1,2,j} z_j - \frac 12 q_j\rgt)+\tilde \MM_{1,1,j} + \MM_{2,1,j}, \notag\\ &\MM_{-,j}:= -\left} \def\rgt{\right(m_{1,2,j} z_j - \frac 12 q_j\rgt)+\tilde \MM_{1,1,j} - \MM_{2,1,j}. \eal It is easy to see that also $\MM_{+,j}$ and $\MM_{-,j}$ satisfy the bound \pref{pacm}. In the physical case $z>0$, one has $|q_j|=q_j$ and then \bal\lb{4:58} Z^+_{j+1} &=Z^+_{1} e^{2j\ln L-\p\G_{j,1}(0)+\frac14 \sum_{k=1}^j q_k +\sum_{k=1}^j m_{+,k}}, \notag\\ Z^-_{j+1} &=Z^-_{1} e^{2j\ln L-\p\G_{j,1}(0)-\frac34\sum_{k=1}^jq_k+\sum_{k=1}^j m_{-,k}}, \eal where \bal &m_{+,j}:=\log\left} \def\rgt{\right(1 +\frac14 q_j+\MM_{+,j}\rgt)-\frac14 q_j, \notag\\ &m_{-,j}:=\log\left} \def\rgt{\right(1 -\frac34 q_j+\MM_{-,j}\rgt)+\frac34 q_j. \eal Hence $|m_{+,j}|$ and $|m_{-,j}|$ satisfy a bound like \pref{pacm}. Therefore, for a $C\=C(L)$ and for three constants $\{\tilde c_{\s}:\s=0,\pm\}$ that are vanishing for $\t, |q_1|\to 0$, one has: for $\s=\pm$, $m_{\s,k}$ is summable and $$ \left} \def\rgt{\right|\sum_{k=1}^j m_{\s,k}-\tilde c_{\s}\rgt| \le C \frac{|q_1| +\t}{\sqrt{1+|q_1|(j-1)}}; $$ while $q_k$ is not summable but $$ \left} \def\rgt{\right|\sum_{k=1}^j q_k-\ln (1+ |q_1| j)- \tilde c_0\rgt|\le C|q_j|. $$ Setting $c_+:=\tilde c_++\tilde c_0$ and $c_-:=\tilde c_-+\tilde c_0$, from \pref{4:58} one finds the explicit formula for $Z^+_j$ and $Z^-_j$ in \pref{leibler1}. \subsection{Case $|\h|<|\bar \h|$} When $0<\h<\frac12$, we expect that the sequence $(Z_j)$ dominates $(\bar Z_j)$; therefore we recast \pref{matrix} as \begin{equation}} \def\ee{\end{equation}\lb{pezzini} \begin{pmatrix} Z_{j+1}\\\\\bar Z_{j+1} \end{pmatrix} =L^2e^{-4\p\h^2\G_j(0)-\h^2 |q_j|+m_{j}} \left} \def\rgt{\right[\begin{pmatrix} 1& 0\\\\ 0&\ell_j \end{pmatrix} +\begin{pmatrix} 0 & m_{-,j} \\\\ m_{+,j}&0 \end{pmatrix}\rgt] \begin{pmatrix} Z_j \\\\ \bar Z_j \end{pmatrix} \ee where \bal &m_{j}:=\ln\left} \def\rgt{\right(1-\h^2 |q_j| +\tilde M_{1,1,j}\rgt)+\h^2 |q_j|, \notag\\ &\ell_j:= e^{-4\p(\bar\h^2-\h^2)\G_j(0)+\h^2|q_j|-m_{j}} \left} \def\rgt{\right(1-\bar \h^2|q_j| +\tilde M_{2,2,j}\rgt), \notag\\ &m_{-,j}:=e^{\h^2|q_j|-m_{j}} \left} \def\rgt{\right(m_{1,2,j} z_j +M_{1,2,j}\rgt), \notag\\ &m_{+,j}:=e^{-4\p(\bar\h^2-\h^2)\G_j(0)+\h^2|q_j|-m_{j}} \left} \def\rgt{\right(m_{2,1,j} z_j + M_{2,1,j}\rgt). \eal For a $C\=C(L)$ and $\s=\pm$ one has \bal\lb{rhea} &|m_j|\le C \left} \def\rgt{\right[|q_1| L^{-\frac14j} +\t \frac{|q_1|}{[1+|q_1|(j-1)]^{\frac32}}\rgt], \notag\\ &|\ell_j|\le L^{-2(\bar \h^2-\h^2)}\left} \def\rgt{\right[1+C|q_j| +CL^{-\frac j4}\rgt], \notag\\ &|m_{\s,j}|\le C |q_j|. \eal The difference with the case $\h=-\bar\h$ is that in \pref{rhea} the coefficient $m_{j}$ is absolutely summable in $j$, while $m_{+,j}$ and $m_{-,j}$ are not. This will be compensated by the presence of several factors of $\ell_j<1$. For $z>0$, the solution of \pref{pezzini} is \bal\lb{pezzini2} \begin{pmatrix} Z_{j+1}\\\\\bar Z_{j+1} \end{pmatrix} &= L^{2j} e^{-4\p \h^2\G_{j,1}(0)-\h^2\sum_{k=1}^j q_k +\sum_{k=1}^j m_k} Q(j,1) \begin{pmatrix} Z_1\\\\ \bar Z_1 \end{pmatrix} \eal where $Q(f,i)$ is a two-by-two matrix parametrized by two integers $f\ge i$: \begin{equation}} \def\ee{\end{equation}\lb{daniela} Q(f,i)=\prod_{n=i}^f \left} \def\rgt{\right[\begin{pmatrix} 1& 0\\\\ 0&\ell_n \end{pmatrix} +\begin{pmatrix} 0 & m_{-,n} \\\\ m_{+,n}&0 \end{pmatrix}\rgt]. \ee From the definition \pref{daniela}, for any $1\le j_0\le j$, we have the factorization $Q(j,1)=Q(j,j_0) Q(j_0-1,1)$. We will take advantage of it by choosing a $j_0$ that is large when the difference $\h^2-\bar \h^2$ is small; and estimating $Q(j,j_0)$ and $Q(j_0-1,1)$ in different ways. This will avoid that $L^{-1}$ (and hence $z$) be vanishing in the limit $\h\to\frac12$. \begin{lemma} If $0\le z\le \frac14$, for every $0\le \h<\frac12$ there exist a scale $j_0\=j_0(\h)$ and a constant $C_0\=C_0(\h)$ such that: \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item Estimates for the entries of $Q(j_0-1,1)$ are \bal\lb{melyssa1} &|Q(j_0-1,1)_{1,1}-1|\le C_0 |q_1|^2, \notag\\ &|Q(j_0-1,1)_{1,2}-\sum_{d=1}^{j_0-1} \ell_{1}\cdots\ell_{d-1}m_{-,d}|\le C_0 |q_1|^3, \notag\\ &|Q(j_0-1,1)_{2,1}|\le C_0 |q_1|, \notag\\ &|Q(j_0-1,1)_{2,2}- \ell_1\cdots \ell_{j_0-1}|\le C_0 |q_1|^2. \eal \item For $m=1,2$ the limits $\bar Q_{1,m}(j_0):=\lim_{j\to \io}Q(j,j_0)_{1,m}$ exist and \bal\lb{stat2} &|\bar Q_{1,1}(j_0)-1|\le C_0 \sqrt{|q_1|}, \notag\\ &|\bar Q(j_0)_{1,2}-\sum_{d\ge j_0}\ell_{j_0}\cdots\ell_{d-1}m_{-,d}|\le C_0 |q_1|^{\frac32}; \eal besides, estimates for the speed of convergence of the above limits are \bal\lb{stat1} &|Q(j,j_0)_{1,1}-\bar Q(j_0)_{1,1}|\le C_0 |q_j|, &|Q(j,j_0)_{1,2}-\bar Q(j_0)_{1,2}|\le C_0 |q_j|, \notag\\ &|Q(j,j_0)_{2,1}|\le C_0 |q_j|,\qquad &|Q(j,j_0)_{2,2}-\ell_{j_0}\cdots\ell_{j}|\le C_0 |q_j||q_1|. \eal \ed In the limit $\h\to\frac12$ the constant $C_0(\h)$ is divergent. \end{lemma} {\it\0Proof.}\ Consider \pref{daniela} and expand the product of the sum. The interpretation of the result can be given in terms of the process of two ``states'', $A_1$ and $A_2$, and one ``particle'': at each time $n=i, i+1,i+2,\ldots, f$ the particle can either hold in one of the two states or jump to the other. The ``cost'' of staying in state $A_1$ or $A_2$ at time $n$ is $1$ and $\ell_n$, respectively. The cost of jumping form $A_1$ to $A_2$ or from $A_2$ to $A_1$ at time $n$ is $m_{+,n}$ and $m_{-,n}$ respectively. Let us denote $$ \sum_{u_1,d_1,\ldots,u_n, d_n}^{*[i,f]} $$ the sum with constraint $i\le u_1< d_1<\cdots <u_n< d_n \le f$. \begin{enumerate}[a.\;\leftmargin=1cm]} \def\ed{\end{enumerate} \item The entry $Q(f,i)_{1,1}$ is the sum of the cost of all the patterns that start and end at $A_1$, $$ Q(f,i)_{1,1}=1+ \sum_{n\ge 1}\sum_{u_1,d_1,\ldots,u_n, d_n}^{*[i,f]} \prod_{s=1}^n m_{+,u_s} \ell_{u_s+1}\cdots\ell_{d_s-1} m_{-,d_s}, $$ where: $n$ is the number of intervals of time that the particle has spent in $A_2$; $u_s$'s are the times in which the particle jumps from $A_1$ to $A_2$, and $d_s$'s are the times in which the particle jumps from $A_2$ to $A_1$. As it is easy to see from \pref{rhea}, there exists a constant $C$ such that, if $0\le z\le \frac14$, then $\ell_n\le C$ and $|m_{+,u_s} m_{-,d_s}|\le C |q_1|^2$. Hence, for a $C_0\=C_0(j_0)$ (and divergent in the limit $j_0\to \io$) \begin{equation}} \def\ee{\end{equation}\lb{urra1a} |Q(j_0-1,1)_{1,1}-1|\le C_0 |q_1|^2. \ee However, if $j_0$ is larger that a $j'_0\=j'_0(\h)$, then one has the better bound $\ell_n\le L^{-(\h^2-\bar\h^2)}$ for every $n\ge j_0$. Therefore, if $d_s-1\ge u_s\ge j_0$, then $|m_{+,u_s} m_{-,d_s}|\le C |q_{u_s}|^2 $; if also $d_s-u_s\ge 2$, then $\ell_{u_s+1}\cdots\ell_{d_s-1}\le L^{-(\bar\h^2 -\h^2)(d_{s}-u_s-2)}$. In this way we obtain that $\lim_{j\to \io}Q(j,j_0)_{1,1}$ exists and \bal \left} \def\rgt{\right|Q(j,j_0)_{1,1}-1\rgt| &\le \sum_{n\ge 1} C^n \left} \def\rgt{\right(\sum_{u\ge j_0} q^2_{u} \sum_{w\ge0 } L^{-(\bar\h^2 -\h^2)w}\rgt)^n \notag\\ &\le \sum_{n\ge 1} \left} \def\rgt{\right(\frac{|q_1|}{1+|q_1| j_0} \frac{\tilde C}{1-L^{-(\bar\h^2 -\h^2)}}\rgt)^n . \eal Since $\frac{|q_1|}{1+|q_1| j_0}\le \sqrt{|q_1|} j_0^{-\frac12}$, if $j_0$ is larger than a $j''_{0}(\h)$ the term of the last series is bounded by $|q_1|^{\frac n2}$. Therefore, for $|q_1|<\frac14$, such series is summable and \begin{equation}} \def\ee{\end{equation}\lb{urra1b} \left} \def\rgt{\right|Q(j,j_0)_{1,1}-1\rgt|\le 2\sqrt{|q_1|}. \ee To study the speed of convergence of $\lim_{j\to \io}Q(j,j_0)_{1,1}$, consider an $f>j$ and the difference \bal Q(f,j_0)_{1,1}-Q(j,j_0)_{1,1}=\sum_{n\ge 1}\sum_{u_1,d_1,\ldots,u_n, d_n\atop d_n\ge j+1}^{*[j_0,f]} \prod_{s=1}^n m_{+,u_s} \ell_{u_s+1}\cdots\ell_{d_s-1} m_{-,d_s}. \eal As $|m_{+,u_n} \ell_{u_n+1}\cdots\ell_{d_n-1} m_{-,d_n}|\le C_0 |q_{d_n}|^2 L^{-\frac12(\bar\h^2 -\h^2)(d_n-u_n-2)}$ for a $C_0\=C_0(\h)$, using a similar argument and the constraint $d_n\ge j+1$, $$ |Q(f,j_0)_{1,1}-Q(j,j_0)_{1,1}|\le C_0 |q_j|. $$ \item The entry $Q(f,i)_{1,2}$ is the sum of the cost of all the patterns that start at $A_2$ and end at $A_1$, \bal Q(f,i)_{1,2} &=\sum_{d=i}^f \ell_{i}\cdots\ell_{d-1}m_{-,d} \left} \def\rgt{\right[1+ \sum_{n\ge 1}\sum_{u_1,d_1,\ldots,u_n,d_n}^{*[d+1,f]} \prod_{s=1}^{n-1} m_{+,u_s} \ell_{u_s+1}\cdots\ell_{d_s-1} m_{-,d_s}\rgt] \notag\\ &=\sum_{d=i}^f \ell_{i}\cdots\ell_{d-1}m_{-,d} Q(f,d+1)_{1,1}. \eal Therefore, for constants $C$ and $C_0\=C_0(j_0)$, we have \bal\lb{urrah2a} &\left} \def\rgt{\right|Q(j_0-1,1)_{1,2}-\sum_{d=1}^{j_0-1} \ell_{1}\cdots\ell_{d-1}m_{-,d}\rgt| \notag\\ &\le j_0 C^{j_0}|q_1|\sup_{d\le j_0-1} |Q(j_0-1,d)_{1,1}-1|\le C_0 |q_1|^3. \eal Besides, $\lim_{j\to \io}Q(j,j_0)_{1,2}$ exists and for a constant $C_1\=C_1(\h,j_0)$ \bal\lb{urrah2b} &\left} \def\rgt{\right|Q(j,j_0)_{1,2}-\sum_{d=j_0}^{j} \ell_{j_0}\cdots\ell_{d-1}m_{-,d}\rgt| \notag\\ &\le \frac{C|q_1|}{1-L^{-(\bar\h^2 -\h^2)}} \sup_{d\ge j_0} |Q(j,d)_{1,1}-1|\le C_1 |q_1|^{\frac32}. \eal To study the speed of convergence of the limit consider an $f\ge j+1$ and the difference \bal Q(f,j_0)_{1,2}-Q(j,j_0)_{1,2} &=\sum_{d=j_0}^j \ell_{i}\cdots\ell_{d-1}m_{-,d} \left} \def\rgt{\right[Q(f,d+1)_{1,1}-Q(j,d+1)_{1,1}\rgt] \notag\\ &+\sum_{d=j+1}^f \ell_{i}\cdots\ell_{d-1}m_{-,d} Q(f,d+1)_{1,1}. \eal Then \bal \left} \def\rgt{\right|Q(f,j_0)_{1,2}-Q(j,j_0)_{1,2}\rgt| &\le \frac{C|q_1|}{1-L^{-(\bar\h^2 -\h^2)}} \sup_{j_0\le d\le j}\left} \def\rgt{\right|Q(f,d+1)_{1,1}-Q(j,d+1)_{1,1}\rgt| \notag\\ &+\frac{C|q_j|}{1-L^{-(\bar\h^2 -\h^2)}} \sup_{j+1\le d\le f}\left} \def\rgt{\right| Q(f,d+1)_{1,1}\rgt| \le C_0|q_j|. \eal \item The entry $Q(f,i)_{2,1}$ is the sum of the cost of all the patterns that start at $A_1$ and end at $A_2$, \bal Q(f,i)_{2,1} &=\sum_{u=i}^f \left} \def\rgt{\right[1+ \sum_{n\ge 1}\sum_{u_1,d_1,\ldots,u_n,d_n}^{*[i,u-1]} \prod_{s=1}^{n-1} m_{+,u_s} \ell_{u_s+1}\cdots\ell_{d_s-1} m_{-,d_s}\rgt]m_{+,u} \ell_{u+1}\cdots\ell_{f} \notag\\ &=\sum_{u=i}^f Q(u-1,i)_{1,1} \;m_{+,u} \ell_{u+1}\cdots\ell_{f}. \eal For constants $C$ and $C_0\=C_0(j_0)$, we have \begin{equation}} \def\ee{\end{equation}\lb{urrah3a} |Q(j_0-1,i)_{2,1}|\le j_0C^{j_0} |q_1| \sup_{u\le j_0-1} |Q(u,i)_{1,1}| \le C_0 |q_1|. \ee If $j-1\ge u\ge j_0$, for a constant $C_1\=C_1(\h)$ we have $|m_{+,u}\ell_{u+1}\cdots\ell_{j}|\le C_1 |q_j| L^{-\frac12(\bar \h^2-\h^2)(j-u)}$; hence \begin{equation}} \def\ee{\end{equation}\lb{urrah3} |Q(j,j_0)_{2,1}|\le 2C_1|q_j|. \ee \item The entry $Q(f,i)_{2,2}$ is the total cost of all the patterns that start and end at $A_2$: \bal Q(f,i)_{2,2} &=\ell_{i}\cdots\ell_{f} \notag\\ &+\sum_{d<u=i}^f \ell_{i}\cdots\ell_{d-1}m_{-,d}\; Q(u-1, d+1)_{1,1}\; m_{+,u} \ell_{u+1}\cdots\ell_{f}. \eal For $C_0\=C_0(j_0)$, \begin{equation}} \def\ee{\end{equation}\lb{urrah4a} |Q(j_0-1,i)_{2,2}-\ell_{1}\cdots\ell_{j_0-1}|\le C_0|q_1|^2\;; \ee besides \begin{equation}} \def\ee{\end{equation}\lb{urrah4b} |Q(j,j_0)_{2,2}-\ell_{j_0}\cdots\ell_{j}|\le C_0|q_1||q_j|. \ee \ed From these formulas, one obtains \pref{melyssa1}, \pref{stat2} and \pref{stat1}. \quad\qed\vskip.8em Combining the bounds of this Lemma, we obtain: \bal\lb{am0} &Q(j,1)_{1,1}=e^{\tilde c_1+ \tilde r_{1,j}}, \notag\\ &Q(j,1)_{1,2}=\sum_{d\ge 1} \ell_1\cdots \ell_{d-1} m_{-,d}e^{\tilde c_2+\tilde s_{1,j}}, \notag\\ &Q(j,1)_{2,1}=\tilde r_{2,j,}, \notag\\ &Q(j,1)_{2,2}=e^{-(\bar\h^2-\h^2)[4\p\G_{j,1}(0)+\sum_{k=1}^j q_k]+ \tilde c_3 + \tilde r_{3,j}} + \tilde r_{4,j}, \eal where: $|\tilde c_{1}|\le C_0 \sqrt{|q_1|}$; $|\tilde c_{2}|, |\tilde c_{3}|\le C_0 |q_1|$; $|\tilde r_{1}|, |\tilde r_{2}|, |\tilde r_{3}|\le C_0 |q_j|$; $|\tilde s_{1,j}|\le C_0 (1+|q_1|j)^{-1}$. Plugging \pref{am0} into \pref{pezzini2}, one obtains point 2. of Theorem \ref{leiblerbis}. Finally, note that the lowest order in $z$ of $\lim_{j\to \io}Q(j,1)_{1,2}$ is \bal\lb{3:13pm} &z\sum_{j\ge 1} e^{-4\p (\bar\h^2-\h^2)\G_{j-1,1}(0)} m_{1,2,j} \notag\\ &=z e^{4\p (\bar\h^2-\h^2)\G_{0}(0)} \left} \def\rgt{\right(\sum_{j\ge 0} e^{-4\p (\bar\h^2-\h^2)\G_{j-1,0}(0)} m_{1,2,j} - m_{1,2,0}\rgt) \notag\\ &:=z e^{4\p (\bar\h^2-\h^2)\G_{0}(0)} \left} \def\rgt{\right(c(\h) - m_{1,2,0}\rgt) \eal From the definition of $m_{1,2,j}$ in \pref{lw2} we find \bal &c(\h)=\sum_{n\ge 0}e^{-4\p (\bar\h^2-\h^2)\G_{\io,0}(0)}\sum_{y\in \mathbb{Z}^2}\bar R^{(\io)}_{1,c,n}(y) \notag\\ &= \sum_{n\ge 0}L^{-2n}e^{-4\p (\bar\h^2-\h^2)\G_{n-1,0}(0)}\sum_{y\in \mathbb{Z}^2} e^{-\bar \h \a^2 \G_{\io, n+1}(y|0)}e^{\bar \h\a^2\G_n(0)} \left} \def\rgt{\right(e^{-\bar \h \a^2\G_n(y)}-1\rgt) \eal where the series in $n$ is summable and strictly positive, as one can verify by the inequality $e^x-1\ge x+\frac12 x^2 e^{-x_0}$, valid for any $x:|x|\le x_0$, and by the fact that $-\bar \h> 0$, $|\G_n(y)|\le \G_n(0)$, $$ \sum_{y\in \mathbb{Z}^2}\G_n(y)\ge 0, \qquad \sum_{y\in \mathbb{Z}^2}\G_n(y)^2>0. $$ \section{Exact asymptotic formulas} \subsection{Proof of Lemma \ref{scott} and Lemma \ref{t3.1b}} \lb{pscott} A key result is the following Lemma, in which we introduce a continuous approximation of the covariances $\G_j$ which has a simpler scaling transformation. \begin{lemma} \lb{seiberg} Consider the set of ``continuous covariances'' $\tilde \G_j(x)$, $j=0,1, \ldots R-1$, defined for $x\in \mathbb{R}^2$ as \begin{equation}} \def\ee{\end{equation}\lb{seiberg0} \tilde \G_j(x):=\int\frac{d^2p}{(2\p)^2}\; e^{i x p}\; \frac{u(L^j p)-u(L^{j+1} p)}{p^2} \ee where $u(p)$ is a differentiable even function such that $u(L^j p)-u(L^{j+1} p)\ge 0$ for every $j$ and \begin{equation}} \def\ee{\end{equation}\lb{seiberg00} \qquad u(0)=1,\qquad |u(p)|\le \frac{C}{1+|p|^4} . \ee There exists a special choice of $u$ and a constant $C>0$ such that, for every $x\in \mathbb{Z}^2$, \mp{forse ci vuole la derivata discreta anche per $\tilde \G_j$} \begin{equation}} \def\ee{\end{equation}\lb{seiberg3} |\G_j(x)-\tilde \G_j(x)|\le C L^{-\frac14j}, \qquad |\dpr^\m \G_j(x)-\tilde \G_j^{,\m}(x)| \le C L^{-\frac54j}, \ee where the upper label $^{,\m}$ indicates the continuous derivative (as opposed to $\dpr^\m$ that is the lattice one). \end{lemma} The proof is in Appendix A.3. of \citep{Fa12}. \pref{seiberg0} and \pref{seiberg0} have important consequences: first, for every $x\in \mathbb{R}^2$ \begin{equation}} \def\ee{\end{equation}\lb{seiberg1} \tilde \G_j(x)=\tilde \G_0(L^{-j}x); \ee second, \begin{equation}} \def\ee{\end{equation}\lb{seiberg4} \tilde \G_j(0)= \int\frac{d^2p}{(2\p)^2}\; \; \frac{u(L^j p)-u(L^{j+1} p)}{p^2} =\frac{1}{2\p} \ln L; \ee finally, $\tilde \G_{\io,0}(x|0)$ is a differentiable function and, asymptotically for large $|x|$, \begin{equation}} \def\ee{\end{equation}\lb{seiberg2} \tilde \G_{\io,0}(x|0):=\sum_{j=0}^\io \left} \def\rgt{\right[\tilde \G_j(x)-\tilde \G_j(0)\rgt]=-\frac1{2\p} \ln |x| + \tilde c_E + o(1) \ee where $o(1)$ is a vanishing term in the limit $|x|\to \io$ and $\tilde c_E$ is a constant. Consider the coefficient $a_j$ in \pref{dfm2}. Let $\tilde R^{(j-1)}_{0,b,n}$ and $\tilde \G_j(0|y)$ be the same function as $R^{(j-1)}_{0,b,n}$ and $\G_j(0|y)$, respectively, but with $\tilde \G_j(x)$ in place of $\G_j(x)$ for any $j$. Using \pref{seiberg3} we have \bal \sum_{y\in \mathbb{Z}^2} |y|^2 \left} \def\rgt{\right|R^{(j)}_{0,b,j}-\tilde R^{(j)}_{0,b,j}(y)\rgt| &\le C(L) L^{-\frac14 j} \notag\\ \sum_{y\in \mathbb{Z}^2} |y|^2 \left} \def\rgt{\right|R^{(j)}_{0,b,n}(y)-R^{(j-1)}_{0,b,n}(y)-\tilde R^{(j)}_{0,b,n}(y)+ \tilde R^{(j-1)}_{0,b,n}(y)\rgt| &\le C(L) L^{-\frac j4 }L^{-\frac34 (j-n)} \eal therefore in the definition of $a_j$ we can replace $R^{(j-1)}_{0,b,n}$ with $\tilde R^{(j-1)}_{0,b,n}$ up to an error $C L^{-\frac j4}$. Besides, $$ \left} \def\rgt{\right|\sum_{y\in \mathbb{Z}^2} |y|^2 \tilde R^{(j)}_{0,b,n}(y)-\int\! d^2y \; |y|^2 \tilde R^{(j)}_{0,b,n}(y)\rgt|\le C(L) L^{-j}; $$ therefore in the formula for $a_j$ replacing the sum with an integral generates and error not larger than $CL^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon j}$ for a $\vartheta} \def\O{\Omega} \def\Y{\Upsilon<1$. In conclusion, an equivalent formula for $a_j$ is, up to an $O(L^{-\frac{j}{4}})$ error, \bal\lb{aup} {\a^2\over 2} \int\!d^2y\; y^2 \left} \def\rgt{\right[\sum_{n=0}^j \tilde R^{(j)}_{0,b,n}(y) -\sum_{n=0}^{j-1} \tilde R^{(j-1)}_{0,b,n}(y)\rgt]. \eal We now take advantage of the exact scale transformation \pref{seiberg1}. We have $\tilde R^{(j-1)}_{n}(y)=L^4 \tilde R^{(j)}_{n+1}(yL)$; hence the two sums in \pref{aup} cancel each others almost completely, and \pref{aup} becomes \bal {\a^2\over 2} \int\!d^2y\; y^2 \tilde R^{(j)}_{0,b,0}(y) &={\a^2\over 2} \int\!d^2y\; y^2 e^{-\a^2\tilde \G_{\io,1}(0|y)} e^{-\a^2\tilde \G_{0}(0)} \left} \def\rgt{\right(e^{\a^2\tilde \G_{0}(y)}-1\rgt) +O(L^{-j}) \notag\\ &=\frac{\a^2}{2} \int\!\frac{d^2y}{y^2}\; \left} \def\rgt{\right[w(y)- w(y L^{-1})L^{4-\frac{\a^2}{2\p}} \rgt] +O(L^{-j}) \eal for $w(y)=y^4e^{-\a^2\tilde \G_{\io,0}(0|y)}$; a new $O(L^{-j})$ error in the first line is due to the replacement of $e^{-\a^2\tilde \G_{j,1}(0|y)}$ with $e^{-\a^2\tilde \G_{\io,1}(0|y)}$. At $\a^2=8 \p$, the last integral can be exactly computed only using the differentiability of $w(y)$ and the boundary values $w(0)=0$ and $\lim_{y\to\io}w(y)=e^{-8\p \tilde c_E}$, see \pref{seiberg2}. This proves the first of \pref{scott1}. Now consider the coefficient $m_{2,1,j}$ in \pref{lw2}. Arguing as done for $a_j$, up to $O(L^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon j})$ corrections, it is given by \begin{equation}} \def\ee{\end{equation}\lb{aup2} \int\!d^2 y \; \left} \def\rgt{\right[\sum_{n=0}^{j} \tilde R^{(j)}_{1,c,n}(y)e^{- \frac{\a^2}{2} (2\h-1)\tilde \G_j(0)}- \sum_{n=0}^{j-1} \tilde R^{(j-1)}_{1,c,n}(y) \rgt] \ee where the formula for $\tilde R^{(j-1)}_{1,c,n}(y)$ is obtained from the formula for $R^{(j-1)}_{1,c,n}(y)$ by replacing $\G_j(x)$ with $\tilde \G_j(x)$. In the case $\h=-\bar \h =\frac12$, by means of the exact scaling $\tilde R^{(j-1)}_{1,c,n}(y)=L^2 \tilde R^{(j)}_{1,c,n+1}(yL)$, \pref{aup2} becomes \bal \int\!d^2 y \; \tilde R^{(j)}_{1,c,0}(y) &= \int\!d^2 y \;e^{-\frac{\a^2}2 \tilde \G_{\io,1}(0|y)} e^{-\frac{\a^2}2 \tilde \G_{0}(0)} \left} \def\rgt{\right(e^{\frac{\a^2}2 \tilde \G_{0}(y)}-1\rgt)+ O(L^{-j}) \notag\\ &=\int\!\frac{d^2y}{y^2}\; \left} \def\rgt{\right[\sqrt{w(y)}- \sqrt{w(y L^{-1})}L^{2-\frac{\a^2}{4\p}} \rgt] +O(L^{-j}) \eal where $w(y)$ is the same function introduced for $a_j$. At $\a^2=8\p$ the last integral can be exactly computed only using the differentiability of $\sqrt{w(y)}$ for $y\neq 0$ and the boundary values $\sqrt{w(0)}=0$ and $\lim_{y\to\io}\sqrt{w(y)}=e^{-4\p \tilde c_E}$. This proves the first of \pref{scott3}; the second is also proven because at $\h=-\h=\frac12$ one has $m_{1,2,j}=m_{2,1,j}$. Finally, consider the coefficient $b_j$ in \pref{dfm2} and $m_{1,1,j}$, $m_{2,2,j}$ in \pref{lw2}. With the same argument used for $a_j$, an equivalent formula the last two coefficients, up to an $\h^2$ or $\bar \h^2$ prefactor, is \bal\lb{aba} &\frac{\a^2}{2}\sum_{\m=\hat e}\int\! d^2y\; \left} \def\rgt{\right[\left} \def\rgt{\right(\tilde \G_{j}^{,\m}(y)\rgt)^2 + 2\tilde \G_{j-1,0}^{,\m}(y)\tilde \G_{j}^{,\m}(y)\rgt] +O(L^{-\frac {j}{4}}) \notag\\ &=\frac{\a^2}{2}\sum_{\m=\hat e}\int\! d^2y\; \left} \def\rgt{\right[\left} \def\rgt{\right(\tilde \G_{j,0}^{,\m}(y)\rgt)^2 - \left} \def\rgt{\right(\tilde \G_{j-1,0}^{,\m}(y)\rgt)^2)\rgt] +O(L^{-\frac {j}{4}}) \eal At $\a^2=8\p$, \pref{aba} is also an equivalent formula for $b_j$. As $\tilde \G_{j-1,0}^{,\m}(y)=L \tilde \G_{j,1}^{,\m}(yL)$, the last integral in \pref{aba} becomes \bal &\frac{\a^2}{2}\sum_{\m=\hat e}\int\! d^2y\; \left} \def\rgt{\right[\left} \def\rgt{\right(\tilde \G_{j,0}^{,\m}(y)\rgt)^2 - \left} \def\rgt{\right(\tilde \G_{j,1}^{,\m}(y)\rgt)^2\rgt] \notag\\ &= \frac{\a^2}{2}\int\!\frac{d^2p}{(2\p)^2} \frac{[u( p)]^2-[u(L p)]^2}{p^2} + \a^2\int\!\frac{d^2p}{(2\p)^2} \frac{[u(p)]-[u(L p)]}{p^2} u(L^{j+1}p) \eal In the last line, the former integral can be exactly computed while the latter, using the boundedness of the derivatives of $u$ is $O(L^{-j})$. This proves \pref{scott2} and the second of \pref{scott1}. \subsection{Proof of Theorem \ref{finale}} From \pref{hofern} we have \begin{equation}} \def\ee{\end{equation}\lb{albanese} w^-_{2,a,R}(y) =\frac12 \sum_{n=0}^{R-1} Z_n^2L^{-4n} e^{\h^2\a^2\G_{R,n+1}(y|0)} e^{-\h^2\a^2\G_n(0)}\left} \def\rgt{\right(e^{\h^2\a^2 \G_n(y)}-1\rgt). \ee Use the inequality $Z_n^2 L^{-4n}\le C_\vartheta} \def\O{\Omega} \def\Y{\Upsilon L^{-\vartheta} \def\O{\Omega} \def\Y{\Upsilon n}$ for a $\vartheta} \def\O{\Omega} \def\Y{\Upsilon< \min\{4\h^2,4\bar \h^2, 1\}$, to replace the function $\G_{R,n+1}(y|0)$ with $\G_{\io,n+1}(y|0)$ up to an $O(L^{-R\vartheta} \def\O{\Omega} \def\Y{\Upsilon})$ error term. Note that each $y$ can be uniquely written as $y=L^{n_0}\t$ for $|\t|\in [1,L)$ and an integer $n_0$; then $e^{\h^2\a^2 \G_n(y)}-1=0$ every time $n\le n_0-1$ so that, in \pref{albanese}, one can actually start the sum from $n=n_0$. Accordingly, a formula for $\lim_{R\to\io}w^-_{2,a,R}(y)$ is \bal\lb{albanese1} w^-_{2,a}(y)=\frac12 \sum_{n\ge n_0} Z_n^2L^{-4n} e^{\h^2\a^2 \G_{\io,n+1}(y|0)} e^{-\h^2\a^2 \G_n(0)}\left} \def\rgt{\right(e^{\h^2\a^2 \G_n(y)}-1\rgt) . \eal Using the same argument, a formula for $w^-_{2,\bar a}(y)=\lim_{R\to\io}\bar w^-_{2,a,R}(y)$ is given by \pref{albanese1} after replacing $Z_j$ and $\h$ with $\bar Z_j$ and $-\bar \h$. Let us consider three different cases. \subsubsection{Case $0<\h<\frac12$.} It is convenient to write \pref{albanese1} as the difference of two convergent series \bal \lb{epifanio} &\frac12 \sum_{n\ge n_0} Z_n^2L^{-4n}e^{\h^2\a^2 \G_{\io,n}(y|0)}-\frac12 \sum_{n\ge n_0} Z_n^2L^{-4n}e^{-\h^2\a^2 \G_n(0)}e^{\h^2\a^2 \G_{\io,n+1}(y|0)}. \eal By replacing in the second series the factor $Z_n^2L^{-4n}e^{-\h^2\a^2 \G_n(0)}$ with the almost identical factor $Z_{n+1}^2L^{-4(n+1)}$, each term of the latter series cancels a term in the former, so that only the term for $n=n_0$ in the first series survives; besides, by definition of $n_0$ we have $\G_{\io,n_0}(y|0)= \G_{\io,0}(y|0)+ \G_{n_0-1,0}(0)$. Hence \bal\lb{zalone0} w^-_{2,a}(y)=&\frac12 Z_{n_0}^2L^{-4n_0}e^{\h^2\a^2 \G_{n_0-1,0}(0)}e^{\h^2\a^2 \G_{\io,0}(y|0)} \notag\\ &-\frac12 \sum_{n=n_0}^{\io} \left} \def\rgt{\right(Z_n^2L^{-4n}e^{-\h^2\a^2 \G_n(0)}-Z_{n+1}^2L^{-4(n+1)}\rgt)e^{\h^2\a^2 \G_{\io,n+1}(y|0)}. \eal The first term in \pref{zalone0} is the leading one. Indeed, from \pref{leibler2} we have \bal\lb{zalone} &Z^2_{j}L^{-4j} =e^{-\h^2\a^2 \G_{j-1,0}(0)}[1+|q_1| (j-1)]^{-2\h^2}e^{\tilde c_1+\tilde r_{1,j}}, \notag\\ &\bar Z^2_{j}L^{-4j} =e^{-\h^2\a^2 \G_{j-1,0}(0)}[1+|q_1| (j-1)]^{-2\h^2} r_{2,j}, \eal where $|r_{m,j}|\le C(1+|q_1|j|)^{-\frac12}$ for any $m=1,2$ and $\tilde c_1$, $\tilde c_2$ are vanishing in the limit $z\to 0$. Therefore the first term of \pref{zalone0} is \begin{equation}} \def\ee{\end{equation}\lb{zalone0C} \frac{e^{8\p \h^2 c_E}}{2|y|^{4\h^2} } (1+ |q_1|\log_L |y|)^{-2\h^2} e^{\tilde c}(1+o(1)), \ee for $o(1)$ a term bounded by $C(\log|y|)^{-\frac12}$. The second term in \pref{zalone0} is subleading by a factor $(\log_L|y|)^{-\frac12}$ at least. Indeed from \pref{6:52pm} we also find \bal &|Z^2_n L^{-4n} e^{-\h^2\a^2 \G_n(0)}-Z^2_{n+1} L^{-4(n+1)}| \notag\\ &\le Z^2_{n+1} L^{-4(n+1)}\left} \def\rgt{\right[\left} \def\rgt{\right(\frac{1+|q_1|n}{1+|q_1| (n-1)}\rgt)^{2\h^2} \left} \def\rgt{\right(\frac{1+\tilde s_{1,n}}{1+\tilde s_{1,n+1}}\rgt)-1\rgt] \notag\\ &\le C \frac{L^{-4\h^2 n_0}}{(1+|q_1| n_0)^{2\h^2+\frac12}}L^{-4\h^2(n-n_0)}. \eal Summing over $n\ge n_0$, one obtains the bound $ C|y|^{-4\h^2}(1+|q_1| n_0)^{-2\h^2-\frac12}$, which is subleading with respect to \pref{zalone0C}. Instead, for $w_{2,\bar a}(y)$ the method used above does not work; however, we can provide an upper bound that shows that $w_{2,\bar a}(y)$ is subleading with respect to $w_{2,a}(y)$: as $$ e^{\bar \h^2\a^2 \G_{\io,n+1}(y|0)} e^{-\bar \h^2\a^2 \G_n(0)}\left} \def\rgt{\right|e^{\bar \h^2\a^2 \G_n(y)}-1\rgt|\le C \bar \h^2\a^2 |\G_n(y)|, $$ $w_{2,a}(y)$ is bounded by $$ C \bar \h^2\a^2 \sum_{n\ge n_0} \bar Z^2_{n} L^{-4n}|\G_n(y)|\le C \frac{L^{-4\h^2 n_0}}{(1+|q_1| n_0)^{2\h^2+\frac12}}. $$ Next consider $w^-_{2,b,R}$ in \pref{hofern}. For any $0\le \vartheta} \def\O{\Omega} \def\Y{\Upsilon <1$ and a corresponding constant $C_\vartheta} \def\O{\Omega} \def\Y{\Upsilon\=C_\vartheta} \def\O{\Omega} \def\Y{\Upsilon(\h)$, we have the bound \bal |w^-_{2,b,R}(y)| &\le \h\bar \h\a^2 C\sum_{n=1}^{R-1} Z_n L^{-2n} \bar Z_n L^{-2n} e^{-\frac{\a^2}{2}\G_{R-1,n+1}(0)} \notag\\ &\le \h\bar \h\a^2 C_\vartheta} \def\O{\Omega} \def\Y{\Upsilon L^{-4\h^2\vartheta} \def\O{\Omega} \def\Y{\Upsilon R}. \eal Hence $\lim_{R\to\io}w^-_{2,b,R}(y)=0$. Finally, consider $w^{-}_{2,c}(y)$. From \pref{hofern} we find \bal |w^{-}_{2,c}(y)|&\le \sum_{k\ge 0} L^{-4k} e^{-L^{-k} |y|} \sum_{n\ge k} 2^{-(n-k)} \notag\\ &\times \left} \def\rgt{\right\{Z_k^2 \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ |\hat K_{2,n}^{(a,k)}\left} \def\rgt{\right(0,\z, X,0, \s, y, -\s\rgt)|\rgt] \rgt. \notag\\ &\quad+\left} \def\rgt{\right. \bar Z^2_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ |\hat K_{2,n}^{(\bar a,k)}\left} \def\rgt{\right(0,\z, X,0, \s, y, -\s\rgt)| \rgt] \rgt. \notag\\ &\hskip-2em+\left} \def\rgt{\right. Z_k\bar Z_k \frac12\sum_{\s=\pm1}\sum_{X\in \SS_n}^{X\ni 0} \mathbb{E}_j\left} \def\rgt{\right[ |\hat K_{2,n}^{(b,k)}\left} \def\rgt{\right(\s,\z, X, 0, \s, y,-\s\rgt)| \rgt]\rgt\} \notag\\ &\hskip-2em\le SA^{-\frac12}k^*_s(\sqrt A,1/2) \sum_{k\ge 0} L^{-4k} e^{-L^{-k} |y|} \sum_{n\ge k} 2^{-(n-k)} \notag\\ &\hskip-2em\times \left} \def\rgt{\right\{Z_k^2 \|K_{2,n}^{(a,k)}\|_{2,h,T_j}+ \bar Z^2_k \|K_{2,n}^{(\bar a,k)}\|_{2,h,T_j}+ Z_k \bar Z_k \|K_{2,n}^{(b,k)}\|_{2,h,T_j}\rgt\}. \eal Hence, from \pref{maldacena1}, we obtain \bal |w^{-}_{2,c}(y)|&\le \frac{|q_1|C}{|y|^{4\h^2}(1+|q_1|\log_L |y|)^{2\h^2+1}}. \eal \subsubsection{Case $\frac12<\h<1$.} The fundamental difference with the previous case is in the formula for the renormalization constants. From \pref{leibler2} we have \bal\lb{checco} &\bar Z^2_{j}L^{-4j}e^{\bar \h^2\a^2 \G_{j-1,0}(0)} =[1+|q_1| (j-1)]^{-2\bar \h^2}c(\h)^2z^2e^{\tilde c_1+\tilde r_{1,j}}, \notag\\ &Z^2_{j}L^{-4j}e^{\bar \h^2\a^2 \G_{j-1,0}(0)} =[1+|q_1| (j-1)]^{-2\bar \h^2}e^{\tilde c_2}\tilde r_{2,j}, \eal where $c(\h)$ is the positive constant in Theorem \ref{leiblerbis}. Now, proceeding with $w_{2,\bar a}(y)$ with the same method that in the previous section was used for $w_{2, a}(y)$ we obtain the formula \begin{equation}} \def\ee{\end{equation}\lb{zalone0D} z^2 c(\h)^2\frac{e^{8\p \bar \h^2 c_E}}{2|y|^{4\bar \h^2} } (1+ |q_1|\log_L |y|)^{-2\bar \h^2} e^{\tilde c}(1+o(1)), \ee for $o(1)$ a term bounded by $C(\log|y|)^{-\frac12}$. Conversely, proceeding with $w_{2,a}(y)$ with the same method that in the previous section was used for $w_{2,\bar a}(y)$ we obtain the bound $$ C \h^2\a^2 \sum_{n\ge n_0} Z^2_{n} L^{-4n}|\G_n(y)|\le C \frac{L^{-4\bar \h^2 n_0}}{(1+|q_1| n_0)^{2\bar \h^2+\frac12}}, $$ which is subleading with respect to \pref{zalone0D}. Finally, with the same arguments of the previous section, $w_{2,a}(y)=0$ and \bal |w^{-}_{2,c}(y)|&\le \frac{|q_1|C}{|y|^{4\bar \h^2}(1+|q_1|\log_L |y|)^{2\bar \h^2+1}}. \eal \subsubsection{Case $\h=\frac12$} From \pref{leibler1}, we have \bal\lb{6:52pm} &Z_{j}=\frac{Z_{j}^++Z_{j}^-}{2}=\frac12 L^{2j} e^{-\p \G_{j-1,0}(0)}(1+|q_1|(j-1))^{\frac14} e^{\tilde c+\tilde s_{j}} \eal where $\tilde c$ vanishes for $z\to0$ and $|\tilde s_{j}|\le \frac{C}{\sqrt{1+|q_1|j}}$. Hence \bal\lb{zalone2} Z^2_{n_0}L^{-4n_0}e^{\h^2\a^2 \G_{n_0-1,0}(0)} =\frac14 (1+|q_1| (n_0-1))^{\frac12}e^{2\tilde c+2\tilde s_{n_0}} \eal and the formula for $w^-_{2,a}(y)$ is \begin{equation}} \def\ee{\end{equation}\lb{zalone0D} \frac{e^{2\p c_E}}{8|y| } (1+ |q_1|\log_L |y|)^{\frac12} e^{\tilde c}(1+o(1)). \ee Now consider $w^-_{2,\bar a}(y)$. Since a formula for $\bar Z_j=(Z^+_j-Z^-_j)/2$ is again given by \pref{6:52pm}, --but for numerically different $\tilde c$ and $\tilde s_j$-- also the formula for $w^-_{2,\bar a}(y)$ is \pref{zalone0D}. Finally, consider $w^{-}_{2,c}(y)$. \bal |w^{-}_{2,c}(y)|&\le \frac{|q_1|C}{|y|(1+|q_1|\log_L |y|)^{\frac12}}. \eal This completes the proof of point 1 of Theorem \ref{finalevero}. \subsection{Proof of Theorem \ref{finale}} For the sake of brevity in this section we denote $\tilde \mathbb{E}_R$ the limiting expectation $\lim_{m\to 0}\mathbb{E}_R$. Let us consider the last term of \pref{finalcorr} \bal\lb{farmers} &e^{-\d E_R|\L|} \tilde \mathbb{E}_R\left} \def\rgt{\right[\frac{\dpr^2 K_{2,R}(\F)}{\dpr J_{x,+}\dpr J_{0,-}}\rgt]_{J=0} =e^{-\d E_R|\L|} \sum_{k=0}^R 2^{-(R-k)} e^{-L^{-k}|x|} L^{-4k} \notag\\ &\qquad\times \Bigg\{Z^2_k\tilde\mathbb{E}_R\left} \def\rgt{\right[K_{2,R}^{(a,k)}\rgt] +\bar Z^2_k\tilde\mathbb{E}_R\left} \def\rgt{\right[K_{2,R}^{(\bar a,k)}\rgt] +Z_k\bar Z_k\tilde\mathbb{E}_R\left} \def\rgt{\right[K_{2,R}^{(b,k)}\rgt]\Bigg\} \eal where we suppressed in $K_{2,R}^{(\d,k)}$ the dependence in $(\z,\L, x,+,0,-)$. Using \pref{r2},\pref{ex6.53bis} and \pref{maldacena1}, an upper bound for the absolute value of \pref{farmers} is $$ C'A^{-1}e^{C|q_R|} |q_R| \sum_{k=0}^R 2^{-(R-k)} e^{-L^{-k}|x|} \left} \def\rgt{\right(L^{-4k}Z_k^2 + L^{-4k}\bar Z_k^2\rgt) \frac{|q_k|}{|q_R|}. $$ In the limit $R\to \io$, this bound is vanishing: indeed $|q_R|\to 0$ while the sum remains bounded by the fact that $Z_k L^{-2k}, \bar Z_k L^{-2k}\le C$. Next, consider the first term in \pref{finalcorr}: expanding the product inside the square brackets, one obtains four terms. Since they can all be studied in similar way, let us consider one of them: \bal\lb{farmers1} &e^{-\d E_R|\L|}\tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} \frac{\dpr V_{1,R}(\F)}{\dpr J_{x,+}} \frac{\dpr V_{1,R}(\F)}{\dpr J_{0,-}}\rgt]_{J=0} \notag\\ &= e^{-\d E_R|\L|}L^{-4R}\tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} \left} \def\rgt{\right(Z_R^2 e^{i\h \a(\z_x-\z_0)}+\bar Z_R^2 e^{i\bar \h \a(\z_x-\z_0)}\rgt)\rgt] \notag\\ &+ e^{-\d E_R|\L|}L^{-4R}\tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} Z_R \bar Z_R \left} \def\rgt{\right(e^{i \a (\h\z_x-\bar\h\z_0)}+e^{i \a (\bar\h\z_x-\h\z_0)}\rgt)\rgt] \eal It is easy to see that, for a $C\=C(\a)$ and for $z$ smaller than a $z(L,\a)$, $$ \|V_{0,R}(\z,\L)\|_{h,T_R(\L)}\le C |q_R|\left} \def\rgt{\right(1+\max_{n=1,2}\|\nabla_R^n \z\|^2_{L^\io(\L)}\rgt) \le C |q_R|+\frac12\ln G^{\rm str}(\z,\L), $$ $$ \|W_{0,R}(\z,\L)\|_{h,T_R(\L)}\le C |q_R|^2\left} \def\rgt{\right(1+\max_{n=1,2}\|\nabla_R^n \z\|^2_{L^\io(\L)}\rgt) \le C |q_R|^2+\frac12\ln G^{\rm str}(\z,\L), $$ $$ \|e^{i \a_1\z_x}\|_{h,T_R(\L)}\le e^{h|\a|}; $$ therefore, for any $\a_1, \a_2\in \mathbb{R}$, \bal &\left} \def\rgt{\right|e^{V_{0,R}(\z) + W_{0,R}(\z)} e^{i (\a_1\z_x+\a_2\z_0)}\rgt| \notag\\ &\le\|e^{V_{0,R}(\z) + W_{0,R}(\z)}\|_{h,T_R(\L)} \;\|e^{i \a_1\z_x}\|_{h,T_R(\L)}\; \|e^{i \a_2\z_0}\|_{h,T_R(\L)} \notag\\ &\le e^{h|\a_1|+h|\a_2|}e^{2C |q_R|}G^{\rm str}(\z,\L). \eal In conclusion, the absolute value of \pref{farmers1} can be bounded by $$ C(\a) e^{C|q_R|}\left} \def\rgt{\right(L^{-4R}Z_R^2 + L^{-4R}\bar Z_R^2\rgt). $$ In the limit $R\to \io$ this bound is vanishing since $|q_R|, L^{-2R} Z_R,L^{-2R} \bar Z_R \to 0$. The remaining term of \pref{farmers1} is the one that gives the right hand side of \pref{canaletto}. To prove this fact, we need to study the difference \bal\lb{am} &e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} \frac{\dpr^2 W_{2,R}(\F)}{\dpr J_{x,+}\dpr J_{0,-}}\rgt]_{J=0} -2\left} \def\rgt{\right[w_{2,a,R}^-(x)+ \bar w_{2,a,R}^-(x) +w_{2,c,R}^-(x)\rgt] \notag\\ &=2w_{2,a,R}^-(x) \left} \def\rgt{\right\{e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} e^{i\h \a(\z_x-\z_0)}\rgt]-1\rgt\} \notag\\ &+2\bar w_{2,a,R}^-(x) \left} \def\rgt{\right\{e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} e^{i\bar \h \a(\z_x-\z_0)}\rgt]-1\rgt\} \notag\\ &+2\bar w_{2,b,R}^-(x) e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} e^{i\a(\h \z_x-\bar \h\z_0)}\rgt] \notag\\ &+2\bar w_{2,b,R}^-(x) e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)} e^{i\a(\bar \h \z_x-\h\z_0)}\rgt]. \notag\\ &+2 w_{2,c,R}^-(x) \left} \def\rgt{\right\{e^{-\d E_R|\L|} \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{V_{0,R}(\z) + W_{0,R}(\z)}\rgt]-1\rgt\}. \eal Observe that, by \pref{p4a} and \pref{p4b} $$ \lim_{R\to \io}\tilde\mathbb{E}_R\left} \def\rgt{\right[ e^{i\h \a(\z_x-\z_0)}\rgt]=1, \qquad \tilde\mathbb{E}_R\left} \def\rgt{\right[e^{i\a(\h \z_x-\bar \h\z_0)}\rgt]=0. $$ From them it is easy to show that \pref{am} is vanishing in the limit $R\to \io$.
\section{Introduction} The ``strange metals'', as realised in high $T_c$ superconductors and related systems, are among the great mysteries of modern physics. \cite{Zaanen:2010yk} It is believed that these are strongly interacting ``quantum soups'' controlled by an emergent temporal scale invariance of their underlying quantum dynamics. \cite{PhysRevLett.63.1996} The general principles governing such strongly interacting quantum critical states could be captured by the holographic dualities discovered in string theory. \cite{Maldacena:1997re} However, an explanation of the iconic linear resistivity \cite{Zaanen:2010yk,PhysRevLett.63.1996} observed experimentally remains elusive. Holographic duality asserts that some strongly interacting quantum field theories are mathematically equivalent to certain classical theories of gravity in higher-dimensional curved spacetimes. \cite{Maldacena:1997re} This is a very useful theoretical tool as it allows us to easily calculate, in a well-controlled way, properties of certain strongly interacting, quantum critical states of matter by analysing their classical gravitational duals. Analyses of this kind have found examples of systems, controlled by quantum critical IR fixed points, with linear resistivities,\cite{Faulkner:2010zz,Donos:2012ra,Hoyos:2013eza} amongst other interesting phenomena. While interesting, many of these results appear to be highly dependent upon the microscopic details of the IR fixed point being considered and, a priori, it is not clear whether these details are generic and could be realised in real electron systems, or whether they are artefacts of the kinds of highly symmetric quantum field theories which are dual to classical theories of gravity. In this paper, we address these issues by describing a simple universal mechanism by which strongly interacting, quantum critical states of matter can acquire a linear resistivity, and propose that this could be at work in the strange metallic phase of the cuprates. This mechanism is at work in certain strongly interacting field theories which have holographic duals (specifically, those controlled by a locally quantum critical IR fixed point). However, it is based on general physical principles which do not rely on the existence of a dual, classical gravitational description of the state. Specifically, a theory whose IR physics is well-described by the laws of hydrodynamics with a ``minimal'' viscosity will have, when weakly coupled to random disorder, a viscous contribution to its resistivity which is proportional to its thermodynamic entropy. A state with a Sommerfeld heat capacity (entropy $S\sim T$), such as the cuprate strange metals, will therefore have a resistivity linear in temperature $T$. The physical principles which govern the transport of charge in highly collective, charged quantum critical states of matter are rather different from those in an ordinary Fermi liquid composed of long-lived electronic quasiparticles (see references \onlinecite{Hartnoll:2007ih,Hartnoll:2008hs,Hartnoll:2012rj,Anantua:2012nj,Mahajan:2013cja,Hartnoll:2014gba} for some recent studies of systems of this type). In particular, the absence of long-lived quasiparticles means that the natural quantities to consider are the collective properties of the system: its total electrical current and momentum. In a system of this type, one intuitively expects that the intrinsic decay rate of the electrical current will generically be large. However, in the absence of a background lattice or impurities, the DC resistivity will vanish. This is because the electrical current carries momentum, which is conserved due to the translational invariance of the system, and thus it cannot decay. If we couple such a system weakly to impurities or a background lattice, such that the intrinsic interaction rate of the electronic fluid is much greater than the rate at which it interacts with the impurities or lattice, the momentum will dissipate slowly and will control the decay of electrical current at late times. The DC resistivity is therefore controlled by the decay rate of the total momentum of the state, in contrast to normal metals in which the decay rate of quasiparticle excitations plays a similar role. We show that, if the collective state obeys the laws of hydrodynamics, the momentum dissipation rate in the presence of impurities is determined by the transport coefficients of the liquid, including its viscosity, which depend upon the microscopic details of the system. It is natural to expect the viscosity of a fluid arising from a strongly interacting quantum critical system to be small because, in a kinetic theory of quasiparticles, the viscosity is proportional to the mean free time between quasiparticle collisions. This can be made much more precise: the ``almost perfect'' fluids formed by strongly interacting, quantum critical systems are expected to have a viscosity of the order of the entropy density of the state, in units of $\hbar/k_B$.\cite{Sachdev,Kovtun:2004de,Iqbal:2008by} This is a universal feature of systems with holographic duals, and has been measured experimentally in strongly interacting, collective systems without such duals.\cite{Schafer:2009dj,Heinz:2013th} In this scenario, there is a universal viscous contribution to the momentum dissipation rate, and hence to the DC resistivity, which is proportional to the entropy of the state. This hydrodynamic description of a metallic state is far removed from the standard Fermi liquid theory of metals. The quasiparticles of a Fermi liquid interact weakly and thus the state has a very high viscosity.\cite{PinesNozieres} This means it cannot reach the local thermal equilibrium required for hydrodynamics to be valid before interactions with the lattice cause the state to lose momentum. We discuss in detail some experimentally observable consequences of this hydrodynamic mechanism and discuss its potential applicability to real systems, in particular the strange metallic phase of the cuprates. For the reader uncomfortable with our assertion that a strongly interacting quantum critical system could behave like a hydrodynamic liquid with a minimal viscosity, we provide a proof of principle by discussing a specific classical gravitational theory, which arises as a low energy limit of string theory, that is dual to a field theory that exhibits this mechanism. This theory is locally quantum critical in the IR, i.e. it has a dynamical critical exponent $z\rightarrow\infty$. Analogous holographic theories controlled by a finite $z$ fixed point do not exhibit this mechanism, for reasons we will discuss.\cite{Lucas:2014zea} We also discuss how one can include the gross effects of the weak interaction of this holographic strange metal state with disorder by the inclusion of a graviton mass.\cite{Vegh:2013sk} The remainder of the paper is structured as follows. In section \ref{sec:hydrocalculation}, we briefly summarise the memory matrix approach to transport in systems in which momentum is the only long-lived quantity and use this to calculate the momentum relaxation time, and hence the resistivity, in a hydrodynamic state weakly coupled to random disorder. We show that this has a viscous contribution which, in a strongly interacting quantum critical state, will be proportional to the entropy density of the state. The presence of this viscous contribution follows from a simple physical argument which can be understood without knowledge of the memory matrix formalism. In section \ref{sec:holographysection}, we then describe a holographic system in which, at sufficiently long distances, hydrodynamics is a good enough description that this mechanism is at work, as well as describing how one can directly include the main effects of quenched disorder in this setup. Finally, in section \ref{sec:discussionsection} we close by summarising our results and explaining their experimental implications, the potential applicability to real systems such as the strange metal phase of the cuprates, as well as some interesting directions for further work. \section{Resistivity of a hydrodynamic fluid} \label{sec:hydrocalculation} As outlined in the introduction, we are interested in highly collective, charged states in which the intrinsic relaxation timescale of the electrical current is small. Despite these short relaxation timescales, the DC electrical resistivity in a translationally invariant theory of this kind will vanish. This is because the current carries a non-zero momentum which cannot decay, as it is a conserved Noether charge of the system. If we allow the momentum to decay slowly, in comparison to the intrinsic decay rate of the electrical current, by breaking the translational invariance of the system -- for example via weak interactions with random impurities or a lattice -- the decay rate of the electrical current at late times, and hence the DC resistivity of the state $\rho_{DC}$, will be proportional to the rate at which momentum dissipates $\Gamma_k$. In the following, we will express this dissipation rate in terms of the characteristic timescale $\tau_k$ over which the momentum decays: $\Gamma\equiv\tau_k^{-1}$. Furthermore, assuming that the interactions which dissipate momentum are weak (such that momentum lives for a long time), then at leading order in a perturbative expansion, $\tau_k^{-1}$ is determined by properties of the translationally invariant state. In a hydrodynamic state in which translational invariance is weakly broken, $\tau_k^{-1}$ depends on the viscosity $\eta$ of the state such that the DC resistivity will have a viscous contribution $\rho_{DC}(T)\sim\eta(T)$. Before we show this explicitly using the memory matrix formalism, we outline a simple physical argument which illustrates why this is the case. In a (relativistic) hydrodynamic liquid, momentum diffuses with a characteristic diffusion constant $D=\eta/\left(\epsilon+P\right)$, where $\epsilon$ and $P$ are the energy density and pressure of the state. In the non-relativistic limit, $\epsilon+P=m_en_e$ where $m_e$ and $n_e$ are the electron mass and number density respectively. \cite{Forster} If translational invariance is broken over a characteristic length scale $l$, then the characteristic timescale $\tau_k$ over which this diffusing momentum relaxes will be (see also references \onlinecite{2005JPhy4.131..255S,PhysRevLett.106.256804}) \begin{equation} \frac{1}{\tau_k} = \frac{D}{l^2} = \frac{\eta}{\left(\epsilon+P\right)l^2}. \label{monrelaxtime} \end{equation} This mechanism produces a DC resistivity $\rho_{DC}=\left(\epsilon+P\right)/\sigma^2\tau_k$, where $\sigma$ is the charge density of the state, or in the non-relativistic case $\rho_{DC}=1/\omega_p^2\tau_{k}$, where $\omega_p$ is the plasma frequency of the electrons. This reasoning is valid for any hydrodynamic liquid, but we now specialise to the case of strongly interacting quantum critical liquids. These have the special property that there is a simple relationship between the shear viscosity and the entropy density $\eta/s=A\hbar/k_B$, where $A$ is a number of order one \cite{Sachdev} that in holographic theories is generically equal to $1/4\pi$. \cite{Kovtun:2004de,Iqbal:2008by} This minimal viscosity ensures that relaxation to local thermal equilibrium takes the minimal amount of time, which may be significantly shorter than the timescale for momentum loss due to Umklapp scattering or interactions with impurities. This is drastically different from an ordinary metal, in which fast Umklapp scattering of quasiparticles prevents the formation of a hydrodynamic state. \cite{PhysRevB.7.2317} Combining these simple results, we arrive at a stunning conclusion: \begin{equation} \rho_{DC} (T) = \frac{A\hbar}{\omega_p^2m_el^2} \frac{S_e(T)}{k_B}. \label{rhovsent} \end{equation} Having assumed only that the hydrodynamic limit is approximately valid and that the viscosity is minimal, we have found that the DC resistivity $\rho_{DC}$ due to this mechanism is proportional to the entropy per electron $S_e$. \subsection{Memory matrix derivation} We will now use a more rigorous memory matrix calculation to confirm our intuition that this effect should be present. Recent overviews of this formalism can be found in references \onlinecite{Hartnoll:2007ih,Hartnoll:2012rj,Mahajan:2013cja}. If the momentum of a system is its only long-lived quantity, and there is overlap between the electrical current and momentum operators in that theory, the DC electrical resistivity is related to the momentum relaxation time $\tau_k$ via\cite{Hartnoll:2012rj} \begin{equation} \rho_{DC}=\frac{\chi_{\vec{P}\vec{P}}}{\chi_{\vec{J}\vec{P}}^2}\;\tau_k^{-1} \end{equation} where $\chi_{\mathcal{O}_A\mathcal{O}_B}$ are the static susceptibilities of the state. The momentum relaxation time $\tau_k$ is determined by the microscopic processes which cause momentum to dissipate. Suppose we perturb a translationally invariant metal by the introduction of a spatially random potential $V\left(\vec{x}\right)$ for an operator $\mathcal{O}$ in the IR Hamiltonian (for definiteness, we have chosen here a (2+1)-dimensional metallic state) \begin{equation} \delta H=\int d^2\vec{x}\;V\left(\vec{x}\right)\mathcal{O}\left(t,\vec{x}\right) \end{equation} where, on statistically averaging over the random potential, \begin{equation} \left<V\left(\vec{x}\right)\right>=0, \;\;\;\;\;\;\; \left<V\left(\vec{x}\right)V\left(\vec{y}\right)\right>=\bar{V}^2\delta^{(2)}\left(\vec{x}-\vec{y}\right). \end{equation} Such sources would naturally arise in the IR theory if there are random impurities present. Provided that the interaction with disorder is weak -- automatically true if the operator $\mathcal{O}$ is irrelevant in the IR -- this will cause momentum to dissipate slowly (so that it is still a long-lived quantity) and we can compute this dissipation rate perturbatively. At leading order, it is determined by the spectral weight of $\mathcal{O}$ at low energies \textit{in the translationally invariant state}\cite{Hartnoll:2007ih,Hartnoll:2008hs} \begin{equation} \label{eq:generalmemorymatrixresultfortau} \tau_k^{-1}=\frac{\bar{V}^2}{2\chi_{\vec{P}\vec{P}}}\int \frac{d^2k}{\left(2\pi\right)^2}\;k^2 \lim_{\omega\rightarrow0}\frac{\text{Im}G_{\mathcal{O}\mathcal{O}}\left(\omega,k\right)}{\omega}\Biggr|_{\bar{V}\rightarrow0} \end{equation} where $G_{\mathcal{O}_A\mathcal{O}_B}$ denotes the retarded Greens function of two operators. This result has a simple physical interpretation: if translational invariance is weakly broken at a characteristic momentum scale $k$ by a source for $\mathcal{O}$, then at leading order in a perturbative expansion the rate at which momentum dissipates due to this should depend on the number of low energy excitations of the translationally invariant state with momentum $k$ that overlap with the operator $\mathcal{O}$, which is what the spectral weight tells us. The result above depends on the integral of the spectral weight over $k$, as the spatially random potential breaks translational invariance at all length scales. It is expected that multiple operators may acquire a spatially dependent source in the low energy theory describing a real metal, in the presence of impurities and a lattice. In this situation, the argument above applies provided that all such operators couple weakly, and the most relevant of these will then provide the leading contribution to $\rho_{DC}$. In this scenario, the microscopic quantities controlling the DC resistivity are the low energy spectral weights of the relevant operators in a given theory. Suppose that the low energy physics of the translationally invariant state is well-described by the laws of hydrodynamics. The important quantities in a (relativistic) hydrodynamic theory are the conserved energy-momentum tensor $T^{\mu\nu}$ and charge current $J^\mu$ of the theory. The form of the retarded Greens functions of these operators are specified by the laws of hydrodynamics (see, for example, reference \onlinecite{Kovtun:2012rj}), and are written in terms of the transport coefficients of the theory, such as the viscosity $\eta$, whose values are dependent upon the microscopic physics underlying the effective hydrodynamic description. The presence of random impurities will give rise to a random source of energy density $T^{tt}$ in the low energy theory. If the coupling to random impurities is weak such that momentum dissipates slowly, this will produce a DC resistivity \begin{equation} \label{eq:neutralmemorymatrixrho} \begin{aligned} \rho_{DC}=\frac{\chi_{\vec{P}\vec{P}}}{\chi_{\vec{J}\vec{P}}^2}\;\tau_k^{-1}&=\frac{\bar{V}^2_{T^{tt}}}{2\chi_{\vec{J}\vec{P}}^2}\int \frac{d^2k}{\left(2\pi\right)^2}k^2\lim_{\omega\rightarrow0}\frac{\text{Im}G_{T^{tt}T^{tt}}\left(\omega,k\right)}{\omega}\Biggr|_{\bar{V}_{T^{tt}}\rightarrow0}\sim\frac{\bar{V}^2_{T^{tt}}}{\sigma^2}\int dk k\left(\eta k^2+\ldots\right), \end{aligned} \end{equation} where we have used the known spectral weight of $T^{tt}$ in a relativistic hydrodynamic theory with a conformally invariant vacuum\cite{Kovtun:2012rj} i.e. the stress-energy tensor is traceless (the pressure is related to the energy density by $\epsilon=2P$, and the bulk viscosity vanishes). This case is relevant for the holographic theories we will discuss shortly. In the final step of equation (\ref{eq:neutralmemorymatrixrho}) we have neglected order 1 numerical prefactors. The leading term in the small $k$ expansion on the right hand side is precisely the viscous contribution identified in equation (\ref{monrelaxtime}), but now derived in a rigorous manner. For a hydrodynamic theory with a minimal viscosity, this produces a DC resistivity $\rho_{DC}\left(T\right)\sim s\left(T\right)$. We are assuming that hydrodynamics is a good effective description down to a temperature-independent microscopic length scale, which acts as a UV cutoff on the momentum integrals, resulting in a controlled hydrodynamic momentum expansion. If the random impurities lead to a significant source of charge density $J^t$ in the low energy theory, there will be additional contributions to the DC resistivity, taking the form \begin{equation} \label{eq:chargedmemorymatrixrho} \rho_{DC}\sim\frac{\bar{V}^2_{J^{t}}}{\sigma^2}\int dkk\Bigl(\frac{1}{\sigma_Q}\left[2\frac{\sigma^2}{\epsilon+P}-\left(\frac{d\sigma}{d\mu}\right)_T\right]^2+k^2\frac{\sigma^2}{\left(\epsilon+P\right)^2}\eta+\ldots\Bigr), \end{equation} where we have again used the spectral weight of $J^t$ in a relativistic hydrodynamic theory with a conformally invariant vacuum.\cite{Kovtun:2012rj} There is also a viscous contribution in this case, in addition to a contribution inversely proportional to the ``universal conductivity'' transport coefficient $\sigma_Q$. Despite its name, the temperature dependence of $\sigma_Q$ is not universal. This illustrates an important point: there are various contributions to the DC resistivity of a hydrodynamic liquid, which will depend upon the microscopic details of the specific theory under consideration. We have highlighted the role of the viscous term because there are good theoretical and experimental reasons to expect it to make a universal contribution $\rho_{DC}\left(T\right)\sim\eta\left(T\right)\sim S\left(T\right)$, independent of such details, in strongly interacting, quantum critical systems. Additional contributions to equations (\ref{eq:neutralmemorymatrixrho}) and (\ref{eq:chargedmemorymatrixrho}) will also arise if one relaxes the requirement of a conformally invariant vacuum, includes higher order terms in the constitutive relations, or includes non-analytic effects.{\cite{Balasubramanian:2013yqa,Kovtun:2011np} Furthermore, if there are inhomogeneities sourced by a more relevant operator in the hydrodynamic theory than the energy density, their effects will be more important than those discussed here. These should be evaluated on a case-by-case basis. We have shown rigorously, using the memory matrix formalism, that there should be a viscous contribution to the DC resistivity $\rho_{DC}\left(T\right)\sim\eta(T)$ when a hydrodynamic state is weakly coupled to disorder. Results of this nature have been derived previously by other methods.\cite{2005JPhy4.131..255S,PhysRevLett.106.256804} We note in particular reference \onlinecite{PhysRevLett.106.256804}, in which an expression similar to equation (\ref{eq:chargedmemorymatrixrho}) was found, but with the leading term inversely proportional to the thermal conductivity. This is consistent, as the thermal conductivity is proportional to the universal conductivity $\sigma_Q$ in the type of theories we have considered here.\cite{Herzog:2009xv} In the cuprates, it is realistic that this term is small (in comparison to the contribution from neutral disorder in equation (\ref{eq:neutralmemorymatrixrho})) due to the chemistry of these systems.\cite{PhysRevLett.106.256804,RevModPhys.81.45} \section{Holographic realisation of this mechanism} \label{sec:holographysection} We have shown that if the low energy physics of a system is well-described by hydrodynamics with a minimal viscosity $\eta\sim s$, there will be a viscous contribution to the DC resistivity such that $\rho_{DC}\sim\eta\sim s$, when this theory is weakly coupled to disorder. To motivate the applicability of these results to strongly correlated electron systems and, in particular, the strange metal phase of the cuprates, we will now describe a well-controlled example of a strongly interacting, quantum critical state of matter with these features. We give this example as a proof of principle, although we do not expect that its specific properties -- in particular, that it can be described by a dual, classical theory of gravity -- are necessary for the existence of the effective hydrodynamic description which is required. For the reader unfamiliar with using holographic techniques to study strongly interacting states, the references \onlinecite{McGreevy:2009xe,Hartnoll:2009sz,Herzog:2009xv,Hartnoll:2011fn} provide informative introductions to this topic. \subsection{Translationally invariant theory} As indicated above, the example we give of a strongly interacting, quantum critical state with these properties is equivalent, by a holographic duality, to a classical theory of gravity in a curved, higher dimensional spacetime. This (2+1)-dimensional `strange metal' is a finite density state which exhibits local quantum criticality at low energies. Local quantum criticality is an emergent quantum temporal scale invariance but with short-ranged spatial correlations, and is closely related to the marginal Fermi liquid phenomenology proposed to explain experimental observations in the cuprates.\cite{PhysRevLett.63.1996} States with this property are common in holography, and they typically exhibit superconducting instabilities,\cite{Gubser:2008px,Hartnoll:2008vx} non-Fermi-liquid behaviour of fermionic two-point functions,\cite{Liu:2009dm,Cubrovic:2009ye,Faulkner:2009wj,Gubser:2009qt} and interesting power laws in the mid-IR optical conductivity in the presence of a lattice.\cite{Horowitz:2012ky,Horowitz:2012gs,Ling:2013nxa} This state is dual to a black brane solution of a (3+1)-dimensional Einstein-Maxwell-Dilaton theory of classical gravity that was first studied in this context by Gubser and Rocha,\cite{Gubser:2009qt} and which can be uplifted to a solution of 11-dimensional supergravity \cite{Cvetic:1999xp} \begin{equation} S_{EMD}=\frac{1}{2\kappa_4^2}\int d^4x\sqrt{-g}\left[\mathcal{R}-\frac{1}{4}e^\phi F_{\mu\nu}F^{\mu\nu}-\frac{3}{2}\partial_\mu\phi\partial^\mu\phi+\frac{6}{L^2}\cosh\phi\right]. \end{equation} The charged black brane solution of this theory is \begin{equation} \label{eq:masslessholographicsolution} \begin{aligned} ds^2&=\frac{r^2g(r)}{L^2}\left(-h(r)dt^2+dx^2+dy^2\right)+\frac{L^2}{r^2g(r)h(r)}dr^2, \;\;\;\;\; h(r)=1-\frac{\left(Q+r_0\right)^3}{\left(Q+r\right)^3},\\ A_t(r)&=\sqrt{\frac{3Q\left(Q+r_0\right)}{L^2}}\left(1-\frac{Q+r_0}{Q+r}\right), \;\;\;\;\;\phi(r)=\frac{1}{3}\log(g(r)),\;\;\;\;\;g(r)=\left(1+\frac{Q}{r}\right)^\frac{3}{2}. \end{aligned} \end{equation} As usual, the ``holographic'' co-ordinate $r$ labels the additional dimension in which the gravitational theory is defined compared to the field theory, and $r_0$ denotes the location of the black brane horizon. Note that in addition to the minimal ingredients required for a holographic dual of a quantum field theory at non-zero density (the metric $g_{\mu\nu}$ and a U(1) gauge field $A_{\mu}$), we include a scalar field $\phi$ (the dilaton) which means that the field theory contains a neutral scalar operator with non-trivial dynamics. Although the gravitational theory is simpler in the absence of the dilaton, the price of this simplicity is that it is dual to a field theory state which is unrealistic: it has a finite zero temperature entropy. The presence of the dilaton changes the ground state, which now has vanishing zero temperature entropy. At finite temperature, it has an entropy density which is linear in temperature $s\sim T$ at low $T$. We note here that there are many other locally quantum critical, holographic examples of this mechanism which produces $\rho_{DC}\sim s$ -- we have chosen to describe a case in which $s\sim T$ as this is the case relevant to the experimental system we are addressing: the strange metal phase of the cuprates. The low energy, local quantum criticality of this state is encoded in the near-horizon geometry of the dual black brane solution at zero temperature.\cite{Iqbal:2011in} This geometry is conformal to AdS$_2\times\mathbb{R}^2$ and therefore transforms covariantly (the line element transforms as $ds^2\rightarrow\lambda^{-1}ds^2$) under a scaling symmetry which acts on the temporal co-ordinate $t$, but not on the spatial co-ordinates $(x,y)$, of the strongly interacting state \begin{equation} \label{eq:scalingsymmetries} t\rightarrow\lambda t,\;\;\;\;\;\;x\rightarrow x,\;\;\;\;\;\;y\rightarrow y,\;\;\;\;\;\;r\rightarrow \lambda^{-2}r. \end{equation} This state violates hyperscaling and in the standard classification in terms of a dynamical critical exponent $z$ and hyperscaling violation exponent $\theta$, this state has $z\rightarrow\infty$ with the ratio $-\theta/z=1$ fixed.\cite{Goldstein:2009cv,Charmousis:2010zz,Hartnoll:2012wm} \subsection{Weak interactions with random disorder} The low energy spectral functions of operators in the thermal state of this theory are of the form\cite{Hartnoll:2012rj,Anantua:2012nj} \begin{equation} \label{eq:holographicgreensfunctionmatching} \lim_{\omega\rightarrow0}\text{Im}G_{\mathcal{O}\mathcal{O}}\left(\omega,k,T\right)\sim H(k)\lim_{\omega\rightarrow0}\text{Im}\mathcal{G}^{IR}_{\mathcal{O}\mathcal{O}}\left(\omega,k,T\right)\sim H(k)\;\omega\;T^{2\nu_\mathcal{O}\left(k\right)-1}. \end{equation} At low $\omega$ and $T$, the $\omega$ and $T$ dependence of the spectral function is contained in the ``IR spectral function'' $\mathcal{G}^{IR}\left(\omega,k,T\right)$, whose form is determined by the near-horizon geometry in the dual gravitational theory. The IR local quantum criticality of this state means that this IR spectral function is a power law in $\omega$ and $T$, and is approximately momentum-independent. The only momentum dependence is in the exponent $\nu_\mathcal{O}\left(k\right)$, which is determined by the mass, in the near-horizon geometry, of the excitation of the field dual to the operator $\mathcal{O}$. To determine the low energy spectral function of an operator in the field theory, one must additionally determine the momentum-dependent ``matching function'' $H(k)$ by matching the solutions of the field equations for excitations in the near-horizon geometry to the asymptotic region of the geometry. The advantage of the decomposition of the Greens function given in equation (\ref{eq:holographicgreensfunctionmatching}) is that, to determine the leading $T$-dependence of the spectral function of $\mathcal{O}$, the matching to the asymptotic region does not need to be explicitly performed. One must only determine the exponent $\nu_\mathcal{O}\left(k\right)$, which is related to the scaling dimension of $\mathcal{O}$ at the locally critical IR fixed point and is fixed by the near-horizon properties of the dual gravitational theory. This property can be used to determine the DC resistivity of these states,\cite{Hartnoll:2012rj,Anantua:2012nj} as we will now review. The translationally invariant state dual to the solution in equation (\ref{eq:masslessholographicsolution}) has a vanishing DC resistivity due to the conservation of momentum. If spatially random sources are added for $\mathcal{O}$, they cause momentum to dissipate at a rate \begin{equation} \Gamma_k\equiv\tau_k^{-1}\sim\frac{\bar{V}_\mathcal{O}^2}{\chi_{\vec{P}\vec{P}}}\int dk\;k^3\; H(k)\; T^{2\nu_\mathcal{O}\left(k\right)-1}, \end{equation} provided that $\mathcal{O}$ is weakly coupled at the low energies and temperatures of interest and we can therefore use equation (\ref{eq:generalmemorymatrixresultfortau}). The introduction of random impurities in a sample will induce a spatially random source of energy density $T^{tt}$, which therefore produces a DC resistivity \begin{equation} \rho_{DC}=\frac{\chi_{\vec{P}\vec{P}}}{\chi_{\vec{J}\vec{P}}^2}\tau_k^{-1}\sim\frac{\bar{V}_{T^{tt}}^2}{\sigma^2}\int dk\;k^3\;H(k)\;T^{2\nu_{T^{tt}}\left(k\right)-1}, \end{equation} where the appropriate exponent for this operator is \begin{equation} \nu_{T^{tt}}\left(k\right)\equiv\nu\left(k\right)=\sqrt{\frac{11}{3}+4\frac{k^2}{\mu^2}-\frac{8}{3}\sqrt{1+3\frac{k^2}{\mu^2}}}, \end{equation} where $\mu=\sqrt{3}Q/L^2$ is the chemical potential of the state. At leading order in $T$, the dominant contribution to this integral comes from the homogeneous $k=0$ mode of the disorder such that \begin{equation} \label{eq:linearresistivityoperatordimensionresult} \rho_{DC}\left(T\right)\sim T^{2\nu\left(0\right)-1}\sim T. \end{equation} Including the weak momentum dependence of $\nu\left(k\right)$ in the integral produces small logarithmic corrections to this leading order result such that $\rho_{DC}\sim T\log\left(\mu/T\right)^{-1}$. For more general locally critical holographic states, it is implicitly contained in the results of reference \onlinecite{Anantua:2012nj} that this calculation yields $\rho_{DC}\left(T\right)\sim s\left(T\right)$ at leading order. The operators $J^t$ and $T^{tt}$ have the same exponent $\nu\left(k\right)$ as their dual fields are coupled in the IR geometry, and thus a spatially random source for charge density will also produce a contribution $\rho_{DC}\left(T\right)\sim s\left(T\right)$. Finally, we note that both of these operators are (marginally) irrelevant at low energies and thus this approach is consistent. When analysed in this way, it appears that the linear resistivity of this state is highly dependent upon the microscopic details of the theory, in particular the dimensions of operators in the low energy theory, which are calculated from masses of field excitations in a higher-dimensional gravitational theory. It is not obvious whether these features could realistically be expected in real electron systems, such as in the strange metal phase of the cuprates, or whether they arise only in the special class of strongly interacting field theories which have dual gravitational descriptions. We can get another perspective on this result by revisiting the spectral functions of these holographic locally critical quantum field theories. By explicitly calculating the ``matching functions'' $H(k)$ numerically, one finds that the Greens functions of these theories are consistent with those of hydrodynamics, with a minimal viscosity, down to length scales $k\lesssim\mu$.\cite{Davison:2013bxa,Davison:2013uha,Tarrio:2013tta} The most explicit verification of this is in the simplest example of the Reissner-Nordstrom-AdS$_4$ solution of Einstein-Maxwell theory where one can analytically compute the matching functions $H(k)$ and determine that hydrodynamics is a good description of this theory, in the limit where the finite $k$ corrections to $\nu\left(k\right)$ can be neglected.\cite{Davison:2013bxa} We expect this to be true more generally in holographic locally critical states, and it would be interesting to explicitly verify this. There is therefore a contribution to the DC resistivity $\rho_{DC}\left(T\right)\sim s\left(T\right)$ via the viscous mechanism described in section \ref{sec:hydrocalculation}. This gives a more physical understanding of the result in equation (\ref{eq:linearresistivityoperatordimensionresult}). Fundamentally, if the state is described by hydrodynamics then the scaling dimensions of $T^{tt}$, $J^t$ etc., which were used to compute $\rho_{DC}$ in reference \onlinecite{Anantua:2012nj}, are not arbitrary numbers but are constrained by the transport coefficients (such as the viscosity) of the hydrodynamic description. The finite $k$ corrections to $\nu\left(k\right)$ in these theories, which are not captured by hydrodynamics, produce small, logarithmic corrections to this result as previously explained. These corrections reflect the fact that at low energies and finite momenta, this holographic state has slightly \textit{less} spectral weight than relativistic hydrodynamics (since $\nu_k\ge\nu_0$) and it would be interesting to search for such corrections experimentally via precision measurements of $\rho_{DC}$ over a large range of $T$. We note that coupling a hydrodynamic liquid to a periodic potential will also give a viscous contribution to the DC resistivity $\rho_{DC}\sim s$ with $l$ in equation (\ref{rhovsent}) given by the lattice spacing. In the state dual to the gravitational solution (\ref{eq:masslessholographicsolution}), one instead finds that the DC resistivity $\rho_{DC}\sim T^{2\nu\left(k_L\right)-1}$ obeys a power law where the power depends upon $k_L$.\cite{Anantua:2012nj} In these circumstances, the non-hydrodynamic finite-$k$ corrections to $\nu\left(k\right)$ are important and thus an effective hydrodynamic description is not accurate enough. An exception to this is when $k_L\ll\mu$, in which case $\nu\left(k_L\right)\sim\nu\left(0\right)$, in which case $\rho_{DC}\sim T$ as was noted in a footnote in reference \onlinecite{Anantua:2012nj}. Finally, we note that although the hydrodynamic underpinning should ensure the universality of our mechanism, its subsequent expression for the resistivity does rely on the additional assumption of local quantum criticality. For critical theories with a finite dynamical critical exponent $z$ -- where locally quantum critical means $z=\infty$ -- the UV-cut-off in the momentum integral in the memory matrix expressions (\ref{eq:neutralmemorymatrixrho}) and (\ref{eq:chargedmemorymatrixrho}) may become temperature dependent, see e.g. reference \onlinecite{Lucas:2014zea} for an example of this in a different set-up. \subsection{Explicit inclusion of momentum dissipation} As we have just reviewed, although random disorder breaks translational invariance on all length scales, the leading contribution to the momentum dissipation rate $\Gamma_k$ and hence to $\rho_{DC}$ in these holographic locally quantum critical states is from the homogeneous $k=0$ mode of this disorder. We can explicitly incorporate its effects on $\rho_{DC}$ in the holographic theory via the inclusion of a graviton mass term in the action\cite{Vegh:2013sk,Davison:2013jba,Blake:2013bqa} \begin{equation} \begin{aligned} \label{eq:specificmassivegravityaction} S=\frac{1}{2\kappa_4^2}\int d^4x\sqrt{-g}\Bigl[\mathcal{R}-\frac{1}{4}e^\phi F_{\mu\nu}F^{\mu\nu}-\frac{3}{2}\partial_\mu\phi\partial^\mu\phi+\frac{6}{L^2}\cosh\phi-\frac{1}{2}m^2\left(\text{Tr}\left(\mathcal{K}\right)^2-\text{Tr}\left(\mathcal{K}^2\right)\right)\Bigr], \end{aligned} \end{equation} where $\mathcal{K}^\mu_\alpha\mathcal{K}^\alpha_\nu=g^{\mu\alpha}f_{\alpha\nu}$, and the non-zero elements of the fixed reference metric $f_{\mu\nu}$ are $f_{xx}=f_{yy}=1$. This explicit breaking of diffeomorphism invariance in the gravitational theory is equivalent to the loss of momentum conservation in the dual field theory. This theory has a planar black hole solution \begin{equation} \begin{aligned} \label{eq:massivegravitysolution} ds^2&=\frac{r^2g(r)}{L^2}\left(-h(r)dt^2+dx^2+dy^2\right)+\frac{L^2}{r^2g(r)h(r)}dr^2,\\ A_t(r)&=\sqrt{\frac{3Q\left(Q+r_0\right)}{L^2}\left(1-\frac{m^2L^4}{2\left(Q+r_0\right)^2}\right)}\left(1-\frac{Q+r_0}{Q+r}\right),\;\;\;\;\; \phi(r)=\frac{1}{3}\log(g(r)),\\ h(r)&=1-\frac{m^2L^4}{2\left(Q+r\right)^2}-\frac{\left(Q+r_0\right)^3}{\left(Q+r\right)^3}\left(1-\frac{m^2L^4}{2\left(Q+r_0\right)^2}\right),\;\;\;\;\;\;\;g(r)=\left(1+\frac{Q}{r}\right)^\frac{3}{2}, \end{aligned} \end{equation} where the radial co-ordinate $r$ spans the range between $r_0$, the location of the horizon, and $\infty$, the boundary of the spacetime. The temperature $T$ and chemical potential $\mu$ of the dual field theory are \begin{equation} \begin{aligned} T=\frac{r_0\left(6\left(1+Q/r_0\right)^2-\frac{m^2L^4}{r_0^2}\right)}{8\pi L^2\left(1+Q/r_0\right)^{3/2}},\;\;\;\;\;\;\;\;\;\;\mu=\frac{\sqrt{3Q\left(Q+r_0\right)\left(1-\frac{m^2L^4}{2\left(Q+r_0\right)^2}\right)}}{L^2}. \end{aligned} \end{equation} Even with $m\ne0$, the zero temperature near-horizon geometry of (\ref{eq:massivegravitysolution}) retains the scaling symmetries (\ref{eq:scalingsymmetries}). For the chemical potential $\mu$ to be real, we require that $m^2L^4\le2\left(Q+r_0\right)^2$, and therefore $T/\mu\propto\sqrt{r_0/Q}$ at low $T/\mu$. The linear entropy density $s$ of the system at low $T$ can be made transparent by writing it as a function of $r_0/Q$ and $\bar{m}\equiv m/\mu$ \begin{equation} \begin{aligned} s/\mu^2=\frac{2\pi L^2}{3\kappa_4^2}\sqrt{r_0/Q}\sqrt{1+r_0/Q}\left(1+\frac{3\bar{m}^2}{2\left(1+r_0/Q\right)}\right)\sim T/\mu \text{ at low } T, \end{aligned} \end{equation} and the charge density $\sigma$ is \begin{equation} \begin{aligned} \sigma/\mu^2=\frac{L^2}{2\sqrt{3}\kappa_4^2}\sqrt{1+r_0/Q}\sqrt{1+\frac{3\bar{m}^2}{2\left(1+r_0/Q\right)}}\sim \left(T/\mu\right)^0 \text{ at low } T, \end{aligned} \end{equation} where we have calculated these from the area of the horizon and Gauss' law respectively. The universal result of Blake and Tong \cite{Blake:2013bqa} then ensures a linear resistivity $\rho_{DC}$ at low temperatures \begin{equation} \label{eq:analyticmassivegravityresistance} \rho_{DC}=\frac{s}{4\pi\sigma^2}m^2=\frac{2\kappa_4^2}{L^2}\frac{1}{\sqrt{1+Q/r_0}}\frac{m^2}{\mu^2}\sim T/\mu \text{ at low } T, \end{equation} as in the cuprate strange metal phase. The graviton mass produces a non-zero $\rho_{DC}$ by coupling the momentum to a uniform operator i.e. one with no characteristic periodic length scale\cite{Davison:2013jba} -- one consequence of this is that the planar black hole solution is homogeneous and isotropic. This operator plays the role of the homogeneous mode sourced by random disorder in $T^{tt}$ (or $J^t$), producing a DC resistivity linear in $T$. The arbitrary dimensionful parameter in the holographic setup (the graviton mass $m$) corresponds to the energy scale characterising the impurities i.e. $l^{-1}$ in equation (\ref{rhovsent}), or $\bar{V}$ in equation (\ref{eq:neutralmemorymatrixrho}). It must be small $m\ll\mu$ for the coupling to impurities to be weak in the IR. There is a lot of evidence that massive theories of gravity arise generically as the effective description of the low energy physics in holographic systems in which momentum dissipates slowly. That is, translational symmetry breaking generates an effective mass for metric components which controls the DC resistivity of the state. This is certainly the case when translational symmetry is broken by sources for a scalar operator.\cite{Andrade:2013gsa,Blake:2013owa,Lucas:2014zea,Gouteraux:2014hca,Donos:2014uba} One way to generate a graviton mass like that above\cite{Andrade:2013gsa} (rather than by inserting it by hand into the action) is to break translational invariance explicitly, in a homogeneous and isotropic way, via the introduction of linear sources for two additional scalar fields $\varphi^i=mx^i$. This effective theory does not capture the subleading effects of the non-homogeneous components of the disorder: see reference \onlinecite{Lucas:2014zea} for work in this direction. It would be interesting to try and generalise these results to the cases where translational invariance is explicitly broken in the UV not by scalars, but by the metric or gauge field. This would constitute an explicit proof that theories of massive gravity are always the relevant low energy effective theory for holographic systems with slow momentum dissipation. \section{Discussion} \label{sec:discussionsection} In summary, we have presented a simple mechanism which can explain how strange metals with a linear in temperature electronic entropy can acquire a linear in temperature resistivity. It relies upon the assumption that the electronic system is well described by the laws of hydrodynamics with a minimal viscosity, a feature that arises naturally in systems with a holographic dual. While experimental tests will ultimately determine whether this explanation is correct, holography has once again proved a valuable tool for thinking about old problems in new ways. We outline below some of the implications of our mechanism and some experimental tests of it. As we have previously mentioned, a hydrodynamic description is not applicable at short distances in a conventional metal, and it would be very interesting to determine further ``smoking gun'' experimental signals of hydrodynamic behaviour in metals. We hope to return to this point in the future. The mechanism we have described offers remarkably simple explanations for some mysterious experimental facts in the high $T_c$ cuprate superconductors. Firstly, the linear resistivity of the strange metal phase is due to its linear electronic entropy.\cite{PhysRevLett.71.1740} Secondly, the vanishing residual resistivity (see, for example, reference \onlinecite{2013PNAS..11012235B}) is due to the vanishing entropy when $T=0$: in this limit $\eta$ vanishes and thus this perfect fluid does not dissipate momentum. Experimentally,\cite{vdmarelpowerlawcuprates} it is known that $1/\tau_k \simeq \hbar / k_B T$ and using equation (\ref{rhovsent}), we estimate that $l \simeq 10^{-9}$ m, remarkably close to the Ioffe-Regel limit ($l\simeq a$, the lattice constant). This sheds new light on the long-standing puzzle of why the strong chemical disorder which is known to be intrinsic to cuprate crystals is not imprinted on their residual resistivity. It may also explain why the effects of electron-phonon couplings appear to be invisible in the electronic transport properties. At non-zero temperatures, there should be a strongly temperature-dependent contribution to momentum relaxation due to inelastic scattering against phonons, but in a dirty metal this is overwhelmed by the temperature-independent elastic scattering. If we can incorporate both effects by an effective mean free path $l$ in equation (\ref{monrelaxtime}), the effects of the electron-phonon interactions will be invisible. The mechanism we have described here may be tested experimentally. The $T^2$ resistivity of the cuprates in the pseudogap regime\cite{2013PNAS..11012235B,2013PNAS..110.5774M} coincides with a reduction in the electronic specific heat to a form which looks quadratic in temperature.\cite{PhysRevLett.71.1740} If the entropy law (\ref{rhovsent}) is controlling this resistivity, then systematic, precision measurements over a large range of $T$ should show a very close correlation between these quantities. The pseudogap regime is associated with ordering phenomena.\cite{2013Natur.498...75S} To have $\rho_{DC}\sim s$ in this regime, we require some novel, emergent, quantum critical degrees of freedom to be present. Although these are alien to the conventional theories of ordering, symmetry breaking is a ubiquitous phenomenon in holographic theories where these IR degrees of freedom abound. A second experimental signature of hydrodynamic behaviour in metals is a strong violation of the Wiedemann-Franz law.\cite{Mahajan:2013cja} To measure this would require suppressing $T_c$ with a large magnetic field in order to isolate the electronic contribution to the thermal conductivity. Finally, let us emphasise one of our key assumptions, and some questions that remain to be addressed. We argued that, unlike a Fermi liquid, a strongly interacting quantum critical system can form a hydrodynamic state before its interactions with a periodic potential become important, due to its minimal viscosity. However, given that the cuprates can form Mott insulators, the effects of a periodic potential are, a priori, expected to be large. For the holographic state studied here, the timescale $\tau^{}_U$ over which momentum is lost due to interactions with a periodic potential is $\tau_U^{-1}\sim T^{2\nu\left(k_L\right)-1}$ and this timescale may be significantly longer than the corresponding timescale due to quenched disorder in equation (\ref{monrelaxtime}).\cite{Hartnoll:2012rj} This is required for our explanation to be valid. Our mechanism cannot explain either the high value of $T_c$ or the $T$ dependence of the Hall angle \cite{PhysRevLett.67.2088,PhysRevLett.67.2092} in the cuprates. In particular, an explanation of the Hall angle may require the presence of an independent relaxation time associated to Hall transport, unlike in the hydrodynamic model we have outlined in which all transport is controlled by the momentum relaxation time. Furthermore, as highlighted in reference \onlinecite{2013Sci...339..804B}, linear resistivities with a ``Planckian" momentum relaxation rate occur in a large variety of systems, including simple metals in the phonon-dominated regime, and heavy fermion-like systems. Phonon domination is the most deadly adversary of the hydrodynamic liquid we have described, while the heavy fermion systems acquire their name from their large specific heats that tend to diverge upon decreasing the temperature towards their quantum critical points. Moreover, these systems are much cleaner than the cuprates and we therefore do not see any reason to believe that our mechanism applies in these cases. \begin{acknowledgments} We are grateful to Mike Blake, Chris Herzog, Andrei Parnachev, Giuseppe Policastro, David Tong, David Vegh and especially Sean Hartnoll and Subir Sachdev for helpful discussions, and to the Isaac Newton Institute, Cambridge for hospitality while this work was completed. This research was supported in part by a VIDI grant and a VICI grant from the Netherlands Organization for Scientific Research (NWO), by the Netherlands Organization for Scientific Research/Ministry of Science and Education (NWO/OCW) and by the Foundation for Research into Fundamental Matter (FOM). This publication was made possible through the support of a grant from the John Templeton Foundation. The opinions expressed in this publication are those of the authors and do not necessarily reflect the views of the John Templeton Foundation. \end{acknowledgments} \bibliographystyle{apsrev}
\section{Introduction} In supersymmetric (SUSY) theories, the $R$ symmetry plays a unique role in suppressing a constant term in the superpotential. Without the $R$ symmetry, the constant term is expected to be at the Planck scale, which requires a SUSY breaking scale to be the Planck scale to achieve the almost flat universe. Thus, there is a strong case for the existence of a spontaneously broken $R$-symmetry if SUSY is the solution to the hierarchy problem~\cite{MaianiLecture,Veltman:1980mj,Witten:1981nf,Kaul:1981wp} between the weak scale and the Planck scale or the scale of the Grand Unified Theory (GUT). One caveat of the $R$ symmetry is that a generation of the appropriate vacuum expectation value (VEV) of the superpotential requires a symmetry breaking field to have a Planck scale $A$-term VEV and a non-vanishing $F$ term VEV at the same time if the symmetry is a continuous one~\cite{Dine:2009sw}. This means that an $R$ symmetry breaking field is nothing but the Polonyi field for the continuous $R$ symmetry. Therefore, by taking the Polonyi problem~\cite{Coughlan:1983ci} seriously, the $R$ symmetry which suppresses the constant term of the superpotential should be a discrete one. Interestingly, the simplest model of spontaneous discrete $R$ symmetry breaking consisting of a single chiral field has a convex but a very flat potential around the origin of the chiral field,% \footnote{ Ref.~\cite{Kumekawa:1994gx} pointed out not only the presence of the so-called $\eta$ problem in supergravity inflation models but also the importance of the $R$ symmetry to have flat potentials necessary for the inflation to occur. } which evokes a scalar potential used in new inflation models~\cite{Linde:1981mu,Albrecht:1982wi}. In fact, the simplest $R$-breaking model satisfies the slow-roll conditions in a wide parameter region, and hence, the $R$-breaking field is a good candidate for an inflaton~\cite{Kumekawa:1994gx,Izawa:1996dv,Izawa:1997df,Ibe:2006fs,Ibe:2006ck,Takahashi:2013cxa}. It is also remarkable that the domain wall problem~\cite{Zeldovich:1974uw} associated with the discrete $R$ symmetry breaking is automatically solved when the $R$ symmetry breaking field plays a role of the inflaton.% \footnote{ This situation is analogous to the original new inflation model~\cite{Linde:1981mu,Albrecht:1982wi}, where an inflaton is identified with a GUT breaking field and the monopole problem is solved. } We here emphasize that new inflation models tend to predict a small tensor fraction due to their small inflation scales~\cite{Lyth:1996im}. This property is fairly supported by the upper limit on the tensor fraction of cosmic perturbations set by the recent observations of the cosmic microwave background (CMB) ~\cite{Hinshaw:2012aka,Ade:2013zuv,Ade:2013uln}.% \footnote{ Simple large field inflation models such as the chaotic inflation models with a quadratic or a quartic potential~\cite{Linde:1983gd} are, on the other hand, now slightly disfavored at least by $1\sigma$ level, which requires some extensions~\cite{Kallosh:2010ug,Takahashi:2010ky,Harigaya:2012pg,Croon:2013ana,Nakayama:2013jka}. } In this paper, we further investigate the compatibility of the $R$-breaking new inflation model with the results of the Planck experiment~\cite{Ade:2013zuv,Ade:2013uln}. As we will see, the $R$-breaking new inflation model is consistent with all cosmological constraints and observations in a wide parameter region. Furthermore, the model predicts a lower bound on the gravitino mass, $m_{3/2}>O(100)$\,TeV. This lower limit on the gravitino mass is consistent with the observed Higgs mass of $126$\,GeV~\cite{Aad:2012tfa,Chatrchyan:2012ufa} in a class of models in which the masses of the stops are of order the gravitino mass~\cite{Okada:1990vk,Ellis:1990nz,Haber:1990aw}. We also show that the baryon asymmetry of the universe as well as the observed dark matter density can be consistently explained along with the $R$-breaking new inflation model. \section{Brief review on the $R$-breaking new inflation model} Let us begin with the simplest model of spontaneous discrete $Z_{NR}$ symmetry breaking consisting of a single chiral field $\phi$~\cite{Kumekawa:1994gx,Izawa:1996dv}. Here, we assume that $\phi$ is a singlet except for the $R$ symmetry with an $R$ charge $2$. Assuming $N=2n$, the superpotential of $\phi$ is given by, \label{eq:super} \begin{eqnarray} W=v^2\phi -\frac{g}{n+1}\phi^{n+1} + \cdots, \end{eqnarray} where the ellipses represent higher power terms of $\phi$. We neglect them throughout this paper, since we are interested in the region with $|\phi|\ll1$. The size of the coupling constant $g$ will be discussed later. Here and hereafter, we take the unit of the reduced Planck scale $M_{PL}\simeq 2.4\times 10^{18}$ GeV being unity. The parameters $v^2$ and $g$ are taken real and positive without loss of generality.% \footnote{ In order for the gravitino mass to be far smaller than the Planck scale, $v^2$ must be suppressed. The suppression can be explained, for example, by assuming an $U(1)_R$ symmetry under which $\phi$ has a charge of $2/(n+1)$, and the $U(1)_R$ symmetry be dynamically broken by a condensation of a (composite) chiral field with an $U(1)_R$ charge of $2-2(n+1)$~\cite{Izawa:1996dv}. } At supersymmetric vacua, the $Z_{2nR}$ symmetry is spontaneously broken down to the $Z_{2R}$ symmetry by the VEV of $\phi$, \begin{eqnarray} \vev{\phi}\simeq \left(\frac{v^2}{g}\right)^{1/n}\times e^{2\pi i m/n}, ~~~~(m = 0,1,\cdots,n-1) \end{eqnarray} which leads to the VEV of the superpotential, \begin{eqnarray} \vev{W}\simeq \frac{n}{n+1}v^2 \left(\frac{v^2}{g}\right)^{1/n} e^{2\pi i m/n} . \end{eqnarray} As we emphasized in the introduction, the scalar potential of this model is convex but very flat around $\phi \sim 0$. Thus, if the initial field value of $\phi$ is set close to its origin by, for example, a positive Hubble induced mass term of pre-inflation~\cite{Izawa:1997df} and the slow-roll conditions are satisfied, $\phi$ automatically brings about the inflation. Therefore, the simplest model of discrete $R$-symmetry breaking is equipped with necessary structures as a model of new inflation. Now, let us discuss details of the new inflation model. For that purpose, let us note that the K\"ahler potential of $\phi$ is given by \begin{eqnarray} \label{eq:Kahler} K = \phi \phi^\dag + \frac{1}{4}k \left(\phi \phi^\dag \right)^2 + \cdots, \end{eqnarray} where the ellipses denote higher power terms of $\phi$, whose contributions to the dynamics of $\phi$ are negligible again. The parameter $k$ is at most of order unity, and we assume $k>0$ so that $\phi = 0$ is a local maximum (see below). From Eqs.~(\ref{eq:super}) and (\ref{eq:Kahler}), the scalar potential of the scalar component of $\phi$ is given by \begin{eqnarray} V(\phi) &=& |v^2-g\phi^n|^2 - k v^4 |\phi|^2 + \cdots\nonumber\\ &=& v^4 -\left(g v^2\phi^n + {\rm h.c.}\right) - kv^4|\phi|^2 \cdots \ . \end{eqnarray} In terms of the radial and the angular components of $\phi$, $\phi = \varphi e^{i\theta}/\sqrt{2}$, the scalar potential is rewritten as, \begin{eqnarray} V(\varphi,\theta) = v^4 -\frac{k}{2}v^4 \varphi^2 -\frac{g}{2^{n/2-1}}v^2 \varphi^n {\rm cos}\left(n\theta\right) +\cdots. \end{eqnarray} It can be seen that for a given $\varphi>0$, the minimum of the potential is provided by $\theta = 2\pi l/n~(l=0,1,\cdots,n-1)$. In the following, the radial component $\varphi$ plays a role of the inflaton in new inflation. As we have mentioned, we assume that the initial condition of $\varphi$ is close to $0$, i.e. $|\varphi| \ll 1$. We further suppose that the initial condition of the angular direction $\theta$ is given by $\theta = 0~({\rm mod}~2\pi/n)$ for the time being. Since $\theta = 0~({\rm mod}~2\pi/n)$ is the minimum of the potential along the angular direction, $\theta = 0~({\rm mod}~2\pi/n)$ is kept during the inflation. Along the inflaton trajectory, the first and the second slow-roll parameters are given by \begin{eqnarray} \epsilon &\equiv& \frac{1}{2}\left(\frac{\partial V/\partial \varphi}{V}\right)^2 = \frac{1}{2}\left(k \varphi+ \frac{ng}{2^{n/2-1}}\frac{\varphi^{n-1}}{v^2}\right)^2, \nonumber\\ \eta &\equiv& \frac{\partial^2 V/\partial \varphi^2}{V} = - k - \frac{n(n-1)g}{2^{n/2-1}}\frac{\varphi^{n-2}}{v^2}. \end{eqnarray} Thus, the slow-roll conditions can be actually satisfied for $|\varphi|\ll 1$ as long as $k\ll 1$. By assuming $|k| \ll 1$, the inflation lasts until the inflaton reaches to \begin{eqnarray} \varphi_{\rm end} = \left(\frac{2^{(n-2)/2}v^2}{n(n-1)g}\right)^{1/(n-2)}\ , \end{eqnarray} at which the slow-roll conditions are violated, $|\eta| \simeq 1$. It should be noted that there is an one-to-one correspondence between the number of $e$-foldings $N_e$ and the field value of $\varphi$ during the inflation via \begin{eqnarray} \label{eq:efold} N_e (\varphi) = \int_{\varphi_{\rm end}}^\varphi \frac{V}{\partial V/\partial \varphi}{\rm d}\varphi \ . \end{eqnarray} Thus, by taking the inverse of Eq.~(\ref{eq:efold}), we obtain \begin{eqnarray} \varphi^{n-2}(N_e) = \frac{2^{(n-2)/2}kv^2}{ng}\left( e^{k(n-2)N_e}-1 + k(n-1)e^{k(n-2)N_e} \right)^{-1}. \end{eqnarray} In order to compare model predictions with CMB observations, let us calculate the properties of the curvature perturbation. The spectrum of the curvature perturbation ${\cal P}_\zeta$ and its spectral index $n_s$ are given by \begin{eqnarray} \label{eq:Pzeta} {\cal P}_\zeta &=& \frac{1}{24\pi^2}\frac{V}{\epsilon}= \frac{1}{24\pi^2}\left( n^2g^2 k^{-2(n-1)}v^{4(n-3)}\left(e^{k(n-2)N_e}-1\right)^{2(n-1)} \right)^{\frac{1}{n-2}}e^{-2k(n-2)N_e}, \\ n_s &=& 1-6\epsilon+2\eta= 1-2k\left( 1 + \frac{n-1}{\left(1+k\left(n-1\right)\right)e^{k(n-2)N_e}-1} \right), \end{eqnarray} respectively. In Fig.~\ref{fig:tilt}, we show the prediction on the spectral index for $n=4,5,6$ and $N_e=50$. The colored region shows a region favored by the Planck experiment, i.e. $n_s = 0.9643\pm 0.012$~\cite{Ade:2013uln} for the pivot scale $k_*=0.002~{\rm Mpc}^{-1}$ at 95\%C.L. It can be seen that the model with $n\leq4$ is disfavored by the Planck experiment for $N_e =50$. For $n=5$, $k\sim 10^{-2}$ is favored. In Fig.~\ref{fig:tilt2}, we show the $N_e$ dependence of the spectral index for $n=4$. The figure shows that the model with $n=4$ is still consistent with the Planck experiment for $N_e \mathop{}_{\textstyle \sim}^{\textstyle >} 56$.% \footnote{ In Ref.~\cite{Takahashi:2013cxa}, it is pointed out that the model with $n=4$ is also consistent with the Planck experiment if there are a small constant term in the superpotential beside the one from the condensation of $\phi$. Since we assume that the $R$ symmetry is broken only by the condensation of $\phi$, that solution is not applicable. } We will discuss impacts of the observed spectral index to the gravitino mass in the next section. Before closing this section, let us discuss more general initial conditions for the inflaton field, $\theta \neq 0~({\rm mod}~2\pi/n)$. In particular, we are interested in how the spectral index is affected, since $n=4$ is severely constrained for $\theta = 0~({\rm mod}~2\pi/n)$ by the Planck results. In Fig.~\ref{fig:potential}, we show a schematic picture of the shape of the inflaton potential for $n=4$. For a better presentation, we show only the region with ${\rm Re}(\phi)>0$. For a fixed number of e-foldings, a non-zero angle $\theta$ leads to a larger corresponding field value for $\varphi$. As a result, the curvature of the inflaton trajectory becomes negatively larger, and the spectral index becomes more red-tilted. Therefore, even if we consider the initial condition with $\theta \neq 0~({\rm mod}~2\pi/n)$, the model with $n=4$ is still disfavored unless $N_e$ is large. \begin{figure}[tb] \begin{center} \includegraphics[width=0.5\linewidth]{tilt.pdf} \end{center} \caption{The spectral index of the curvature perturbation $n_s$ for $n=4,5,6$ with $N_e=50$. A colored region show the 95\% C.L. favored region by the Planck experiment, $n_s = 0.9643\pm 0.012$~\cite{Ade:2013uln}. } \label{fig:tilt} \end{figure} \begin{figure}[tb] \begin{center} \includegraphics[width=0.5\linewidth]{tilt2.pdf} \end{center} \caption{The spectral index of the curvature perturbation $n_s$ for $n=4$ with various $N_e$. The colored region show the 95\% C.L. limit from the Planck experiment, $n_s = 0.9643\pm 0.012$~\cite{Ade:2013uln}. } \label{fig:tilt2} \end{figure} \begin{figure}[tb] \begin{center} \includegraphics[width=0.4\linewidth]{potential.pdf} \end{center} \caption{A schematic picture for the scalar potential of $\phi$. The two lines show the trajectories of the inflaton with angular initial condition with either $\theta = 0$ or $\theta \neq 0$. The later trajectory feels steeper potential, and hence, the spectral index becomes more red-tilted. } \label{fig:potential} \end{figure} \section{Lower bound on the gravitino mass} In this section, we put a lower bound on the gravitino mass $m_{3/2}$ in the $R$-breaking new inflation models based on the results obtained in the previous section. From Eq.~(\ref{eq:Pzeta}), the parameter $v^2$ is expressed by the curvature perturbation, ${\cal P}_\zeta \simeq 2.2\times10^{-9}$~\cite{Ade:2013uln}, as \begin{eqnarray} \label{eq:v} v^2 &=& \left( \left(24\pi^2{\cal P}_\zeta\right)^{n-2} \left(ng\right)^{-2} \left(\frac{k}{e^{k(n-2)N_e}-1}\right)^{2(n-1)} e^{2k(n-2)^2N_e} \right)^{\frac{1}{2(n-3)}}\ , \end{eqnarray} which leads to \begin{eqnarray} \label{eq:v} v &\simeq& \begin{cases} 9.0\times10^{11}~{\rm GeV}~g^{-1/2}&(n=4,k=0.01,N_e=56),\\ 6.2\times10^{13}~{\rm GeV}~g^{-1/4}&(n=5,k=0.01,N_e=50),\\ 2.5\times10^{14}~{\rm GeV}~g^{-1/6}&(n=6,k=0.01,N_e=50). \end{cases} \end{eqnarray} It should be noted that $v$ does not depend on $k$ significantly. As a result, the gravitino mass $m_{3/2}$ is given by \begin{eqnarray} \label{eq:m32} m_{3/2} &=& \frac{ng}{n+1} \left(\frac{v^2}{g}\right)^{\frac{n+1}{n}}\nonumber\\ &\simeq& \begin{cases} 1.6\times10^{2}~{\rm GeV}~g^{-3/2}&(n=4,k=0.01,N_e=56),\\ 2.0\times10^{7}~{\rm GeV}~g^{-4/5}&(n=5,k=0.01,N_e=50),\\ 1.1\times10^{9}~{\rm GeV}~g^{-5/9}&(n=6,k=0.01,N_e=50). \end{cases} \end{eqnarray} As we have shown in the previous section, the model with $n=4$ is consistent with the Planck experiment only if $N_e \mathop{}_{\textstyle \sim}^{\textstyle >} 56$. This requires a very large $v^2$, which in turn puts a lower bound on the gravitino mass. To see this, let us remind ourselves that $N_e$ is given by the inflation scale as~\cite{Liddle:1993fq} \begin{eqnarray} \label{eq:Ne} N_e = 52-{\rm ln}\left(\frac{10^{12}~{\rm GeV}}{v}\right)\ , \end{eqnarray} for the pivot scale $k_*=0.002~{\rm Mpc}^{-1}$. Here, we have assumed an instantaneous reheating after the inflation, which brings about the largest $N_e$ for a fixed inflation scale. From Eqs.~(\ref{eq:v}), (\ref{eq:m32}) and (\ref{eq:Ne}), we obtain a relation between $m_{3/2}$ and $N_e$, which is shown in Fig.~\ref{fig:m32Ne}. From the figure and the constraint $N_e \mathop{}_{\textstyle \sim}^{\textstyle >} 56$, we obtain a lower bound on the gravitino mass, $m_{3/2}>{\cal O}(10^8)$ GeV. Next, let us discuss the model with $n>4$. In Fig.~\ref{fig:m32}, we show the gravitino mass for $n=5,6$ with $N_e=50$, $k=0.01$. In can be seen that the larger and smaller $n$ and $g$ are, the larger the gravitino mass is. Hence, we can derive a lower bound on $m_{3/2}$ from a upper bound on $g$ for the model with $n=5$. It should be noted that there is an upper bound on $g$ from the unitarity limit, which can be extracted by considering the leading radiative correction to the K\"ahler potential due to the coupling $g$, \begin{eqnarray} \delta K \simeq \frac{5!}{(16\pi^2)^4}g^2M_*^6 \phi \phi^\dag, \end{eqnarray} where $M_*$ is the cutoff of the loop integration. By requiring the unitarity up to the Planck scale, i.e. $M_* \simeq M_{\rm PL}$, the unitarity limit, $|\delta K| \mathop{}_{\textstyle \sim}^{\textstyle <} \phi\phi^\dag$, leads to an upper bound on $g$,% \footnote{ This requirement based on $M_* =1 $ is equivalent to the Born unitarity up to the Planck scale. } \begin{eqnarray} \label{eq:uplimg} g \mathop{}_{\textstyle \sim}^{\textstyle <} (16\pi^2)^2/\sqrt{5!}\simeq 2000. \end{eqnarray} By substituting this upper limit into Eqs.~(\ref{eq:m32}) and (\ref{eq:uplimg}), we obtain a lower bound on the gravitino mass, $m_{3/2} \mathop{}_{\textstyle \sim}^{\textstyle >} 100~{\rm TeV}$ for $n > 4$. In summary, we find that the lower bound on the gravitino mass; \begin{eqnarray} \label{eq:gravitino mass} m_{3/2} \mathop{}_{\textstyle \sim}^{\textstyle >} 100~{\rm TeV}, \end{eqnarray} in the $R$-breaking new inflation model. For $n =4$, the (much higher) lower limit on the gravitino mass is obtained to achieve the observed spectral index, while the milder limit for $n>4$ is obtained from the size of the curvature perturbation. As stressed in the introduction, this lower bound is consistent with the observed Higgs mass of $125$ GeV~\cite{Okada:1990vk,Ellis:1990nz,Haber:1990aw}. \begin{figure}[tb] \begin{center} \includegraphics[width=0.5\linewidth]{m32Ne.pdf} \end{center} \caption{A relation between $m_{3/2}$ and $N_e$ for $n=4$. } \label{fig:m32Ne} \end{figure} \begin{figure}[tb] \begin{center} \includegraphics[width=0.5\linewidth]{m32.pdf} \end{center} \caption{The gravitino mass for $n=5,6$ with $N_e=50$, $k=0.01$. } \label{fig:m32} \end{figure} \section{Baryon asymmetry and dark matter density} In this section, we argue that the baryon asymmetry as well as the dark matter density in the present universe can be explained consistently with the $R$-breaking inflation model. In the following, we concentrate on the model with $n=5$, $k\simeq 0.01$ and $N_e=50$. \subsection{Baryon asymmetry} \subsubsection*{Thermal leptogenesis} Let us first discuss whether the thermal leptogenesis~\cite{Fukugita:1986hr} can be achieved in the $R$-breaking new inflation model, that is, whether a reheating temperature $T_R$ can be high enough, $T_R \mathop{}_{\textstyle \sim}^{\textstyle >} 10^9$ GeV~\cite{Buchmuller:2004nz}. First, let us consider an inflaton decay via Planck-suppressed dimension five interactions% \footnote{ For example, a K\"ahler interaction $K= \lambda\phi^\dag QQ$, where $Q$ is some chiral field lighter than the inflaton, provides such decay channel.} in which the decay width of the inflaton $\Gamma_{\phi,{\rm dim}-5}$ is as large as $m_{\phi}^3$, where $m_{\phi}$ is the inflaton mass around the vacuum, \begin{eqnarray} m_\phi = 5 g \left(\frac{v^2}{g}\right)^{4/5} \simeq 1.4\times 10^{11}~{\rm GeV} \left(\frac{g}{1000}\right)^{-1/5}. \end{eqnarray} In this case, a reheating temperature $T_R$ is as large as \begin{eqnarray} \label{eq:TR_dim5} T_R\sim \sqrt{\Gamma_{\phi,{\rm dim}-5}} \sim 10^7~{\rm GeV}\left(\frac{g}{1000}\right)^{-3/10}\ll 10^9~{\rm GeV}. \end{eqnarray} Therefore, for a successful thermal leptogenesis, we are lead to introduce unsuppressed interactions.% \footnote{ If the dimension five interaction saturates the unitarity bound, $\lambda\sim 4\pi$, $T_R$ is as large as $10^8$ GeV. When the right-handed neutrinos have a non-hierarchical mass spectrum and the neutrino Yukawa matrix is rather tuned, the thermal leptogenesis is possible~\cite{Flanz:1996fb,Pilaftsis:1997jf,Blanchet:2008pw}. } In order to enhance the decay rate of the inflaton, let us consider a superpotential \begin{eqnarray} \label{eq:super_decay} W = \frac{y}{2\ell} \phi^\ell QQ, \end{eqnarray} where $Q$ is some chiral field lighter than the inflaton and $y$ is a coupling constant. Due to large $\vev{\phi}$, \begin{eqnarray} \vev{\phi}=\left(\frac{v^2}{g}\right)^{1/5}\simeq 2\times 10^{-3}\left(\frac{g}{1000}\right)^{-3/10}, \end{eqnarray} the decay of the inflaton by this interaction is effective even if $\ell>1$. The decay width of $\phi$ by this operator is given by \begin{eqnarray} \Gamma_\phi = \frac{1}{8\pi} y^2 |\vev{\phi}|^{2\ell-2}m_{\phi} =\frac{\ell^2}{8\pi}\frac{m_Q^2}{|\vev{\phi}|^2} m_{\phi}, \end{eqnarray} where $m_Q$ is the mass of $Q$. A reheating temperature is given by \begin{eqnarray} \label{eq:TR} T_R \simeq \left(\frac{90}{\pi^2 g_*}\right)^{1/4}\sqrt{\Gamma_\phi}= 1.8\times~10^9 {\rm GeV} \left(\frac{m_Q}{5\times10^{10}{\rm GeV}}\right) \left(\frac{\ell}{3}\right) \left(\frac{g}{1000}\right)^{1/5} \left(\frac{g_*}{200}\right)^{-1/4}, \end{eqnarray} where $g_*$ is the effective degree of freedom of the radiations. It can be seen that the thermal leptogenesis is marginally possible. In the mentioned above reheating scenario, we have introduced a matter field $Q$. Note that we cannot identify $Q$ with the minimal supersymmetric standard model (MSSM) higgs doublets, since a Dirac mass term of the MSSM higgs doublets, the so-called $\mu$ term, should be as small as the gravitino mass, and hence a reheating temperature is not high enough (see Eq.~(\ref{eq:TR})). An interesting idea is to identify $Q$ with the right-handed neutrinos, $N_i~(i=1,2,3)$~\cite{Kumekawa:1994gx}. In this case, the masses of the right-handed neutrinos, which should be far smaller than the Planck scale in order to obtain the observed masses of the left-handed neutrinos by the seesaw mechanism~\cite{seesaw}, are controlled by the $Z_{2nR}$ symmetry rather than the $B-L$ symmetry. For example, let us arrange the right-handed neutrinos by their masses; $m_{N_1}\leq m_{N_2}\leq m_{N_3}$. The inflaton decays mostly into the heaviest right-handed neutrino as long as the decay is kinematically allowed, that is, $2m_{N_i}<m_{\phi}$. If the inflaton decays mostly into $N_2$ or $N_3$ and the resulting reheating temperature is larger enough than $m_{N_1}$, the thermal leptogenesis is marginally possible. \subsubsection*{Non-thermal leptogenesis} We have shown that the thermal leptogenesis is marginally possible in the $R$-breaking new inflation model with $n=5$. Interestingly, when we identify $Q$ with the right-handed neutrinos, a possibility of the non-thermal leptogenesis scenario~\cite{Ibe:2006fs} is also opened.% \footnote{ If we introduce a K\"ahler interaction $K=\phi^\dag NN$ instead of the superpotential given by Eq.~(\ref{eq:super_decay}), a reheating temperature is as large as $10^7$ GeV (Eq.~(\ref{eq:TR_dim5})) and the non-thermal leptogenesis is possible. In this case, the right-handed neutrinos has an $R$ charge of one, and the masses of the right-handed neutrinos are in general of order the Planck scale. In order to obtain $m_N<m_\phi$ as well as the observed masses of the left-handed neutrinos, some tunings are necessary. If we further assume that the scale $v^2$ is given by a breaking of some charged field, the masses of the right-handed neutrinos are also given by breaking of the charged field and hence is naturally small. } There, the inflaton decays into right-handed neutrinos and the non-equilibrium decay of the right-handed neutrinos with a $CP$ violation generates lepton numbers. For simplicity, let us assume that the inflaton decays mostly into the lightest right-handed neutrino $N_1$. The entropy yield of the baryon number is given by~\cite{Ibe:2005jf} \begin{eqnarray} \eta_B\equiv \frac{n_B}{s}= 9\times 10^{-11} \left(\frac{T_R}{10^6{\rm GeV}}\right) \left( \frac{2m_{N_1}}{m_\phi}\right) \left( \frac{m_{\nu3}}{0.05{\rm eV}}\right) \frac{1}{{\rm sin}^2\beta}\delta_{\rm eff}, \end{eqnarray} where $m_{\nu3}$ is the mass of the heaviest left-handed neutrino, and $\beta$ is defined by the vacuum expectation values of the up-type and down-type higgs doublets, $H_u$ and $H_d$, as ${\rm tan}\beta = \vev{H_u}/\vev{H_d}$. $\delta_{\rm eff}$ represents a degree of the $CP$ violation, which is given by the Yukawa couplings of the right-handed neutrinos, and expected be of order one. Compared with the observed value, $\eta_{B-{\rm obs}} \simeq 8.5\times 10^{-11}$~\cite{Ade:2013zuv}, an appropriate baryon asymmetry can be generated in the non-thermal leptogenesis scenario. \subsection{Dark matter density} In the MSSM, there is a candidate for dark matter, the lightest supersymmetric particle (LSP). Here, we assume pure gravity mediation models/minimal split SUSY models~\cite{Ibe:2006de,minimalSplit}, in which the gaugino masses are generated only by one-loop effects and hence smaller in comparison with the gravitino, higgsino and sfermion masses, and the wino is the LSP. The wino mass $M_2$ is given by~\cite{Giudice:1998xp} \begin{eqnarray} \label{eq:AMSB} M_2 = \frac{g_2^2}{16\pi^2}\left(m_{3/2}+L\right), \end{eqnarray} where $g_2$ is the $SU(2)$ gauge coupling constant. The first term originates from an anomaly mediated effect~\cite{Dine:1992yw,Randall:1998uk,Giudice:1998xp}, while the second term, $L$, parametrizes a higgsino threshold correction.% \footnote{ If there is a vector-like matter in addition to the MSSM fields, the gaugino masses receive a one-loop correction further~\cite{Nelson:2002sa,Nakayama:2013uta}. For a comprehensive discussion on the phenomenology of the gauginos in that case, see Ref.~\cite{Harigaya:2013asa}. } As shown in Ref.\,\cite{Ibe:2006de}, $L$ is expected to be of order the gravitino mass in pure gravity mediation models/minimal split SUSY models. There are three sources for wino productions, a thermal wino relic, non-thermal production of gravitinos from a thermal bath, and gravitino production from the inflaton decay. We explain them in the following. \subsubsection*{Thermal wino relic} Since the wino has an $SU(2)$ gauge interaction, it is in a thermal equilibrium in the early universe. As the temperature of the universe decreases, the wino abundance freezes out and remains as a dark matter since the wino is the LSP. This is nothing but the conventional WIMP scenario. In order for the thermal abundance not to excess the observed cold dark matter value, $\Omega_c h^2=0.1196\pm0.0031$~\cite{Ade:2013zuv}, it is required that~\cite{Hisano:2006nn} \begin{eqnarray} \label{eq:M2th} M_2 \mathop{}_{\textstyle \sim}^{\textstyle <} 3~{\rm TeV}. \end{eqnarray} \subsubsection*{Gravitino scattered from thermal bath} Since the gravitino interacts with another light fields only through Planck-suppressed interactions, once it is scattered from a thermal bath, it does not interact with the thermal bath again, and eventually decays into the wino. A contribution to the wino abundance from this process is given by~\cite{Kawasaki:1994af,Gherghetta:1999sw,Ibe:2004tg} \begin{eqnarray} \label{eq:M2sc} \Omega_{\rm wino,sc}h^2 \simeq 0.12\left( \frac{M_2}{200~{\rm GeV}}\right)\left(\frac{T_R}{10^{10}~{\rm GeV}}\right). \end{eqnarray} \subsubsection*{Gravitino from inflaton decay} After SUSY breaking, there is no remaining symmetry which prevents a mixing between the inflaton field and the SUSY breaking field at the vacuum. This effect induces an inflaton decay into gravitinos~\cite{Kawasaki:2006gs,Asaka:2006bv,Dine:2006ii,Endo:2006tf,Kawasaki:2006hm,Endo:2006qk,Endo:2007ih,Endo:2007sz}, which provides another source of non-thermal wino dark matter. As an example, let us take the following effective superpotential for the SUSY breaking field $Z$, \begin{eqnarray} W_{\rm eff}= \Lambda^2 Z, \end{eqnarray} where $\Lambda^2$ is a SUSY breaking scale, which should satisfy $\Lambda^2=\sqrt{3}m_{3/2}$ in our flat universe.% \footnote{We have assumed that $|\vev{Z}|\ll1$ to avoid the Polonyi problem.} By calculating the scalar potential of the scalar components of $Z$ and $\delta\phi\equiv \phi-\vev{\phi}$ including supergravity effects, we obtain a mixing term, \begin{eqnarray} \label{eq:mixing} V_{\rm mix} = \sqrt{3}(1-b) m_{\phi} \vev{\phi} m_{3/2}\delta\phi Z^\dag + {\rm h.c.}, \end{eqnarray} where $b$ is a coupling constant in the K\"ahler potential, $K\supset b ZZ^\dag \phi \phi^\dag$. A mixing angle $\epsilon$ between the scalar components of $Z$ and $\delta\phi$ is given by \begin{eqnarray} \epsilon = \sqrt{3}(1-b) m_{\phi} \vev{\phi} m_{3/2}/m_Z^2, \end{eqnarray} where $m_Z$ is the mass of the SUSY breaking field. Here it is assumed that $m_Z\gg m_{\phi}$, which is the case with typical dynamical SUSY breaking models.% \footnote{If not, an inflaton decay into gravitinos is suppressed~\cite{Dine:2006ii,Endo:2006tf}. An inflaton decay into SUSY breaking sector fields, which are expected to exist in general dynamical SUSY breaking models, can be also suppressed by separating the dynamical scale and the mass of $Z$, $m_Z\ll \Lambda$~\cite{Nakayama:2012hy}. } A coupling between the scalar component of $Z$ and its fermionic component $\psi$, the goldstino, is provided by the following K\"ahler potential which gives a mass to the scalar component of the SUSY breaking field~\cite{Dine:2006ii,Nakayama:2012hy}, \begin{eqnarray} \label{eq:mZ Kahler} K\supset -\frac{m_Z^2}{12m_{3/2}^2}ZZ^\dag Z Z^\dag. \end{eqnarray} The $D$ term of Eq.~(\ref{eq:mZ Kahler}) yields \begin{eqnarray} {\cal L}\supset -\frac{\sqrt{3}}{6}\frac{m_Z^2}{m_{3/2}}Z^\dag \psi \psi + {\rm h.c.} \end{eqnarray} From Eqs~(\ref{eq:mixing}) and (\ref{eq:mZ Kahler}), the decay rate is given by \begin{eqnarray} \Gamma_{3/2}\equiv \Gamma_{\phi\rightarrow 2\psi_{3/2}} \simeq \Gamma_{Z\rightarrow 2\psi,m_Z=m_{\phi}}|\epsilon|^2 = \frac{(b-1)^2}{32\pi} m_{\phi}^3 \vev{\phi}^2. \end{eqnarray} The entropy yield of the gravitino after the inflaton decay, $Y_{3/2}$, is estimated as \begin{eqnarray} Y_{3/2} = 2\times \frac{\Gamma_{3/2}}{\Gamma_{\rm tot}} \frac{3 T_R}{4 m_{\phi}} = \frac{3}{2} \sqrt{\frac{\pi^2 g_*}{90}}\frac{\Gamma_{3/2}}{m_{\phi}T_R}, \end{eqnarray} where $\Gamma_{\rm tot}$ is a total decay width of the inflaton. The wino abundance is given by \begin{eqnarray} \label{eq:M2de} \Omega_{\rm wino,dec}h^2 = \left(\frac{M_2}{\rm 3.5\times 10^{-9}~{\rm GeV}}\right)\times Y_{3/2}. \end{eqnarray} In Fig.~\ref{fig:constraint}, we show constraints on the gravitino mass and the reheating temperature from the wino abundance in a $(m_{3/2},T_R)$ plane, which is obtained by Eqs.~(\ref{eq:M2th}), (\ref{eq:M2sc}) and (\ref{eq:M2de}). Here, we have assumed that the wino mass is given by the purely anomaly mediated effect, $M_2\simeq 3\times 10^{-3} m_{3/2}$. The figure shows that the observed dark matter density is mainly explained by the non-thermal contributions. If the coupling constant in the K\"ahler potential, $b$, is close to unity, the mixing between the SUSY breaking field and the inflaton is suppressed and hence the contribution from the inflaton decay is small. We have also shown constraints from the baron asymmetry in the non-thermal leptogenesis scenario. The reheating temperature is identified with the one given in Eq.~(\ref{eq:TR}). In the lowest colored region, the generated baryon asymmetry is smaller than the observed value even if the $CP$ violation is maximum, $\delta_{\rm eff}=1$. The result is insensitive to ${\rm tan \beta}$ as long as ${\rm tan \beta} \mathop{}_{\textstyle \sim}^{\textstyle >} 1$. It can be seen that there is a portion of parameter space in which the baryon asymmetry as well as the dark matter density in the present universe is explained. \begin{figure}[tb] \begin{center} \includegraphics[width=0.7\linewidth]{constraint.pdf} \end{center} \caption{ Constraint on the gravitino mass and the reheating temperature from the wino abundance and the successful non-thermal leptogenesis scenario. Here, we have assumed the wino mass $M_2$ in Eq.~(\ref{eq:AMSB}) with $L=0$. } \label{fig:constraint} \end{figure} \section{Summary and discussion} In this paper, we have investigated a compatibility of the supersymmetric $R$-breaking new inflation model with the results of the Planck experiment. We have shown that a lower bound on the gravitino mass, $m_{3/2}>{\cal O}(100)$ TeV, is obtained from the result of the Planck experiment. We have also shown that the baryon asymmetry as well as the dark matter density in the present universe can be explained consistently with the $R$-breaking inflation model. As a final remark, let us interpret the gravitino mass from the landscape point of view~\cite{Bousso:2000x a,Kachru:2003aw,Susskind:2003kw,Denef:2004ze}. In the landscape of vacua, it is possible that the gravitino mass is biased to low energy scales in order to obtain the electroweak scale as naturally as possible. In this case, the nature should choose the gravitino mass which saturates the lower bound given by Eq.~(\ref{eq:gravitino mass}). Therefore, the gravitino mass, $m_{3/2}\simeq100$\,TeV, is a prediction in the $R$-breaking new inflation model in the landscape point of view.% \footnote{ If there is a severer bound on $g$ than the unitarity bound, a larger gravino mass, such as PeV, is predicted from the landscape point of view. This arugement may support the explanation of the PeV IceCube neutrino events~\cite{Aartsen:2013bka} by decaying gravitino dark matter\cite{Feldstein:2013kka}. } It should be cautioned that there is a hidden parameter in this argument, $k$, which has been fixed $k\simeq 0.01$ to account for the observed spectral index. From the anthropic point of view, however, there seems no reason for the spectral index to be close to unity as observed. If we allow for a spectral index as large as $0.8$, for example, then the gravitino mass is lowered down to, \begin{eqnarray} m_{3/2}\simeq 1.9\times 10^3~{\rm GeV}\times\left(\frac{g}{2000}\right)^{-4/5}~~ (n=5,k=0.1,N_e=50), \end{eqnarray} which is much smaller than $100$ TeV. This shows that our landscape argument is self-consistent only if the parameter $k$ is fixed to be close to $0.01$ by some underlying theory. If not, the landscape argument predicts that $k\sim 0.1$ and $m_{3/2}\sim1$ TeV, in which the electroweak scale is obtained much more naturally than the case with $k\sim 0.01$ and $m_{3/2}\sim 100$ TeV, and the prediction already contradicts with the observed value of the spectral index. This situation is similar to anthropic arguments~\cite{Weinberg:1987dv} on the electroweak scale. It is argued that an electroweak scale of the one realized in the nature is required for the people to exist in the universe~\cite{Agrawal:1998xa,Jeltema:1999na}. There, other parameters other than the Higgs boson mass in the standard model such as the gauge coupling constants and the Yukawa couplings are fixed to the observed value. The anthropic prediction on the electroweak scale is viable only if all such couplings are consider to be fixed by some underlying theory. Instead of fixing the parameter $k$, we may move ahead with the landscape point of view under an additional assumption. Suppose that the parameter with the positive mass dimension in the superpotential, $v$, is strongly biased to larger mass scales. However, $v$ is anthropically required to be sufficiently small in order to generate a small cosmological perturbation, ${\cal P}_\zeta\sim 10^{-9}$. Consequently, the maximum $v$ on the hyper-surface of the parameter space corresponding to ${\cal P}_\zeta \sim 10^{-9}$ would have been chosen anthropically. In Fig.~\ref{fig:kVSv}, we show a line in a $k-v$ space in which ${\cal P}_\zeta = 2.2 \times 10^{-9}$. It can be seen that $k\sim 10^{-2}$, which is consistent with the observed spectral index, gives the maximum $v$. Note that the result is insensitive to the parameter $g$. It is remarkable that a high energy biased $v$ explains the reason why the spectral index $n_s$ is not too small such as $0.8$ but close to the observed value, i.e. $n_s\sim0.96$. \begin{figure}[tb] \begin{center} \includegraphics[width=0.5\linewidth]{kVSv.pdf} \end{center} \caption{ A line in a $k-v$ space in which ${\cal P}_\zeta =2.2\times 10^{-9}$. } \label{fig:kVSv} \end{figure} \section*{Acknowledgments} This work is supported by Grant-in-Aid for Scientific research from the Ministry of Education, Science, Sports, and Culture (MEXT), Japan, No.\ 22244021 (T.T.Y.), No.\ 24740151 (M.I), and also by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. The work of K.H. is supported in part by a JSPS Research Fellowships for Young Scientists.
\section{Introduction}\label{sectionintroduction} Causal sets are discrete order-theoretic models of classical spacetime. A {\it causal set} is a set $C$, assumed to be {\it countable}, whose {\it elements correspond to spacetime events,} together with a binary relation $\prec$ on $C$, satisfying the additional axioms of {\it transitivity}, {\it irreflexivity}, and {\it interval finiteness}. The binary relation $\prec$ defines an {\it interval finite partial order}\footnotemark\footnotetext{Section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} clarifies how the {\it irreflexive relation} $\prec$ corresponds to a {\it partial order} in the usual sense of the term.} on $C$, called the {\it causal order}, with the physical interpretation that $x\prec y$ in $C$ if and only if the event represented by $x$ exerts causal influence on the event represented by $y$. The physical interpretation of $C$ is completed by fixing a {\it discrete measure}\footnotemark\footnotetext{Here, {\it discrete measure} simply means that singletons have positive measure.} $\mu$ on $C$, assigning to each subset of $C$ a volume equal to its number of elements in fundamental units, up to {\it Poisson-type fluctuations.}\footnotemark\footnotetext{See the remarks on page 2 of Ahmed, Dodelson, Greene, and Sorkin's paper {\it Everpresent $\Lambda$} \cite{SorkinEverPresentLambda04}.} The role of $\prec$ and $\mu$ in modeling classical spacetime is summarized by the phrase, {\it ``order plus number equals geometry,"} coined by Rafael Sorkin, the foremost architect and advocate of causal set theory. This phrase represents a special case of what I refer to as the {\it causal metric hypothesis} (\hyperref[cmh]{CMH}), which is the idea that the properties of the physical universe, and in particular, the metric properties of classical spacetime, {\it arise from causal structure at the fundamental scale.} Approaches to fundamental physics involving some form of the causal metric hypothesis may be collectively referred to as {\bf causal theory}. Besides causal set theory, these include other discrete causal theories, such as {\it causal dynamical triangulations} and {\it causal nets,} as well as theories involving {\it interpolative} causal models of spacetime, such as {\it domain theory.} The purpose of this paper is to offer potentially critical improvements to the causal set program, and to causal theory in general. This effort is based on the conviction that causal theory is among the best-motivated existing approaches to the outstanding problems of fundamental physics, and that causal set theory is perhaps the cleanest and best-balanced existing version of causal theory. Despite important conceptual distinctions, and critical technical differences, between the existing formulation of causal set theory and the approach I offer in this paper, the two are close enough to be considered part of the same broad program. For example, they are closer to each other than they are to the other versions of causal theory mentioned above. In this introductory section, I sketch a broad conceptual context for the more specific material to follow. Section \hyperref[subsectionapproach]{\ref{subsectionapproach}} is a brief overview of causal set theory, describing its origins, outlining its principal methods, and providing a glimpse of its current state of development. I have included references to more thorough treatments, carefully selected for reliability and quality of exposition. Section \hyperref[subsectionpurpose]{\ref{subsectionpurpose}} outlines in more detail the purposes of this paper, its method and approach, and its intended audience. Section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} introduces general principles of natural philosophy used throughout the paper. Section \hyperref[subsectionnotation]{\ref{subsectionnotation}} describes the notation and conventions of the paper in general terms; appendix \hyperref[appendixindex]{A} is much more thorough. Section \hyperref[subsectionoutline]{\ref{subsectionoutline}} provides a hyperlinked topical outline of the succeeding material, more detailed than that provided by the table of contents above. \subsection{Overview of the Causal Sets Program}\label{subsectionapproach} {\bf { Historical Antecedents.}} The study of causality long predates any formal mathematical notion of order, and it is important to resist viewing relevant early developments in a {\it teleological} sense, as mere steps along a path toward present convention. For example, Zeno's {\it dichotomy paradox,} proposed around 450 B.C., is often described today in terms of the subdivision of continuum intervals, and ``resolved" by the convergence of geometric series. However, this particular version of Zeno's paradox was originally stated in a {\it physical} context, in terms of {\it sequences of events}, with the physical issue at stake being whether or not fundamental causal structure is {\it interpolative.} From this perspective, Zeno's paradox is as relevant today as it was when it was first proposed. Finiteness axioms, such as the causal set axiom of {\it interval finiteness} (\hyperref[if]{IF}), stated in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}}, or the alternative axiom of {\it local finiteness} (\hyperref[lf]{LF}), presented in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, offer {\it noninterpolative} resolutions of Zeno's paradox. Prior to the discover of relativity and quantum theory, the overwhelming success of continuum methods in physics relegated most alternative ideas about the basic structure of the universe to the periphery of scientific thought. Great thinkers such as Leibniz and Riemann considered such ideas, but the scientific community as a whole followed more pragmatic lines. Relativity brought fresh scrutiny to {\it structural issues,} elevating causality to its present central role. For historical background on the relativistic roots of causal set theory, it is difficult to improve upon the efforts of Sorkin. For example, in his 2005 article {\it Causal Sets: Discrete Gravity} \cite{SorkinCausalSetsDiscreteGravity05}, Sorkin details the isolation and elevation of causal structure in the relativity literature as early as the 1930's, and even quotes Einstein's misgivings about continuum geometry, and Riemann's early remarks about {\it discrete manifolds.} David Finkelstein also provides useful historical context reaching back before 1950, in his 1988 paper {\it ``Superconducting" Causal Nets} \cite{Finkelstein88}. Finkelstein's 1969 paper {\it Space-Time Code} \cite{Finkelstein69} is perhaps the earliest ``modern" paper on discrete causal theory, described by Sorkin \cite{SorkinCausalSetsDiscreteGravity05} as having ``adumbrated'' the causal sets program. Results of Hawking and Malament in the late 1970's, discussed further below, provided new evidence of the quintessential role of causal structure in determining the properties of classical spacetime. Soon after this, ideas very similar to causal set theory appeared in Myrheim's 1978 CERN preprint {\it Statistical Geometry}, \cite{Myrheim78} and 't Hooft's 1978 notes {\it Quantum Gravity: A Fundamental Problem and some Radical Ideas} \cite{tHooft78}. These efforts were spurred by the continued failure of continuum methods to provide a viable avenue toward a theory of quantum gravity, and the growing expectation that fundamental spacetime structure is non-smooth. Finally, causal set theory was introduced by Bombelli, Lee, Meyer, and Sorkin, in their 1987 paper {\it Space-Time as a Causal Set} \cite{Sorkinetal87}.\footnotemark\footnotetext{Crediting Finkelstein, Myrheim, and 't Hooft's earlier efforts, Bombelli, Lee, Meyer, and Sorkin write that {\it ``The picture of space-time as a causal set is by no means new,"} but go on to explain that Myrheim and 't Hooft's proposals are {\it ``undeveloped,"} and that Finkelstein's {\it ``led to formulations in which the issues we would like to address here were not dealt with."}} {\bf Malament's Theorem; Metric Recovery.} An important foundational result directly preceding the emergence of causal set theory is a 1977 theorem of David Malament, relating the causal and conformal structures of classical spacetime in the context of general relativity \cite{Malament77}. Hawking, King, and McCarthy \cite{Hawking76} had already established that {\it topological structure determines conformal structure} under suitable hypotheses; Malament showed that causal structure, in turn, determines topological structure. Combining these two results, Malament concluded that the only obstruction to the recovery of pseudo-Riemannian spacetime geometry from causal structure is a {\it lack of appropriate data about scale;} i.e., a missing conformal factor. More precisely, let $X$ and $X'$ be connected, four-dimensional, smooth manifolds without boundary, and let $g$ and $g'$ be smooth pseudo-Riemannian metrics of Lorentz signature on $X$ and $X'$, respectively. In this context, Malament proved that {\it if $f:X\rightarrow X'$ is a bijection such that $f$ and $f^{-1}$ preserve future-directed continuous timelike curves,\footnotemark\footnotetext{In general relativity, a {\it causal curve} is a curve that is either timelike or null. Malament's Lemma 1 (\cite{Malament77}, page 1400), building on Hawking's theorem \cite{Hawking76}, establishes that timelike curves suffice in this context.} then $f$ is a conformal isometry.} Subsequent results in causal set theory and domain theory have clarified and broadened this picture. For example, Luca Bombelli and David Meyer's 1989 paper {\it The origin of Lorentzian geometry} \cite{Bombelli89}, and Keye Martin and Prakash Panangaden's 2006 paper {\it A Domain of Spacetime Intervals in General Relativity} \cite{Martin06}, both contain results pertaining to the recovery of geometry from the causal structure of a {\it countable dense subset} of classical spacetime. {\bf Sorkin: ``Order Plus Number Equals Geometry."} Malament's theorem provides clear motivation for Sorkin's phrase, {\it ``order plus number equals geometry,"} which encapsulates Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). Here, {\it order} is represented by the binary relation $\prec$ on a causal set $C$, which corresponds to the family of causal curves on a pseudo-Riemannian manifold in Malament's theorem, while {\it number} is endowed with volume via the discrete measure $\mu$ on $C$, which corresponds to Malament's missing conformal factor. The idea of using the {\it counting measure} as a proxy for volume data appeared almost immediately after Malament's paper \cite{Malament77}. In his 1978 preprint \cite{Myrheim78}, Myrheim writes, \begin{quotation}\noindent{\it``If spacetime is assumed to be discrete, then the counting measure is the natural measure, and the causal ordering is the only structure needed. Coordinates and metric may be derived as secondary, statistical concepts."} (page 1) \end{quotation} In their 1987 paper \cite{Sorkinetal87}, Bombelli, Lee, Meyer, and Sorkin write, \begin{quotation}\noindent{\it``In this view volume is number, and macroscopic causality reflects a deeper notion of order in terms of which all the ``geometrical" structures of space-time must find their ultimate expression."} (page 522) \end{quotation} These statements, which implicitly {\it identify} the discrete volume measure $\mu$ with the counting measure, precede the appearance of the caveat {\it ``up to Poisson-type fluctuations"} in the later literature. Section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} contains a more thorough discussion of this point, along with appropriate quotations. Due to the axiom of interval finiteness (\hyperref[if]{IF}), the information encoded in a causal set can only {\it approximately} supply the causal and conformal ingredients of Malament's theorem. It is important to emphasize that {\it causal set theory treats relativistic spacetime as an approximation of a causal set, not vice versa.} More precisely, continuum geometry is viewed, from the causal set perspective, as a {\it large-scale smoothing} of the fundamental structure encoded by $C$, $\prec$ and $\mu$. The extent to which a ``manifold-like" causal set {\it uniquely} determines large-scale manifold structure is, as far as I know, still an outstanding issue.\footnotemark\footnotetext{In particular, Martin and Panangaden's ``discrete metric recovery" result \cite{Martin06} involves a {\it dense} subset, and therefore has no {\it effective volume gap} for the ``size" of elements.} The conjecture that an appropriate uniqueness result exists is called the causal set {\it hauptvermutung}, or {\it fundamental conjecture}, a historical reference to an analogous conjecture in geometric topology.\footnotemark\footnotetext{Namely, that any two triangulations of a triangulable space admit a common refinement. John Milnor proved in 1961 that this conjecture is {\it false}.} Sorkin discusses the causal set {\it hauptvermutung} in \cite{SorkinCausalSetsDiscreteGravity05}. {\bf Manifold Embedding; Sprinkling.} Most causal sets bear little resemblance to manifolds, and are therefore irrelevant to classical spacetime structure. Here, the quantifier ``most" is made precise by the {\it asymptotic enumeration} of Kleitman and Rothschild \cite{KleitmanRothschild75}, discussed in more detail in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}}. An {\it effective realization} of the causal metric hypothesis (\hyperref[cmh]{CMH}), whereby one may {\it derive} the emergence of approximate manifold structure at appropriate scales, beginning with just the axioms of causal set theory and suitable dynamical laws, has not yet been discovered. A useful intermediate step is to study causal sets {\it induced} by ``sprinkling" elements discretely on a pseudo-Riemannian manifold, according to a {\it Poisson process.} One may thereby obtain physically relevant causal sets for investigative purposes, without the necessity of knowing how these sets might arise dynamically. The method of sprinkling was introduced at the dawn of causal set theory in Bombelli, Lee, Meyer, and Sorkin's inaugural paper \cite{Sorkinetal87}. Sprinkling is illustrated schematically in figure \hyperref[sprinkling]{\ref{sprinkling}} below. In figure \hyperref[sprinkling]{\ref{sprinkling}}a, a region of a classical spacetime manifold $X$ is shown, together with a distinguished event $x$ on $X$, and a portion of its light cone. In figure \hyperref[sprinkling]{\ref{sprinkling}}b, a family $C$ of elements is sprinkled on $X$. In figure \hyperref[sprinkling]{\ref{sprinkling}}c, $X$ is removed from the picture, and only the abstract causal structure of $C$, induced by $X$, remains. Each edge in the figure represents a causal relation between its endpoints, directed upward. This is an example of a {\it Hasse diagram}, defined more generally below. Note that the diagram omits {\it reducible} relations; i.e., relations between pairs of elements of $C$ connected by complex sequences of other relations. This practice is permissible in causal set theory, essentially because causal sets are {\it transitive.} However, it {\it cannot} be applied to most of the alternative models of causal structure appearing in this paper, without destroying information. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(0,8.5)}{\pgfbox[left,center]{{\bf }}} \pgfputat{\pgfxy(.15,5.2)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,5.2)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,5.2)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.5)(16.5,5.5) \pgfxyline(0,-.1)(0,5.5) \pgfxyline(5.5,-.1)(5.5,5.5) \pgfxyline(11,-.1)(11,5.5) \pgfxyline(16.5,-.1)(16.5,5.5) \begin{pgftranslate}{\pgfpoint{.4cm}{-.35cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \begin{pgftranslate}{\pgfpoint{0cm}{.3cm}} \begin{colormixin}{30!white} \pgfmoveto{\pgfxy(2.5,2.6)} \pgflineto{\pgfxy(5.4,5.5)} \pgflineto{\pgfxy(.2,5.5)} \pgflineto{\pgfxy(.2,4.9)} \pgflineto{\pgfxy(2.5,2.6)} \pgffill \pgfmoveto{\pgfxy(2.5,2.6)} \pgflineto{\pgfxy(4.6,.5)} \pgflineto{\pgfxy(.4,.5)} \pgflineto{\pgfxy(2.5,2.6)} \pgffill \end{colormixin} \pgfputat{\pgfxy(3.2,4.8)}{\pgfbox[center,center]{\large{$X$}}} \pgfputat{\pgfxy(2.8,2.6)}{\pgfbox[center,center]{\large{$x$}}} \pgfputat{\pgfxy(4.9,.8)}{\pgfbox[center,center]{space}} \pgfputat{\pgfxy(.8,5.3)}{\pgfbox[center,center]{time}} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(.2,.2)(.2,5.5) \pgfxyline(-.1,.5)(5.5,.5) \end{pgftranslate} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.85cm}{-.35cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \begin{pgftranslate}{\pgfpoint{0cm}{.3cm}} \begin{colormixin}{30!white} \pgfmoveto{\pgfxy(2.5,2.6)} \pgflineto{\pgfxy(5.4,5.5)} \pgflineto{\pgfxy(.2,5.5)} \pgflineto{\pgfxy(.2,4.9)} \pgflineto{\pgfxy(2.5,2.6)} \pgffill \pgfmoveto{\pgfxy(2.5,2.6)} \pgflineto{\pgfxy(4.6,.5)} \pgflineto{\pgfxy(.4,.5)} \pgflineto{\pgfxy(2.5,2.6)} \pgffill \end{colormixin} \pgfputat{\pgfxy(3.2,4.8)}{\pgfbox[center,center]{\large{$X$}}} \pgfputat{\pgfxy(4.5,2.6)}{\pgfbox[center,center]{\large{$C$}}} \pgfputat{\pgfxy(2.8,2.6)}{\pgfbox[center,center]{\large{$x$}}} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.7)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,2.8)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,4.7)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.4)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.1,2.2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.1,5.2)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,3.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.6,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.5)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.2)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3.3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.7,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3,1.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.6,5.1)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4.1,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.4,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.4,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.8)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(.2,.2)(.2,5.5) \pgfxyline(-.1,.5)(5.5,.5) \end{pgftranslate} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.25cm}{-.35cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \begin{pgftranslate}{\pgfpoint{0cm}{.3cm}} \pgfputat{\pgfxy(4.5,2.6)}{\pgfbox[center,center]{\large{$C$}}} \pgfputat{\pgfxy(2.8,2.6)}{\pgfbox[center,center]{\large{$x$}}} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.7)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,2.8)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,4.7)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.4)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.1,2.2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.1,5.2)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,3.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.6,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.5)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.2)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3.3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.7,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3,1.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.6,5.1)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4.1,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.4,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.4,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.8)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node13} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node12} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node6}{Node15} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node8}{Node7} \pgfnodeconnline{Node8}{Node12} \pgfnodeconnline{Node9}{Node5} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node11} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node6} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node6} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node8} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node19} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node18} \pgfnodeconnline{Node17}{Node35} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node27} \pgfnodeconnline{Node19}{Node15} \pgfnodeconnline{Node20}{Node11} \pgfnodeconnline{Node20}{Node29} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node22}{Node30} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node24}{Node25} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node26} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node22} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node26}{Node34} \pgfnodeconnline{Node27}{Node23} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node17} \pgfnodeconnline{Node31}{Node26} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{pgftranslate} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify} \end{pgfpicture} \caption{a) Classical spacetime manifold $X$; b) family $C$ of elements sprinkled on $X$; c) abstract causal structure of $C$, induced by $X$.} \label{sprinkling} \end{figure} \vspace*{-.5cm} {\bf Quantum Theory.} The quantum theory of causal sets may be broadly divided into {\it background-dependent} and {\it background-independent} approaches. The former involve the study of auxiliary structures, such as particles and fields, existing {\it on} a causal set, while the latter involve the quantum nature of spacetime itself. Background dependent approaches are the easier of the two, and have been extensively developed in the existing literature. While interesting in their own right, most of these approaches differ too much in emphasis from the approach developed in this paper to provide helpful sources. An exception is the work of Ioannis Raptis, beginning with his 2000 paper {\it Algebraic Quantization of Causal Sets} \cite{RaptisAlgebraicQuantization00}, which provides useful context for the theory of {\it causal path algebras} appearing in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}, and the {\it abstract quantum causal theory} developed in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}, of this paper. Background independent quantum causal theory is much more interesting and difficult. The familiar {\it histories approach to quantum theory} offers promising general methods in this context. This approach may be traced back to Richard Feynman's method of {\it path integration,} first introduced in his 1948 paper {\it Space-Time Approach to Non-Relativistic Quantum Mechanics} \cite{FeynmanSOH48}, and subsequently applied with great success to a host of important problems throughout modern physics. Chris Isham's category-theoretic treatment of histories, introduced in his 2005 paper {\it Quantising on a Category} \cite{IshamQuantisingI05}, provides part of the motivation for the theory of {\it co-relative histories} and {\it kinematic schemes} in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper. Also relevant is Sorkin's work on the general topic of {\it quantum measure theory,} including his 2012 paper {\it Toward a Fundamental Theorem of Quantal Measure Theory} \cite{SorkinQuantalMeasure12}, which discusses quantum-theoretic measures on configuration spaces of histories, both for causal sets and discrete lattices. These ideas foreshadow the theory of {\it co-relative kinematics,} and {\it completions of kinematic schemes,} both discussed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below. {\bf Dynamics.} Causal set theory presently lacks dynamical laws describing the evolution of suitable manifold-like causal sets, or configuration spaces of such sets. Nevertheless, suggestive ``toy models" have been examined. The most prominent of these is a {\it classical stochastic model} called {\bf sequential growth dynamics}, introduced by Sorkin and his student David Rideout in their 1999 paper {\it Classical sequential growth dynamics for causal sets} \cite{SorkinSequentialGrowthDynamics99}. In sequential growth dynamics, a causal set is ``built up one element at a time," with each ``evolutionary step" represented by a special causal set {\it morphism,} called a {\it transition}.\footnotemark\footnotetext{In fact, a transition only {\it represents} a physically more fundamental relationship called a {\it co-relative history,} as explained in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}.} Transitions organize the class of causal sets, considered up to isomorphism,\footnotemark\footnotetext{Throughout this paper, structured sets are usually considered only up to isomorphism, since differences between members of an isomorphism class are physically immaterial.} into a ``higher-level" structure $\ms{K}$, which I refer to as a {\it kinematic scheme}. The theory and terminology of kinematic schemes is discussed in more detail in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} of this paper. $\ms{K}$ has the abstract structure of a {\it multidirected set}.\footnotemark\footnotetext{In particular, it is better {\it not} to view $\ms{K}$ as a partially ordered set, since there exist ``inequivalent" transitions with the same source and target. See section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} for an example of this, due to Brendan McKay.} It is analogous to a mathematical {\it category,} with its member causal sets corresponding to {\it objects,} and transitions between pairs of these causal sets corresponding to {\it morphisms.}\footnotemark\footnotetext{Of course, transitions {\it are} morphisms in the category of causal sets; the point is that they serve a similar structural role in $\ms{K}$, which is {\it not} a category. More precisely, they {\it represent} the ``morphisms" of $\ms{K}$, which are co-relative histories, as explained in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}}.} This analogy has been developed in a striking way by Chris Isham and his collaborators. In section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper, I offer a somewhat different perspective regarding such constructions. Figure \hyperref[sequential]{\ref{sequential}}a below illustrates a portion of the kinematic scheme $\ms{K}$. Individual causal sets appear in the large open nodes, and transitions are represented by the edges. To each transition is associated an {\it initial causal set} $C_i$, and a {\it terminal causal set} $C_t$, given by adding a single element to $C_i$, along with a family of relations. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{6.1cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,6)(16.5,6) \pgfxyline(0,-.1)(0,6) \pgfxyline(11,-.1)(11,6) \pgfxyline(16.5,-.1)(16.5,6) \pgfputat{\pgfxy(.15,5.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(11.15,5.7)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{-2.75cm}{0cm}} \pgfnodecircle{Node0}[stroke]{\pgfxy(8.25,.7)}{0.6cm} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,2.4)}{0.6cm} \pgfnodecircle{Node2}[stroke]{\pgfxy(6.25,3.5)}{0.6cm} \pgfnodecircle{Node3}[stroke]{\pgfxy(10.25,3.5)}{0.6cm} \pgfnodecircle{Node4}[stroke]{\pgfxy(4.25,5.2)}{0.6cm} \pgfnodecircle{Node5}[stroke]{\pgfxy(6.25,5.2)}{0.6cm} \pgfnodecircle{Node6}[stroke]{\pgfxy(8.25,5.2)}{0.6cm} \pgfnodecircle{Node7}[stroke]{\pgfxy(10.25,5.2)}{0.6cm} \pgfnodecircle{Node8}[stroke]{\pgfxy(12.25,5.2)}{0.6cm} \pgfnodeconnline{Node0}{Node1} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node2}{Node5} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \pgfnodecircle{Node01}[stroke]{\pgfxy(8.25,.7)}{0.13cm} \pgfxyline(8.12,.57)(8.38,.83) \pgfnodecircle{Node11}[fill]{\pgfxy(8.25,2.4)}{0.1cm} \pgfnodecircle{Node21}[fill]{\pgfxy(6,3.5)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(6.5,3.5)}{0.1cm} \pgfnodecircle{Node31}[fill]{\pgfxy(10.25,3.25)}{0.1cm} \pgfnodecircle{Node32}[fill]{\pgfxy(10.25,3.75)}{0.1cm} \pgfnodeconnline{Node31}{Node32} \pgfnodecircle{Node41}[fill]{\pgfxy(3.85,5.2)}{0.1cm} \pgfnodecircle{Node42}[fill]{\pgfxy(4.25,5.2)}{0.1cm} \pgfnodecircle{Node43}[fill]{\pgfxy(4.65,5.2)}{0.1cm} \pgfnodecircle{Node51}[fill]{\pgfxy(6,5)}{0.1cm} \pgfnodecircle{Node52}[fill]{\pgfxy(6.5,5)}{0.1cm} \pgfnodecircle{Node53}[fill]{\pgfxy(6.25,5.5)}{0.1cm} \pgfnodeconnline{Node51}{Node53} \pgfnodeconnline{Node52}{Node53} \pgfnodecircle{Node61}[fill]{\pgfxy(8,4.95)}{0.1cm} \pgfnodecircle{Node62}[fill]{\pgfxy(8.5,4.95)}{0.1cm} \pgfnodecircle{Node63}[fill]{\pgfxy(8,5.45)}{0.1cm} \pgfnodeconnline{Node61}{Node63} \pgfnodecircle{Node71}[fill]{\pgfxy(10,5.4)}{0.1cm} \pgfnodecircle{Node72}[fill]{\pgfxy(10.5,5.4)}{0.1cm} \pgfnodecircle{Node73}[fill]{\pgfxy(10.25,4.9)}{0.1cm} \pgfnodeconnline{Node71}{Node73} \pgfnodeconnline{Node72}{Node73} \pgfnodecircle{Node81}[fill]{\pgfxy(12.1,4.8)}{0.1cm} \pgfnodecircle{Node82}[fill]{\pgfxy(12.1,5.6)}{0.1cm} \pgfnodecircle{Node83}[fill]{\pgfxy(12.6,5.2)}{0.1cm} \pgfnodeconnline{Node81}{Node82} \pgfnodeconnline{Node81}{Node83} \pgfnodeconnline{Node83}{Node82} \pgfputat{\pgfxy(4.5,2)}{\pgfbox[center,center]{transitions}} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(4.6,2.3)(5.2,4) \pgfxyline(5.6,2.1)(7.1,2.7) \pgfxyline(5.6,1.8)(7.9,1.5) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.25cm}{1.5cm}} \pgfnodecircle{Node100}[fill]{\pgfxy(14,.3)}{0.1cm} \pgfnodecircle{Node101}[fill]{\pgfxy(14,1)}{0.1cm} \pgfnodecircle{Node102}[fill]{\pgfxy(13.2,1.5)}{0.1cm} \pgfnodecircle{Node103}[fill]{\pgfxy(14.8,1.5)}{0.1cm} \pgfnodecircle{Node104}[fill]{\pgfxy(12.4,2.2)}{0.1cm} \pgfnodecircle{Node105}[fill]{\pgfxy(13.2,2.2)}{0.1cm} \pgfnodecircle{Node106}[fill]{\pgfxy(14,2.2)}{0.1cm} \pgfnodecircle{Node107}[fill]{\pgfxy(14.8,2.2)}{0.1cm} \pgfnodecircle{Node108}[fill]{\pgfxy(15.6,2.2)}{0.1cm} \pgfnodeconnline{Node100}{Node101} \pgfnodeconnline{Node101}{Node102} \pgfnodeconnline{Node101}{Node103} \pgfnodeconnline{Node102}{Node104} \pgfnodeconnline{Node102}{Node105} \pgfnodeconnline{Node102}{Node106} \pgfnodeconnline{Node103}{Node106} \pgfnodeconnline{Node103}{Node107} \pgfnodeconnline{Node103}{Node108} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Portion of Sorkin and Rideout's kinematic scheme $\ms{K}$ describing sequential growth of causal sets; b) abstract structure of this portion of $\ms{K}$. } \label{sequential} \end{figure} \vspace*{-.5cm} The kinematic scheme $\ms{K}$ only describes {\it possible} evolutionary behavior of causal sets. The actual {\it dynamics} of sequential growth is supplied by assigning a {\it weight} or {\it phase} $\theta(\tau)$ to each transition $\tau$ in $\ms{K}$, encoding the {\it ``likelihood that $C_i$ transitions to $C_t$."}\footnotemark\footnotetext{This deliberately vague description is intended to cover a range of possible choices for the values of $\theta$, from real probabilities, to complex amplitudes, to more exotic possibilities.} The ``likelihood" of evolution between any pair of causal sets $C$ and $C'$ in $\ms{K}$ may then be expressed in terms of the weights of all transitions along all upward-directed sequences of transitions from $C$ to $C'$ in $\ms{K}$. In the specific context of classical sequential growth dynamics, Sorkin and Rideout take these weights to be real probabilities, but $\ms{K}$ may also be viewed in a quantum-theoretic context, as in \cite{SorkinQuantalMeasure12}. Each upward-directed sequence of transitions from $C$ to $C'$ in $\ms{K}$ corresponds to a sequential labeling of the elements of $C'-C$. If $C$ is empty, such a sequence defines a {\it total labeling} $\{x_0,x_1,x_2,...\}$ of $C'$, which is equivalent to a {\it linear extension} of the causal order, or alternatively, to a {\it bijective order morphism} of $C'$ into a linear suborder of the nonnegative integers.\footnotemark\footnotetext{This procedure is unambiguous if and only if $\ms{K}$ is viewed as a (higher-level) multidirected set. In particular, a total labeling of an ``initial" finite causal set $C_i$, together with the isomorphism class of the ``next" causal set $C_t$, generally does {\it not} determine a class of labelings of $C_t$ related by automorphisms of $C_t$.} Such a sequence is analogous to a relativistic {\it frame of reference,} supplementing the causal structure of its terminal causal set with arbitrary nonphysical information, just as relativistic coordinates assign arbitrary time orders to pairs of spacelike-separated events. In section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} of this paper, I generalize and formalize such sequences as {\it co-relative kinematics.} {\bf Phenomenology.} Like other approaches to fundamental spacetime structure, causal set theory cannot yet boast a robust record of phenomenological success. However, the difficulty of wringing any definitive and experimentally accessible prediction out of a relatively young fundamental theory makes it reasonable to exercise considerable patience in evaluating the ongoing viability of the causal sets program, given its plausibility and motivation. The most well-known phenomenological ``achievement" of causal set theory to date is an approximate prediction of the size of {\it fluctuations} of the cosmological constant, as discussed in \cite{SorkinEverPresentLambda04}. Assuming fluctuations centered about zero, this prediction falls within the range suggested by observation. However, this result relies on assumptions about the interpretation of volume, the emergent dimensionality of spacetime, and the large-scale structure of the universe, that one would {\it a priori} prefer to avoid. For a fine overview of this and other topics in causal set phenomenology, I refer the reader to Sumati Surya's recent review article {\it Directions in Causal Set Quantum Gravity} \cite{Surya11}. Surya references a wide variety of recent developments, including Dowker, Henson, and Sorkin's work on {\it diffusion in a causal set background} \cite{SorkinLorentz04}, examination of {\it causal set curvature} and analogues of the {\it Einstein-Hilbert action} by Dowker and her collaborators \cite{DowkerScalarCurvature10}, and the construction of a {\it scalar field Feynman propagator} by Johnston \cite{JohnstonFeynman09}. Causal set phenomenology has thus far been directed more toward working out the preliminaries of {\it how causal set theory might make contact with the experimental realm} than toward making specific, falsifiable predictions. Most of the topics cited above involve background-dependent versions of the theory, and are therefore of only limited relevance to this paper. An important basic phenomenological question is how nongravitational matter might arise in causal set theory. Sorkin \cite{SorkinSequentialGrowthDynamics99} advocates {\it ``a kind of ``effective action" for a field of Ising spins living on the relations of the causal set."} In certain cases, such a field may be {\it attributed to the underlying causal structure.} See, for example, the analogous discussion of {\it valence fields} in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} of this paper. \subsection{Purposes; Viewpoint and Style; Intended Audience}\label{subsectionpurpose} Three outstanding problems of modern theoretical physics are: 1) {\it description of fundamental spacetime structure;} 2) {\it development of a suitable theory of quantum gravity;} and 3) {\it unification of physical law.} These problems are, of course, intertwined. Ever since the discovery of general relativity, gravity has been understood to be {\it structural} in nature, and it is difficult to imagine a successful approach to quantum gravity without a deeper structural understanding of spacetime. Quantum theory, whatever its interpretation, is believed by most serious theorists to represent a fundamental aspect of nature, so any successful unification of physical law is expected to be quantum-theoretic. It seems, then, that two common ingredients necessary for the solution of these three problems are {\it a suitable notion of structure} and {\it a suitable notion of quantum theory.} This paper represents an attempt to help identify such notions. {\bf Purposes.} Under the causal metric hypothesis (\hyperref[cmh]{CMH}), the ``suitable notion of structure" is {\it causal structure,} and the question becomes {\it how this structure should be modeled.} The primary purpose of this paper is to address this question, beginning at the classical level. I focus attention on {\it discrete models satisfying a local finiteness condition,} motivated by the common expectation that a finite fundamental scale exists in physics, by technical results such as Malament's theorem, and by the mathematical tractability and interest of such models. Turning to the quantum realm, the histories approach to quantum theory, mentioned above, is a general and flexible candidate for a ``suitable notion of quantum theory," and is also particularly amenable to discrete causal theory. An important secondary purpose of this paper, therefore, is to {\it outline how classical causal structure may be combined with the histories approach to quantum theory.} This leads to the theory of {\it co-relative histories,} described in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} below. The fact that the resulting version of quantum causal theory seems to have attractive properties reinforces the preceding choices. {\bf Viewpoint and Style.} Causal set theory already occupies similar philosophical and theoretical ground, seeking to combine discrete causal structure with the histories approach to quantum theory. It is therefore reasonable to take causal set theory as a starting point, and propose modifications as needed. Other similar theories also play important roles: Finkelstein's {\it causal nets} are very similar, at the level of individual objects, to the {\it directed sets}\footnotemark\footnotetext{The term {\it directed set} has a meaning equivalent to {\it small, simple directed graph} in this paper, which is a more general meaning than it has in the usual order-theoretic settings. See the discussion of nonstandard terminology in section \hyperref[subsectionnotation]{\ref{subsectionnotation}}, and the definition of directed sets in section \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} below.} used in this paper as models of classical spacetime, and Isham's category-theoretic methods supply important aspects of higher-level structure and hierarchy. However, the overall approach developed here is closer, from a physical perspective, to Sorkin's causal set theory, than to any other existing approach. Mathematically, this paper is significantly different from most of the causal set literature, incorporating substantial influences from modern algebra and algebraic geometry. However, these influences are manifested less by direct use of existing algebraic notions, such as those contributing to Isham's approach, than by {\it analogy.} For example, the multidirected structures of principal interest in this paper are not categories, but {\it resemble} categories in useful ways. The length and style of this paper is motivated by the conviction that one cannot expect others to seriously consider an alternative approach to fundamental physics without a substantial account of its underlying motivations and potential applications. Hence, in writing this paper, I felt compelled to venture beyond merely advocating changes to the axioms of causal set theory, at least as far as introducing the rudiments of the resulting quantum causal theory, outlined in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below. Anything less would not have provided sufficient evidence that the ideas presented here actually lead to any significant improvement of existing theory. {\bf Intended Audience.} Every author wishes for a broad audience. While lengthy concessions to accessibility risk exasperating expert readers, the opposite error of excessive terseness is far worse. In writing this paper, I felt that modest expository detours, such as an extra paragraph reminding the reader what a category is, or an extension of the discussion of path summation to include comparison with the continuum analogue, were well justified. My own background as a student of algebraic geometry naturally colors the presentation in a manner most likely to appeal to those with a taste for structure and generality, but hopefully not in an overwhelming fashion. \subsection{Underlying Natural Philosophy}\label{naturalphilosophy} A significant portion of this paper consists of analyzing the {\it physical plausibility} of various axiomatic systems for encoding causal structure. Einstein once remarked that the deep physical principles underlying the natural world cannot be {\it logically deduced,} but must be reached by {\it ``intuition, resting on sympathetic understanding of experience"} \cite{EinsteinEssays34}. In recent years, Einstein has sometimes been cited as a misleading exception to the ``rule" that close grappling with experimental evidence generally yields greater success than intuition. This rule is probably valid if theoretical success is credited solely to the {\it first} satisfactory explanation of phenomena, and if intuition is allowed to operate in a vacuum. However, if secondary explanations, and the {\it ``sympathetic understanding of experience,"} are admitted, then intuition can lay claim to the Lagrangian and Hamiltonian formulations of classical mechanics, the histories approach to quantum theory, and many other seminal ideas besides general relativity. Hence, in this broader context, even intuition has its merits.\footnotemark\footnotetext{The mathematical reader will recall here Gordan's reluctant endorsement of Hilbert's ``theology."} With this in mind, I offer six philosophical principles underlying the intuitive viewpoint of this paper, together with a few contextual remarks. I make no attempt here to be original or definitive. {\bf 1. Physics should seek not to prescribe what may be, but to describe what is.}\footnotemark\footnotetext{Idiotic interpretations of this statement, involving general phobia of assumptions or predictions, are easily avoided.} This distinction is illustrated by comparing general relativity with causal theory. The former theory views spacetime structure as {\it prescribing which pairs of events may be causally related,}\footnotemark\footnotetext{Some, such as Carlo Rovelli \cite{Rovelli04}, might argue that general relativity is merely {\it misinterpreted} along these lines.} while the latter approach views this structure as {\it describing which pairs of events are causally related.} The descriptive philosophy avoids meaningless inconsistencies, such as ``time travel paradoxes," and unjustified assumptions, such as the presumed steady state of the universe before Hubble's observations. The causal set axioms of transitivity (\hyperref[tr]{TR}) and interval finiteness (\hyperref[if]{IF}) are objectionably prescriptive. {\bf 2. Mathematics and physics are distinct; each informs the other.} Mathematical structures in physics should be chosen for their conceptual merits, rather than their familiarity or convenience.\footnotemark\footnotetext{{\it ``Concept over convenience"} recalls Gauss' {\it ``notions over notations."}} This leads not only to good physics, but also to interesting mathematics. Mathematics returns the favor by producing concepts and methods whose physical significance is only appreciated later. Care must be taken to distinguish between mathematical and physical properties. An unfortunate byproduct of the long-standing success of continuum theory in physics has been the automatic and unjustified attribution of {\it mathematical} properties of the continuum, such as the {\it interpolative property,} and {\it Cauchy completeness,} to physical spacetime. The reasoning behind the {\it independence convention} (\hyperref[ic]{IC}), introduced in section \hyperref[subsectionchains]{\ref{subsectionchains}} of this paper, provides an example of proper distinction between mathematical and physical properties; in this case, {\it irreducibility} and {\it independence} of relations between pairs of elements in a directed or multidirected set. {\bf 3. Basic structural concepts are crucial.} It is instructive to consider the poverty of structural alternatives to continuum geometry available to physicists during the early 20th century, when relativity and quantum theory were being developed. Many physically suggestive ideas from fields such as order theory, graph theory, information theory, computer science, category theory, algebra, and algebraic geometry, were not yet known. Even group theory faced a difficult reception.\footnotemark\footnotetext{For example, Schr\"{o}dinger's {\it gruppenpest}, as related by Wigner.} The twenty-first century scientific community is much better equipped, at least in theory, to follow up Einstein's intuition that physics is essentially structural in nature, brilliantly vindicated by general relativity, but unconsummated thereafter. Many of the structural ideas appearing in this paper have roots in modern algebra, particularly in the work of Alexander Grothendieck. Especially important is Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}), discussed further in section \hyperref[settheoretic]{\ref{settheoretic}} below. {\bf 4. Local and global properties must be properly distinguished.} The history of physics is littered with errors resulting from specious local-to-global reasoning and dubious extrapolation across scales. Often such errors arise from failure to recognize the limitations of ``obvious" observations, such as the apparent motionlessness of the earth, or the apparent flatness of spacetime in the vicinity of the earth, or the apparent possibility of assigning definite values of position and momentum simultaneously to macroscopic material bodies. Particularly troublesome are ``local" conditions adopted without recognition of their global consequences. For example, the causal set axiom of interval finiteness (\hyperref[if]{IF}), called {\it local finiteness} in the literature, unjustifiably prescribes {\it global} structure, due to an improper distinction between local and global properties. {\bf 5. The nature of experimentation has a bearing on the nature of theory, not merely on the details of what theory can hope to predict.} Besides attempting to explain specific experimental results, theorists must consider {\it general demands and prohibitions} associated with the experimental method. For example, the {\it directed} relationship between experimental conditions and experimental results demands the application of something very similar to causal theory at our closest interface with nature. The nature of experimentation also favors axiomatizing local rather than global properties, since the latter are often experimentally inaccessible. For example, in the context of continuum geometry, it is reasonable to assume that classical spacetime is four-dimensional, since dimension is defined locally. It is not reasonable, barring unlikely observational scenarios such as ``circles in the sky," to assume that classical spacetime has any particular global topology, although conclusions about global structure could conceivably be {\it derived} on dynamical grounds. {\bf 6. Censor the fatal, not the merely unexpected.} A reasonable facet of theory-building is to impose conditions censoring properties so qualitatively contrary to observation that any theory exhibiting them is immediately discredited. More succinctly, it is reasonable to ``ignore the irrelevant." However, this must be done with great care, due to the limitations of human judgment and imagination. For example, Planck's approach to black-body radiation, eliminating all but a discrete set of emission frequencies, to avoid the fatal ultraviolet catastrophe, is an example of justified censorship. However, Einstein's fixing of the cosmological constant, to achieve his {\it expectation} of a steady-state universe, is not. The causal set axioms of transitivity (\hyperref[tr]{TR}) and interval finiteness (\hyperref[if]{IF}) both censor nonfatal phenomena; the former by ignoring distinctions among certain modes of influence between pairs of events, and the latter by drastically constraining the global structure of classical spacetime. \subsection{Notation and Conventions; Figures}\label{subsectionnotation} {\bf Notation and Conventions.} The mathematical objects of principal interest in this paper are three types of {\it structured sets} called {\it acyclic directed sets,} {\it directed sets} and {\it multidirected sets,} respectively. These are defined in section \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} below. They may be viewed as increasingly broad generalizations of causal sets. In almost all cases, only the {\it isomorphism class} of a structured set is significant, but most constructions are described in terms of particular representatives. Individual structured sets are usually denoted by capital letters such as $A$, $D$ or $M$, and their individual elements by lower-case letters such as $w,x, y,$ and $z$. Whenever a more explicit treatment is necessary, an acyclic directed set or directed set may be represented by an ordered pair $(A,\prec)$ or $(D,\prec)$, where $\prec$ is an appropriate binary relation, and a multidirected set may be represented by an ordered quadruple $(M,R,i,t)$, where $R$ is a set of relations, and where $i$ and $t$ are {\it initial} and {\it terminal element maps,} described in more detail in section \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}}. The {\it precursor symbol} $\prec$ is used for three distinct but related purposes: first, to represent a {\it binary relation} on a set, as mentioned above; second, to represent an {\it individual relation} between two elements of a set; third, to indicate that {\it at least one such relation exists.} For an acyclic directed set or directed set, an individual relation $x\prec y$ is an {\it element} of the binary relation $\prec$. For a multidirected set, there may exist multiple relations between a given pair of elements in either or both directions. In this case, the expression $x\prec y$ either refers to a particular relation with initial element $x$ and terminal element $y$, or indicates that at least one such relation exists. Categories, and other structures filling similar hierarchical roles, such as kinematic schemes, are denoted by mathscript letters such as $\ms{C}$, for the category of causal sets, or a general category; $\ms{A}$, for the category of acyclic directed sets; $\ms{D}$, for the category of directed sets;, or $\ms{M}$, for the category of multidirected sets. Functors, and other similar ``higher-level maps," are assigned notation {\it suggestively,} rather than in any systematic fashion; for example, the {\it transitive closure} functor on the category of directed sets is denoted by $\tn{tr}$, while the {\it relation space} functor on the category of multidirected sets is denoted by $\ms{R}$. Familiar number systems, such as the nonnegative integers, integers, rationals, reals, and complex numbers, are denoted by the usual blackboard bold symbols $\mathbb{N}, \mathbb{Z}, \mathbb{Q}, \mathbb{R}, \mathbb{C}$. I use ``nonnegative integers," rather than ``natural numbers," to avoid confusion about the status of zero. The additive group of integers modulo a positive integer $n$ is denoted by $\mathbb{Z}_n$. The term {\it space} usually means {\it structured set;} the structure involved is often {\it a priori} order-theoretic, or more generally, multidirected, rather than topological, though this structure may be used to {\it define} natural topologies. Appendix \hyperref[appendixindex]{A} provides a thorough index of notation. {\bf Nonstandard Terminology.} This paper contains notable disagreements with standard mathematical terminology. Most of these disagreements arise because the contexts in which certain existing structures and methods appear in this paper are much different than their original contexts, thereby demanding different viewpoints. In some cases, existing terminology is actually ``wrong," in the sense that it contradicts ``sacred notions." For example, the condition of interval finiteness (\hyperref[if]{IF}), called {\it local finiteness} in the causal set literature, {\it is not a local condition,} and placing it in such a role leads to serious problems. In other cases, terms whose plain sense evokes a useful general idea have acquired inconvenient proprietary meanings. For example, the term {\it directed set} conventionally signifies a set equipped with a reflexive, transitive binary relation, satisfying a common successor property. Since the alternative term {\it filtered set} has the same conventional meaning, but conveys it better, I reclaim the term {\it directed set} to mean simply a set equipped with a binary relation. Similarly, I use the term (causal) {\it preorder} to mean simply a binary relation generating a causal order under the operation of transitive closure, even though the term {\it preorder} conventionally signifies reflexivity and transitivity. The alternative term {\it quasiorder} has the same conventional meaning, and conveys it just as well. {\bf Highlighting and Hyperlinks.} Important technical terms, headings of definitions, axioms, lemmas, theorems, and important arguments and principles, are introduced using {\bf bold font}. Most definitions are embedded in the text, but particularly important definitions are formally set off from the text and numbered. {\it Italic font} is used for the definition or introduction of other technical terms, bodies of definitions, axioms, lemmas, theorems, and quotations, non-English terms and phrases, and ordinary English emphasis. It is also used with discretion to emphasize important technical terms referenced {\it before or after} their bold-font definitions. Hyperlinks are included throughout the text to enable easy navigation. Axioms, conventions, and principles of sufficient importance to demand frequent reference are labeled by short acronyms. The axiom of {\it transitivity} (\hyperref[tr]{TR}) and the {\it order extension principle} (\hyperref[oep]{OEP}) are examples. These acronyms serve as tags for hyperlinks to the original definitions. Their function is opposite to the common usage of acronyms as ``labor-saving devices," to the exasperation of the reader. {\bf Illustrative Figures.} Though many of the constructions in this paper apply to general multidirected sets, most of the accompanying figures, such as figure \hyperref[sprinkling]{\ref{sprinkling}}c in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, depict plane diagrams of acyclic directed sets, called {\bf generalized Hasse diagrams,} analogous to Minkowski diagrams of relativistic spacetime. In a generalized Hasse diagram, elements of an acyclic directed set are represented by nodes, and relations between pairs of elements are represented by edges between pairs of nodes. The directions of relations are inferred via the vertical coordinates assigned to their initial and terminal nodes. More specifically, if $x$ and $y$ are two elements of an acyclic directed set $(A,\prec)$, then by convention, $x\prec y$ in $A$ if and only if there is an edge between the corresponding nodes in a generalized Hasse diagram for $A$, such that the vertical coordinate of the node corresponding to $x$ is less than the vertical coordinate of the node corresponding to $y$. Pairs of elements whose corresponding nodes have no edge between them are assumed to be unrelated, regardless of the coordinates of these nodes. Heuristically, {\it causal influence flows upward along edges.} The qualifier {\it generalized} is included because ``standard" Hasse diagrams represent only {\it irreducible} relations, in the sense of part 3 of definition \hyperref[definitionchains]{\ref{definitionchains}} below, while the diagrams appearing in this paper also admit reducible relations. {\it Acyclic} multidirected sets may be represented by generalized Hasse diagrams in an analogous way, but directed sets and multidirected sets involving cycles cannot be so represented. Regardless of the actual properties of the structured sets appearing in a figure, these sets are usually labeled and referenced at the same level of generality at which the corresponding constructions or examples are intended to apply. For example, an acyclic directed set appearing in a figure may be labeled $M$, rather than $A$, if the figure is intended to apply to multidirected sets in general. Conversely, if the purpose of the figure is to demonstrate that a certain property manifests itself even in the special case of acyclic directed sets, then the set appearing in the figure may be labeled $A$, rather than $M$. Figures are numbered separately from definitions, lemmas and theorems, since they are auxiliary to the logical development of the paper. \subsection{Summary of Contents; Outline of Sections}\label{subsectionoutline} {\bf Summary of Contents.} This paper has seven sections and one appendix. Section \hyperref[sectionintroduction]{\ref{sectionintroduction}} is the present introduction. Section \hyperref[sectionaxioms]{\ref{sectionaxioms}} presents the axioms of causal set theory, and provides necessary terminology and definitions. Sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} are devoted to careful analysis of the causal set axioms of transitivity (\hyperref[tr]{TR}), interval finiteness (\hyperref[if]{IF}), and the binary axiom (\hyperref[b]{B}). The axiom of irreflexivity (\hyperref[ir]{IR}) is involved indirectly, since it interacts with transitivity to imply the condition of acyclicity (\hyperref[ac]{AC}). The principal conclusions of these sections are the following: \begin{itemize} \item The transitive causal orders of causal set theory are information-theoretically deficient as discrete models of classical spacetime. Generally nontransitive binary relations called {\it causal preorders} should be viewed as more fundamental. In this context, causal cycles must be ruled out explicitly if they are considered undesirable. Hence, for conservative models of classical spacetime, in which ``closed timelike curves" are forbidden, transitivity and irreflexivity should be replaced with the single axiom of acyclicity. For more general models, allowing ``closed timelike curves," transitivity and irreflexivity should be simply discarded without replacement. In either case, causal sets should be replaced with a more general class of directed sets. \item Interval finiteness suffers from multiple pathologies. It does not rule out locally infinite behavior, and imposes unjustified restrictions on the global structure of classical spacetime. For these reasons, interval finiteness should be replaced with the alternative axiom of {\it local finiteness} (\hyperref[lf]{LF}). This replacement has important consequences both for the directed sets used to model classical spacetime, and for the multidirected sets arising in quantum causal theory. \item The useful practice of encoding spacetime events as elements of a directed set, formalized by the binary axiom (\hyperref[b]{B}), may be generalized to take advantage of natural directed and multidirected structures whose elements correspond to more complex physical entities than individual spacetime events, such as causal relations between pairs of events, or sequences of such relations. For example, viewing relations between elements of a multidirected set $M$ as elements of a directed set $\ms{R}(M)$, called its {\it relation space,} enables a path-summation approach to quantum causal theory with no direct analogue in ordinary element space. \end{itemize} Section 6 demonstrates how the developments of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} may be applied to the quantum theory of spacetime. Combining Grothendieck's relative viewpoint (\hyperref[rv]{RV}) with the histories approach to quantum theory leads to the theory of {\it co-relative histories} and {\it kinematic schemes.} Applying the theory of relation space, developed in section \hyperref[sectionbinary]{\ref{sectionbinary}}, leads to the derivation of {\it causal Schr\"{o}dinger-type equations,} which describe the dynamics of quantum spacetime in the discrete causal setting. An example of such an equation is the following: \[\psi_{R;\theta}^-(r)=\theta(r)\sum_{r^-\prec r}\psi_{R;\theta}^-(r^-).\] The notation appearing in this equation is explained in sections \hyperref[sectionbinary]{\ref{sectionbinary}} and \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}. Briefly, $\psi_{R;\theta}^-$ is a causal analogue of Schr\"{o}dinger's wave function, $R$ is a subset of relation space analogous to a region of spacetime, $\theta$ is a {\it phase map} analogous to a Lagrangian, and $r$ and $r^-$ are relations. Section \hyperref[sectionconclusions]{\ref{sectionconclusions}} gathers together the conclusions of preceding sections, and proposes specific changes to the axioms of causal set theory. This section also includes brief discussions of omitted material, outlines directions for future research, and describes potential connections with other areas of physics and mathematics. Appendix \hyperref[appendixindex]{A} provides a detailed index of notation. {\bf Topical Outline.} Following is a hyperlinked topical outline of sections 2-7: {\bf Section \hyperref[sectionaxioms]{\ref{sectionaxioms}}: Axioms and Definitions.} {\bf \hyperref[subsectioncausalmetric]{\ref{subsectioncausalmetric}}: }{\bf The Causal Metric Hypothesis.} \begin{itemize} \item \hyperref[cmh]{Causal metric hypothesis (CMH): {\it ``The properties of the physical universe are manifestations of}} \hyperref[cmh]{{\it causal structure."} Combines ``physics is structural" with familiar building block for structure.} \item \hyperref[ccmh]{Classical causal metric hypothesis (CCMH): {\it ``Classical spacetime arises from the structure of a directed set."}} \item \hyperref[strongform]{Scope of the CMH; e.g., strong form: {\it ``All physical phenomena may be explained in terms of}} \hyperref[strongform]{{\it causal structure."}} \hyperref[strongform]{Too ambitious?} \item \hyperref[technicalimplementation]{Technical implementations of the CMH.} \item \hyperref[quantumcmh]{Quantum causal metric hypothesis (QCMH). Revisited in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}}.} \end{itemize} {\bf \hyperref[subsectionaxioms]{\ref{subsectionaxioms}}: }{\bf The Axioms of Causal Set Theory.} \begin{itemize} \item \hyperref[subsectionaxioms]{Two formulations: irreflexive formulation and partial order formulation.} \item \hyperref[irreflexive]{Irreflexive formulation: axioms.} \item \hyperref[defcausalset]{Formal definition of a causal set.} \item \hyperref[partialorder]{Partial order formulation: disadvantages.} \item \hyperref[axiomsinliterature]{Treatment in the literature.} \end{itemize} {\bf \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}}: }{\bf Acyclic Directed Sets; Directed Sets; Multidirected Sets.} \begin{itemize} \item \hyperref[subsectionacyclicdirected]{Generalizations of causal sets.} \item \hyperref[acyclicdirected]{Acyclic directed sets: conservative models of classical spacetime.} \item \hyperref[graph]{Graph-theoretic viewpoint.} \item \hyperref[directedsets]{Directed sets: can model causal analogues of closed timelike curves.} \item \hyperref[multidirectedsets]{Multidirected sets: necessary for quantum theory.} \end{itemize} {\bf \hyperref[subsectionchains]{\ref{subsectionchains}}: }{\bf Chains; Antichains; Irreducibility; Independence.} \begin{itemize} \item \hyperref[chains]{Chains; antichains; mathematical properties such as reducibility and irreducibility.} \item \hyperref[conditionsmulti]{Conditions on multidirected sets: generalizing transitivity, interval finiteness, etc.} \item \hyperref[independence]{Independence: a physical property.} \item \hyperref[ic]{Independence convention (IC): {\it ``Relations are independent unless stated otherwise."}} \item \hyperref[interpolativecomparison]{Comparison to interpolative case; e.g., domain theory.} \end{itemize} {\bf \hyperref[subsecpredecessors]{\ref{subsecpredecessors}}: }{\bf Domains of Influence; Predecessors and Successors; Boundary and Interior.} \begin{itemize} \item \hyperref[relpastfuture]{Relativistic notions; e.g., light cone; chronological and causal pasts and futures.} \item \hyperref[newbehavior]{New behavior in causal theory due to irreducibility and independence.} \item \hyperref[domains]{Predecessors and successors; domains of influence in multidirected sets.} \item \hyperref[predecessors]{Direct pasts and futures; maximal past; minimal future.} \item \hyperref[boundaryint]{Boundary and interior of a multidirected set.} \end{itemize} {\bf \hyperref[settheoretic]{\ref{settheoretic}}: }{\bf Order Theory; Category Theory; Influence of Grothendieck.} \begin{itemize} \item \hyperref[isomclassesfinite]{Counting (isomorphism classes of) finite acyclic directed sets.} \item \hyperref[isomclassescountable]{Counting (isomorphism classes of) countable acyclic directed sets.} \item \hyperref[orderext]{Linear orders; order extensions.} \item \hyperref[oep]{Order extension principle (OEP): {\it ``Every acyclic binary relation extends to a linear order."}} \item \hyperref[cat]{Category theory. Two roles: framework and analogy.} \item \hyperref[ishamtopos]{Isham's topos theory;} \hyperref[differentnotions]{different notions suggested by causal theory.} \item \hyperref[groth]{Influence of Grothendieck: relative viewpoint (RV); hidden structure (HS).} \end{itemize} \vspace*{.5cm} {\bf Section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}: Transitivity, Independence, and the Causal Preorder.} {\bf \hyperref[subsectionmodes]{\ref{subsectionmodes}}: }{\bf Independent Modes of Influence.} \begin{itemize} \item \hyperref[transproblem]{Basic problem: transitive relations fail to distinguish among independent modes of influence.} \item \hyperref[infoanalogies]{Na\"{i}ve analogies from information theory.} \item \hyperref[finkelstein]{Nontransitive relations in the literature: Finkelstein's causal nets.} \item \hyperref[raptisquant]{Raptis' algebraic quantization of causal sets. } \item \hyperref[sorkincurrent]{Sorkin's current view.} \item \hyperref[independencefundamental]{Independence at the fundamental scale.} \end{itemize} {\bf \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}}: }{\bf Six Arguments that Transitive Binary Relations are Deficient.} \begin{itemize} \item 1) \hyperref[cauchyarg]{Cauchy surface analogy.} \item 2) \hyperref[futurearg]{Elements need not ``know their futures."} \item 3) \hyperref[irredindarg]{Irreducibility and independence are distinct.} \item 4) \hyperref[configpath]{Configuration space pathologies (Kleitman-Rothschild).} \item 5) \hyperref[commcat]{Analogy with commutativity in a category.} \item 6) \hyperref[otherimprovements]{Nontransitivity facilitates other improvements; e.g. local finiteness.} \end{itemize} {\bf \hyperref[subsectionpreorder]{\ref{subsectionpreorder}}: }{\bf The Causal Preorder.} \begin{itemize} \item \hyperref[nontransgen]{Nontransitive binary relation generating the causal order (both may have cycles).} \item \hyperref[finkelsteinconnect]{Similar to Finkelstein's ``causal connection" relation.} \item \hyperref[notusualpreorder]{Not a preorder in usual sense.} \end{itemize} {\bf \hyperref[subsectiontransitiveclosure]{\ref{subsectiontransitiveclosure}}: }{\bf Transitive Closure; Skeleton; Degeneracy; Functoriality.} \begin{itemize} \item \hyperref[transclosure]{Two operations: transitive closure adds reducible relations;} \hyperref[skeleton]{skeleton deletes reducible relations.} \item \hyperref[skelnottransred]{Skeleton distinct from transitive reduction.} \item \hyperref[causordtranspreorder]{Causal order is transitive closure of causal preorder.} \item \hyperref[degenconfig]{Degeneracy; configuration space implications.} \item \hyperref[transfunctorial]{Functorial properties: transitive closure a functor, adjoint to inclusion; skeleton not a functor.} \end{itemize} \vspace*{.5cm} {\bf Section \hyperref[sectioninterval]{\ref{sectioninterval}}: Interval Finiteness versus Local Finiteness.} {\bf \hyperref[subsectiontopology]{\ref{subsectiontopology}}: }{\bf Local Conditions; Topology.} \begin{itemize} \item \hyperref[localcond]{Local conditions: detectable in arbitrarily small neighborhoods.} \item \hyperref[relnotionstop]{Relativistic notions: topology and the metric.} \item \hyperref[causallocality]{Causal locality.} \item \hyperref[gentopology]{General topology;} \hyperref[hiddentop]{topological hidden structure: ``extra elements;"} \hyperref[toplocfin]{topological local finiteness.} \item \hyperref[fourtopspaces]{Four topologies:} \hyperref[discretetop]{discrete topology;} \hyperref[inttop]{interval topology;} \hyperref[conttop]{continuum topology;} \hyperref[startop]{star topology.} \item \hyperref[cardvalscalar]{Cardinal-valued scalar fields.} \end{itemize} {\bf \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}: }{\bf Interval Finiteness versus Local Finiteness.} \begin{itemize} \item \hyperref[intcriticisms]{Initial criticisms of interval finiteness.} \item \hyperref[lf]{Local finiteness (LF).} \item \hyperref[locfintoplocfinstar]{Lemma 4.2.1: {\it ``Local finiteness coincides with topological local finiteness in the star topology."}} \item \hyperref[varlocfin]{Variations on local finiteness for directed sets:} \hyperref[lft]{Local finiteness in the transitive closure (LFT).} \hyperref[lfs]{Local finiteness in the skeleton (LFS).} \item \hyperref[intfinlocfinincomp]{Interval finiteness and local finiteness are incomparable.} \item \hyperref[interactiontrans]{Interaction with transitivity;} \hyperref[interactionmeasure]{interaction with the measure axiom.} \end{itemize} {\bf \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}: }{\bf Relative Multidirected Sets over a Fixed Base.} \begin{itemize} \item \hyperref[infintmeaning]{Infinite intervals: physically meaningful?} \item \hyperref[causetrelz]{Causal sets as relative directed sets over $\mathbb{Z}$.} \item \hyperref[theoremcausalsetrelint]{Theorem 4.3.1: {\it ``Any countable, interval finite acyclic directed set is relative over $\mathbb{Z}$."}} \item \hyperref[relconseq]{Connection to sequential growth dynamics.} \item \hyperref[relacdirarb]{Relative multidirected sets over an arbitrary base.} \item \hyperref[democracy]{Democracy of bases: why restrict to relative directed sets over $\mathbb{Z}$?} \item \hyperref[infkleitroth]{Temporal versus spatial size: ``countably infinite Kleitman-Rothschild-type pathology."} \end{itemize} {\bf \hyperref[subsectionintervalfinitenessdeficient]{\ref{subsectionintervalfinitenessdeficient}}: }{\bf Eight Arguments against Interval Finiteness and Similar Conditions.} \begin{itemize} \item 1) \hyperref[intfinnonloc]{Interval finiteness is causally nonlocal.} \item 2) \hyperref[nonlocininttop]{Interval finiteness is nonlocal in the interval topology.} \item 3) \hyperref[toplocfinintnofinnbhd]{Topological local finiteness in the interval topology does not imply the existence of} \\\hspace*{.45cm}\hyperref[toplocfinintnofinnbhd]{finite interval neighborhoods.} \item 4) \hyperref[intnotcapture]{Open intervals fail to capture local multidirected structure.} \item 5) \hyperref[intnotimplyloc]{Interval finiteness does not imply local finiteness.} \item 6) \hyperref[locnotimplyint]{It does not follow from local finiteness.} \item 7) \hyperref[fatalsorkin]{It permits fatal local behavior under Sorkin's version of the CMH.} \item 8) \hyperref[unjustglobal]{It imposes unjustified global restrictions on classical spacetime structure.} \end{itemize} {\bf \hyperref[subsectionarglocfin]{\ref{subsectionarglocfin}}: }{\bf Six Arguments for Local Finiteness.} \begin{itemize} \item 1) \hyperref[locfincausloc]{It is causally local.} \item 2) \hyperref[locfintoploc]{It is topologically local.} \item 3) \hyperref[locfincaptures]{It captures and isolates local multidirected structure.} \item 4) \hyperref[compatsorkin]{It is compatible with Sorkin's version of the CMH.} \item 5) \hyperref[notunjustglobal]{It does not impose unjustified global restrictions.} \item 6) \hyperref[naturalnontrans]{It is natural in the nontransitive setting.} \end{itemize} {\bf \hyperref[subsectionhierarchyfiniteness]{\ref{subsectionhierarchyfiniteness}}: }{\bf Hierarchy of Finiteness Conditions.} \begin{itemize} \item \hyperref[eltfin]{Other finiteness conditions:} \hyperref[eltfin]{element finiteness (EF);} \hyperref[loceltfin]{local element finiteness (LEF);} \\ \hyperref[relfin]{relation finiteness (RF);} \hyperref[parrelfin]{pairwise relation finiteness (PRF);} \hyperref[chainfin]{chain finiteness (CF);} \\\hyperref[antichainfin]{antichain finiteness (AF).} \item \hyperref[discussfincond]{Discussion regarding finiteness conditions for classical spacetime models.} \item \hyperref[theoremhierarchyfiniteness]{Theorem 4.6.1: logical implications among finiteness conditions for acyclic directed sets.} \item \hyperref[theoremhierarchyfinitenessmulti]{Theorem 4.6.2: logical implications among finiteness conditions for multidirected sets.} \end{itemize} \vspace*{.5cm} {\bf Section \hyperref[sectionbinary]{\ref{sectionbinary}}: The Binary Axiom: Events versus Elements.} {\bf \hyperref[subsectionrelation]{\ref{subsectionrelation}}: }{\bf Relation Space over a Multidirected Set.} \begin{itemize} \item \hyperref[spacesinphys]{Importance of ``spaces" other than ordinary spacetime in physics.} \item \hyperref[defirelationspace]{Definition of relation space.} \item \hyperref[thmrelspacefunctor]{Theorem 5.1.2: {\it ``Passage to relation space is a functor $\ms{R}$ with special properties.}} \item \hyperref[acyclicinterpolativediscussion]{Discussion of acyclic directed and interpolative cases.} \item \hyperref[abeltspace]{Inverse problem: abstract element space.} \item \hyperref[theoremelementfunctor]{Theorem 5.1.4: {\it ``Passage to abstract element space is a functor $\ms{E}$."}} \item \hyperref[theoreminterior]{Theorem 5.1.5: {\it ``$\ms{R}$ preserves information except at the boundary."}} \item \hyperref[cauchyrel]{Cauchy surfaces: impermeability.} \hyperref[eltspaceperm]{Maximal antichains of elements are generally permeable.} \item \hyperref[rideoutperm]{The literature on permeability: Major, Rideout, Surya.} \item \hyperref[theoremrelimpermeable]{Theorem 5.1.7: {\it ``Maximal antichains in relation space are impermeable."}} \item\hyperref[preferrelspace]{Preferability of relation space for spatial notions.} \item \hyperref[analogymorphism]{Analogy between $\ms{R}(M)$ and morphism categories.} \item \hyperref[starrevisited]{The star model revisited.} \item \hyperref[inducedtwoelt]{Induced binary relation on two-element subsets.} \end{itemize} {\bf \hyperref[subsectionpowerset]{\ref{subsectionpowerset}}: }{\bf Power Spaces.} \begin{itemize} \item \hyperref[powerspaces]{Built from ``relations between subsets of arbitrary size."} \item \hyperref[inducedpowerspaces]{Induced power spaces;} \hyperref[holisticpowerspaces]{holistic power spaces.} \item \hyperref[higherinduced]{Higher induced relations;} \hyperref[splicerel]{splice relations.} \item \hyperref[causalatoms]{Causal atoms;} \hyperref[atomictop]{atomic topologies;} \hyperref[causalatomicdec]{causal atomic decomposition;} \hyperref[causalatomicres]{causal atomic resolution.} \item \hyperref[classicalhol]{Top-down causation, classical holism;} \hyperref[degreeshol]{Degrees of holism.} \item \hyperref[twistor]{Analogies with twistor theory;} \hyperref[shapedynamics]{shape dynamics.} \end{itemize} \newpage {\bf \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}: }{\bf Causal Path Spaces.} \begin{itemize} \item \hyperref[causalpaths]{Causal paths.} \item \hyperref[pathspacephys]{Path spaces in mathematics and physics.} \item \hyperref[pathmorph]{Paths as morphisms.} \hyperref[linsource]{Linear directed sets: sources of paths.} \item \hyperref[pathsets]{Causal path sets;} \hyperref[pathprod]{products of paths; causal path spaces.} \item \hyperref[concatprod]{Concatenation product for directed sets;} \hyperref[dirprod]{directed product for multidirected sets;} \\\hyperref[spliceprod]{splice products: more general.} \item \hyperref[catsemicat]{Categories and semicategories;} \hyperref[causalpathsemicat]{causal path semicategories.} \item \hyperref[causalpathalg]{Causal path algebras;} \hyperref[conventionalpath]{conventional applications of path algebras;} \hyperref[raptisincidence]{Raptis' incidence algebra.} \end{itemize} {\bf \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}}: }{\bf Path Summation over a Multidirected Set.} \begin{itemize} \item \hyperref[pathfunctcont]{Path functionals; motivation from continuum theory;} \hyperref[lagrangehamilton]{Lagrangian and Hamiltonian;} \\\hyperref[lagrangehamilton]{Hamilton's principle.} \item \hyperref[pathfunctac]{Path functionals for multidirected sets;} \hyperref[pathsumac]{path summation over a multidirected set.} \end{itemize} \vspace*{.5cm} {\bf Section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}: Quantum Causal Theory.} {\bf \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}}: }{\bf Quantum Preliminaries; Iteration of Structure; Co-Relative Histories.} \begin{itemize} \item \hyperref[superposition]{Superposition; path integration; histories approach to quantum theory in general.} \item \hyperref[historiesquantumcausal]{Histories approach to quantum causal theory;} \hyperref[backgrounddependentqct]{background dependent quantum causal theory;} \\\hyperref[backgroundindependentqct]{background independent quantum causal theory.} \item \hyperref[ishamquantcat]{The literature:} \hyperref[ishamquantcat]{Isham's quantization on a category;} \hyperref[sorkinquantal]{Sorkin's quantum measure theory.} \item \hyperref[iteration]{Iteration of structure (IS): {\it ``Multidirected sets whose elements are directed sets."}} \item \hyperref[transitions]{Transitions.} \hyperref[benderrobinson]{Generally too specific to be fundamental:} \hyperref[benderrobinson]{Bender and Robinson's rigidity result.} \item \hyperref[symmetry]{Symmetry considerations;} \hyperref[galois]{causal Galois groups.} \item \hyperref[labelcorelative]{Co-relative histories: induce higher level multidirected structure.} \hyperref[corelmulti]{McKay's example.} \item \hyperref[deficorelative]{Formal definition of co-relative histories;} \hyperref[corelativesubtle]{category-theoretic subtleties.} \end{itemize} {\bf \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}: }{\bf Abstract Quantum Causal Theory via Path Summation.} \begin{itemize} \item \hyperref[adaptinghistories]{Adapting the histories approach;} \hyperref[backgrounddepadapt]{background-dependent adaptations;} \\ \hyperref[backgroundindepadapt]{ingredients for background independent approach.} \item \hyperref[appliestoboth]{Path summation principle (PS): {\it ``The same abstract theory applies to both background-dependent}} \hyperref[appliestoboth]{{\it and background-independent cases."}} \item \hyperref[contpathintegral]{Feynman's continuum path integral;} \hyperref[feynpost]{Feynman's postulates;} \hyperref[feynamp]{Feynman's quantum amplitude.} \item \hyperref[causalpathint]{Causal analogues of Feynman's path integral;} \hyperref[dependsimperm]{importance of impermeability;} \\\hyperref[abstractamp]{abstract quantum amplitudes.} \end{itemize} {\bf \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}}: }{\bf Schr\"{o}dinger-Type Equations.} \begin{itemize} \item \hyperref[schrodhistory]{Schr\"{o}dinger-type equations via the histories approach in general.} \item \hyperref[contwavefunc]{Continuum wave functions;} \hyperref[conschrodequ]{continuum Schr\"{o}dinger equation via path integrals.} \item \hyperref[abstractpathfunct]{Abstract chain functionals;} \hyperref[abstractwavefunct]{abstract wave functions.} \item \hyperref[causalschrodequ]{Causal Schr\"{o}dinger-type equations;} \hyperref[causalfeynamp]{causal analogues of Feynman's inner product formulas.} \end{itemize} {\bf \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}}: }{\bf Kinematic Schemes.} \begin{itemize} \item \hyperref[pathsumbackground]{Path summation in the background independent context.} \item \hyperref[kinversusdyn]{Kinematics versus dynamics;} \hyperref[kinpreschemes]{kinematic preschemes;} \\\hyperref[underlyingdirclass]{Underlying directed sets and multidirected sets.} \item \hyperref[kinschemes]{Kinematic schemes;} \hyperref[hereditary]{hereditary property (H);} \hyperref[weakaccessibility]{weak accessibility (WA).} \item \hyperref[quantumcausalmetric]{Quantum causal metric hypothesis (QCMH): {\it ``The properties of quantum spacetime arise from a kinematic scheme of directed sets."}} \item\hyperref[pathsumkinscheme]{Path summation over a kinematic scheme;} \hyperref[corelkin]{co-relative kinematics.} \item \hyperref[possequkinscheme]{Positive sequential kinematic scheme;} \hyperref[relsorkinrideout]{comparison to Sorkin and Rideout's kinematic scheme} \hyperref[relsorkinrideout]{describing sequential growth of causal sets.} \item \hyperref[genkinematics]{Generational kinematics: analogous to relativistic frames of reference.} \item \hyperref[completions]{Completions of kinematic schemes;} \hyperref[kinschemecountablyinf]{kinematic schemes of countable acyclic directed sets.} \item \hyperref[morphkinscheme]{``Functors" of kinematic schemes;} \hyperref[numberanalogies]{analogies with familiar number systems.} \item \hyperref[universalkinschemes]{Universal kinematic schemes; kinematic spaces.} \end{itemize} \vspace*{.5cm} {\bf Section \hyperref[sectionconclusions]{\ref{sectionconclusions}}: Conclusions.} {\bf \hyperref[subsectionalternative]{\ref{subsectionalternative}}: }{\bf New Axioms, Perspectives, and Technical Methods.} \begin{itemize} \item \hyperref[axiomaticanalysis]{Summary of axiomatic analysis.} \item \hyperref[suggestedalt]{Suggested alternative axioms;} \hyperref[conservativealt]{conservative alternative;} \hyperref[radicalalt]{radical alternatives.} \item \hyperref[summaryperspectives]{Summary of new perspectives and technical methods.} \end{itemize} {\bf \hyperref[subsectionomitted]{\ref{subsectionomitted}}: }{\bf Omitted Topics and Future Research Directions.} \begin{itemize} \item \hyperref[remainingax]{Further remarks on countability, irreflexivity, and acyclicity.} \item \hyperref[covariance]{Covariance.} \item \hyperref[algstructure]{Algebraic structure and hierarchy.} \item \hyperref[phasetheory]{Phase theory: what is the target object of the phase map?} \item \hyperref[randomgraph]{Random graph dynamics; phase transitions.} \item \hyperref[holalt]{Alternatives to power spaces.} \item \hyperref[othertheories]{Connections with other physical theories;} \hyperref[othermath]{connections with other mathematical topics.} \end{itemize} {\bf \hyperref[subsectionacknowledgements]{\ref{subsectionacknowledgements}}: }{\bf Acknowledgements; Personal Notes.} \begin{itemize} \item \hyperref[ack]{Acknowledgements.} \item \hyperref[personal]{Personal notes.} \end{itemize} \newpage \section{Axioms and Definitions}\label{sectionaxioms} Causal set theory is based on Rafael Sorkin's version of the {\it causal metric hypothesis} (\hyperref[cmh]{CMH}), summarized by the phrase, {\it ``order plus number equals geometry."} The general philosophy of the causal metric hypothesis is that the observed properties of the physical universe, including the structure of spacetime, and the dynamical behavior of matter and energy, are ultimately just manifestations of cause and effect. The axioms of causal set theory represent one of many possible ways to distill from this broad idea a specific quantitative approach to fundamental physics. In this section, I introduce these axioms, in a sufficiently general context to support their analysis in succeeding sections. I also define other important structures and methods used throughout the remainder of this paper. In section \hyperref[subsectioncausalmetric]{\ref{subsectioncausalmetric}} below, I explain the general reasoning behind the causal metric hypothesis, and place Sorkin's specific version in a broader context. In section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}}, I present the axioms of causal set theory, using the {\it irreflexive formulation,} and explain why this formulation is structurally preferable to the alternative {\it partial order formulation.} In section \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}}, I introduce three increasingly general analogues of causal sets, which I call {\it acyclic directed sets,} {\it directed sets,} and {\it multidirected sets.} Acyclic directed sets are equivalent to {\it small, simple, acyclic directed graphs,} but are viewed in order-theoretic, rather than graph-theoretic, terms. They serve as ``conservative" models of classical spacetime structure. Directed sets, not necessarily acyclic, appear in the study of more general models of classical spacetime, analogous to solutions of general relativity admitting closed timelike curves. They are equivalent to {\it small, simple directed graphs.} The reader should be aware that this is a more general meaning than the standard one. Multidirected sets arise naturally in the theory of {\it configuration spaces of directed sets,} and play an important role in the version of quantum causal theory developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper. They are equivalent to {\it small directed multigraphs,} or equivalently, {\it small quivers.} In section \hyperref[subsectionchains]{\ref{subsectionchains}}, I introduce {\it chains} and {\it antichains} in directed sets and multidirected sets. I also discuss distinctions among mathematical and physical properties associated with chains, and more specifically, with individual relations between pairs of elements. Particularly important is the distinction between {\it irreducibility,} a mathematical property, and {\it independence,} a physical property. I state the {\it independence convention} (\hyperref[ic]{IC}), which is crucial to the interpretation of directed sets and multidirected sets throughout this paper. In section \hyperref[subsecpredecessors]{\ref{subsecpredecessors}}, I discuss {\it domains of influence}, which are generalizations of chronological and causal pasts and futures in general relativity. They are distinct from the {\it domains} appearing in the related field of domain theory. In particular, I discuss {\it direct pasts and futures,} {\it maximal pasts,} and {\it minimal futures,} which are vital for analyzing the axioms of transitivity (\hyperref[tr]{TR}) and interval finiteness (\hyperref[if]{IF}) in sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}} below. In section \hyperref[settheoretic]{\ref{settheoretic}}, I introduce general structural ideas from order theory and category theory, such as the {\it order extension principle} (\hyperref[oep]{OEP}), and Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}). \subsection{The Causal Metric Hypothesis}\label{subsectioncausalmetric} {\bf Background and Definition.} Causal theory is founded on a single fundamental idea called the {\it causal metric hypothesis}. The special case of causal set theory represents one possible way of formalizing Sorkin's specific version of this hypothesis. The philosophical content of the causal metric hypothesis is that the observed properties of the physical universe arise from causal relationships between pairs of events, or more generally, between pairs of families of events.\footnotemark\footnotetext{This caveat is included to allow for the possibility of {\it classical holism,} though I focus on classically reductionist models in this paper. See section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} for further discussion.} \hspace*{.3cm} CMH.\refstepcounter{textlabels}\label{cmh} {\bf Causal Metric Hypothesis.} {\it The properties of the physical universe are manifestations \\ \hspace*{1.45cm} of causal structure.} The causal metric hypothesis fleshes out the longstanding idea, going back to Gauss, Riemann, Einstein, Kaluza and Klein, Weyl, and many others, that physics is {\it essentially structural in nature,} by proposing the familiar relationship between cause and effect as the fundamental building block of this structure. Focusing on the classical case, the principal philosophical difference between the causal metric hypothesis and the viewpoint of general relativity is one of {\it description versus prescription,} as described in section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}}. Whereas general relativity treats spacetime geometry as a {\it constraint} on causal structure, the causal metric hypothesis treats it as an emergent manifestation thereof. The causal-theoretic rejoinder to John Wheeler's statement that {\it ``spacetime tells matter how to move; matter tells spacetime how to curve"} \cite{WheelerQuantum98}, is {\it ``things happen; `spacetime' and `matter' are ways of describing them."} The viewpoint afforded by the causal metric hypothesis must be expressed mathematically before its physical consequences may be explored in a precise manner. Still working in the classical context, the basic mathematical structure associated with a {\it particular instance} of cause and effect is an {\it ordered pair of abstract elements,} the first representing the cause, and the second representing the effect. In causal theory, the {\it arrow of time} arises, ultimately, from many instances of this primitive order. For an entire collection $D$ of events, or families of events, one must consider a corresponding collection $\prec$ of ordered pairs of elements of $D$. Viewing $D$ as a set, the collection $\prec$ is, by definition, a subset of the Cartesian product $D\times D$; i.e., a {\it binary relation} on $D$. In this paper, the pair $(D,\prec)$ is called a {\it directed set.} A more precise {\it classical version} of the causal metric hypothesis may then be stated as follows: \hspace*{.3cm} CCMH.\refstepcounter{textlabels}\label{ccmh} {\bf Classical Causal Metric Hypothesis.} {\it The properties of classical spacetime arise \\ \hspace*{1.7cm} from the structure of a directed set.} This is still a very general idea, whose practical application requires narrowing the focus to ``physically relevant directed sets." This involves nontrivial choices. First, the apparently {\it unidirectional} nature of time suggests, at least at a na\"{i}ve level, that the binary relation $\prec$ should be {\it acyclic;} i.e., that events do not contribute to their own causes, either directly or indirectly. An acyclic binary relation {\it generates a partial order}, in a sense described in sections \hyperref[subsectionaxioms]{\ref{subsectionaxioms}}, \hyperref[subsectionpreorder]{\ref{subsectionpreorder}}, and \hyperref[subsectiontransitiveclosure]{\ref{subsectiontransitiveclosure}}. This lends plausibility to the appearance of {\it order} in Sorkin's version of the causal metric hypothesis, without reference to geometric notions. General relativity permits {\it closed timelike curves,} so acyclicity is a significant assumption. Second, experimental evidence over the last century reveals the fundamental {\it discreteness} of numerous physical quantities at ``small" scales. Causal set theory takes the bold step of incorporating this discreteness at the {\it classical} level, via a {\it discrete measure} $\mu$ that, roughly speaking, ``counts fundamental volume units." This lends plausibility to the appearance of {\it number} in Sorkin's version of the causal metric hypothesis, without reference to ``quantizing spacetime." Since familiar geometric notions are expected to emerge only at relatively large scales in causal set theory, the persistence of volume as a meaningful concept down to the fundamental scale is again a significant assumption. \refstepcounter{textlabels}\label{strongform} {\bf Scope of the Causal Metric Hypothesis.} The proper scope of the causal metric hypothesis is debatable. A conservative approach is to view causal structure as merely a {\it ``replacement for relativistic spacetime,"} without attempting to explain ``particles" and ``fields" by means of the same structure. Versions of causal theory following this approach are {\it theories of gravity,} rather than {\it unified theories.} At the opposite extreme is the {\bf strong form of the causal metric hypothesis,} which seeks to explain {\it all physical phenomena} in terms of causal structure. This is the most ambitious and optimistic form of the causal metric hypothesis, but also the most pleasing at a philosophical level, since it removes any possibility of tension between ``material bodies" and ``background structure," whether static or dynamical. In this sense, the strong form of the causal metric hypothesis achieves perfect {\it background independence.} \refstepcounter{textlabels}\label{technicalimplementation} {\bf Technical Implementations.} Many different technical implementations of the causal metric hypothesis are possible. For example, rather than assigning the {\it same} volume to each element of a directed set, except possibly for small statistical fluctuations, as in Sorkin's approach, one might choose to take account of {\it local causal structure,} as described in section \hyperref[sectioninterval]{\ref{sectioninterval}}, in the assignment of volume. Such an approach accords with the general philosophy that familiar properties of spacetime emerge from causal structure at appropriate scales, without necessarily admitting meaningful extension down to the fundamental scale. For example, it would be absurd, for models as general as causal sets, to expect properties such as {\it dimension} and {\it curvature} to retain their usual geometric meanings in this regime.\footnotemark\footnotetext{Causal dynamical triangulations {\it does} assume a fundamental dimension, but this is a much less general approach than causal set theory.} \refstepcounter{textlabels}\label{quantumcmh} {\bf Quantum Version.} Finally, there is a {\it quantum theoretic version} of the causal metric hypothesis (\hyperref[qcmh]{QCMH}), discussed further in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below, in which the role of classical causal structure is superseded by {\it higher-level multidirected structures} on configuration spaces of directed sets, called {\it kinematic schemes.} In this context, the quantum causal metric hypothesis states that {\it the properties of quantum spacetime arise from the structure of a kinematic scheme of directed sets.} As mentioned in section \hyperref[subsectionapproach]{\ref{subsectionapproach}}, this approach is partly motivated by Isham's {\it quantization on a category} \cite{IshamQuantisingI05}, and Sorkin's {\it quantum measure theory} \cite{SorkinQuantalMeasure12}. \subsection{The Axioms of Causal Set Theory}\label{subsectionaxioms} The causal set literature contains two {\it almost equivalent} formulations of causal set theory, which I refer to as the {\it irreflexive formulation,} and the {\it partial order formulation.} In this section, I introduce and discuss both. Subsequently, the irreflexive formulation is assumed throughout the paper, unless stated otherwise. I occasionally borrow convenient terminology from the partial order formulation, however. \refstepcounter{textlabels}\label{irreflexive} {\bf Irreflexive Formulation.} The irreflexive formulation of causal set theory is defined in terms of irreflexive binary relations, analogous to the familiar {\it less than} relation on the integers. This formulation may be expressed by the following six axioms: \refstepcounter{textlabels}\label{b} \hspace*{.4cm}B. \hspace*{.25cm}{\bf Binary Axiom}: {\it Classical spacetime may be modeled as a set $C$, whose elements \\ \hspace*{1.1cm} represent spacetime events, together with a binary relation $\prec $ on $C$, whose elements \\ \hspace*{1.1cm} represent causal relations between pairs of spacetime events.} \refstepcounter{textlabels}\label{m} \hspace*{.4cm}M. \hspace*{.2cm}{\bf Measure Axiom}: {\it The volume of a spacetime region corresponding to a subset \\ \hspace*{1.1cm} $S$ of $C$ is equal to the cardinality of $S$ in fundamental units, up to Poisson-type \\ \hspace*{1.1cm} fluctuations.} \refstepcounter{textlabels}\label{c} \hspace*{.4cm}C. \hspace*{.25cm}{\bf Countability}: {\it $C$ is countable.} \refstepcounter{textlabels}\label{tr} \hspace*{.25cm}TR. \hspace*{.1cm}{\bf Transitivity}: {\it Given three elements $x,y,$ and $z$ in $C$, if $x\prec y\prec z$, then $x\prec z$.} \refstepcounter{textlabels}\label{if} \hspace*{.3cm}IF. \hspace*{.18cm}{\bf Interval Finiteness}: {\it For every pair of elements $x$ and $z$ in $C$, the open interval \\\hspace*{1.1cm}$\llangle x,z\rrangle:=\{y\in C\hspace*{.1cm}|\hspace*{.1cm} x\prec y\prec z\}$ has finite cardinality.} \refstepcounter{textlabels}\label{ir} \hspace*{.3cm}IR. \hspace*{.2cm}{{\bf Irreflexivity}: {\it Elements of $C$ are not self-related with respect to $\prec $; i.e., $x\nprec x$.} The measure axiom may be expressed more precisely in terms of a {\it discrete measure} $\mu:\ms{P}(C)\rightarrow\mathbb{R}^+$, where $\ms{P}(C)$ is the {\it power set} of $C$; i.e., the set of all subsets of $C$, and where $\mathbb{R}^+$ is the set of positive real numbers.\footnotemark\footnotetext{Use of the real numbers here is merely for convenience; any ``sufficiently large" extension of the positive integers suffices. No essential properties of the continuum are necessary.} Transitivity and irreflexivity together imply the condition of {\it acyclicity}, discussed further below, which rules out causal analogues of closed timelike curves. Interval finiteness is called {\it local finiteness} in the literature, but it is not a local condition in any suitable sense. I reserve the term {\it local finiteness} for a different, genuinely local condition (\hyperref[lf]{LF}), introduced in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} below. Cardinality conditions rarely appear explicitly in the causal set literature, but countability is usually implicit.\footnotemark\footnotetext{For example, Sorkin and Rideout's theory of sequential growth dynamics involves specific enumerations of causal sets. More generally, interval finite but uncountable ``causal sets" are ``physically ridiculous." } For ease of reference, I include a formal definition of causal sets in terms of the above axioms:\\ \begin{defi}\label{defcausalset} A {\bf causal set} is a countable set $C$, equipped with a transitive, interval finite, irreflexive binary relation $\prec$, physically interpreted according to the binary axiom and the measure axiom. \end{defi} It is sometimes necessary to make the binary relation $\prec$ on a causal set $C$ explicit. In such cases, a causal set may be introduced as a pair $(C,\prec)$. Alternatively, it may be denoted by the single symbol $C$, with the understanding that this is an abbreviation. A subset $S$ of a causal set $C=(C,\prec)$ inherits from $C$ a binary relation, called the {\bf subset relation}, also denoted by $\prec$, where $x\prec y$ in $S$ if and only if $x\prec y$ in $C$. It is often useful to view the class of causal sets, together with the class of structure-preserving maps between pairs of causal sets, as a {\it category} $\ms{C}$, called the {\bf category of causal sets}. Basic category-theoretic notions are outlined in section \hyperref[settheoretic]{\ref{settheoretic}}. For the present, it is sufficient to think of a category as a collection of {\it objects}, in this case causal sets, together with a collection of {\it morphisms} between pairs of objects, in this case {\it structure-preserving maps} between pairs of causal sets. To be explicit, a morphism between causal sets $C=(C,\prec)$ and $C'=(C',\prec')$ is a set map $\phi:C\rightarrow C'$ that {\it preserves structure,} in the sense that $\phi(x)\prec' \phi(y)$ in $C'$ whenever $x\prec y$ in $C$. Here, $C$ is called the {\bf source} of $\phi$, and $C'$ is called the {\bf target} of $\phi$. The {\bf index} of $\phi$ is the supremum of the cardinalities of its fibers, where the {\bf fiber} $\phi^{-1}(x')$ of $\phi$ over $x'$ in $C'$ is the set of elements of $C$ that map to $x'$ under $\phi$. An {\bf isomorphism} is an invertible morphism. In almost all cases, only the isomorphism class of a causal set is significant. A bijective morphism is generally {\it not} an isomorphism, since the target may have ``extra relations," extending the binary relation of the source. Such bijective morphisms play an important role in sections \hyperref[settheoretic]{\ref{settheoretic}} and \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}. A {\bf monomorphism} is an injective morphism. I sometimes refer to an injective morphism $\phi:C\rightarrow C'$ as {\it embedding its source as a subobject of its target.} {\bf Partial Order Formulation.}\refstepcounter{textlabels}\label{partialorder} A common alternative formulation of causal set theory uses {\it interval finite partial orders} in the place of irreflexive binary relations. A {\bf partial order} $\preceq$ on a set $P$ is a {\it reflexive, antisymmetric, transitive, binary relation} on $P$, where {\bf reflexivity} means that $x\preceq x$ for every $x$ in $P$, and {\bf antisymmetry} means that if $x\preceq y$ and $y\preceq x$ for two elements $x$ and $y$ in $P$, then $x=y$. Transitivity and antisymmetry together imply that $P$ has no cycles {\it except for} the reflexive cycles $x\preceq x$. As the notation suggests, a partial order $\preceq$ is analogous to the familiar {\it less than or equal to} relation on the integers. Interval finiteness is an extra condition, not part of the definition of a partial order. Interval finite partial orders are special cases of {\bf discrete orders,} which are partial orders in which every nonextremal element has a {\it maximal predecessor} and {\it minimal successor.}\footnotemark\footnotetext{See the quote of Sorkin in section \hyperref[subsectionmodes]{\ref{subsectionmodes}} for an example of this terminology. The reader should also be aware, however, that the term {\it discrete order} is sometimes assigned other, mutually contradictory meanings, such as a discrete {\it linear} order, or a ``trivial" order involving only reflexive relations.} Predecessors and successors are discussed in more detail in section \hyperref[subsecpredecessors]{\ref{subsecpredecessors}} below. The irreflexive and partial order formulations of causal set theory are {\it equivalent at the level of objects} in the following sense: a causal set $(C,\prec)$ may be viewed as a countable interval finite partially ordered set, by taking $x\preceq y$ if $x\prec y$ or $x=y$; conversely, a countable interval finite partially ordered set $(P,\preceq)$ may be viewed as a causal set, by taking $x\prec y$ if $x\preceq y$ and $x\neq y$. The partial order formulation of causal set theory has the superficial advantage of familiarity. For example, in general relativity, the {\it causal relation} is usually taken to be reflexive,\footnotemark\footnotetext{In my view, this is simply an unfortunate choice of definition.} while the {\it chronological relation} is taken to be irreflexive.\footnotemark\footnotetext{{\it Neither} the irreflexive binary relation $\prec$ nor the partial order $\preceq$ corresponds to the chronological relation, which ``excludes the entire boundaries of light cones."} Also, partial orders are more popular than irreflexive relations in many mathematical contexts. On structural grounds, however, the partial order formulation is inconvenient. Partial orders are technically not {\it acyclic,} due to the existence of reflexive relations $x\le x$. Hence, the {\it categories} of causal sets and interval finite partially ordered sets are not equivalent, since the latter category admits ``structure-destroying morphisms" that wrap nontrivial relations around reflexive cycles. These inconveniences, of course, do not reflect any essential difference in physically relevant {\it information content} between the two formulations. Though I use the irreflexive formulation throughout this paper, I sometimes abuse terminology and {\it refer} to the binary relation on a causal set as a partial order, via the object-level equivalence mentioned above. I also apply familiar techniques from order theory to causal sets, such as the {\it order extension principle} (\hyperref[oep]{OEP}), introduced in section \hyperref[settheoretic]{\ref{settheoretic}} below. {\bf Treatment in the Literature.}\refstepcounter{textlabels}\label{axiomsinliterature} The binary axiom (\hyperref[b]{B}), the measure axiom (\hyperref[m]{M}), and countability (\hyperref[c]{C}), do not appear as {\it axioms} in the causal set literature. However, the {\it content} of the binary axiom and measure axiom appear explicitly, and countability is implicit in the focus and methods. Treatment of the remaining axioms depends on whether the irreflexive formulation or the partial order formulation is being used. Here, I reproduce a few representative quotations to illustrate this treatment. In their inaugural paper {\it Space-Time as a Causal Set} \cite{Sorkinetal87}, Bombelli, Lee, Meyer, and Sorkin use the partial order formulation of causal set theory, with the exceptions that an explicit statement of countability, and the nuance of {\it Poisson-type fluctuations,} do not appear: \begin{quotation}\noindent{\it ``...when we measure the volume of a region of space-time, we are merely indirectly counting the number of ``point events" it contains... ...volume is number, and macroscopic causality reflects a deeper notion of order in terms of which all the ``geometrical" structures of spacetime must find their ultimate expression... ...Before proceeding any further, let us put the notion of a causal set into mathematically precise language. A {\it partially ordered set}... ...is a set... ...provided with an order relation, which is transitive... ...noncircular... ...\tn{[and]} reflexive. A partial ordering is {\it locally finite} if every ``Alexandroff set"... ...contains a finite number of elements... ...a {\it causal set} is then by definition a locally finite, partially ordered set."} (page 522) \end{quotation} Here, {\it ``noncircular"} means {\it acyclic} (\hyperref[ac]{AC}) {\it except for reflexive cycles.} An {\it ``Alexandroff set"} is an {\it open interval} $\llangle x,z\rrangle:=\{y\in C\hspace*{.1cm}|\hspace*{.1cm} x\prec y\prec z\}$; hence, {\it ``local finiteness"} means {\it interval finiteness} (\hyperref[if]{IF}) in this context. Later papers are split between the irreflexive formulation and the partial order formulation. For example, in {\it Classical sequential growth dynamics for causal sets} \cite{SorkinSequentialGrowthDynamics99}, Sorkin and Rideout use the irreflexive formulation (see page 2). Later papers also amend the interpretation of volume to allow for Poisson-type fluctuations. For example, in {\it Everpresent $\Lambda$} \cite{SorkinEverPresentLambda04}, Ahmed, Dodelson, Greene, and Sorkin write, \begin{quotation}\noindent{\it ``In order to do justice to local Lorentz invariance, the correspondence between number and volume... ...must be subject to Poisson-type fluctuations..."} (page 2) \end{quotation} Absence of explicit cardinality conditions in the causal set literature may be attributed to the fact that countability is {\it automatic} from the viewpoint of {\it recovering a spacetime manifold from a discrete subset.} More fundamentally, allowing uncountable ``causal sets" greatly exacerbates the existing ``imbalance" of the category of causal sets toward ``large spatial and small temporal size," as indicated by the {\it Kleitman-Rothschild pathology,} discussed in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}}, and its ``countable analogue," discussed in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}} below. These pathologies arise, ultimately, because of the axioms of transitivity (\hyperref[tr]{TR}) and interval finiteness (\hyperref[if]{IF}). \subsection{Acyclic Directed Sets; Directed Sets; Multidirected Sets}\label{subsectionacyclicdirected} It is useful to study causal sets in the context of more general structured sets, which I refer to as {\it acyclic directed sets,} {\it directed sets,} and {\it multidirected sets}, respectively. Acyclic directed sets represent a ``blank canvas for conservative models of classical spacetime;" i.e., models without causal analogues of closed timelike curves. Directed sets and multidirected sets are even more general. While {\it possibly} unnecessary for modeling classical spacetime, directed sets and multidirected sets arise unavoidably in the theory of configuration spaces of classical spacetime models in causal theory. As elaborated in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper, such configuration spaces are of central importance in the histories approach to quantum causal theory. \refstepcounter{textlabels}\label{acyclicdirected} {\bf Acyclic Directed Sets.} In this paper, the term {\it acyclic directed set} means merely a set $A$ equipped with an acyclic binary relation. The meaning of acyclicity is already evident from the foregoing discussion of the causal metric hypothesis and the axioms of causal set theory, but it is convenient, for future reference, to spell out the notion here. A {\bf cycle} in a set equipped with a binary relation is a sequence of relations with the same initial and terminal element: $x=x_0\prec x_1\prec...\prec x_{n-1}\prec x_n=x$. A ``trivial example" of a cycle is a {\it reflexive relation} $x\prec x$. The set $A$ is called acyclic if its binary relation is acyclic; i.e., has no cycles. It is useful to state this condition as an axiom: \refstepcounter{textlabels}\label{ac} \hspace*{.3cm} AC. {\bf Acyclicity}: {\it The binary relation on $A$ is acyclic.} It is also useful, though redundant, to include a formal definition of acyclic directed sets:\\ \begin{defi} An {\bf acyclic directed set} is a set $A$ equipped with an acyclic binary relation $\prec$. \end{defi} Irreflexivity (\hyperref[ir]{IR}) alone does not imply acyclicity; for example, cycles of the form $x\prec y\prec x$ are allowed under irreflexivity. However, irreflexivity and transitivity (\hyperref[tr]{TR}) together {\it do} imply acyclicity. This distinction is important in this paper, since most of the binary relations considered here are not assumed to be transitive. An acyclic directed set may be introduced explicitly as a pair $(A,\prec)$, or it may be denoted in abbreviated fashion by a single symbol $A$. {\it Transitive} acyclic directed sets are equivalent {\it as objects} to partially ordered sets, extending the correspondence introduced in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} in the countable case. A subset $S$ of an acyclic directed set $(A,\prec)$ inherits a natural {\bf subset relation}, given by restricting $\prec$ to $S$. The {\bf category of acyclic directed sets} is the category $\ms{A}$ whose objects are acyclic directed sets, and whose morphisms $(A,\prec)\rightarrow (A',\prec')$ are set maps $\phi:A\rightarrow A'$, such that $\phi(x)\prec' \phi(y)$ in $A'$ whenever $x\prec y$ in $A$. Related notions, such as sources, targets, fibers, indices, isomorphisms, and monomorphisms, generalize from the case of causal sets in an obvious way. Because of the choice to define the category $\ms{C}$ of causal sets using the irreflexive formulation of causal set theory, causal sets may be viewed as objects of either $\ms{C}$ or $\ms{A}$ without ambiguity concerning their morphism classes. Technically, this means that $\ms{C}$ {\it embeds into $\ms{A}$ as a full subcategory.} Heuristically, it means that pairs of causal sets share the same relationships whether viewed as objects of $\ms{C}$ or $\ms{A}$.\refstepcounter{textlabels}\label{conventionads} \refstepcounter{textlabels}\label{graph} Acyclic directed sets are equivalent to {\it small, simple, acyclic directed graphs}, under the correspondence sending elements to vertices and relations to directed edges. Here {\it small} means that the vertex class of the graph under consideration is a set, rather than a {\it proper class,} and {\it simple} means that for any pair of vertices $x$ and $y$, there is at most one edge from $x$ to $y$. In this case, the equivalence extends to the level of categories. Both sides of this equivalence are useful: the set-theoretic side for its convenient terminology, and the graph-theoretic side for the visual perspective afforded by generalized Hasse diagrams. \refstepcounter{textlabels}\label{directedsets} {\bf Directed Sets.} As mentioned above, general relativity admits solutions involving {\it closed timelike curves,} whose causal analogues are the {\it cycles} discussed above. In fact, such curves appear in rather generic situations, such as the Kerr black hole. Acyclic directed sets are inadequate to model causal analogues of such spacetimes; more general objects, which I refer to as {\it directed sets,} are needed. In this paper, the term {\it directed set} means merely a set $D$ equipped with a binary relation. However, the reader should keep in mind that the term {\it directed set} has a different conventional meaning in the context of order theory, as explained in the discussion of nonstandard terminology in section \hyperref[subsectionnotation]{\ref{subsectionnotation}} above. It is useful to include a formal definition of this term as it is used in this paper: \refstepcounter{textlabels}\label{defdirected} \vspace*{.2cm} \begin{defi} A {\bf directed set} is a set $D$ equipped with a binary relation $\prec$. \end{defi} A directed set may be introduced explicitly as a pair $(D,\prec)$, or it may be denoted in abbreviated fashion by a single symbol $D$. A subset $S$ of a directed set $(D,\prec)$ inherits a natural {\bf subset relation}, given by restricting $\prec$ to $S$. The {\bf category of directed sets} is the category $\ms{D}$ whose elements are directed sets, and whose morphisms $(D,\prec)\rightarrow (D',\prec')$ are set maps $\phi:D\rightarrow D'$, such that $\phi(x)\prec' \phi(y)$ in $D'$ whenever $x\prec y$ in $D$. Related notions, such as sources, targets, fibers, indices, isomorphisms, and monomorphisms, generalize from the case acyclic directed sets in an obvious way. The category $\ms{A}$ of acyclic directed sets embeds into $\ms{D}$ as a full subcategory.\footnotemark\footnotetext{It is tempting to invent suggestive terminology for special classes of directed sets in the context of causal theory. The best term, of course, is ``causal sets," which is already appropriated. Finkelstein \cite{Finkelstein88} uses the term {\it causal nets.} Benincasa and Dowker \cite{DowkerScalarCurvature10}, following Riemann, use the term {\it discrete manifolds} in a general context, including discrete causal models as a special case. Other possible choices are {\it causal graphs} and {\it discrete chronofolds,} though the latter is the name of an Apple smartphone application.} Like acyclic directed sets, directed sets may also be understood in graph-theoretic terms, as {\it small, simple, directed graphs.} Note that the definition of a {\it simple} directed graph allows for the coexistence of ``reciprocal edges" $x\prec y$ and $y\prec x$, but prohibits multiple edges between $x$ and $y$ in the same direction. Besides their use as models of classical spacetime, directed sets arise naturally in more abstract contexts, in which their elements represent {\it more general entities} than structureless spacetime events. Examples include the {\it relation spaces} and {\it power spaces} introduced in section \hyperref[sectionbinary]{\ref{sectionbinary}}. \refstepcounter{textlabels}\label{multidirectedsets} {\bf Multidirected Sets.} The most general structured sets enjoying broad use in this paper are {\it multidirected sets,} which arise in the study of configuration spaces of acyclic directed sets and directed sets. Special configuration spaces, called kinematic schemes, arise naturally in the histories approach to quantum causal theory, developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below. The multidirected sets of principal interest in this paper arise by {\it``decategorifying"} kinematic schemes; i.e., demoting them to a lower level of algebraic hierarchy by forgetting the internal structure of their member sets. Possible {\it classical} applications of multidirected sets are mentioned very briefly in section \hyperref[subsectionomitted]{\ref{subsectionomitted}} below. \refstepcounter{textlabels}\label{defmultidirected} \vspace*{.2cm} \begin{defi} A {\bf multidirected set} consists of a set of elements $M$, a set of relations $R$, and {\bf initial} and {\bf terminal element maps} $i:R\rightarrow M$ and $t:R\rightarrow M$, assigning to each relation initial and terminal elements. \end{defi} A multidirected set for which each ordered pair $(x,y)$ of elements has at most one relation $r$ satisfying the conditions that $i(r)=x$ and $t(r)=y$, may be viewed as a directed set, whose binary relation $\prec$ is defined by setting $x\prec y$ whenever there exists a relation $r$ such that $i(r)=x$ and $t(r)=y$. However, the structure of a general multidirected set {\it cannot} be expressed by a single binary relation, since an ordered pair $(x,y)$ of elements of $M$ generally does not {\it uniquely identify} a relation from $x$ to $y$. In this context, the expression $x\prec y$ either refers to a {\it particular} relation with initial element $x$ and terminal element $y$, or indicates that {\it at least one such relation exists.} A multidirected set may be introduced as a quadruple $(M,R,i,t)$, or it may be denoted in abbreviated fashion by a single symbol $M$. A {\bf subobject} of a multidirected set $M=(M,R,i,t)$ consists of a subset of $M$, together with subset of $R$, chosen in such a way that the initial and terminal element maps $i$ and $t$ map the latter subset into the former. Subsets of causal sets, acyclic directed sets, and directed sets, together with their subset relations, are all special cases of subobjects of multidirected sets. \newpage Abstractly, multidirected sets are structurally analogous to {\it small categories;} i.e., categories whose object classes are sets. As in the case of directed sets, the elements of a multidirected set are often taken to represent entities more complex than events. From a graph-theoretic perspective, multidirected sets are equivalent to {\it small directed multigraphs,} or equivalently, {\it small quivers}.\footnotemark\footnotetext{This term deliberately evokes the ``quiver of arrows" carried by an archer.} Use of the term {\it quiver} often has {\it algebraic connotations;} for example, the {\it concatenation algebra} over a multidirected set, introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below, is a special type of {\it quiver algebra}.\footnotemark\footnotetext{The use of multidirected sets as {\it algebraic substrata} goes back at least to {\it Gabriel's theorem} in 1972, which classifies {\it connected quivers of finite type} and their representations in terms of {\it Dynkin diagrams} and {\it root systems.}} As I have defined it, a multidirected set is precisely what Abrams and Pino call a ``directed graph" in \cite{AbramsPathAlgebra05}. The terminology of {\it multidirected sets} is the most convenient choice for this paper, for at least five reasons: 1) it is descriptive; 2) it is congenial to the causal set viewpoint; 3) it is relatively neutral in its connotations; 4) it is amenable to generalized order-theoretic notions and terminology; and 5) it eliminates potential confusion about proper classes. The {\bf category of multidirected sets} is the category $\ms{M}$ whose objects are multidirected sets, and whose morphisms are pairs of maps of elements and relations that respect initial and terminal element maps. More precisely, a morphism $\phi$ between two multidirected sets $(M,R,i,t)$ and $(M',R',i',t')$ consists of {\it two} maps: a {\bf map of elements} $\phi_{\tn{\fsz{elt}}}:M\rightarrow M'$, and a {\bf map of relations} $\phi_{\tn{\fsz{rel}}}:R\rightarrow R'$, satisfying the conditions that $\phi_{\tn{\fsz{elt}}}(i(r))=i'(\phi_{\tn{\fsz{rel}}}(r))$ and $\phi_{\tn{\fsz{elt}}}(t(r))=t'(\phi_{\tn{\fsz{rel}}}(r))$ for each relation $r$ in $R$.\footnotemark\footnotetext{The student of algebraic geometry will recall that a morphism of algebraic schemes also consists of two maps: a map of topological spaces, and a map of structure sheaves. Of course, the latter map is ``in the opposite direction." Nevertheless, there is a loose analogy between $R$ and the {\it tangent sheaf} in geometry.} Morphisms of multidirected sets play a prominent role in section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, where I use them to analyze the properties of the {\it relation space functor} and the {\it abstract element space functor.} Related notions, such as sources, targets, fibers, indices, isomorphisms, and monomorphisms, may be defined by applying familiar notions to {\it both} maps $\phi_{\tn{\fsz{elt}}}$ and $\phi_{\tn{\fsz{rel}}}$. Figure \hyperref[figmultidirected]{\ref{figmultidirected}} below illustrates the differences among {\it causal sets,} {\it acyclic directed sets,} {\it directed sets,} {\it acyclic multidirected sets,} and {\it multidirected sets.} Figure \hyperref[figmultidirected]{\ref{figmultidirected}}a shows a causal set. Note the transitivity of the binary relation. Figure \hyperref[figmultidirected]{\ref{figmultidirected}}b shows an acyclic directed set. Here, only {\it some} of the relations ``implied by transitivity" are present; for example, there is a relation $w\prec z$ corresponding to the pair of relations $w\prec x\prec z$, but no relation $w\prec y$ corresponding to the pair of relations $w\prec x\prec y$. Figure \hyperref[figmultidirected]{\ref{figmultidirected}}c shows a directed set. Note that reflexive relations, such as $t\prec t$, and {\it reciprocal relations,} such as $u\prec v$ and $v\prec u$, are allowed. Figure \hyperref[figmultidirected]{\ref{figmultidirected}}d shows an acyclic multidirected set. While acyclicity is defined above only for binary relations, a straightforward generalization to multidirected sets is given in section \hyperref[subsectionchains]{\ref{subsectionchains}} below. Informally, multiple independent relations between a given pair of elements are allowed in this case, but only in one direction. For example, there are two independent relations from $w$ to $x$ and from $x$ to $z$. Figure \hyperref[figmultidirected]{\ref{figmultidirected}}e shows a general multidirected set. Figures \hyperref[figmultidirected]{\ref{figmultidirected}}c and \hyperref[figmultidirected]{\ref{figmultidirected}}e illustrate the fact that cycles cannot be represented by generalized Hasse diagrams. Whenever cycles are present, the directions of relations are indicated explicitly by arrows. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{3.6cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,3.3)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(3.45,3.3)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(6.75,3.3)}{\pgfbox[left,center]{c)}} \pgfputat{\pgfxy(10.05,3.3)}{\pgfbox[left,center]{d)}} \pgfputat{\pgfxy(13.35,3.3)}{\pgfbox[left,center]{e)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,3.6)(16.5,3.6) \pgfxyline(0,-.1)(0,3.6) \pgfxyline(3.3,-.1)(3.3,3.6) \pgfxyline(6.6,-.1)(6.6,3.6) \pgfxyline(9.9,-.1)(9.9,3.6) \pgfxyline(13.2,-.1)(13.2,3.6) \pgfxyline(16.5,-.1)(16.5,3.6) \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfputat{\pgfxy(.5,.1)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.55,1.15)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.2,3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.8,2.5)}{\pgfbox[center,center]{$y$}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.133cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,3.1)}{0.133cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.8,2.1)}{0.133cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.7,3.9)}{0.133cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.9,1.7)}{0.133cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.5,3.2)}{0.133cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,.7)}{0.133cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,3.8)}{0.133cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.3,1.9)}{0.133cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,.5)}{0.133cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.1,3.1)}{0.133cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node5}{Node3} \pgfnodeconnline{Node5}{Node4} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node7}{Node6} \pgfnodeconnline{Node10}{Node8} \pgfnodeconnline{Node10}{Node11} \pgfnodeconnline{Node11}{Node8} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.35cm}{0cm}} \pgfputat{\pgfxy(.5,.1)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.55,1.15)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.2,3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.8,2.5)}{\pgfbox[center,center]{$y$}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.133cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,3.1)}{0.133cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.8,2.1)}{0.133cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.7,3.9)}{0.133cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.9,1.7)}{0.133cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.5,3.2)}{0.133cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,.7)}{0.133cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,3.8)}{0.133cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.3,1.9)}{0.133cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,.5)}{0.133cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.1,3.1)}{0.133cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node5}{Node3} \pgfnodeconnline{Node5}{Node4} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node11}{Node8} \pgfnodeconnline{Node11}{Node10} \pgfnodeconnline{Node7}{Node6} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.55cm}{0cm}} \pgfputat{\pgfxy(2.3,1.6)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(2,2.9)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3,2.4)}{\pgfbox[center,center]{$v$}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.133cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,3.1)}{0.133cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.8,2.1)}{0.133cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.7,3.9)}{0.133cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.9,1.7)}{0.133cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.5,3.2)}{0.133cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,.7)}{0.133cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,3.8)}{0.133cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.3,1.9)}{0.133cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,.5)}{0.133cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.1,3.1)}{0.133cm} \pgfmoveto{\pgfxy(1.3,1.1)} \pgflineto{\pgfxy(1.9,1.7)} \pgfstroke \pgfmoveto{\pgfxy(3.3,1.5)} \pgfcurveto{\pgfxy(3.5,1.5)}{\pgfxy(3.8,1.8)}{\pgfxy(3.3,1.9)} \pgfstroke \pgfmoveto{\pgfxy(.6,1.8)} \pgflineto{\pgfxy(.5,3.1)} \pgfstroke \pgfmoveto{\pgfxy(1.1,3.5)} \pgflineto{\pgfxy(1.7,3.9)} \pgfstroke \pgfmoveto{\pgfxy(1.25,3)} \pgflineto{\pgfxy(1.7,3.9)} \pgfstroke \pgfmoveto{\pgfxy(1.35,1.9)} \pgflineto{\pgfxy(.8,2.1)} \pgfstroke \pgfmoveto{\pgfxy(2.1,3.55)} \pgflineto{\pgfxy(2.5,3.2)} \pgfstroke \pgfmoveto{\pgfxy(2.2,2.45)} \pgflineto{\pgfxy(1.9,1.7)} \pgfstroke \pgfmoveto{\pgfxy(2.75,1.95)} \pgflineto{\pgfxy(2.5,3.2)} \pgfstroke \pgfmoveto{\pgfxy(4.05,1.8)} \pgflineto{\pgfxy(4.1,3.1)} \pgfstroke \pgfmoveto{\pgfxy(3.4,3.3)} \pgfcurveto{\pgfxy(3.2,3.4)}{\pgfxy(3.1,3.7)}{\pgfxy(3,3.8)} \pgfstroke \pgfmoveto{\pgfxy(3.7,3.7)} \pgfcurveto{\pgfxy(3.8,3.6)}{\pgfxy(4.1,3.3)}{\pgfxy(4.1,3.1)} \pgfstroke \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfmoveto{\pgfxy(.7,.5)} \pgflineto{\pgfxy(1.3,1.1)} \pgfstroke \pgfmoveto{\pgfxy(3.3,1.9)} \pgfcurveto{\pgfxy(3,1.9)}{\pgfxy(2.7,1.4)}{\pgfxy(3.3,1.5)} \pgfstroke \pgfmoveto{\pgfxy(.7,.5)} \pgflineto{\pgfxy(.6,1.8)} \pgfstroke \pgfmoveto{\pgfxy(.5,3.1)} \pgflineto{\pgfxy(1.1,3.5)} \pgfstroke \pgfmoveto{\pgfxy(.8,2.1)} \pgflineto{\pgfxy(1.25,3)} \pgfstroke \pgfmoveto{\pgfxy(1.9,1.7)} \pgflineto{\pgfxy(1.35,1.9)} \pgfstroke \pgfmoveto{\pgfxy(1.7,3.9)} \pgflineto{\pgfxy(2.1,3.55)} \pgfstroke \pgfmoveto{\pgfxy(2.5,3.2)} \pgflineto{\pgfxy(2.2,2.45)} \pgfstroke \pgfmoveto{\pgfxy(4.1,3.1)} \pgfcurveto{\pgfxy(3.8,3.1)}{\pgfxy(3.7,3.1)}{\pgfxy(3.4,3.3)} \pgfstroke \pgfmoveto{\pgfxy(3,3.8)} \pgfcurveto{\pgfxy(3.2,3.9)}{\pgfxy(3.4,3.8)}{\pgfxy(3.7,3.7)} \pgfstroke \pgfmoveto{\pgfxy(3,.7)} \pgflineto{\pgfxy(2.75,1.95)} \pgfstroke \pgfmoveto{\pgfxy(4,.5)} \pgflineto{\pgfxy(4.05,1.8)} \pgfstroke \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{9.9cm}{0cm}} \pgfputat{\pgfxy(.5,.1)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.55,1.15)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.2,3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.8,2.5)}{\pgfbox[center,center]{$y$}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.133cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,3.1)}{0.133cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.8,2.1)}{0.133cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.7,3.9)}{0.133cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.9,1.7)}{0.133cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.5,3.2)}{0.133cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,.7)}{0.133cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,3.8)}{0.133cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.3,1.9)}{0.133cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,.5)}{0.133cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.1,3.1)}{0.133cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node5}{Node3} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node10}{Node11} \pgfnodeconnline{Node11}{Node8} \pgfnodeconnline{Node7}{Node6} \pgfnodeconncurve{Node5}{Node4}{70}{300}{.5cm}{.5cm} \pgfnodeconncurve{Node5}{Node4}{110}{260}{.5cm}{.5cm} \pgfnodeconncurve{Node11}{Node8}{120}{15}{.5cm}{.5cm} \pgfnodeconncurve{Node11}{Node8}{170}{280}{.5cm}{.5cm} \pgfnodeconncurve{Node1}{Node5}{20}{250}{.5cm}{.5cm} \pgfnodeconncurve{Node1}{Node5}{70}{200}{.5cm}{.5cm} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{13.15cm}{0cm}} \pgfputat{\pgfxy(.5,.1)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.55,1.15)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.2,3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.8,2.5)}{\pgfbox[center,center]{$y$}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.133cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,3.1)}{0.133cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.8,2.1)}{0.133cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.7,3.9)}{0.133cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.9,1.7)}{0.133cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.5,3.2)}{0.133cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,.7)}{0.133cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,3.8)}{0.133cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.3,1.9)}{0.133cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,.5)}{0.133cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.1,3.1)}{0.133cm} \pgfmoveto{\pgfxy(1.2,1.3)} \pgfcurveto{\pgfxy(1.4,1.5)}{\pgfxy(1.7,1.6)}{\pgfxy(1.9,1.7)} \pgfstroke \pgfmoveto{\pgfxy(1.3,1.1)} \pgflineto{\pgfxy(1.9,1.7)} \pgfstroke \pgfmoveto{\pgfxy(1.35,.8)} \pgfcurveto{\pgfxy(1.15,.65)}{\pgfxy(1,.55)}{\pgfxy(.7,.5)} \pgfstroke \pgfmoveto{\pgfxy(3.2,2.5)} \pgfcurveto{\pgfxy(3.4,2.6)}{\pgfxy(3.7,2)}{\pgfxy(3.3,1.9)} \pgfstroke \pgfmoveto{\pgfxy(3.3,1.5)} \pgfcurveto{\pgfxy(3.5,1.5)}{\pgfxy(3.8,1.8)}{\pgfxy(3.3,1.9)} \pgfstroke \pgfmoveto{\pgfxy(3,4.3)} \pgfcurveto{\pgfxy(3.5,4.3)}{\pgfxy(3.5,3.8)}{\pgfxy(3,3.8)} \pgfstroke \pgfmoveto{\pgfxy(3.8,1.9)} \pgfcurveto{\pgfxy(3.85,2.3)}{\pgfxy(3.9,2.7)}{\pgfxy(4.1,3.1)} \pgfstroke \pgfmoveto{\pgfxy(4.3,1.9)} \pgfcurveto{\pgfxy(4.3,2.3)}{\pgfxy(4.2,2.7)}{\pgfxy(4.1,3.1)} \pgfstroke \pgfmoveto{\pgfxy(.6,1.8)} \pgflineto{\pgfxy(.5,3.1)} \pgfstroke \pgfmoveto{\pgfxy(1.1,3.5)} \pgflineto{\pgfxy(1.7,3.9)} \pgfstroke \pgfmoveto{\pgfxy(1.25,3)} \pgflineto{\pgfxy(1.7,3.9)} \pgfstroke \pgfmoveto{\pgfxy(1.35,1.9)} \pgflineto{\pgfxy(.8,2.1)} \pgfstroke \pgfmoveto{\pgfxy(2.1,3.55)} \pgflineto{\pgfxy(2.5,3.2)} \pgfstroke \pgfmoveto{\pgfxy(2.2,2.45)} \pgflineto{\pgfxy(1.9,1.7)} \pgfstroke \pgfmoveto{\pgfxy(3.55,3.45)} \pgflineto{\pgfxy(3,3.8)} \pgfstroke \pgfmoveto{\pgfxy(2.75,1.95)} \pgflineto{\pgfxy(2.5,3.2)} \pgfstroke \pgfmoveto{\pgfxy(4.05,1.8)} \pgflineto{\pgfxy(4.1,3.1)} \pgfstroke \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfmoveto{\pgfxy(.7,.5)} \pgfcurveto{\pgfxy(.8,.7)}{\pgfxy(.9,1)}{\pgfxy(1.2,1.3)} \pgfstroke \pgfmoveto{\pgfxy(.7,.5)} \pgflineto{\pgfxy(1.3,1.1)} \pgfstroke \pgfmoveto{\pgfxy(1.9,1.7)} \pgfcurveto{\pgfxy(1.8,1.5)}{\pgfxy(1.65,1.1)}{\pgfxy(1.35,.8)} \pgfstroke \pgfmoveto{\pgfxy(3.3,1.9)} \pgfcurveto{\pgfxy(3,1.9)}{\pgfxy(2.6,2.4)}{\pgfxy(3.2,2.5)} \pgfstroke \pgfmoveto{\pgfxy(3.3,1.9)} \pgfcurveto{\pgfxy(3,1.9)}{\pgfxy(2.7,1.4)}{\pgfxy(3.3,1.5)} \pgfstroke \pgfmoveto{\pgfxy(3,3.8)} \pgfcurveto{\pgfxy(2.5,3.8)}{\pgfxy(2.5,4.3)}{\pgfxy(3,4.3)} \pgfstroke \pgfmoveto{\pgfxy(4,.5)} \pgfcurveto{\pgfxy(3.8,1)}{\pgfxy(3.8,1.3)}{\pgfxy(3.8,1.9)} \pgfstroke \pgfmoveto{\pgfxy(4,.5)} \pgfcurveto{\pgfxy(4.2,1)}{\pgfxy(4.3,1.3)}{\pgfxy(4.3,1.9)} \pgfstroke \pgfmoveto{\pgfxy(.7,.5)} \pgflineto{\pgfxy(.6,1.8)} \pgfstroke \pgfmoveto{\pgfxy(.5,3.1)} \pgflineto{\pgfxy(1.1,3.5)} \pgfstroke \pgfmoveto{\pgfxy(.8,2.1)} \pgflineto{\pgfxy(1.25,3)} \pgfstroke \pgfmoveto{\pgfxy(1.9,1.7)} \pgflineto{\pgfxy(1.35,1.9)} \pgfstroke \pgfmoveto{\pgfxy(1.7,3.9)} \pgflineto{\pgfxy(2.1,3.55)} \pgfstroke \pgfmoveto{\pgfxy(2.5,3.2)} \pgflineto{\pgfxy(2.2,2.45)} \pgfstroke \pgfmoveto{\pgfxy(4.1,3.1)} \pgflineto{\pgfxy(3.55,3.45)} \pgfstroke \pgfmoveto{\pgfxy(3,.7)} \pgflineto{\pgfxy(2.75,1.95)} \pgfstroke \pgfmoveto{\pgfxy(4,.5)} \pgflineto{\pgfxy(4.05,1.8)} \pgfstroke \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Causal set; b) acyclic directed set; c) directed set; d) acyclic multidirected set; e) multidirected set with cycles.} \label{figmultidirected} \end{figure} \vspace*{-.5cm} \subsection{Chains; Antichains; Irreducibility; Independence}\label{subsectionchains} Many important properties of multidirected sets may be described in terms of sequences of consecutive elements and relations, called {\it chains.} For example, a multidirected set $M$ is acyclic if and only if no chain in $M$ has the same initial and terminal element. Subsets of multidirected sets admitting no chains between pairs of elements are also important; these are called {\it antichains}. Both chains and antichains are formally introduced in definition \hyperref[definitionchains]{\ref{definitionchains}} below. Chains are used in definition \hyperref[deficonditionsmulti]{\ref{deficonditionsmulti}} to generalize familiar conditions on directed sets, originally defined in terms of binary relations, to the case of multidirected sets. When a multidirected set is assigned a physical interpretation, for example, as a model of information flow or causal structure, it is vital to distinguish between its {\it absolute mathematical properties,} and the {\it physical characteristics subjectively ascribed to it.} An important example is the distinction between mathematical {\it reducibility} or {\it irreducibility} of a relation between a pair of elements $x$ and $y$, and physical {\it dependence} or {\it independence} of the information encoded by this relation, with respect to the information encoded by other chains between $x$ and $y$. Definition \hyperref[definitionindependence]{\ref{definitionindependence}} introduces the property of independence in the more general context of families of chains between pairs of elements. This leads to the {\it independence convention} (\hyperref[ic]{IC}), which is crucial to the physical interpretation of directed sets and multidirected sets throughout the remainder of the paper. {\bf Chains; Antichains; Mathematical Properties.}\refstepcounter{textlabels}\label{chains} Definition \hyperref[definitionchains]{\ref{definitionchains}} introduces chains and antichains. The reader should note that chains are special cases of {\it paths,} discussed in more detail in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. \\ \begin{defi}\label{definitionchains} Let $M=(M,R,i,t)$ be a multidirected set, and let $x$ and $y$ be elements of $M$. \begin{enumerate} \item A {\bf chain} $\gamma$ in $M$ is a sequence of elements and relations of the form $...\prec x_0\prec x_1\prec...$ in $M$, where the notation $x_n\prec x_{n+1}$ refers to a particular relation $r$ in $R$ such that $x_n=i(r)$ and $x_{n+1}=t(r)$. The {\bf chain set} $\tn{Ch}(M)$ of $M$ is the set of all chains in $M$. \item A {\bf chain of length} $n$, or {\bf $n$-chain}, from $x$ to $y$ in $M$, is a chain of the form $x=x_0\prec x_1\prec...\prec x_{n-1}\prec x_n=y$. The {\bf set of $n$-chains} $\tn{Ch}_n(M)$ in $M$ is the subset of $\tn{Ch}(M)$ consisting of chains of length $n$. A {\bf complex chain} is a chain of length at least two. \item\refstepcounter{textlabels}\label{irred} A relation $r$ in $R$ is called {\bf reducible} if there exists a complex chain from its initial element to its terminal element. Such a chain is called a {\bf reducing chain} for $r$. If $r$ is not reducible, it is called {\bf irreducible}. $M$ itself is called {\bf irreducible} if all its relations are irreducible. \item An {\bf antichain} $\sigma$ in $M$ is a subset of $M$ admitting no chain in $M$ between any pair of its elements. \end{enumerate} \end{defi} In the context of classical causal structure, a chain is analogous to a relativistic {\it world line,} while an antichain corresponds to a {\it spacelike section} of spacetime; i.e., a {\it Cauchy surface}. A $0$-chain in $M$ is just an element of $M$, and a $1$-chain in $M$ is just an element of the relation set $R$ of $M$. $R$ is the underlying set of the {\it relation space} $\ms{R}(M)$ over $M$, studied in detail in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below. Similarly, {\it chain spaces} over $M$, whose underlying sets are the chain sets $\tn{Ch}_n(M)$ and $\tn{Ch}(M)$, are introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}. A chain in $M$ may include a given element or relation more than once; when this occurs, it indicates the existence of a cycle in $M$, as spelled out in definition \hyperref[deficonditionsmulti]{\ref{deficonditionsmulti}} below. Any relation $r$ in $M$ belonging to a cycle is reducible, via any chain beginning and ending with $r$. The reason for defining chains in terms of {\it elements and relations,} with the awkward use of the expression $x_n\prec x_{n+1}$ to denote a {\it particular} relation, is to include individual elements in the definition, as $0$-chains. For most other purposes, it is more convenient to define chains purely in terms of relations. \refstepcounter{textlabels}\label{conditionsmulti} {\bf Conditions on Multidirected Sets.} Conditions on directed sets, such as transitivity (\hyperref[tr]{TR}), interval finiteness (\hyperref[if]{IF}), irreflexivity (\hyperref[ir]{IR}), and acyclicity (\hyperref[ac]{AC}), defined in sections \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} and \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} above in terms of binary relations, are also useful in the more general context of multidirected sets. It is therefore necessary to re-express these conditions in terms of relation sets and initial and terminal element maps. Chains are particularly convenient for this purpose.\\ \begin{defi}\label{deficonditionsmulti} Let $M=(M,R,i,t)$ be a multidirected set. \begin{enumerate} \item $M$ is {\bf transitive} if every complex chain in $M$ is a reducing chain for a relation $r$ in $R$. \item Let $x$ and $z$ be two elements of $M$. The {\bf open interval} $\llangle x,z\rrangle$ in $M$ is the subset of $M$ consisting of all elements $y$ admitting a chain from $x$ to $y$ and a chain from $y$ to $z$. $M$ is {\bf interval finite} if every open interval in $M$ is finite. \item A relation $r$ in $R$ is called a {\bf reflexive relation} at $x$ in $M$ if $i(r)=t(r)=x$. $M$ is {\bf irreflexive} if $R$ includes no reflexive relations. \item A {\bf cycle} in $M$ is a chain $x_0\prec x_1\prec...\prec x_{n-1}\prec x_n$ such that $x_0=x_n$. $M$ is {\bf acyclic} if it contains no cycles. \end{enumerate} \end{defi} Other important conditions on multidirected sets, such as local finiteness (\hyperref[lf]{LF}), are examined and compared in section \hyperref[sectioninterval]{\ref{sectioninterval}} below. \refstepcounter{textlabels}\label{independence} {\bf Independence: A Physical Property.} Definitions \hyperref[definitionchains]{\ref{definitionchains}} and \hyperref[deficonditionsmulti]{\ref{deficonditionsmulti}} above are {\it purely mathematical,} making no reference to the physical interpretation of the multidirected set $M$. Definition \hyperref[definitionindependence]{\ref{definitionindependence}} below introduces the notion of {\it independence} of a family of chains in a multidirected set, which is a physical property subjectively associated with such a family. The mathematical reader should be careful to observe that this notion of independence is different than the notions of independence appearing in mathematical fields such as linear algebra and matroid theory, and that this difference is not merely one of distinction between physical and mathematical properties. In such mathematical contexts, independence is, broadly speaking, an {\it absolute, ``internal" property, signifying absence of redundancy,} while in the present physical context, it is a {\it relative property, signifying uniqueness of information content.} For example, the linear independence of a set $S$ of vectors in a vector space $V$ does not depend on the complement $V-S$; in particular, one cannot ``spoil" the linear independence of $S$ by adding new dimensions to $V$. Further, $S$ has no ``monopoly" on the information about $V$ it contains; any basis for the span of $S$ in $V$ contains the same information about $V$. By contrast, independence of a family $\Gamma$ of chains in a multidirected set $M$ means that $\Gamma$ encodes, possibly in a redundant fashion, {\it at least some information} that cannot be recovered from any subobject of $M$ not containing $\Gamma$. Definition \hyperref[definitionindependence]{\ref{definitionindependence}} makes this idea precise for the special case of a family of chains between a particular pair of elements of $M$. \\ \begin{defi}\label{definitionindependence} Let $M=(M,R,i,t)$ be a multidirected set, interpreted as a model of information flow or causal structure, and let $x$ and $y$ be elements of $M$. \begin{enumerate} \item A family $\Gamma$ of chains between $x$ and $y$ in $M$ is called {\bf dependent} if there exists another such family $\Gamma'$, not containing $\Gamma$, encoding all information or causal influence encoded by $\Gamma$. \item In particular, a chain $\gamma$ from $x$ to $y$ in $M$ is called {\bf dependent} if there exists a family $\Gamma'$ of chains from $x$ to $y$, not including $\gamma$, encoding all information or causal influence encoded by $\gamma$. \item If a chain or family of chains is not dependent, it is called {\bf independent}. \end{enumerate} \end{defi} Dependence and independence may be easily generalized to the case of families of chains between pairs of {\it subsets} of $M$, but definition \hyperref[definitionindependence]{\ref{definitionindependence}} is sufficient for the purposes of this paper. Independence is constrained, but not determined, by the mathematical structure of $M$. Of principal interest is the {\it dependence or independence of individual relations;} i.e., $1$-chains. For example, {\it irreducible relations in directed sets are necessarily independent,} since no other chains exist between their initial and terminal elements. Reducible relations, on the other hand, may be interpreted as either dependent or independent. For multidirected sets, multiple relations may share the same initial and terminal elements, so even irreducible relations may be interpreted as dependent in this context. These ambiguities illustrate the need to fix {\it conventions} regarding dependence and independence. Causal set theory implicitly treats reducible relations as {\it dependent}, as reflected in the practice of representing causal sets by standard Hasse diagrams. However, {\it this convention leads to information-theoretic shortcomings in causal set theory,} as explained in section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}. Since the multidirected sets of principal interest in this paper satisfy a local finiteness condition (\hyperref[lf]{LF}) rendering dependent relations information-theoretically superfluous, I adopt the following convention: \refstepcounter{textlabels}\label{ic} \hspace*{.3cm} IC. {\bf Independence Convention}: {\it Every relation in a multidirected set is interpreted as \\ \hspace*{1.1cm}independent unless stated otherwise.} \refstepcounter{textlabels}\label{interpolativecomparison} It is reasonable to ask why one would even {\it consider} including dependent relations in a physical model. In the locally finite case, dependent relations only add redundancy, but nontrivial {\it locally infinite} multidirected sets exist in which {\it every relation is both reducible and dependent.} The principal example is, of course, relativistic spacetime, which exhibits this behavior due to the {\bf interpolative property} of the continuum. Under the interpolative property, every relation $x\prec z$ admits an {\bf interpolating element} $y$ such that $x\prec y\prec z$.\footnotemark\footnotetext{For the continuum itself, the interpolative property may be realized by {\it averaging}.} The interpolative property has been studied since antiquity; for example, it forms the crux of Zeno's {\it dichotomy paradox,} mentioned in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above. In a more modern context, the interpolative property is closely related to the use of {\it Cauchy surfaces} to separate ``past" and ``future" regions of spacetime in relativistic cosmology. Suitable analogues of Cauchy surfaces for multidirected sets are introduced in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below. {\it Domain theory} is a branch of causal theory involving interpolative causal structures.\footnotemark\footnotetext{As noted in the introduction to section \hyperref[sectionaxioms]{\ref{sectionaxioms}}, the domains studied in domain theory are distinct from the {\it domains of influence} introduced in section \hyperref[subsecpredecessors]{\ref{subsecpredecessors}}.} A useful reference for domain theory is the paper of Martin and Panangaden \cite{Martin06}, cited in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} in the discussion of Malament's metric recovery theorem. The independence convention makes physical sense only if one asserts the freedom to use relations between pairs of elements to encode {\it actual influence or information flow}. The causal set axiom of transitivity (\hyperref[tr]{TR}) removes this freedom by {\it prescribing} the existence of a relation between two elements in a causal set whenever there is a chain between them. This eliminates the option of representing {\it dependent} influence solely by means of complex chains, thereby limiting the variety of causal structures that can be modeled. This is the root of the {\it Kleitman-Rothschild configuration space pathology,} revisited in greater depth in section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} below. However, a greater problem is the way in which the presence of dependent relations clouds the structural picture in causal set theory, hampering important perspectives and technical methods. \subsection{Domains of Influence; Predecessors and Successors; Boundary and Interior}\label{subsecpredecessors} \refstepcounter{textlabels}\label{relpastfuture} {\bf Domains of Influence; Predecessors and Successors.} An iconic feature of Minkowski spacetime diagrams in special relativity is the {\it light cone} of an event $x$, which is the union of its past and future. In terms of causal structure, the light cone of $x$ is the set of all events in Minkowski spacetime that can either influence $x$ or be influenced by $x$. In general relativity, the light cone is superseded in this role by the {\it total domain of influence} $J(x)$ of $x$. The uneasy distinction between which events {\it can} influence $x$ or be influenced by $x$, and which events actually {\it do} exert or experience such influence, arises from the {\it prescriptive} nature of spacetime geometry in general relativity, as mentioned in section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} above. This distinction disappears under the causal metric hypothesis (\hyperref[cmh]{CMH}). More specific domains of influence are defined by taking subsets of $J(x)$. For example, the {\it chronological future} $I^+(x)$ of $x$ is the set of events that may be reached from $x$ by a {\it future-directed timelike curve,} while the {\it causal future} $J^+(x)$ of $x$ is the set of events that may be reached from $x$ by a {\it future-directed causal curve;} i.e., a future-directed curve that is either timelike or null. Chronological pasts $I^-(x)$ and causal pasts $J^-(x)$ are defined in a similar way.\footnotemark\footnotetext{This is standard relativistic notation. In the causal context, I replace the letter $J$ by $D$, for ``domain." Distinct analogues of chronological pasts and futures are not needed in this paper; see, however, the discussion of the {\it interval topology} in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} below.} Relativistic domains of influence play a central role in Malament's metric recovery theorem, discussed in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above. \refstepcounter{textlabels}\label{newbehavior} Despite the conceptual tension between potential and actual influence in general relativity, the mathematical classification of relativistic domains of influence is comparatively simple, due to the reducibility and dependence of relations between events in relativistic spacetime, discussed in section \hyperref[subsectionchains]{\ref{subsectionchains}} above. For more general directed sets and multidirected sets, including those of principal interest in discrete causal theory, the possibilities of irreducibility and independence introduce qualitatively new features, giving rise to more nuanced domains of influence. These include {\it direct pasts and futures,} {\it maximal pasts,} and {\it minimal futures,} introduced in definition \hyperref[predecessors]{\ref{predecessors}} below. More complicated domains of influence may be defined in terms of {\it filtrations} of the past and future of a general subset of a multidirected set, but the simple pointwise definitions given here suffice for the purposes of this paper. The physically suggestive notions of ``past" and ``future" are useful for general multidirected sets, even though the classical spacetime models of principal interest in this paper are merely directed sets. \\ \begin{defi}\label{domains} Let $M=(M,R,i,t)$ be a multidirected set, and let $w$, $x$, and $y$ be elements of $M$. \begin{enumerate} \item If there exists a chain from $w$ to $x$ in $M$, then $w$ is called a {\bf predecessor} of $x$. The set $D^-(x)$ of all predecessors of $x$ is called the {\bf past} of $x$. \item If there exists a chain from $x$ to $y$ in $M$, then $y$ is called a {\bf successor} of $x$. The set $D^+(x)$ of all successors of $x$ is called the {\bf future} of $x$. \item The {\bf total domain of influence} $D(x)$ of $x$ in $M$ is the union $D^-(x)\cup D^+(x)$ of the past and future of $x$. \item A {\bf domain of influence} of $x$ is a subset of $D(x)$. \end{enumerate} \end{defi} Particularly important domains of influence are defined in terms of {\it direct predecessors} and {\it successors, maximal predecessors,} and {\it minimal successors.}\\ \begin{defi}\label{predecessors} Let $M=(M,R,i,t)$ be a multidirected set, and let $x$ be an element of $M$. Let $w$ be a predecessor of $x$, and let $y$ be a successor of $x$. \begin{enumerate} \item If there exists a relation $r$ in $R$ such that $i(r)=w$ and $t(r)=x$, then $w$ is called a {\bf direct predecessor} of $x$. If $r$ is irreducible, then $w$ is called a {\bf maximal predecessor} of $x$. \item The sets of direct predecessors and maximal predecessors of $x$ are denoted by $D_0^-(x)$ and $D_{\text{\tn{\fsz{max}}}}^-(x)$, and are called the {\bf direct past} and the {\bf maximal past} of $x$, respectively. \item If there exists a relation $r$ in $R$ such that $i(r)=x$ and $t(r)=y$, then $y$ is called a {\bf direct successor} of $x$. If $r$ is irreducible, then $y$ is called a {\bf minimal successor} of $x$. \item The sets of direct successors and minimal successors of $x$ are denoted by $D_0^+(x)$ and $D_{\text{\tn{\fsz{min}}}}^+(x)$, and are called the {\bf direct future} and the {\bf minimal future} of $x$, respectively. \end{enumerate} \end{defi} It follows immediately from definition \hyperref[predecessors]{\ref{predecessors}} that $D_{\text{\tn{\fsz{max}}}}^-(x)\subset D_0^-(x)\subset D^-(x)$, and that $D_{\text{\tn{\fsz{min}}}}^+(x)\subset D_0^+(x)\subset D^+(x)$. The direct past and the past of $x$ coincide in the transitive case, by the definition of transitivity, and similarly for the direct future and the future of $x$. In the interpolative case, maximal pasts and minimal futures are always empty. For multidirected sets including cycles, pasts and futures are not necessarily disjoint. For example, if $M$ includes a reflexive relation $x\prec x$, then the element $x$ belongs to its own past and future. The physical interpretation of direct pasts and futures requires the independence convention (\hyperref[ic]{IC}), or a suitable alternative, since without such a convention the physical significance of reducible relations is ambiguous. Figure \hyperref[successors]{\ref{successors}} below illustrates the future, direct future, and minimal future of an element $x$ in a multidirected set, which in this case is a nontransitive acyclic directed set with $35$ elements. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.3)(16.5,5.3) \pgfxyline(0,-.1)(0,5.3) \pgfxyline(5.5,-.1)(5.5,5.3) \pgfxyline(11,-.1)(11,5.3) \pgfxyline(16.5,-.1)(16.5,5.3) \pgfputat{\pgfxy(.15,5)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,5)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,5)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{-.1cm}} \begin{pgftranslate}{\pgfpoint{0.3cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.10cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \color{white} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \color{black} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.11cm} \pgfputat{\pgfxy(2.2,2.45)}{\pgfbox[center,center]{\large{$x$}}} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.8cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.10cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \color{white} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \color{black} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.11cm} \pgfputat{\pgfxy(2.2,2.45)}{\pgfbox[center,center]{\large{$x$}}} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.3cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.10cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node18} \color{white} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \color{black} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.11cm} \pgfputat{\pgfxy(2.2,2.45)}{\pgfbox[center,center]{\large{$x$}}} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Future of an element $x$; b) direct future of $x$; c) minimal future of $x$.} \label{successors} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{boundaryint} {\bf Boundary and Interior.} An element $x$ in a multidirected set $M$ is called an {\bf extremal element} if either its past or its future is empty. A ``Big Bang-like" or ``Big Crunch-like" event might be modeled as an extremal element. ``Free agency" might also be represented in this way. In the context of classical spacetime structure, the set of all extremal elements serves as a {\it temporal boundary.} Spatial boundaries, by contrast, are not fundamental in causal theory. The {\it interior} of a multidirected set is the complement of its boundary. In the context of classical spacetime structure, the interior is where ``ordinary physics" occurs: every event in the interior has causes, and produces effects. These concepts are made precise in definition \hyperref[defboundary]{\ref{defboundary}}. \\ \begin{defi}\label{defboundary} Let $M=(M,R,i,t)$ be a multidirected set. \begin{enumerate} \item The {\bf boundary} $\partial M$ of $M$ is the subset of all extremal elements of $M$; i.e., elements with either empty past or empty future. \item The {\bf interior} $\tn{Int}(M)$ of $M$ is the subset of all nonextremal elements of $M$; i.e., the complement of the boundary of $M$. \end{enumerate} \end{defi} Figure \hyperref[boundary]{\ref{boundary}}a below illustrates the boundary of a multidirected set $M$. Figure \hyperref[boundary]{\ref{boundary}}b illustrates the interior of $M$. It is sometimes useful to regard domains of influence of elements in a multidirected $M$, as well as its boundary and interior, as {\it subobjects} of $M$, rather than merely structureless subsets of $M$. This requires specifying sets of relations to complement the sets of elements specified in definitions \hyperref[predecessors]{\ref{predecessors}} and \hyperref[defboundary]{\ref{defboundary}}. Unless stated otherwise, {\it all} relations between pairs of elements in the specified subset are included;\footnotemark\footnotetext{Of course, the only possible relations in the boundary $\partial M$ are relations from minimal to maximal elements.} the resulting subobjects are called {\bf full subobjects.} For example, the interior $\tn{Int}(M)$ of $M=(M,R,i,t)$, viewed as a subobject of $M$, consists of all nonextremal elements of $M$, {\it together with} all relations in $M$ between pairs of nonextremal elements. This convention is important for interpreting some of the results of later sections, such as theorem \hyperref[theoreminterior]{\ref{theoreminterior}}. Figure \hyperref[boundary]{\ref{boundary}}c below illustrates the interior of a multidirected set $M$, viewed as a full subobject. The domains of influence illustrated in figure \hyperref[successors]{\ref{successors}} above are also shown as full subobjects. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.3)(16.5,5.3) \pgfxyline(0,-.1)(0,5.3) \pgfxyline(5.5,-.1)(5.5,5.3) \pgfxyline(11,-.1)(11,5.3) \pgfxyline(16.5,-.1)(16.5,5.3) \pgfputat{\pgfxy(.15,5)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,5)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,5)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{-.1cm}} \begin{pgftranslate}{\pgfpoint{0.3cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.8cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.3cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Boundary $\partial M$ of a multidirected set $M$; b) interior $\tn{Int}(M)$ of $M$; c) $\tn{Int}(M)$ as a full subobject of $M$.} \label{boundary} \end{figure} \vspace*{-.5cm} \subsection{Order Theory; Category Theory; Influence of Grothendieck}\label{settheoretic} Two broad structural paradigms used in this paper are {\it generalized order theory} and {\it category theory.} Generalized order theory includes all the structured sets considered in this paper, including ordered sets, partially ordered sets, acyclic directed sets, directed sets, and multidirected sets. In this section, I discuss a few order-theoretic subtleties involving the cardinalities of certain isomorphism classes of acyclic directed sets, and the properties of order extensions and linear orders. These subtleties are relevant to the theory of {\it relative multidirected sets over a fixed base,} introduced in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}, and the theory of {\it kinematic schemes,} introduced in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. Category theory, conceived in a pure mathematical context, plays a dual role in this paper, both as a general mathematical framework, and as a novel structural analogy to causal theory. Important aspects of this analogy already appear in Chris Isham's {\it topos-theoretic approach} to quantum gravity. Causal theory returns the favor to pure mathematics by suggesting structural paradigms similar to, but distinct from, category theory. This section concludes with a discussion of how the mathematical work of Alexander Grothendieck influences the conceptual framework of this paper, focusing on the {\it relative viewpoint} (\hyperref[rv]{RV}), and the notion of {\it hidden structure} (\hyperref[hs]{HS}). {\bf Counting Finite Acyclic Directed Sets; Countable Acyclic Directed Sets.}\refstepcounter{textlabels}\label{isomclassesfinite} I begin this section by establishing the cardinalities of a few distinguished classes of acyclic directed sets, considered up to isomorphism. These classes serve as the ``object" classes of the ``smallest physically relevant kinematic schemes," considered in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper. This type of computation may be easily generalized to determine the cardinalities of isomorphism classes of multidirected sets with cardinality bounds on their sets of elements and relations, as well as the cardinalities of isomorphism classes of {\it power spaces,} discussed in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} below. \refstepcounter{textlabels}\label{isomclassescountable} Since a finite set admits only a finite number of binary relations, the class $\ms{A}_{\tn{\fsz{fin}}}$ of finite acyclic directed sets, considered up to isomorphism, is a set of cardinality $\aleph_0$, where $\aleph_0$ is the unique countably infinite cardinal. The class $\ms{A}_{\aleph_0}$ of countable acyclic directed sets, considered up to isomorphism, is a set of continuum cardinality. To see this, first note that the unit continuum interval $[0,1]$, and hence the continuum itself, may be placed in bijective correspondence with a subclass of $\ms{A}_{\aleph_0}$. One way of doing this is illustrated in figure \hyperref[encodingpi]{\ref{encodingpi}} below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{2.4cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,2.5)(16.5,2.5) \pgfxyline(0,-.1)(0,2.5) \pgfxyline(16.5,-.1)(16.5,2.5) \begin{pgftranslate}{\pgfpoint{0cm}{-.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4,.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(7,.5)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(10,.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(13,.5)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(16,.5)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(2.5,1.25)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(5.5,1.25)}{0.10cm} \pgfnodecircle{Node9}[fill]{\pgfxy(8.5,1.25)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(11.5,1.25)}{0.10cm} \pgfnodecircle{Node11}[fill]{\pgfxy(14.5,1.25)}{0.10cm} \pgfnodecircle{Node011}[fill]{\pgfxy(1,1.5)}{0.10cm} \pgfnodecircle{Node021}[fill]{\pgfxy(3.1,1.5)}{0.10cm} \pgfnodecircle{Node022}[fill]{\pgfxy(3.7,1.5)}{0.10cm} \pgfnodecircle{Node023}[fill]{\pgfxy(4.3,1.5)}{0.10cm} \pgfnodecircle{Node024}[fill]{\pgfxy(4.9,1.5)}{0.10cm} \pgfnodecircle{Node031}[fill]{\pgfxy(7,1.5)}{0.10cm} \pgfnodecircle{Node041}[fill]{\pgfxy(9,1.5)}{0.10cm} \pgfnodecircle{Node042}[fill]{\pgfxy(9.5,1.5)}{0.10cm} \pgfnodecircle{Node043}[fill]{\pgfxy(10,1.5)}{0.10cm} \pgfnodecircle{Node044}[fill]{\pgfxy(10.5,1.5)}{0.10cm} \pgfnodecircle{Node045}[fill]{\pgfxy(11,1.5)}{0.10cm} \pgfnodecircle{Node051}[fill]{\pgfxy(11.8,1.5)}{0.10cm} \pgfnodecircle{Node052}[fill]{\pgfxy(12.1,1.5)}{0.10cm} \pgfnodecircle{Node053}[fill]{\pgfxy(12.4,1.5)}{0.10cm} \pgfnodecircle{Node054}[fill]{\pgfxy(12.7,1.5)}{0.10cm} \pgfnodecircle{Node055}[fill]{\pgfxy(13,1.5)}{0.10cm} \pgfnodecircle{Node056}[fill]{\pgfxy(13.3,1.5)}{0.10cm} \pgfnodecircle{Node057}[fill]{\pgfxy(13.6,1.5)}{0.10cm} \pgfnodecircle{Node058}[fill]{\pgfxy(13.9,1.5)}{0.10cm} \pgfnodecircle{Node059}[fill]{\pgfxy(14.2,1.5)}{0.10cm} \pgfnodecircle{Node00}[fill]{\pgfxy(15.4,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(15.55,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(15.7,1.25)}{0.03cm} \pgfnodeconnline{Node1}{Node011} \pgfnodeconnline{Node2}{Node021} \pgfnodeconnline{Node2}{Node022} \pgfnodeconnline{Node2}{Node023} \pgfnodeconnline{Node2}{Node024} \pgfnodeconnline{Node3}{Node031} \pgfnodeconnline{Node4}{Node041} \pgfnodeconnline{Node4}{Node042} \pgfnodeconnline{Node4}{Node043} \pgfnodeconnline{Node4}{Node044} \pgfnodeconnline{Node4}{Node045} \pgfnodeconnline{Node5}{Node051} \pgfnodeconnline{Node5}{Node052} \pgfnodeconnline{Node5}{Node053} \pgfnodeconnline{Node5}{Node054} \pgfnodeconnline{Node5}{Node055} \pgfnodeconnline{Node5}{Node056} \pgfnodeconnline{Node5}{Node057} \pgfnodeconnline{Node5}{Node058} \pgfnodeconnline{Node5}{Node059} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node2}{Node8} \pgfnodeconnline{Node3}{Node8} \pgfnodeconnline{Node3}{Node9} \pgfnodeconnline{Node4}{Node9} \pgfnodeconnline{Node4}{Node10} \pgfnodeconnline{Node5}{Node10} \pgfnodeconnline{Node5}{Node11} \pgfnodeconnline{Node6}{Node11} \pgfmoveto{\pgfxy(.6,1.7)} \pgfcurveto{\pgfxy(.5,2)}{\pgfxy(1,1.6)}{\pgfxy(1,2)} \pgfcurveto{\pgfxy(1,1.6)}{\pgfxy(1.5,2)}{\pgfxy(1.4,1.7)} \pgfstroke \pgfputat{\pgfxy(1,2.2)}{\pgfbox[center,center]{$1$}} \pgfmoveto{\pgfxy(2.9,1.7)} \pgfcurveto{\pgfxy(2.8,2)}{\pgfxy(4,1.6)}{\pgfxy(4,2)} \pgfcurveto{\pgfxy(4,1.6)}{\pgfxy(5.2,2)}{\pgfxy(5.1,1.7)} \pgfstroke \pgfputat{\pgfxy(4,2.2)}{\pgfbox[center,center]{$4$}} \pgfmoveto{\pgfxy(6.6,1.7)} \pgfcurveto{\pgfxy(6.5,2)}{\pgfxy(7,1.6)}{\pgfxy(7,2)} \pgfcurveto{\pgfxy(7,1.6)}{\pgfxy(7.5,2)}{\pgfxy(7.4,1.7)} \pgfstroke \pgfputat{\pgfxy(7,2.2)}{\pgfbox[center,center]{$1$}} \pgfmoveto{\pgfxy(8.8,1.7)} \pgfcurveto{\pgfxy(8.7,2)}{\pgfxy(10,1.6)}{\pgfxy(10,2)} \pgfcurveto{\pgfxy(10,1.6)}{\pgfxy(11.3,2)}{\pgfxy(11.2,1.7)} \pgfstroke \pgfputat{\pgfxy(10,2.2)}{\pgfbox[center,center]{$5$}} \pgfmoveto{\pgfxy(11.6,1.7)} \pgfcurveto{\pgfxy(11.5,2)}{\pgfxy(13,1.6)}{\pgfxy(13,2)} \pgfcurveto{\pgfxy(13,1.6)}{\pgfxy(14.5,2)}{\pgfxy(14.4,1.7)} \pgfstroke \pgfputat{\pgfxy(13,2.2)}{\pgfbox[center,center]{$9$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{Portion of a locally finite causal set encoding the decimal $\pi-3=.14159...$} \label{encodingpi} \end{figure} \vspace*{-.5cm} Anticipating the introduction of {\it local finiteness} (\hyperref[lf]{LF}) in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, the construction illustrated in figure \hyperref[encodingpi]{\ref{encodingpi}} involves only the class $\ms{C}_{\tn{\fsz LF}}$ of isomorphism classes of {\it locally finite causal sets}. The larger classes $\ms{A}_{\aleph_0}$, $\ms{C}$, and $\ms{A}_{\aleph_0,\tn{\fsz LF}}$ of isomorphism classes of countable acyclic directed sets, causal sets, and countable locally finite acyclic directed sets, respectively, are therefore {\it at least} as large as the continuum. Now let $S_{\aleph_0}$ be a {\it fixed countably infinite set}. The class of binary relations on $S_{\aleph_0}$ includes a representative of every infinite member of $\ms{A}_{\aleph_0}$; the finite members may be ignored since they are countable in number. Hence, $\ms{A}_{\aleph_0}$ corresponds bijectively with a subset of the power set of $S_{\aleph_0}\times S_{\aleph_0}$, which has cardinality $2^{\aleph_0}$, the cardinality of the continuum.\footnotemark\footnotetext{Under the {\it continuum hypothesis,} $2^{\aleph_0}$ is equal to the smallest uncountable cardinal $\aleph_1$.} Therefore, the four classes $\ms{C}_{\tn{\fsz LF}}$, $\ms{A}_{\aleph_0}$, $\ms{C}$, and $\ms{A}_{\aleph_0,\tn{\fsz LF}}$ also have cardinality {\it at most} as large as the continuum. \refstepcounter{textlabels}\label{orderext} {\bf Linear Orders; Order Extensions.} {\it Linear orders} are, in a sense, the simplest nontrivial orders. For this reason, it is often useful to {\it compare} multidirected sets, particularly acyclic directed sets, to linearly ordered sets. The theory of {\it order extensions} provides a means of formalizing such comparisons, by extending a partial order to a linear order. Sorkin and Rideout's theory of sequential growth dynamics of causal sets \cite{SorkinSequentialGrowthDynamics99}, already mentioned in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, illustrates the importance of linear extensions, and their corresponding morphisms, in causal set theory. In particular, a {\it total labeling} $\{x_0,x_1,x_2,...\}$ of a causal set $(C,\prec)$ is equivalent to a bijective morphism from $(C,\prec)$ into a linearly ordered subset of the nonnegative integers $\mathbb{N}$. In this paper, it is necessary to compare acyclic directed sets to more general linearly ordered sets, particularly those with {\it discrete linear orders.} Such sets are {\it locally isomorphic} to the integers, except possibly at extremal elements. This topic is discussed in more detail in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}} below. Let $(A,\prec)$ be an acyclic directed set. The transitive closure of $A$ is equivalent to a partially ordered set, via the equivalence introduced in sections \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} and \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} above. Regarding the binary relation $\prec$ on $A$ as a subset of the Cartesian product $A\times A$, an {\bf extension} of $\prec$ is a subset $\prec'$ of $A\times A$ containing $\prec$. In terms of individual relations between pairs of elements of $A$, $\prec'$ is an extension of $\prec$ if and only if each relation $x\prec y$ implies the relation $x\prec'y$. Abusing terminology in order to remain in the acyclic context, a {\bf linear order} on $A$ is an transitive acyclic binary relation $\prec''$ on $A$, such that either $x\prec''y$ or $y\prec''x$ for every pair of {\it distinct} elements $x$ and $y$ in $A$. Equivalently, a linear order is a {\it total} order, in the sense that it admits no extension. If $(A,\prec)$ is any acyclic directed set, an extension of $\prec$ to a linear order $\prec''$ on $A$ is equivalent to a bijective morphism $(A,\prec)\rightarrow (A,\prec'')$. The existence of linear extensions of acyclic binary relations is surprisingly subtle. The transitive analogue of the following {\it order-extension principle} is sometimes used as an {\it axiom} of modern set theory: \refstepcounter{textlabels}\label{oep} \hspace*{.3cm} OEP. {\bf Order Extension Principle:} {\it Every acyclic binary relation $\prec$ on a set $A$ \\ \hspace*{1.4cm} may be extended to a linear order on $A$.} In Zermelo-Fraenkel set theory, the order extension principle may be {\it proven} using the power set axiom and Zorn's lemma. The {\it power set axiom} states that the power set $\ms{P}(A)$ of a set $A$ is itself a set, rather than a {\it proper class}; i.e., a class ``too large" to be a set.\footnotemark\footnotetext{As von Neumann and others have pointed out, ``size" is a poor way to understand the distinction between sets and proper classes, but no great sophistication is needed in the present context.} {\it Zorn's lemma} is an {\it axiom} closely related to the {\it axiom of choice;} it states that any partially ordered set satisfying the property that every chain is bounded above has at least one maximal element. Zorn's lemma may be applied to the class of extensions of a transitive acyclic binary relation $\prec$ on $A$, ordered by inclusion, which is {\it itself} a set, by an elementary application of the power set axiom. The union of the elements of any chain of extensions of $\prec$ serves as an upper bound for the chain, and the maximal element guaranteed by Zorn's lemma is a linear order on $A$, generally {\it not} unique. {\bf Category Theory.}\refstepcounter{textlabels}\label{cat} In sections \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} and \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} of this paper, I introduced the specific categories of causal sets, acyclic directed sets, directed sets, and multidirected sets. Here, I briefly discuss categories in general, and outline the use of category theory in this paper. A {\bf category} $\ms{C}$ consists of a class of objects $\tn{Obj}(\ms{C})$, and a class of morphisms $\tn{Mor}(\ms{C})$, such that every morphism $\gamma$ has an initial object $\mbf{i}(\gamma)$ and a terminal object $\mbf{t}(\gamma)$, every object $C$ has an identity morphism $\gamma_C$,\footnotemark\footnotetext{The property of $\gamma_C$ distinguishing it as an identity morphism is that it is {\it absorbed under composition;} i.e., for any morphisms $\gamma:B\rightarrow C$ and $\gamma':C\rightarrow D$, $\gamma_C\circ\gamma=\gamma$ and $\gamma'\circ\gamma_C=\gamma'$.} and morphisms compose associatively. For many categories, objects possess {\it internal structure,} and morphisms are chosen to {\it preserve structure} in some way. For example, morphisms of multidirected sets preserve initial and terminal elements. However, there are also important cases in which the objects of a category do {\it not} possess internal structure, and all the structure of the category resides in its morphisms. An example is the category-theoretic approach to {\it groups}, in which group elements are represented by endomorphisms of a single structureless object. \refstepcounter{textlabels}\label{ishamtopos}Category theory plays two distinct roles in this paper, the first conventional, and the second relatively novel. First, category theory serves as a {\it general structural paradigm;} i.e., a convenient method of organizing information about mathematical objects and relationships between pairs of objects. Category theory is used in this way in many, if not most, mathematical papers involving modern algebra. Second, structures of particular interest in causal theory, including multidirected sets and kinematic schemes, {\it closely resemble categories.} This leads to interesting interplay between causal theory and category theory. Chris Isham and his collaborators have carried this analogy very far in a particular direction, organizing families of causal sets, as well as many other types of physical models, into special categories called {\it topoi,} in an effort to build up a general {\it topos-theoretic} framework for physics. Though Isham's work is one of the major motivations for this paper, my perspective on the role of category theory in physics is somewhat different. Consider the following two questions: \hspace*{.3cm}{\bf Question 1:} {\it How can existing structural notions in mathematics, such as category theory, \\\hspace*{.3cm}be applied to physics?} \hspace*{.3cm}{\bf Question 2:} {\it Does existing mathematical theory supply adequate structural notions for \\\hspace*{.3cm} theoretical physics, and if not, what more can physics teach us about structural notions?} \refstepcounter{textlabels}\label{differentnotions}Isham's program seems to focus mainly on the first question, with the principal ``existing structural notion" being topos theory. To be sure, Isham considers the second question as well; for example, in his recent article {\it Topos Methods in the Foundations of Physics} \cite{Isham11}, he points out the dangers of taking for granted} the real and complex numbers in the context of quantum gravity, proposes replacing the notion of probabilities with {\it ``generalized truth values,"} and suggests that {\it ``conventional quantum formalism is inadequate to the task of quantum gravity."} All these points are well taken. However, Isham advocates topos-theoretic solutions to these issues, rather than new, physically motivated, structural paradigms. My own view is that category theory and topos theory may not be ideally suited to fundamental physics, and that causal theory suggests {\it essentially different} structural notions that are are worth studying in their own right. Category theory and topos theory arose from the fields of homological algebra and algebraic geometry between 1940 and 1960, in response to specific mathematical issues only distantly related to fundamental spacetime structure. These issues include the essential uniqueness of certain {\it cohomology theories,} and the assignment of algebraic structure to generalized topological spaces. While sufficiently rich and general to {\it accommodate} any conceivable physical theory, categories and topoi may not always be {\it optimal} or {\it natural} for this purpose. Indeed, the clear physical relevance of essentially different structures playing similar roles, such as kinematic schemes in causal theory, suggests that alternative structural notions may be more appropriate in certain physical contexts. Such alternatives already appear, in a variety of mathematical settings, as ``enriched" or ``categorified" graphs or quivers, but are seldom accorded broad structural significance in their own right. However, the legacy of interaction between mathematics and physics suggests than any concept of fundamental physical significance should be taken very seriously, even from a purely mathematical perspective.\footnotemark\footnotetext{One of many instances of the continued mathematical potency of physically motivated methods, setting aside the question of their actual physical success, is the interplay between string theory and algebraic geometry/topology, for example, in {\it Gromov-Witten theory,} {\it ``Monstrous Moonshine,"} {\it mirror symmetry,} and so on.} This paper contains several examples of ``category-like" structures, arising naturally in causal theory, for which na\"{i}ve category-theoretic intuition {\it leads in the wrong direction.} Examples include {\it``objects without identity morphisms,"} and {\it ``morphisms which do not compose,"} appearing repeatedly in the theory of multidirected sets, relation spaces, and power spaces; {\it ``morphisms without initial and terminal objects,"} appearing in the theory of {\it causal path semicategories,} discussed in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}; and {\it ``morphisms whose compositions are families of morphisms rather than single morphisms,"} appearing in the theory of {\it co-relative histories,} discussed in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}}. These examples provide evidence that the analysis of basic structural questions is still very interesting and open-ended in theoretical physics, not merely an exercise in choosing and applying familiar mathematical tools. {\bf Influence of Grothendieck.}\refstepcounter{textlabels}\label{groth} Many of the structural ideas appearing in this paper, both conventional and novel, have {\it algebraic} roots in the work of Alexander Grothendieck, the father of modern algebraic geometry. It was Grothendieck, along with his student Verdier, who introduced topos theory around 1960. Most of the topics to which Grothendieck applied his methods play little direct role in discrete causal theory,\footnotemark\footnotetext{{\it Noncommutative} geometry provides methods for studying the {\it causal path algebras} introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}. Ioannis Raptis discusses the application of noncommutative geometry to {\it incidence algebras} of causal sets. However, noncommutative geometry took its modern shape beginning around 1980, after Grothendieck's most productive period.} but his {\it broad structural concepts} may be applied in a variety of useful ways in this context. Two such concepts used repeatedly in this paper are the {\it relative viewpoint,} and the principle of {\it hidden structure.}\footnotemark\footnotetext{The relative viewpoint is well-known as such. {\it Hidden structure} is my own terminology.} \refstepcounter{textlabels}\label{rv} \hspace*{.3cm}RV. \hspace*{.1cm}{\bf Relative Viewpoint.} {\it It is often more natural, and more informative, to study relationships \\\hspace*{1.1cm}between objects of a certain type, than to study such objects individually.} Grothendieck applied the relative viewpoint to his work on {\it commutative algebraic varieties and schemes,} which are geometric objects defined locally by the vanishing of polynomials. In this context, one fixes a {\it base scheme} $Z$, and considers the {\it category of schemes over} $Z$; i.e., the category whose objects are morphisms of schemes $X\rightarrow Z$. The signature achievement of this method was Grothendieck's proof of the {\it Grothendieck-Riemann-Roch theorem.}\footnotemark\footnotetext{This theorem may be viewed as a special case of the {\it Atiyah-Singer index theorem.} It has become so prominent in string theory that at least one string theorist has proposed renaming it the ``Atiyah-Singer string index theorem."} It has since become ubiquitous in modern algebra. The first significant application of the relative viewpoint in this paper comes in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}, where I use it to study {\it relative multidirected sets over a fixed base.} For example, causal sets may be viewed as relative directed sets over the integers. In section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, the relative viewpoint underlies the theory of {\it relation space,} which leads to the technical methods of abstract quantum causal theory introduced in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}. In section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}}, the relative viewpoint serves as the conceptual foundation of the theory of {\it co-relative histories,} which formalize the idea of ``spacetime evolution." Although the relative viewpoint often involves the study of morphisms in some category, more nuanced versions are sometimes useful. For example, co-relative histories are {\it not} morphisms in the category of directed sets. The relative viewpoint is generally distinct from the long-standing physical and metaphysical philosophy of {\it relationism,} but nontrivial connections between the two ideas arise in the special context of causal theory, particularly in the theory of relation space. \refstepcounter{textlabels}\label{hs} \hspace*{.3cm}HS. \hspace*{.1cm}{\bf Hidden Structure.} {\it Many objects presented as structured sets may be more profitably viewed \\\hspace*{1.1cm}as enlarged sets with ``extra elements," naturally induced by their native structures.} Grothendieck used the idea of hidden structure to construct algebraic schemes, by adding new elements of positive dimension to the sets of zero-dimensional points of algebraic varieties. Since schemes are more natural and better behaved than varieties in many respects, the ``promotion of varieties to schemes" may be viewed as a belated recognition of hidden structure latent in varieties, unnoticed for more than two millennia. In section \hyperref[subsectiontopology]{\ref{subsectiontopology}} of this paper, I treat the {\it relations} in a multidirected set $M=(M,R,i,t)$ as elements of a topological space called the {\it star model} of $M$, which plays an important role in the study of {\it causal locality.} As a set, the star model is the {\it union} of $M$ and its relation set $R$. In section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, I elevate the relation set $R$ to the {\it relation space} $\ms{R}(M)$ over $M$, already mentioned above in the context of the relative viewpoint. Hidden structure is further exploited in the context of power spaces, causal path spaces, and kinematic schemes, discussed in sections \hyperref[sectionbinary]{\ref{sectionbinary}} and \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below. Much of Grothendieck's mathematical output came during a period of relative estrangement between fundamental physics and abstract mathematics, which continued until the string theory revolutions, Alain Connes' work on noncommutative geometry, and other developments during the 1980's and 1990's. With a few exceptions,\footnotemark\footnotetext{For example, Roger Penrose introduced twistor theory in the middle 1960's.} physicists during this interregnum relied on mathematical notions conceived in the century from Riemann to Weyl. Even today, the structural ideas of Grothendieck and his contemporaries are not as broadly appreciated in the physics community as one might wish.\footnotemark\footnotetext{As late as 2005, Chris Isham felt compelled to essentially {\it apologize} for using methods as abstract as category theory. See \cite{IshamQuantisingI05}, pages 3, 10.} This unfortunate state of affairs may be attributed partly to a legacy of physically unilluminating exposition emanating from certain mathematical sources,\footnotemark\footnotetext{For example, the {\it Bourbaki} group, to which Grothendieck briefly contributed. One of the principal scribes for {\it Boubaki} was Jean Dieudonn\'{e}, who also penned Grothendieck's monumental {\it \'{E}l\'{e}ments de G\'{e}om\'{e}trie Alg\'{e}brique} (EGA).} and partly to ``pragmatism" in the physics community during the heyday of experimental particle physics. \newpage \section{Transitivity, Independence, and the Causal Preorder}\label{sectiontransitivity} Transitivity (\hyperref[tr]{TR}) is deeply ingrained in ordinary thought as part of the {\it meaning} of classical causal structure. For example, the relation between grandparent and grandchild, though usually indirect, is recognized as causal, due to transitivity, as illustrated by its role in the so-called {\it grandfather paradox.} Challenging this convention would only incite terminological controversy, to no good purpose. Hence, rather than disputing the notion of a transitive causal order, I introduce a more fundamental {\it nontransitive} binary relation, which I refer to as the {\it causal preorder}, from which the transitive causal order may be derived. Here, it is expedient to use the term {\it order} in a generalized sense, as explained at the beginning of section \hyperref[settheoretic]{\ref{settheoretic}} above. In particular, ``causal orders" may contain cycles, analogous to closed timelike curves in general relativity. The causal preorder encodes only {\it independent} influences between pairs of events, and {\it generates} the causal order under the operation of {\it transitive closure.} Using the independence convention (\hyperref[ic]{IC}), a causal preorder may be unambiguously represented by a directed set in the context of discrete causal theory. In section \hyperref[subsectionmodes]{\ref{subsectionmodes}} below, I present the basic problem of transitive binary relations: they fail to distinguish among classical causal structures involving {\it multiple independent modes of influence} between pairs of events. I discuss two of the few treatments of this issue appearing in the literature, due to David Finkelstein and Ioannis Raptis, and relate Rafael Sorkin's current view of the subject, via private communication. In section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}}, I give six arguments that transitivity represents a major information-theoretic deficiency in causal set theory. In section \hyperref[subsectionpreorder]{\ref{subsectionpreorder}}, I introduce the causal preorder. Finally, in section \hyperref[subsectiontransitiveclosure]{\ref{subsectiontransitiveclosure}}, I introduce the {\it transitive closure functor,} which converts a causal preorder into a transitive causal order, and the {\it skeleton operation,} a ``partial inverse" of the transitive closure, which annihilates reducible relations. Throughout section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, I focus on the special case of directed sets, since the intended context is the structure of classical spacetime. In particular, every consideration of significant importance in this section manifests itself already in the acyclic directed case. The principal conclusion is that {\it causal sets are information-theoretically inadequate to model classical spacetime structure in discrete causal theory, even under the assumption of acyclicity.} It is safe, in any case, to omit multidirected sets from the present discussion, since the existence of multiple independent ``modes of influence" is obvious from the very beginning in the settings in which such sets arise in discrete causal theory, such as in the {\it ``decategorification"} of a kinematic scheme. These concepts are clarified and expanded in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper. \subsection{Independent Modes of Influence}\label{subsectionmodes} \refstepcounter{textlabels}\label{transproblem} {\bf The Basic Problem.} Recognition of a nontransitive binary relation underlying the transitive causal order is motivated by the possibility of multiple independent ways in which one event may influence another, called {\bf modes of influence}. A simple example of this is illustrated in figure \ref{transitivity} below. Consider three events, labeled $x$, $y$, and $z$. If $x$ influences $z$, this influence may be either exclusively independent of $y$, as illustrated in \ref{transitivity}a, exclusively mediated by $y$, as illustrated in \ref{transitivity}b, or it may involve {\it both independent influence and influence mediated by} $y$, as illustrated in \ref{transitivity}c. In accordance with the independence convention (\hyperref[ic]{IC}), the relations indicated by solid lines in the figure all represent independent influences; the dashed line in figure \ref{transitivity}b indicates the {\it hypothetical} relation $x\prec z$ prescribed under the axiom of transitivity (\hyperref[tr]{TR}). Causal set theory, which does {\it not} employ the independence convention, and which {\it does} impose transitivity, requires the presence of this relation, and therefore cannot distinguish between the scenarios illustrated in figures \ref{transitivity}b and \ref{transitivity}c. {\it A priori}, this is a serious information-theoretic deficiency. This situation extends in an obvious way to any relation in a multidirected set admitting a reducing chain. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{4cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,4)(16.5,4) \pgfxyline(0,-.1)(0,4) \pgfxyline(5,-.1)(5,4) \pgfxyline(11.5,-.1)(11.5,4) \pgfxyline(16.5,-.1)(16.5,4) \pgfputat{\pgfxy(.15,3.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.15,3.7)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.65,3.7)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(3,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4,1.75)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(3,3)}{0.1cm} \pgfputat{\pgfxy(3,.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(4.3,1.75)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(3,3.3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.5,2.15)}{\pgfbox[center,center]{\small{influence}}} \pgfputat{\pgfxy(1.5,1.72)}{\pgfbox[center,center]{\small{independent}}} \pgfputat{\pgfxy(1.5,1.35)}{\pgfbox[center,center]{\small{of $y$}}} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.75cm}{.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(3.2,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4.2,1.75)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(3.2,3)}{0.1cm} \pgfputat{\pgfxy(3.2,.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(4.5,1.75)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(3.2,3.3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1.5,2.2)}{\pgfbox[center,center]{\small{relation}}} \pgfputat{\pgfxy(1.5,1.8)}{\pgfbox[center,center]{\small{prescribed by}}} \pgfputat{\pgfxy(1.5,1.4)}{\pgfbox[center,center]{\small{transitivity}}} \pgfputat{\pgfxy(5.8,2.2)}{\pgfbox[center,center]{\small{influence}}} \pgfputat{\pgfxy(5.8,1.8)}{\pgfbox[center,center]{\small{mediated}}} \pgfputat{\pgfxy(5.8,1.4)}{\pgfbox[center,center]{\small{by $y$}}} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfmoveto{\pgfxy(4.5,3)} \pgfcurveto{\pgfxy(5,3.1)}{\pgfxy(4.4,1.75)}{\pgfxy(4.9,1.75)} \pgfstroke \pgfmoveto{\pgfxy(4.9,1.75)} \pgfcurveto{\pgfxy(4.4,1.75)}{\pgfxy(5,.4)}{\pgfxy(4.5,.5)} \pgfstroke \begin{pgfscope} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfxyline(3.2,.5)(3.2,3) \end{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.6,1.75)(3.1,1.75) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10.4cm}{.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(4,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(5,1.75)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(4,3)}{0.1cm} \pgfputat{\pgfxy(4,.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(5.3,1.75)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(4,3.3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(2.6,1.95)}{\pgfbox[center,center]{\small{both modes}}} \pgfputat{\pgfxy(2.6,1.55)}{\pgfbox[center,center]{\small{of influence}}} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{Modes of influence between related elements in a three-element directed set.} \label{transitivity} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{infoanalogies} {\bf Na\"{i}ve Analogies from Information Theory.} A cryptographic analogy is illustrative of this deficiency. Consider the problem of sending a secret message from $x$ to a remote receiver at $z$, either independently, as in figure \hyperref[transitivity]{\ref{transitivity}}a, or via an intermediary at $y$, as in figure \hyperref[transitivity]{\ref{transitivity}}b. The intermediary can prevent third-party tampering, but may not be trustworthy. One possible strategy is to send part of the message independently, and the other part via the intermediary, as in figure \hyperref[transitivity]{\ref{transitivity}}c, in such a way that its meaning can only be deciphered by combining both parts. Transitive relations are clearly deficient for describing these three scenarios, since they fail to distinguish between the scenarios illustrated in figures \hyperref[transitivity]{\ref{transitivity}}b and \hyperref[transitivity]{\ref{transitivity}}c. Transitive relations can encode {\it which} pairs of events communicate, but cannot encode {\it how} they communicate. This problem only becomes worse for larger families of events, in which there may be many independent modes of influence between a given pair of events. The consequences of ignoring such independence may be illustrated by imagining an academic conference, at which the chair announces that only {\it one} member need return the following year, since the information brought to the second meeting by all the other members is sure to be redundant! \refstepcounter{textlabels}\label{finkelstein} {\bf Finkelstein's Causal Nets.} The foregoing ideas are so elementary, and the reasonable alternatives so few, that the same basic steps have surely been followed many times previously. However, it is surprisingly difficult to find explicit treatments of this issue in the literature, at least in the context of classical spacetime structure. Here I briefly detail a few instances that come as close as any I could find to addressing this problem. David Finkelstein, in his 1988 paper {\it ``Superconducting" Causal Nets} \cite{Finkelstein88}, writes, \begin{quotation} {\noindent\it``...the \tn{[transitive]} causal relation $x\mbf{C}y$ is not local, but may hold for events as far apart as the birth and death of the universe. Since we have committed ourselves to local variables, we abandon $\mbf{C}$ for a local causal relation $\mbf{c}$..." }(page 476) \end{quotation} and later, \begin{quotation} {\noindent\it``We call $\mbf{c}$ the (causal) connection relation, understanding that this connection is... ...immediate ($x\mbf{c}y$ and $y\mbf{c}z$ do not necessarily imply $x\mbf{c} z$). Events in a continuum theory have no immediate causal relations..."}(page 477) \end{quotation} and again, \begin{quotation} {\noindent\it``Every causal set $\mbf{C}$ may also be regarded as a causal net $\mbf{c}$, with the connection $x\mbf{c}y$ defined to hold if and only if \tn{[the relation $x\mbf{C}y$ holds]} and no event $z$ exists with $x\mbf{C}z\mbf{C}y$... ... but \tn{[causal nets]} are more general than causal sets."}(page 477) \end{quotation} These remarks suggest that Finkelstein's principal objections to transitive binary relations were\footnotemark\footnotetext{I say ``were," rather than ``are," because Finkelstein suggested to me, via private communication \cite{FinkelsteinPrivate13}, that his views on the topic have changed considerably since 1988.} based on issues of {\it locality} rather than independence, although the greater generality afforded by dropping the axiom of transitivity (\hyperref[tr]{TR}), noted in the final quote above, is information-theoretically meaningful only if reducible relations can encode independent influence. Finkelstein's locality-based critique of transitivity is valid, and in my opinion, unanswerable, though the structural problems in causal set theory involving locality do not arise from transitivity alone. This topic is elaborated in section \hyperref[sectioninterval]{\ref{sectioninterval}} below. Due to Finkelstein's use of a local causal relation, his causal nets share a greater abstract similarity with the directed sets used in this paper to model classical spacetime than Sorkin's causal sets do, at least at the level of individual objects. However, the overall approach developed here is much closer to causal set theory, particularly in the quantum-theoretic context. \refstepcounter{textlabels}\label{raptisquant} {\bf Raptis' Algebraic Quantization of Causal Sets.} A few relatively recent papers examining the use of nontransitive causal relations in the study of fundamental spacetime structure may also be found in the literature. Ioannis Raptis, in his 2000 paper {\it Algebraic Quantization of Causal Sets} \cite{RaptisAlgebraicQuantization00}, deprecates transitive binary relations as unsuitable for modeling {\it quantum} spacetime, writing that {\it ``the physical connection between events in the quantum deep should be one connecting nearest-neighboring events."} He cites Finkelstein's paper \cite{Finkelstein88} as the source of this idea. However, Raptis uses {\it irreducible} relations in his ``quantum causal sets," thereby sidestepping the issue of independence. Also, the objections to transitivity outlined above apply not only in the {\it ``quantum deep,"} but also at the classical level. \refstepcounter{textlabels}\label{sorkincurrent} {\bf Sorkin's Current View.} An interesting {\it historical} point, regarding the alternative between transitive and nontransitive causal relations in the study of fundamental spacetime structure, is that Finkelstein \cite{Finkelstein88} {\it credits Sorkin} with suggesting the idea of the nontransitive local ``causal connection relation" $\mbf{c}$ appearing in Finkelstein's paper, via private communication in 1987! This indicates that Sorkin has considered, and to some extent encouraged, the development of {\it both} approaches. Sorkin recently expressed to me his current view of the matter in the following remarks, reproduced here with his permission: \begin{quotation} {\noindent\it``I suppose the main reason \tn{[not to express causal structure in terms of nontransitive binary relations]} is that \tn{[doing so appears]} to add nothing to the theory, insofar as it tries to explain how a Lorentzian spacetime can emerge from an underlying discrete order. In the continuum one has well defined causal relations like \tn{[the relativistic ``causal" relation]} $J^+$, and they are automatically transitive, whence suited to correspond to a discrete transitive relation of precedence. But what in the continuum would correspond to a more general... ...not-necessarily transitive relation...?"} \cite{SorkinPrivate13} \end{quotation} These remarks suggest, at least to me, that the choice involved is largely one of focus and ambition. If one seeks merely to {\it recover} relativistic spacetime, at a suitable level of approximation, from a binary relation on a ``discrete set," then it may indeed {\it ``add nothing to the theory"} to consider the independence of reducible relations.\footnotemark\footnotetext{However, if one relaxes the measure axiom to admit volume dependence on {\it local causal structure,} as discussed at the end of section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, then independence becomes metrically relevant.} There is something humorous about attaching the adjective ``merely" to a research objective which would represent a monumental advance in theoretical physics, should it succeed. However, if one does aspire to go further, and to recover ``matter and energy" along with spacetime, the possible role of independence becomes clearer. Sorkin goes on to suggest this very idea, while expressing open-minded reservations about its viability: \begin{quotation} {\noindent ``[the question is] \it whether such additional structure could play a useful kinematical or dynamical role in the theory, perhaps as a kind of ``matter" living on the \tn{[causal set]} (a rudimentary scalar field). At present that looks unlikely to me, but of course I'd never say it was impossible."} \cite{SorkinPrivate13} \end{quotation} My own view is that, given the extreme simplicity of binary relations, compared, for example, to continuum manifolds, any approach to fundamental physics based on binary relations should {\it a priori} attempt to leverage every bit of natural structure available. This is particularly true if one can ``obtain more structure by assuming less;" in this case, obtaining a ``scalar field" by replacing the axioms of transitivity and irreflexivity with the weaker axiom of acyclicity, or omitting these axioms altogether. I do not know the details of Sorkin's reasons for doubting that such a generalization can play a {\it ``useful kinematical or dynamical role"} in discrete causal theory, but sections \hyperref[sectioninterval]{\ref{sectioninterval}}, \hyperref[sectionbinary]{\ref{sectionbinary}}, and \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper present, in my belief, strong reasons to believe that it can. Hence, I view the omission of nontransitive binary relations in causal set theory as an unfortunate example of {\it censoring the merely unexpected,} in the sense of section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} above. \refstepcounter{textlabels}\label{independencefundamental} {\bf Independence at the Fundamental Scale.} Regardless of the opinions of the experts, the na\"{i}ve information-theoretic analogies given above make the deficiency of transitive binary relations in discrete causal theory seem more obvious than it actually is. For example, the intuition of ``breaking a message into multiple parts" is dubious in the context of fundamental spacetime structure, though the notion of independent modes of influence does not depend on this intuition. Arguments for transitivity may also be made by drawing analogies with relativistic spacetime geometry, but adopting any such argument endangers both the correctness and the explanatory power of the resulting theory. The safe position is to abstain from transitivity, since this leads to a more general theory. For this reason, I do not discuss {\it specific} defenses of this axiom. However, I emphasize that any such defense must establish that {\it only one of the two scenarios} illustrated in figures \hyperref[transitivity]{\ref{transitivity}}b and \hyperref[transitivity]{\ref{transitivity}}c above is physically relevant at the fundamental scale, since otherwise the need to distinguish between them is unavoidable. Hence, to rescue transitivity, one must argue either that independent modes of influence along a relation and a reducing chain {\it never exist} at the fundamental scale, or, alternatively, that they {\it always exist.} The latter argument is absurd, since it leads to the conclusion that, whenever one event influences another, however long the chain of intervening influences, {\it independent direct influence} must also propagate between the two events; e.g., between the {\it ``birth and death of the universe,"} as Finkelstein puts it. In particular, this means that the independence convention (\hyperref[ic]{IC}) makes no sense in the transitive paradigm, as already noted at the end of section \hyperref[subsectionchains]{\ref{subsectionchains}} above. Eliminating this choice leaves two possibilities: either independent modes of influence along a relation and a reducing chain never exist at the fundamental scale, or transitive binary relations are deficient for modeling the fundamental structure of classical spacetime in discrete causal theory. \subsection{Six Arguments that Transitive Binary Relations are Deficient}\label{subsectiontransitivitydeficient} I now present six arguments that the scenario illustrated in figure \hyperref[transitivity]{\ref{transitivity}}c above, in which independent modes of influence exist along a relation and a reducing chain between two events, {\it should not be ruled out at any scale.} As discussed at the end of section \hyperref[subsectionmodes]{\ref{subsectionmodes}}, this leads immediately to the conclusion that transitive binary relations are deficient for modeling the fundamental structure of classical spacetime in discrete causal theory. Some of the arguments presented below are direct, while others merely make note of suggestive structural analogies. These arguments are supplemented by several of the general principles discussed in section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} above. \refstepcounter{textlabels}\label{cauchyarg} {\bf 1.} {\bf Multiple independent modes of influence between pairs of events are ubiquitous in conventional physics.} An important example of this is illustrated by the widespread use of {\it Cauchy surfaces} in continuum theory, both in the classical and quantum settings. In the context of general relativity, a Cauchy surface is a spacelike hypersurface uniquely intersecting {\it every causal path} between its past and future, in a given region of spacetime. The prominence of Cauchy surfaces is based on the assumption that different causal paths from past to future can carry independent information. Extension of this assumption to the fundamental scale allows the scenario illustrated in figure \hyperref[transitivity]{\ref{transitivity}}c above as a special case. Figure \hyperref[transitivityarguments]{\ref{transitivityarguments}}a below illustrates ``Cauchy surfaces" intersecting the two independent paths from $x$ to $z$ appearing in figure \hyperref[transitivity]{\ref{transitivity}}c. These ``Cauchy surfaces" may be viewed as primitive examples of {\it maximal antichains of relations.} Such maximal antichains play a major role in the theory of {\it relation space,} introduced in section \hyperref[sectionbinary]{\ref{sectionbinary}} of this paper, and in the quantum causal theory developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{4cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,4)(16.5,4) \pgfxyline(0,-.1)(0,4) \pgfxyline(5,-.1)(5,4) \pgfxyline(9,-.1)(9,4) \pgfxyline(13.75,-.1)(13.75,4) \pgfxyline(16.5,-.1)(16.5,4) \pgfputat{\pgfxy(.15,3.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.15,3.7)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(9.15,3.7)}{\pgfbox[left,center]{c)}} \pgfputat{\pgfxy(13.9,3.7)}{\pgfbox[left,center]{d)}} \begin{pgftranslate}{\pgfpoint{-1cm}{.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(2,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3,1.75)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(2,3)}{0.1cm} \pgfputat{\pgfxy(2,.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(3.3,1.75)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2,3.3)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(4.3,1.95)}{\pgfbox[left,center]{\small{``Cauchy}}} \pgfputat{\pgfxy(4.3,1.55)}{\pgfbox[left,center]{\small{surfaces"}}} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(4.2,2.05)(3.7,2.35) \pgfxyline(4.2,1.45)(3.7,1.15) \pgfclearendarrow \pgfsetdash{{2pt}{2pt}}{0pt} \pgfxyline(1.3,2.45)(3.5,2.45) \pgfxyline(1.3,1.05)(3.5,1.05) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.8cm}{.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(2,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2.7,1)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(3.5,1.5)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,2)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(3.7,2.4)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(2.4,2.7)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3,2.9)}{0.10cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,3.25)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(.6,2)}{0.10cm} \pgfnodecircle{Node11}[fill]{\pgfxy(1.4,2)}{0.10cm} \pgfnodecircle{Node12}[fill]{\pgfxy(3,.5)}{0.10cm} \pgfputat{\pgfxy(2,.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(3,3.2)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2,3.55)}{\pgfbox[center,center]{$z$}} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node2}{Node5} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node6}{Node8} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node12}{Node3} \pgfnodeconnline{Node4}{Node10} \pgfnodeconnline{Node4}{Node11} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfnodeconnline{Node8}{Node9} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7cm}{.1cm}} \pgfnodecircle{Node11}[fill]{\pgfxy(3,.85)}{0.1cm} \pgfnodecircle{Node12}[fill]{\pgfxy(3.3,.8)}{0.1cm} \pgfnodecircle{Node13}[fill]{\pgfxy(3.7,.8)}{0.1cm} \pgfnodecircle{Node14}[fill]{\pgfxy(4.1,.9)}{0.1cm} \pgfnodecircle{Node15}[fill]{\pgfxy(4.6,.8)}{0.1cm} \pgfnodecircle{Node16}[fill]{\pgfxy(5,.8)}{0.1cm} \pgfnodecircle{Node17}[fill]{\pgfxy(5.4,.7)}{0.1cm} \pgfnodecircle{Node18}[fill]{\pgfxy(5.7,.8)}{0.1cm} \pgfnodecircle{Node20}[fill]{\pgfxy(2.2,1.75)}{0.1cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2.55,1.7)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(2.8,1.7)}{0.1cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.15,1.75)}{0.1cm} \pgfnodecircle{Node24}[fill]{\pgfxy(3.4,1.7)}{0.1cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.7,1.7)}{0.1cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,1.7)}{0.1cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4.3,1.8)}{0.1cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.65,1.7)}{0.1cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.9,1.7)}{0.1cm} \pgfnodecircle{Node210}[fill]{\pgfxy(5.2,1.65)}{0.1cm} \pgfnodecircle{Node211}[fill]{\pgfxy(5.45,1.7)}{0.1cm} \pgfnodecircle{Node212}[fill]{\pgfxy(5.8,1.6)}{0.1cm} \pgfnodecircle{Node213}[fill]{\pgfxy(6.15,1.7)}{0.1cm} \pgfnodecircle{Node214}[fill]{\pgfxy(6.4,1.7)}{0.1cm} \pgfnodecircle{Node31}[fill]{\pgfxy(3.1,2.56)}{0.1cm} \pgfnodecircle{Node32}[fill]{\pgfxy(3.35,2.6)}{0.1cm} \pgfnodecircle{Node33}[fill]{\pgfxy(3.6,2.6)}{0.1cm} \pgfnodecircle{Node34}[fill]{\pgfxy(4,2.65)}{0.1cm} \pgfnodecircle{Node35}[fill]{\pgfxy(4.6,2.6)}{0.1cm} \pgfnodecircle{Node36}[fill]{\pgfxy(5.1,2.65)}{0.1cm} \pgfnodecircle{Node37}[fill]{\pgfxy(5.4,2.6)}{0.1cm} \pgfnodeconnline{Node11}{Node20} \pgfnodeconnline{Node11}{Node21} \pgfnodeconnline{Node11}{Node23} \pgfnodeconnline{Node11}{Node24} \pgfnodeconnline{Node11}{Node25} \pgfnodeconnline{Node11}{Node27} \pgfnodeconnline{Node12}{Node20} \pgfnodeconnline{Node12}{Node22} \pgfnodeconnline{Node12}{Node23} \pgfnodeconnline{Node12}{Node24} \pgfnodeconnline{Node12}{Node26} \pgfnodeconnline{Node12}{Node28} \pgfnodeconnline{Node12}{Node29} \pgfnodeconnline{Node13}{Node21} \pgfnodeconnline{Node13}{Node22} \pgfnodeconnline{Node13}{Node23} \pgfnodeconnline{Node13}{Node25} \pgfnodeconnline{Node13}{Node26} \pgfnodeconnline{Node13}{Node27} \pgfnodeconnline{Node13}{Node29} \pgfnodeconnline{Node14}{Node20} \pgfnodeconnline{Node14}{Node23} \pgfnodeconnline{Node14}{Node24} \pgfnodeconnline{Node14}{Node25} \pgfnodeconnline{Node14}{Node27} \pgfnodeconnline{Node14}{Node29} \pgfnodeconnline{Node14}{Node210} \pgfnodeconnline{Node14}{Node212} \pgfnodeconnline{Node15}{Node21} \pgfnodeconnline{Node15}{Node22} \pgfnodeconnline{Node15}{Node25} \pgfnodeconnline{Node15}{Node26} \pgfnodeconnline{Node15}{Node28} \pgfnodeconnline{Node15}{Node29} \pgfnodeconnline{Node15}{Node211} \pgfnodeconnline{Node15}{Node212} \pgfnodeconnline{Node16}{Node26} \pgfnodeconnline{Node16}{Node27} \pgfnodeconnline{Node16}{Node210} \pgfnodeconnline{Node16}{Node211} \pgfnodeconnline{Node16}{Node213} \pgfnodeconnline{Node16}{Node214} \pgfnodeconnline{Node17}{Node22} \pgfnodeconnline{Node17}{Node24} \pgfnodeconnline{Node17}{Node25} \pgfnodeconnline{Node17}{Node27} \pgfnodeconnline{Node17}{Node29} \pgfnodeconnline{Node17}{Node211} \pgfnodeconnline{Node17}{Node213} \pgfnodeconnline{Node17}{Node214} \pgfnodeconnline{Node18}{Node24} \pgfnodeconnline{Node18}{Node27} \pgfnodeconnline{Node18}{Node28} \pgfnodeconnline{Node18}{Node29} \pgfnodeconnline{Node18}{Node210} \pgfnodeconnline{Node18}{Node212} \pgfnodeconnline{Node18}{Node213} \pgfnodeconnline{Node18}{Node214} \pgfnodeconnline{Node20}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node23}{Node31} \pgfnodeconnline{Node24}{Node31} \pgfnodeconnline{Node25}{Node31} \pgfnodeconnline{Node27}{Node31} \pgfnodeconnline{Node28}{Node31} \pgfnodeconnline{Node210}{Node31} \pgfnodeconnline{Node25}{Node37} \pgfnodeconnline{Node26}{Node37} \pgfnodeconnline{Node27}{Node37} \pgfnodeconnline{Node28}{Node37} \pgfnodeconnline{Node29}{Node37} \pgfnodeconnline{Node211}{Node37} \pgfnodeconnline{Node213}{Node37} \pgfnodeconnline{Node214}{Node37} \pgfnodeconnline{Node21}{Node32} \pgfnodeconnline{Node22}{Node32} \pgfnodeconnline{Node23}{Node32} \pgfnodeconnline{Node26}{Node32} \pgfnodeconnline{Node27}{Node32} \pgfnodeconnline{Node24}{Node36} \pgfnodeconnline{Node25}{Node36} \pgfnodeconnline{Node27}{Node36} \pgfnodeconnline{Node28}{Node36} \pgfnodeconnline{Node210}{Node36} \pgfnodeconnline{Node211}{Node36} \pgfnodeconnline{Node212}{Node36} \pgfnodeconnline{Node214}{Node36} \pgfnodeconnline{Node20}{Node33} \pgfnodeconnline{Node24}{Node33} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node33} \pgfnodeconnline{Node27}{Node33} \pgfnodeconnline{Node29}{Node33} \pgfnodeconnline{Node210}{Node33} \pgfnodeconnline{Node211}{Node33} \pgfnodeconnline{Node28}{Node35} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node210}{Node35} \pgfnodeconnline{Node211}{Node35} \pgfnodeconnline{Node212}{Node35} \pgfnodeconnline{Node20}{Node34} \pgfnodeconnline{Node21}{Node34} \pgfnodeconnline{Node22}{Node34} \pgfnodeconnline{Node28}{Node34} \pgfnodeconnline{Node29}{Node34} \pgfnodeconnline{Node210}{Node34} \pgfnodeconnline{Node213}{Node34} \pgfnodeconnline{Node214}{Node34} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{9.3cm}{.1cm}} \pgfputat{\pgfxy(5.5,.5)}{\pgfbox[center,center]{$X$}} \pgfputat{\pgfxy(6.5,1.75)}{\pgfbox[center,center]{$Y$}} \pgfputat{\pgfxy(5.5,3)}{\pgfbox[center,center]{$Z$}} \pgfputat{\pgfxy(6.2,1.05)}{\pgfbox[center,center]{$f$}} \pgfputat{\pgfxy(6.2,2.55)}{\pgfbox[center,center]{$g$}} \pgfputat{\pgfxy(5.3,1.75)}{\pgfbox[center,center]{$h$}} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(5.5,.8)(5.5,2.7) \pgfxyline(5.7,.7)(6.35,1.5) \pgfxyline(6.35,2)(5.7,2.8) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) ``Cauchy surfaces" in the directed set first appearing in figure 3.1.1c; b) hypothetical relation in the ``distant future" of an element $x$; c) Kleitman-Rothschild-type partial order (irreducible relations shown); d) category-theoretic analogue of figure 3.1.1b or 3.1.1c.} \label{transitivityarguments} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{futurearg} {\bf 2.} {\bf Influences exerted by an event should not be constrained by details of its future.} This is particularly true in the acyclic case, in which such constraints may legitimately be viewed as pathological, rather than merely dubious. In particular, given events $x$, $y$, and $z$, together with the knowledge that $x$ influences both $y$ and $z$, the question of whether or not $x$ {\it directly} influences $z$, independently of $y$, should not depend on the existence of a relation $y\prec z$ lying in the {\it future} of $x$. For example, figure \hyperref[transitivityarguments]{\ref{transitivityarguments}}b above illustrates an acyclic directed set in which the existence of a reducing chain for the relation $x\prec z$ depends on the existence of a relation $y\prec z$ in the ``distant future" of $x$. It is intuitively obvious that the presence or absence of this relation should impose no ``retroactive" constraint on influences emanating from $x$. \refstepcounter{textlabels}\label{irredindarg} {\bf 3.} {\bf Irreducibility and independence of relations between pairs of elements are {\it a priori} distinct conditions.} This logical distinction was pointed out in section \hyperref[subsectionchains]{\ref{subsectionchains}} above. Irreducibility is an absolute mathematical condition, while independence is a subjective physical condition. In a directed set, irreducible relations between pairs of elements are necessarily independent, but the converse is false. The question then becomes whether or not there are compelling physical reasons to equate the two conditions in the context of discrete causal theory. As far as I know, the answer to this question is negative. It is therefore natural to encode discrete causal structure by means of independent relations between pairs of elements, following the independence convention (\hyperref[ic]{IC}). \refstepcounter{textlabels}\label{configpath} {\bf 4.} {\bf Configuration spaces of transitive binary relations are physically pathological.} A classic example of such a pathology occurs in the acyclic directed case, in which the distribution of finite partial orders is asymptotically dominated by objects manifestly unsuitable as models of classical spacetime. This has been known since before the advent of causal set theory, due to Kleitman and Rothschild's 1975 paper {\it Asymptotic Enumeration of Partial Orders on a Finite Set} \cite{KleitmanRothschild75}. This paper shows that a ``generic large" partially ordered set of cardinality $n$ has just three generations, of cardinalities roughly $n/4$, $n/2$, and $n/4$, respectively, with the number of irreducible relations beginning at a typical nonmaximal element depending linearly on $n$. A Kleitman-Rothschild-type partial order is illustrated in figure \hyperref[transitivityarguments]{\ref{transitivityarguments}}c above. In the context of classical spacetime structure, such a partial order represents a universe of {\it large spatial and negligible temporal size,} with an unreasonably large luminal velocity. {\it A priori,} this raises the concern that the histories approach to quantum causal set theory might be dominated by such objects. This is sometimes referred to as an {\it entropy problem.}\footnotemark\footnotetext{This terminology comes from statistical thermodynamics, in which the large-scale state of a system is ``determined" by the multiplicities of microscopic states sharing the same large-scale properties.} Avoiding problems of this nature was one of the principal motivations for Sorkin and Rideout's theory of sequential growth dynamics, as explained on page 4 of \cite{SorkinSequentialGrowthDynamics99}.\footnotemark\footnotetext{Sorkin and Rideout write, {\it ``Maybe this is not so different from \tn{[Feynman's path integral],} where the smooth paths, which form a set of measure zero... ...dominate the sum over histories..."} If this is so, the most reasonable explanation might be that {\it both situations are cluttered by physically irrelevant structure;} in the first case, the continuum; in the second case, dependent relations due to transitivity.} It is certainly possible, and in a sense, easy, to avoid such problems dynamically, but it is more promising to have a class of models that does not present such troubling features in the first place. Often, at least in mathematics, generic behavior of this type indicates some form of {\it degeneracy,} or {\it lack of naturality} in the class of structures under consideration. In contrast, the asymptotic behavior of (generally nontransitive) acyclic directed sets seems to be much more reasonable, as indicated by Brendan McKay's 1989 paper {\it On the shape of a random acyclic digraph} \cite{McKayDigraphs89}, and Stephan Wagner's 2013 paper {\it Asymptotic Enumeration of Extensional Acyclic Digraphs} \cite{WagnerDigraphs13}. In particular, the configuration space of finite acyclic directed sets is not asymptotically dominated by {\it either} ``near antichains," such as the Kleitman-Rothschild orders, or by ``near-chains." \refstepcounter{textlabels}\label{commcat} {\bf 5.} {\bf Structural notions from mathematics motivate independent modes of influence}. For example, there is a useful and far-reaching analogy between plane diagrams of multidirected sets, including the generalized Hasse diagrams appearing in the acyclic directed case, and category-theoretic diagrams of objects and morphisms. This analogy is illustrated by the triangular diagram in figure \hyperref[transitivityarguments]{\ref{transitivityarguments}}d above, where $X$, $Y$, and $Z$ are objects of a category, and where $f$, $g$, and $h$ are morphisms. This diagram may be viewed as analogous to {\it either} of the two physical scenarios illustrated in figures \ref{transitivity}b and \ref{transitivity}c above, depending on whether or not it {\it commutes}; i.e., whether or not the morphism $h$ is equal to the composition $g\circ f$. If the diagram commutes, then $h$ is ``merely a consequence of $f$ and $g$," encoding no independent information. This case is analogous to the scenario illustrated in figure \ref{transitivity}b. If the diagram does not commute, then $h$ involves information independent of $f$ and $g$. This case is analogous to the scenario illustrated in figure \ref{transitivity}c. In a sense, then, considering only transitive binary relations in discrete causal theory is analogous to considering only categories in which {\it every diagram commutes}, which is a very restrictive condition. The reader familiar with ``concrete" categories, in which objects possess underlying sets, may argue in favor of such ``commutativity" in the case of directed sets, on the grounds that the unique map of singleton sets from $\{x\}$ to $\{z\}$ is equal to the composition of unique maps from $\{x\}$ to $\{y\}$ and from $\{y\}$ to $\{z\}$. This observation is true but irrelevant; the appropriate analog of a ``morphism" in this case is not a set map. Indeed, there is a unique set map between {\it any} pair of singletons in a directed or multidirected set. However, as noted at the end of section \hyperref[settheoretic]{\ref{settheoretic}} above, there are many important examples of categories whose objects have no internal structure, but which nonetheless admit multiple independent morphisms between pairs of objects. I have already cited the example of groups, whose elements may be represented by endomorphisms of a single structureless object. Analogous but more general constructions, with important applications in this paper, arise in the theory of {\it semicategories,} discussed in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. \refstepcounter{textlabels}\label{otherimprovements} {\bf 6.} {\bf Nontransitive relations lead naturally to other improvements in causal theory.} As pointed out by Finkelstein, nontransitive relations enable an improved treatment of {\it causal locality.} This notion is discussed in detail in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} below. The same basic reasoning leads to superior notions of local conditions for multidirected sets in general. Of particular importance is {\it local finiteness} (\hyperref[lf]{LF}), introduced in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, which replaces the pathological causal set axiom of interval finiteness (\hyperref[if]{IF}). Building on these improvements is the theory of {\it relation space,} introduced in section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, which is almost unrecognizable in the transitive paradigm. In particular, the use of maximal antichains of relations as analogues of Cauchy surfaces, which avoids the generic {\it permeability problem} presented by maximal antichains of elements, depends crucially on the independence convention (\hyperref[ic]{IC}), and therefore makes sense only in the nontransitive context. Similarly, much of the algebraic machinery associated with the theory of path summation over a multidirected set, introduced in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}}, is seriously hampered if dependent relations are forced into the picture. Ultimately, the causal analogues of Feynman's path integral and Schr\"{o}dinger's equation, appearing in sections \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} and \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}}, find full expression only in the nontransitive setting. \subsection{The Causal Preorder}\label{subsectionpreorder} \refstepcounter{textlabels}\label{nontransgen} \refstepcounter{textlabels}\label{finkelsteinconnect} \refstepcounter{textlabels}\label{adscausalpreorder} \refstepcounter{textlabels}\label{notusualpreorder} To remedy the shortcomings of transitive binary relations in modeling classical spacetime structure in discrete causal theory, I propose that the transitive causal order should be viewed as a {\it derivative construct,} generated by a more fundamental, generally nontransitive, binary relation called the {\it causal preorder.} Referring again to Finkelstein \cite{Finkelstein88}, the causal preorder corresponds to the {\it ``causal connection relation"} $\mbf{c}$, while the causal order corresponds to the {\it ``transitive causal relation"} $\mbf{C}$. The reader should be aware that the term {\it preorder} already has an inconvenient proprietary definition in the context of order theory, as explained in the discussion of nonstandard terminology in section \hyperref[subsectionnotation]{\ref{subsectionnotation}}. For the duration of this paper, I reclaim this term as follows: \\ \begin{defi}\label{defpreorder} Let $D=(D,\prec)$ be a directed set, viewed as a model of causal structure under the independence convention \tn{(\hyperref[ic]{IC})}. In this context, the binary relation $\prec$ on $D$ is called the {\bf causal preorder} on $D$. \end{defi} The independence convention gives an unambiguous physical meaning to every relation appearing in a causal preorder: each relation encodes independent influence. Dependent influence, meanwhile, is encoded by complex chains. For example, the causal preorder illustrated in figure \hyperref[transitivity]{\ref{transitivity}}b above does {\it not} include the relation $x\prec z$, since $x$ does not influence $z$ independently of $y$, while the causal preorder illustrated in figure \hyperref[transitivity]{\ref{transitivity}}c {\it does} include the relation $x\prec z$. The causal {\it orders} are the same in both cases. \subsection{Transitive Closure; Skeleton; Degeneracy; Functoriality}\label{subsectiontransitiveclosure} {\bf Transitive Closure; Skeleton.} Two useful operations on directed sets are the {\it transitive closure} functor and the {\it skeleton} operation. The transitive closure functor is particularly important because it {\it generates causal orders from causal preorders.} These two operations realize the opposite extremes of transitivity and irreducibility for binary relations. The transitive closure is defined by adding reducible relations between pairs of elements, while the skeleton is defined by deleting such relations. A variety of generalizations to the case of multidirected sets are possible, but are not needed in this paper.\\ \refstepcounter{textlabels}\label{transclosure} \refstepcounter{textlabels}\label{skeleton} \begin{defi}\label{deftransclosure}Let $D=(D,\prec)$ be a directed set. \begin{enumerate} \item The {\bf transitive closure} of $D$ is the directed set $\text{\tn{tr}}(D):=(D,\prec_{\text{\tn{\fsz{tr}}}})$ whose binary relation $\prec_{\text{\tn{\fsz{tr}}}}$ is defined by setting $x\prec_{\text{\tn{\fsz{tr}}}}y$ if and only if there is a chain from $x$ to $y$ in $D$. \item The {\bf skeleton} of $D$ is the acyclic directed set $\text{\tn{sk}}(D):=(D,\prec_{\text{\tn{\fsz{sk}}}})$ whose binary relation $\prec_{\text{\tn{\fsz{sk}}}}$ is defined by setting $x\prec_{\text{\tn{\fsz{sk}}}}y$ if and only if $x\prec y$ is an irreducible relation in $D$. \end{enumerate} \end{defi} \refstepcounter{textlabels}\label{skelnottransred} The skeleton $\tn{sk}(D)$ of any directed set $D$ is automatically acyclic, since any relation in a cycle admits a reducing chain given by going around the cycle. The skeleton is generally not the same as the {\it transitive reduction,} familiar from graph theory, which {\it preserves accessibility.} However, the two coincide for finite acyclic directed sets. Since every relation is itself a chain, the transitive closure of a directed set $D$ has at least as many relations as $D$, with equality occurring if and only if $D$ is transitive. Similarly, since every irreducible relation is a relation, the skeleton of $D$ has at most as many relations as $D$, with equality occurring if and only if $D$ is irreducible. The transitive closure of a directed set is closely related to its {\it chain space,} discussed in a more general context in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. This relationship arises from the fact that every relation in $\tn{tr}(D)$ corresponds to a chain in $D$. This suggests one way to generalize the transitive closure to multidirected sets: by {\it adding a relation corresponding to each complex chain.} Figure \ref{transitiveskel} below illustrates the transitive closure and the skeleton of a directed set. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.2cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,4.9)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,4.9)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,4.9)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.2)(16.5,5.2) \pgfxyline(0,-.1)(0,5.2) \pgfxyline(5.5,-.1)(5.5,5.2) \pgfxyline(11,-.1)(11,5.2) \pgfxyline(16.5,-.1)(16.5,5.2) \begin{pgftranslate}{\pgfpoint{0.25cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.75cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node1}{Node13} \pgfnodeconnline{Node1}{Node12} \pgfnodeconnline{Node1}{Node14} \pgfnodeconnline{Node1}{Node15} \pgfnodeconnline{Node1}{Node18} \pgfnodeconnline{Node1}{Node19} \pgfnodeconnline{Node1}{Node22} \pgfnodeconnline{Node1}{Node23} \pgfnodeconnline{Node1}{Node30} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node2}{Node8} \pgfnodeconnline{Node4}{Node2} \pgfnodeconnline{Node4}{Node3} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node6} \pgfnodeconnline{Node4}{Node7} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node4}{Node12} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node4}{Node15} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node3} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node12} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node5}{Node15} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node2} \pgfnodeconnline{Node9}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node5} \pgfnodeconnline{Node9}{Node6} \pgfnodeconnline{Node9}{Node7} \pgfnodeconnline{Node9}{Node8} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node12} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node14} \pgfnodeconnline{Node9}{Node15} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node9}{Node18} \pgfnodeconnline{Node9}{Node19} \pgfnodeconnline{Node9}{Node22} \pgfnodeconnline{Node9}{Node23} \pgfnodeconnline{Node9}{Node30} \pgfnodeconnline{Node10}{Node3} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node8} \pgfnodeconnline{Node10}{Node12} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node10}{Node14} \pgfnodeconnline{Node10}{Node15} \pgfnodeconnline{Node10}{Node18} \pgfnodeconnline{Node10}{Node19} \pgfnodeconnline{Node10}{Node22} \pgfnodeconnline{Node10}{Node23} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node11}{Node3} \pgfnodeconnline{Node11}{Node7} \pgfnodeconnline{Node11}{Node8} \pgfnodeconnline{Node11}{Node12} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node11}{Node14} \pgfnodeconnline{Node11}{Node15} \pgfnodeconnline{Node11}{Node18} \pgfnodeconnline{Node11}{Node19} \pgfnodeconnline{Node11}{Node22} \pgfnodeconnline{Node11}{Node23} \pgfnodeconnline{Node11}{Node30} \pgfnodeconnline{Node12}{Node3} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node3} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node8} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node15} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node13}{Node22} \pgfnodeconnline{Node13}{Node23} \pgfnodeconnline{Node13}{Node30} \pgfnodeconnline{Node14}{Node3} \pgfnodeconnline{Node14}{Node8} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node3} \pgfnodeconnline{Node16}{Node7} \pgfnodeconnline{Node16}{Node8} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node12} \pgfnodeconnline{Node16}{Node13} \pgfnodeconnline{Node16}{Node14} \pgfnodeconnline{Node16}{Node15} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node18} \pgfnodeconnline{Node16}{Node19} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node16}{Node21} \pgfnodeconnline{Node16}{Node22} \pgfnodeconnline{Node16}{Node23} \pgfnodeconnline{Node16}{Node26} \pgfnodeconnline{Node16}{Node27} \pgfnodeconnline{Node16}{Node29} \pgfnodeconnline{Node16}{Node30} \pgfnodeconnline{Node16}{Node35} \pgfnodeconnline{Node17}{Node3} \pgfnodeconnline{Node17}{Node8} \pgfnodeconnline{Node17}{Node12} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node17}{Node15} \pgfnodeconnline{Node18}{Node3} \pgfnodeconnline{Node18}{Node8} \pgfnodeconnline{Node18}{Node12} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node15} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node18}{Node30} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node3} \pgfnodeconnline{Node20}{Node7} \pgfnodeconnline{Node20}{Node8} \pgfnodeconnline{Node20}{Node12} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node14} \pgfnodeconnline{Node20}{Node15} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node19} \pgfnodeconnline{Node20}{Node21} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node23} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node20}{Node27} \pgfnodeconnline{Node20}{Node29} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node20}{Node35} \pgfnodeconnline{Node21}{Node19} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node21}{Node30} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node22}{Node30} \pgfnodeconnline{Node24}{Node3} \pgfnodeconnline{Node24}{Node7} \pgfnodeconnline{Node24}{Node8} \pgfnodeconnline{Node24}{Node12} \pgfnodeconnline{Node24}{Node13} \pgfnodeconnline{Node24}{Node14} \pgfnodeconnline{Node24}{Node15} \pgfnodeconnline{Node24}{Node17} \pgfnodeconnline{Node24}{Node18} \pgfnodeconnline{Node24}{Node19} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node24}{Node21} \pgfnodeconnline{Node24}{Node22} \pgfnodeconnline{Node24}{Node23} \pgfnodeconnline{Node24}{Node26} \pgfnodeconnline{Node24}{Node27} \pgfnodeconnline{Node24}{Node29} \pgfnodeconnline{Node24}{Node30} \pgfnodeconnline{Node24}{Node35} \pgfnodeconnline{Node25}{Node3} \pgfnodeconnline{Node25}{Node8} \pgfnodeconnline{Node25}{Node12} \pgfnodeconnline{Node25}{Node14} \pgfnodeconnline{Node25}{Node15} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node19} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node22} \pgfnodeconnline{Node25}{Node23} \pgfnodeconnline{Node25}{Node27} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node30} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node25}{Node34} \pgfnodeconnline{Node25}{Node35} \pgfnodeconnline{Node26}{Node19} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node22} \pgfnodeconnline{Node26}{Node23} \pgfnodeconnline{Node26}{Node27} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node26}{Node30} \pgfnodeconnline{Node26}{Node35} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node27} \pgfnodeconnline{Node28}{Node29} \pgfnodeconnline{Node28}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node28}{Node34} \pgfnodeconnline{Node28}{Node35} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node3} \pgfnodeconnline{Node31}{Node8} \pgfnodeconnline{Node31}{Node12} \pgfnodeconnline{Node31}{Node14} \pgfnodeconnline{Node31}{Node15} \pgfnodeconnline{Node31}{Node17} \pgfnodeconnline{Node31}{Node19} \pgfnodeconnline{Node31}{Node21} \pgfnodeconnline{Node31}{Node22} \pgfnodeconnline{Node31}{Node23} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node26} \pgfnodeconnline{Node31}{Node27} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node29} \pgfnodeconnline{Node31}{Node30} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node31}{Node33} \pgfnodeconnline{Node31}{Node34} \pgfnodeconnline{Node31}{Node35} \pgfnodeconnline{Node32}{Node19} \pgfnodeconnline{Node32}{Node21} \pgfnodeconnline{Node32}{Node22} \pgfnodeconnline{Node32}{Node23} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node27} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node30} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node32}{Node34} \pgfnodeconnline{Node32}{Node35} \pgfnodeconnline{Node33}{Node27} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node30} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node34}{Node30} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.25cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Directed set $D$; b) transitive closure $\tn{tr}(D)$ of $D$; c) skeleton $\tn{sk}(D)$ of $D$.} \label{transitiveskel} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{causordtranspreorder} Definition \hyperref[defipreorderorder]{\ref{defipreorderorder}} below gives the precise relationship between the causal preorder on a directed set and the corresponding causal order. I emphasize again that the term {\it order} is used here in a generalized sense. In particular, the transitive closure of a causal preorder including cycles also includes cycles, and is therefore not an order in the usual sense. \\ \begin{defi}\label{defipreorderorder} Let $D$ be a directed set, viewed as a model of causal structure, with causal preorder $\prec$. Then the binary relation $\prec_{\text{\tn{\fsz{tr}}}}$ on the transitive closure $(D,\prec_{\text{\tn{\fsz{tr}}}})$ is called the {\bf causal order} on $D$. \end{defi} \refstepcounter{textlabels}\label{degenconfig} {\bf Degeneracy.} In general, many different directed sets share the same transitive closure and skeleton, up to isomorphism. For example, the transitive closure and skeleton of the directed set $D$ illustrated in figure \hyperref[transitiveskel]{\ref{transitiveskel}}a above differ by about $200$ reducible relations. Adding distinct families of these relations to $\tn{sk}(D)$ produces distinct directed sets; hence, there are roughly $2^{200}$ {\it information-theoretically distinct} directed sets sharing $\tn{tr}(D)$ and $\tn{sk}(D)$ as their transitive closures and skeletons, respectively. The {\bf transitive degeneracy class} of a directed set $D$ is the set of isomorphism classes of directed sets with transitive closure isomorphic to $\tn{tr}(D)$, and the {\bf skeletal degeneracy class} of $D$ is the set of isomorphism classes of directed sets with skeleton isomorphic to $\tn{sk}(D)$. The two classes coincide for {\it finite acyclic directed sets,} since the transitive closure of such a set may be recovered from its skeleton, and vice versa. The {\bf transitive} and {\bf skeletal degeneracies} of $D$ are the cardinalities of the corresponding degeneracy classes. Degeneracy plays an important role in the structure of configuration spaces of directed sets. In particular, degeneracy distinguishes configuration spaces of acyclic directed sets from the corresponding configuration spaces of partial orders, whose elements correspond to transitive degeneracy classes of acyclic directed sets. This clarifies the source of the {\it Kleitman-Rothschild pathology} for finite partial orders, discussed in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}} above, in which ``near-antichains" dominate asymptotically: enumeration of partial orders counts only {\it one} member of each degeneracy class. Referring again to the directed set illustrated in figure \hyperref[transitiveskel]{\ref{transitiveskel}}a, a Kleitman-Rothschild-type order of the same cardinality has roughly $80$ reducible relations, and hence a transitive degeneracy of roughly $2^{80}$, about $120$ powers of $2$ less than the ``more physically realistic" directed set in the figure. This illustrates how enumeration of partial orders {\it relatively overcounts} ``near-antichains" by many of orders of magnitude, even for small cardinalities. \refstepcounter{textlabels}\label{transfunctorial} {\bf Functorial Properties.} The transitive closure and the skeleton are {\it idempotent} operations on the category $\ms{D}$ of directed sets. This means that repeated application of these operations produces nothing new: $\tn{tr}\big(\tn{tr}(D)\big)=\tn{tr}(D)$ and $\tn{sk}\big(\tn{sk}(D)\big)=\tn{sk}(D)$. In the finite acyclic directed case, the two operations are ``roughly inverse" to each other, but this is not true in general; for example, the skeleton of the rational numbers $\mathbb{Q}$, with its usual order, has {\it no relations}, since every relation in $\mathbb{Q}$ is reducible. Similarly, the skeleton of a reflexive relation is a singleton with no relations, since a reflexive relation is reduced by any chain ``going around it multiple times." Taking morphisms into account, the transitive closure extends to a {\it functor} from the category $\ms{D}$ of directed sets to its subcategory $\ms{D}_{\tn{\fsz TR}}$ of transitive directed sets, essentially because chains map to chains under morphisms. There is an inclusion functor $\ms{D}_{\tn{\fsz TR}}\rightarrow \ms{D}$ in the opposite direction, and the two functors are {\it adjoint.} This means that for any directed set $(D,\prec)$, and any transitive directed set $(D',\prec')$, there is a natural identification of morphism sets \refstepcounter{textlabels}\label{transadj} \refstepcounter{textlabels}\label{skelnotfunctor} \[\tn{Mor}_{\ms{D}}\big(D,D'\big)\cong \tn{Mor}_{\ms{D}_{\tn{\fsz TR}}}\big(\tn{tr}(D),D'\big),\] given by identifying morphisms with the same underlying set maps. The skeleton operation does {\it not} define a functor, since irreducible relations do not always map to irreducible relations under morphisms. For example, the inclusion $\mathbb{Z}\rightarrow\mathbb{Q}$, with the usual orders, is a morphism in $\ms{D}$, but does not induce a morphism $\tn{sk}(\mathbb{Z})\rightarrow\tn{sk}(\mathbb{Q})$. \newpage \section{Interval Finiteness versus Local Finiteness}\label{sectioninterval} Interval finiteness (\hyperref[if]{IF}), called {\it local finiteness} in the causal set literature, is described by Rafael Sorkin as {\it``a formal way of saying that a} [causal set] {\it is discrete"} (\cite{SorkinCausalSetsDiscreteGravity05}, page 309). The discrete {\it topology} is useless in causal theory, since it ignores causal structure, but order-theoretic discreteness and measure-theoretic discreteness are central to the emergence of geometry under Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). Local finiteness of some form is an obvious requirement in this setting, at least if one wishes to avoid manifestly nonphysical scenarios, such as the instantaneous emergence of an infinite volume of spacetime from a single event. Unfortunately, interval finiteness permits this very scenario. The need for a better local finiteness condition, and for better treatment of local conditions in general, brings topology back into the picture. Many different topological spaces may be associated with a directed or multidirected set, with each topology inducing its own local conditions. The suitability of these conditions, and of the topologies that induce them, must be judged on {\it physical} grounds. In particular, for directed sets, viewed as models of classical spacetime under the classical causal metric hypothesis (\hyperref[ccmh]{CCMH}), {\it local} should mean {\it causally local} in a suitable sense. Interval finiteness is not even close to being a causally local condition, and related topological conditions, such as {\it topological local finiteness in the interval topology,} are little better. As an alternative, I propose the axiom of {\it local finiteness} (\hyperref[lf]{LF}). Besides being causally local, this axiom admits a topological description that facilitates a better and more comprehensive understanding of discrete causal structure. In section \hyperref[subsectiontopology]{\ref{subsectiontopology}} below, I discuss local conditions and topology in general terms. I define {\it causal locality} for physical systems, in essential agreement with Finkelstein's \cite{Finkelstein88} and Raptis' \cite{RaptisAlgebraicQuantization00} notions of locality. I briefly review standard topological material. After this, I discuss an important application of the principle of {\it hidden structure} (\hyperref[hs]{HS}), introduced in section \hyperref[settheoretic]{\ref{settheoretic}} above, in which multidirected sets induce {\it extra elements} in associated topological spaces. I then discuss four different topological spaces associated with a multidirected set $M=(M,R,i,t)$. The most useful of these I call the {\it star model} of $M$; its underlying set is the union of $M$ and its relation set $R$, and its topology is called the {\it star topology.} In section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, I discuss the shortcomings of interval finiteness, and propose local finiteness as an alternative. I prove that local finiteness coincides with {\it topological local finiteness for the star topology.} I show that interval finiteness and local finiteness are {\it incomparable conditions;} i.e., neither implies the other. In section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}, I present the theory of {\it relative multidirected sets over a fixed base,} a generalized order-theoretic application of Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}). This leads to reinterpretation of interval finite acyclic directed sets, including causal sets, as relative directed sets over the integers. The alternative axiom of local finiteness identifies a much richer class of relative directed sets. In section \hyperref[subsectionintervalfinitenessdeficient]{\ref{subsectionintervalfinitenessdeficient}}, I present eight arguments against interval finiteness as a local finiteness condition for multidirected sets, focusing on the case of directed sets in the context of classical spacetime structure. In section \hyperref[subsectionarglocfin]{\ref{subsectionarglocfin}}, I present six arguments in favor of local finiteness as a suitable alternative. Finally, in section \hyperref[subsectionhierarchyfiniteness]{\ref{subsectionhierarchyfiniteness}}, I introduce a few other finiteness conditions, and prove some results on the hierarchy of finiteness conditions in causal theory. \subsection{Local Conditions; Topology}\label{subsectiontopology} \refstepcounter{textlabels}\label{localcond} {\bf Local Conditions.} A condition on a physical system or mathematical structure is called {\bf local} if it may be checked by examining {\it arbitrarily small neighborhoods of events or elements.} In general mathematical settings, the notion of a small neighborhood is usually made precise in terms of a {\it topology,} as described below. In classical continuum physics, including general relativity, neighborhoods may be defined in {\it metric} terms; i.e., in terms of ``distances" in space or spacetime. Under the causal metric hypothesis (\hyperref[cmh]{CMH}), the roles are reversed: local conditions are more fundamental than metric notions in this context. This requires a suitable definition of locality for multidirected sets. \newpage \refstepcounter{textlabels}\label{relnotionstop} {\bf Relativistic Notions.} Before proceeding, I briefly review a few relativistic notions for contextual purposes. The word {\it metric} has a different meaning in differential geometry, and hence in general relativity, than in the abstract mathematical setting of a {\it metric space.} In the latter setting, a metric is just a {\it global distance function} on a set $S$. Such a metric induces a {\it metric topology} on $S$. For example, the {\it Euclidean metric}, which generalizes the Pythagorean theorem, induces the {\it Euclidean metric topology} on $\mathbb{R}^n$, which in turn induces a natural topology on any $n$-dimensional continuum manifold $X$, via its coordinate charts. However, $X$ generally does not inherit any preferred notion of conformal structure or scale from $\mathbb{R}^n$. A {\it pseudo-Riemannian metric} $g$ on $X$, meanwhile, is a special additional structure on a {\it smooth} continuum manifold $X$, assigning to the tangent spaces of $X$ smoothly varying inner products. The metric $g$ endows $X$ with specific conformal and scale data. One {\it may} express topological properties of $X$ in terms of $g$, but it is generally {\it not} possible to express metric or conformal structure in topological terms. Results such as the theorems of Hawking \cite{Hawking76} and Malament \cite{Malament77}, cited in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, give special conditions under which this may be done; modern results such as those of Martin and Panangaden \cite{Martin06}, dispense with the smoothness assumption. Other useful topologies may also be defined on pseudo-Riemannian manifolds. One example is the {\it Alexandrov topology,} mentioned below in the context of the {\it interval topology} on a multidirected set. \refstepcounter{textlabels}\label{causallocality} {\bf Causal Locality.} For directed sets, viewed as models of classical spacetime under the classical causal metric hypothesis (\hyperref[ccmh]{CCMH}), no pseudo-Riemannian metric exists at the fundamental level, and topological notions must be treated more abstractly. Since the appropriate choice of topology depends on physical considerations in this context, I begin by specifying a {\it physical} notion of what it means for a condition to be local in a causal sense.\\ \begin{defi}\label{definitioncausallocality} A condition on a physical system is {\bf causally local} if it may be checked by examining all independent causes and effects associated with each event. \end{defi} This definition accords with Finkelstein's \cite{Finkelstein88} and Raptis' \cite{RaptisAlgebraicQuantization00} notions of locality. The same intuition is useful in the more general case of multidirected sets, although the physical interpretation is generally different in this context. Use of the terms {\it physical system} and {\it event} in definition \hyperref[definitioncausallocality]{\ref{definitioncausallocality}} is intended to convey the idea that the physical notion of causally locality {\it precedes any specific choice of mathematical model.} However, this definition is only applicable to models for which the independence convention (\hyperref[ic]{IC}) makes sense.\footnotemark\footnotetext{In the context of classical spacetime, this includes the general class of {\it discrete manifolds,} as described by Benincasa and Dowker \cite{DowkerScalarCurvature10}, following Riemann.} For example, interpolative models such as domains require modified definitions. \refstepcounter{textlabels}\label{gentopology} {\bf General Topology.} In mathematics, local conditions are formalized by the theory of {\it topology.} A {\bf topology} $\ms{T}$ on a set $S$, with or without any additional structure, is a special family of subsets of $S$, which may be viewed as generalizations of open intervals on the real line. Elements of $\ms{T}$ are called {\it open sets} of $S$. An open set containing a given element $x$ of $S$ is called a {\it neighborhood} of $x$. To qualify as a topology, $\ms{T}$ must satisfy the axioms that the ``total set" $S$ and the empty set $\oslash$ are open, that finite intersections of open sets are open, and that arbitrary unions of open sets are open. A {\it basis} for $\ms{T}$ is a family of open sets such that every open set of $S$ is a union of sets in the basis. A {\it subbasis} for $\ms{T}$ is a {\it generating family} of open sets for $\ms{T}$, meaning that any topology on $S$ containing the subbasis contains all of $\ms{T}$. More concretely, this means that every open set of $S$, with the possible exceptions of $S$ and $\oslash$, is a union of finite intersections of sets in the subbasis. A {\it topological space} is a set $S$ together with a topology $\ms{T}$ on $S$. A morphism in the category of topological spaces is called a {\it continuous map}; an isomorphism is called a {\it homeomorphism}. This paper does not make explicit use of topological morphisms, however.\footnotemark\footnotetext{The alternative notion of {\it Scott continuity} is often used in order theory, and particularly in domain theory.} A set $S$ may be endowed with multiple different topologies, which may be compared in the following way: a topology $\ms{T}$ on $S$ is called {\it finer} than an alternative topology $\ms{T}'$ on $S$ if every element of $\ms{T}'$ contains an element of $\ms{T}$. If the converse is true, then $\ms{T}$ is called {\it coarser} than $\ms{T}'$. If neither condition is true, then $\ms{T}$ and $\ms{T}'$ are called {\it incomparable}. Topologies most suitable for encoding useful local conditions are those that incorporate a careful balance between fineness and coarseness, with special respect for any auxiliary structure on $S$, such as a binary relation. If a topology is too fine, then arbitrarily small neighborhoods of an element $x$ are generally too small to capture all the behavior one would like to classify as occurring {\it near} $x$. If a topology is too coarse, it generally does not adequately {\it resolve} structure near $x$; i.e., every neighborhood of $x$ is so large that it contains behavior ``irrelevant to $x$" in some sense. Among the topologies associated with multidirected sets discussed later in this section, the {\it discrete topology} is too fine, while the {\it interval topology} is too coarse. \refstepcounter{textlabels}\label{hiddentop} {\bf Topological Hidden Structure.} Not every interesting topological space that may be associated with a multidirected set $M=(M,R,i,t)$ involves a topology on its underlying set. Also of interest are topologies on {\it enlarged sets} containing $M$, whose {\it extra elements} facilitate the analysis of $M$ itself. The {\it continuum model} $M_\mathbb{R}$ of $M$, and the {\it star model} $M_\star$ of $M$, both defined below, are examples of such enlarged sets. The continuum model of $M$ contains an uncountable number of extra elements, drawn from an {\it extrinsic} source; namely, the unit continuum interval $[0,1]$. This limits its relevance, since the resulting topological space is not actually {\it induced} by $M$. By contrast, the extra elements appearing in the star model of $M$ correspond to the elements of the relation set $R$ of $M$, and are therefore an {\it intrinsic} part of the multidirected structure. These elements represent {\it topological hidden structure} associated with $M$. As mentioned in section \hyperref[settheoretic]{\ref{settheoretic}} above, the principle of hidden structure (\hyperref[hs]{HS}) is inspired by Alexander Grothendieck's approach to commutative algebraic geometry, in which an algebraic variety is augmented by the addition of extra elements corresponding to its closed subsets, thereby producing an algebraic scheme. In the context of discrete causal theory, the star model $M_\star$ of a multidirected set $M$ represents a primitive example of an analogous, and equally useful, type of construction, in which $M$ is augmented by the addition of extra elements corresponding to its relations, chains etc. Going in the opposite direction, it is sometimes useful to adjoin to $M$ extra elements ``smaller than ordinary elements." This situation arises, for example, in the theory of {\it abstract element space,} introduced in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below. Similar ideas appear in the theory of {\it causal atoms,} introduced in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}}. Causal atoms may be viewed as ``generalized elements" of a multidirected set. Ordinary elements are sometimes ``smaller than any causal atom" in the cyclic case. \refstepcounter{textlabels}\label{toplocfin} {\bf Topological Local Finiteness.} I now turn to the specific topic of local finiteness conditions. Each topology on a set $S$ is equipped with its own notion of local finiteness. Since it is necessary, in the context of discrete causal theory, to consider the case of topologies on enlarged sets containing a multidirected set, I formalize topological local finiteness in the following generalized way:\\ \begin{defi}\label{topologicallocalfiniteness} Let $S$ and $\overline{S}$ be sets, with $S$ a subset of $\overline{S}$, and let $\overline{\ms{T}}$ be a topology on $\overline{S}$. Then $S$ is called {\bf topologically locally finite} with respect to the pair $(\overline{S},\overline{\ms{T}})$ if every element of $S$ is contained in an element of $\overline{\ms{T}}$ of finite cardinality. \end{defi} Topological local finiteness may be informally reduced to the statement that {\it every element has a finite neighborhood.} However, the definition of ``neighborhood" depends on the choice of topology, and neighborhoods are generally subsets of the enlarged set $\overline{S}$, rather than subsets of $S$ itself.\footnotemark\footnotetext{More generally, one may define a local condition {\it at a particular element} $x$ in $S$, then say that $S$ satisfies the condition (globally) if it satisfies the condition at every element. For example, a generic spacetime in general relativity is smooth almost everywhere, but is {\it not} globally smooth because of the singularity theorems.} \refstepcounter{textlabels}\label{fourtopspaces} {\bf Four Illustrative Topologies.} {\it Many} different topological spaces may be associated with a multidirected set $M=(M,R,i,t)$. Some involve topologies on the underlying set of $M$, while others involve topologies on an enlarged set containing the underlying set of $M$. Here, I discuss just four illustrative examples of such spaces. Their topologies are called the {\it discrete}, {\it interval}, {\it continuum}, and {\it star topologies}.\footnotemark\footnotetext{The interval and continuum topologies are each analogous to the Euclidean metric topology on the real line, but in different ways. The star topology is distinct from the identically named notion appearing in the study of computer networks, although the underlying idea is similar.} The discrete topology and the interval topology are topologies on the underlying set of $M$, while the continuum topology and the star topology are topologies on enlarged sets $M_\mathbb{R}$ and $M_\star$ containing the underlying set of $M$, called the {\it continuum model} and the {\it star model} of $M$, respectively. In topological terms, the goals of this section are modest, and indeed primitive, compared to what appears in the graph-theoretic and order-theoretic literature. The four topological spaces discussed here are chosen specifically to facilitate and clarify the study of local finiteness in causal theory. The star topology on the star model of a multidirected set $M$ is the most useful of these four spaces for defining causally local properties in the context of classical spacetime structure. In this setting, {\it topological local finiteness} for the star topology coincides with both {\it causal local finiteness} in the sense of definition \hyperref[definitioncausallocality]{\ref{definitioncausallocality}} above, as shown in lemma \hyperref[localfinitenesslemma]{\ref{localfinitenesslemma}} below, and the axiom of {\it local finiteness} (\hyperref[lf]{LF}), introduced in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} below, which is {\it chosen} to generalize classical causal local finiteness. For a more detailed topological perspective in the special case of causal sets, the sources cited in Sumati Surya's review article \cite{Surya11} provide an excellent starting point. From a broader perspective, it is instructive to study the various topologies arising in domain theory, such as those discussed by Martin and Panangaden in \cite{Martin06}. Figure \hyperref[topologies]{\ref{topologies}} below illustrates small neighborhoods of an element $x$ in a multidirected set $M$, for each of the four topologies listed above. The unusual appearance of the diagrams in figures \hyperref[topologies]{\ref{topologies}}c and \hyperref[topologies]{\ref{topologies}}d is due to the fact that these diagrams represent the enlarged sets $M_\mathbb{R}$ and $M_\star$; the precise meaning of these diagrams is explained below. As usual, these figures only involve acyclic directed sets, but the associated definitions apply to multidirected sets in general. The reader should keep in mind that ``strange things can happen" when cycles are admitted; for example, an open interval of the form $\llangle x,x \rrangle$ may be nonempty in this context. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{4.9cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(0.15,4.5)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(4.275,4.5)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(8.4,4.5)}{\pgfbox[left,center]{c)}} \pgfputat{\pgfxy(12.525,4.5)}{\pgfbox[left,center]{d)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,4.8)(16.5,4.8) \pgfxyline(0,-.1)(0,4.8) \pgfxyline(4.125,-.1)(4.125,4.8) \pgfxyline(8.25,-.1)(8.25,4.8) \pgfxyline(12.375,-.1)(12.375,4.8) \pgfxyline(16.5,-.1)(16.5,4.8) \begin{pgftranslate}{\pgfpoint{-.1cm}{0cm}} \pgfputat{\pgfxy(1.35,2.6)}{\pgfbox[center,center]{$x$}} \begin{pgfmagnify}{.75}{.75} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \color{black} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.13cm} \color{black} \end{pgfmagnify} \pgfputat{\pgfxy(2.6,2)}{\pgfbox[center,center]{$M$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.025cm}{0cm}} \pgfputat{\pgfxy(1.6,1.8)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.35,2.6)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(2.5,2.5)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(1.1,3.7)}{\pgfbox[center,center]{$y$}} \begin{pgfmagnify}{.75}{.75} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.13cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.10cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.13cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.10cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.13cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.13cm} \pgfsetlinewidth{1pt} \color{white} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.13cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.13cm} \color{black} \pgfsetlinewidth{1pt} \pgfnodecircle{Node8}[stroke]{\pgfxy(1.6,4.6)}{0.13cm} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.13cm} \end{pgfmagnify} \pgfputat{\pgfxy(2.6,2)}{\pgfbox[center,center]{$M$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.15cm}{0cm}} \pgfputat{\pgfxy(1.35,2.6)}{\pgfbox[center,center]{$x$}} \begin{pgfmagnify}{.75}{.75} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[virtual]{\pgfxy(.7,.5)}{0cm} \pgfnodecircle{Node2}[virtual]{\pgfxy(.6,2.5)}{0cm} \pgfnodecircle{Node3}[virtual]{\pgfxy(.5,5)}{0cm} \pgfnodecircle{Node4}[virtual]{\pgfxy(1.1,1.5)}{0cm} \pgfnodecircle{Node5}[virtual]{\pgfxy(.9,2)}{0cm} \pgfnodecircle{Node6}[virtual]{\pgfxy(1,3.7)}{0cm} \pgfnodecircle{Node7}[virtual]{\pgfxy(1.5,3.4)}{0cm} \pgfnodecircle{Node8}[virtual]{\pgfxy(1.6,4.6)}{0cm} \pgfnodecircle{Node9}[virtual]{\pgfxy(2,.9)}{0cm} \pgfnodecircle{Node10}[virtual]{\pgfxy(1.9,1.5)}{0cm} \pgfnodecircle{Node11}[virtual]{\pgfxy(2.5,2)}{0cm} \pgfnodecircle{Node12}[virtual]{\pgfxy(2,4.2)}{0cm} \pgfnodecircle{Node13}[virtual]{\pgfxy(2.5,2.6)}{0cm} \pgfnodecircle{Node14}[virtual]{\pgfxy(2.4,4)}{0cm} \pgfnodecircle{Node15}[virtual]{\pgfxy(2.3,5.1)}{0cm} \pgfnodecircle{Node16}[virtual]{\pgfxy(3,.7)}{0cm} \pgfnodecircle{Node17}[virtual]{\pgfxy(3,2.4)}{0cm} \pgfnodecircle{Node18}[virtual]{\pgfxy(3,3.3)}{0cm} \pgfnodecircle{Node19}[virtual]{\pgfxy(2.9,4.8)}{0cm} \pgfnodecircle{Node20}[virtual]{\pgfxy(3.4,1.4)}{0cm} \pgfnodecircle{Node21}[virtual]{\pgfxy(3.7,3.5)}{0cm} \pgfnodecircle{Node22}[virtual]{\pgfxy(3.4,4)}{0cm} \pgfnodecircle{Node23}[virtual]{\pgfxy(3.3,4.7)}{0cm} \pgfnodecircle{Node24}[virtual]{\pgfxy(4,.5)}{0cm} \pgfnodecircle{Node25}[virtual]{\pgfxy(3.9,1.3)}{0cm} \pgfnodecircle{Node26}[virtual]{\pgfxy(4,2.5)}{0cm} \pgfnodecircle{Node27}[virtual]{\pgfxy(4,4.3)}{0cm} \pgfnodecircle{Node28}[virtual]{\pgfxy(4.4,1.5)}{0cm} \pgfnodecircle{Node29}[virtual]{\pgfxy(4.5,3.3)}{0cm} \pgfnodecircle{Node30}[virtual]{\pgfxy(4.5,5.2)}{0cm} \pgfnodecircle{Node31}[virtual]{\pgfxy(5,.6)}{0cm} \pgfnodecircle{Node32}[virtual]{\pgfxy(5.1,1.5)}{0cm} \pgfnodecircle{Node33}[virtual]{\pgfxy(5,2.5)}{0cm} \pgfnodecircle{Node34}[virtual]{\pgfxy(5.2,3.5)}{0cm} \pgfnodecircle{Node35}[virtual]{\pgfxy(5,4.5)}{0cm} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \color{white} \pgfsetlinewidth{3pt} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \color{black} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfxyline(2.5,2.6)(2.53,2.73) \pgfxyline(1.1,1.5)(1.14,1.57) \pgfxyline(2,.9)(2.02,.95) \pgfxyline(2.5,2.6)(2.5,2.48) \pgfxyline(2.5,2.6)(2.53,2.66) \pgfxyline(2.4,4)(2.36,3.95) \pgfxyline(2.4,4)(2.38,3.92) \pgfxyline(3,2.4)(2.97,2.35) \pgfxyline(3,3.3)(2.97,3.25) \pgfxyline(3,2.4)(3,2.3) \pgfxyline(3.4,1.4)(3.4,1.5) \pgfxyline(3.4,1.4)(3.43,1.45) \pgfxyline(3.9,1.3)(3.95,1.45) \pgfxyline(3.9,1.3)(3.95,1.36) \pgfxyline(4,2.5)(3.97,2.45) \pgfxyline(4.5,3.3)(4.46,3.2) \end{colormixin} \pgfsetlinewidth{1pt} \pgfxyline(1.5,3.4)(1,4.2) \pgfxyline(1.5,3.4)(1.55,4) \pgfxyline(1.5,3.4)(1.05,2.95) \pgfxyline(1.5,3.4)(1.2,2.7) \pgfxyline(1.5,3.4)(1.7,2.45) \pgfxyline(1.5,3.4)(2,3) \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.13cm} \color{black} \pgfsetlinewidth{1pt} \end{pgfmagnify} \pgfputat{\pgfxy(2.6,2)}{\pgfbox[center,center]{$M_{\mathbb{R}}$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{12.275cm}{0cm}} \pgfputat{\pgfxy(1.35,2.6)}{\pgfbox[center,center]{$x$}} \begin{pgfmagnify}{.75}{.75} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[virtual]{\pgfxy(.7,.5)}{0.15cm} \pgfnodecircle{Node2}[virtual]{\pgfxy(.6,2.5)}{0.15cm} \pgfnodecircle{Node3}[virtual]{\pgfxy(.5,5)}{0.15cm} \pgfnodecircle{Node4}[virtual]{\pgfxy(1.1,1.5)}{0.15cm} \pgfnodecircle{Node5}[virtual]{\pgfxy(.9,2)}{0.15cm} \pgfnodecircle{Node6}[virtual]{\pgfxy(1,3.7)}{0.15cm} \pgfnodecircle{Node7}[virtual]{\pgfxy(1.5,3.4)}{0.15cm} \pgfnodecircle{Node8}[virtual]{\pgfxy(1.6,4.6)}{0.15cm} \pgfnodecircle{Node9}[virtual]{\pgfxy(2,.9)}{0.15cm} \pgfnodecircle{Node10}[virtual]{\pgfxy(1.9,1.5)}{0.15cm} \pgfnodecircle{Node11}[virtual]{\pgfxy(2.5,2)}{0.15cm} \pgfnodecircle{Node12}[virtual]{\pgfxy(2,4.2)}{0.15cm} \pgfnodecircle{Node13}[virtual]{\pgfxy(2.5,2.6)}{0.15cm} \pgfnodecircle{Node14}[virtual]{\pgfxy(2.4,4)}{0.15cm} \pgfnodecircle{Node15}[virtual]{\pgfxy(2.3,5.1)}{0.15cm} \pgfnodecircle{Node16}[virtual]{\pgfxy(3,.7)}{0.15cm} \pgfnodecircle{Node17}[virtual]{\pgfxy(3,2.4)}{0.15cm} \pgfnodecircle{Node18}[virtual]{\pgfxy(3,3.3)}{0.15cm} \pgfnodecircle{Node19}[virtual]{\pgfxy(2.9,4.8)}{0.15cm} \pgfnodecircle{Node20}[virtual]{\pgfxy(3.4,1.4)}{0.15cm} \pgfnodecircle{Node21}[virtual]{\pgfxy(3.7,3.5)}{0.15cm} \pgfnodecircle{Node22}[virtual]{\pgfxy(3.4,4)}{0.15cm} \pgfnodecircle{Node23}[virtual]{\pgfxy(3.3,4.7)}{0.15cm} \pgfnodecircle{Node24}[virtual]{\pgfxy(4,.5)}{0.15cm} \pgfnodecircle{Node25}[virtual]{\pgfxy(3.9,1.3)}{0.15cm} \pgfnodecircle{Node26}[virtual]{\pgfxy(4,2.5)}{0.15cm} \pgfnodecircle{Node27}[virtual]{\pgfxy(4,4.3)}{0.15cm} \pgfnodecircle{Node28}[virtual]{\pgfxy(4.4,1.5)}{0.15cm} \pgfnodecircle{Node29}[virtual]{\pgfxy(4.5,3.3)}{0.15cm} \pgfnodecircle{Node30}[virtual]{\pgfxy(4.5,5.2)}{0.15cm} \pgfnodecircle{Node31}[virtual]{\pgfxy(5,.6)}{0.15cm} \pgfnodecircle{Node32}[virtual]{\pgfxy(5.1,1.5)}{0.15cm} \pgfnodecircle{Node33}[virtual]{\pgfxy(5,2.5)}{0.15cm} \pgfnodecircle{Node34}[virtual]{\pgfxy(5.2,3.5)}{0.15cm} \pgfnodecircle{Node35}[virtual]{\pgfxy(5,4.5)}{0.15cm} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \end{colormixin} \pgfsetlinewidth{1pt} \pgfxyline(1.5,3.4)(1.58,4.48) \pgfxyline(1.5,3.4)(.57,4.9) \pgfxyline(1.5,3.4)(.69,2.58) \pgfxyline(1.5,3.4)(.95,2.12) \pgfxyline(1.5,3.4)(1.88,1.62) \pgfxyline(1.5,3.4)(2.4,2.66) \color{white} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.16cm} \color{black} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.13cm} \color{black} \pgfsetlinewidth{1pt} \end{pgfmagnify} \pgfputat{\pgfxy(2.6,2)}{\pgfbox[center,center]{$M_\star$}} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{Small neighborhood of an element $x$ in various topologies: a) discrete topology on $M$; b) interval topology on $M$; c) continuum topology on $M_\mathbb{R}$; d) star topology on $M_\star$.} \label{topologies} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{discretetop} {\bf Discrete Topology.} The discrete topology $\ms{T}_{\tn{\fsz{dis}}}$ on a multidirected set $M$ takes the individual elements of $M$ to be open sets. Since arbitrary unions of open sets are open, {\it every} subset of $M$ is open in the discrete topology. The discrete topology is too fine to be useful in causal theory. Indeed, as illustrated in figure \hyperref[topologies]{\ref{topologies}}a, every element $x$ of $M$ defines a singleton neighborhood of itself. Hence, one can decipher {\it nothing} about local multidirected structure near $x$ by examining arbitrarily small neighborhoods of $x$ in the discrete topology on $M$. \refstepcounter{textlabels}\label{inttop} {\bf Interval Topology.} The interval topology $\ms{T}_{\tn{\fsz{int}}}$ on a multidirected set $M$ is perhaps the most natural topology to study in the special case when $M$ is a causal set. In particular, it is the obvious topology to use for examining the local attributes of the axiom of interval finiteness (\hyperref[if]{IF}). A convenient subbasis for the interval topology on $M$ is the family of pasts and futures $D^\pm(x)$ of elements $x$ of $M$. An open interval $\llangle w,y\rrangle$, for two elements $w$ and $y$ in $M$, may be expressed as the intersection $D^+(w)\cap D^-(y)$ of subbasis elements.\footnotemark\footnotetext{One reason for defining the interval topology in this way, instead of using open intervals directly, is to avoid the nuisance of explicitly identifying ``half-open intervals," which include extremal elements, as members of the subbasis.} As noted above, ``strange properties" arise in the cyclic case; for example, the open interval $\llangle x,x \rrangle$ is nonempty whenever $x$ lies on a cycle. The reader should note that the term {\it interval topology} has many alternative meanings besides the one used here, even in the specific context of partially ordered sets.\footnotemark\footnotetext{Consider, for example, the ``interval topologies" discussed by Marcel Ern\'{e} in \cite{ErneTopologies80}.} The interval topology is one possible causal analogue of the {\it Alexandrov topology} on a pseudo-Riemannian manifold $X$ in the context of general relativity,\footnotemark\footnotetext{This is the source of the term {\it ``Alexandroff set,"} appearing in the quote of Bombelli, Lee, Meyer and Sorkin in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} above. The term {\it Alexandrov topology} has a slightly different meaning in modern mathematics, but the above usage is too deeply rooted to amend, appearing in classic relativity texts such as Hawking and Ellis \cite{HawkingEllis73}.} which has subbasis consisting of the {\it chronological} pasts and futures $I^\pm(x)$ of spacetime events $x$ in $X$. The Alexandrov topology is generally coarser than the manifold topology on $X$ induced by the Euclidean metric topology, but the two coincide for {\it strongly causal spacetimes,} which is one reason why the interval topology is {\it a priori} interesting in causal set theory. Nontrivial choices are involved in abstracting the Alexandrov topology to the discrete causal context,\footnotemark\footnotetext{For example, the chronological pasts and futures $I^\pm(x)$ exclude the boundaries of the corresponding causal pasts and futures $J^\pm(x)$ in the relativistic context. One way to abstract this condition to the discrete causal context would be to exclude $y$ from $D^{+}(x)$ whenever $x\prec y$ is an {\it irreducible} relation, and similarly for the past.} but these choices do not significantly affect the conclusions necessary for this paper. Although the interval topology respects multidirected structure in an obvious sense, it is nonetheless a poor choice for defining local properties of multidirected sets. For example, in the context of classical spacetime, the interval topology on a directed set generally fails to capture and isolate local causal structure near a given event. This is partly because the interval topology is too coarse, but it also has other deficiencies. These are elaborated in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}}, and illustrated in the examples appearing in section \hyperref[subsectionintervalfinitenessdeficient]{\ref{subsectionintervalfinitenessdeficient}} below. Figure \hyperref[topologies]{\ref{topologies}}b above illustrates a small neighborhood of an element $x$ in the interval topology on $M$, consisting of the four elements indicated by the dark nodes. In this case, the chosen neighborhood is itself an open interval. Observe that this neighborhood contains elements, such as $u$, that are {\it causally unrelated} to $x$. This is a generic problem for the interval topology. For future reference, note that {\it interval finiteness is not the same as topological local finiteness for the interval topology,} since interval finiteness requires that {\it every} open interval must be finite, not merely that arbitrarily small intersections of open intervals must be finite. Also note that an open set in the interval topology does not always contain an open interval; i.e., open intervals do not form a basis for the interval topology. Note, by contrast, that the Alexandrov sets $I^+(w)\cap I^-(y)$ {\it do} form a basis for the Alexandrov topology in the relativistic context. \refstepcounter{textlabels}\label{conttop} {\bf Continuum Topology.} A ``concrete," though cumbersome, enlarged set containing a multidirected set $M=(M,R,i,t)$, may be constructed by identifying each element of the relation set $R$ of $M$ with a {\it continuum line segment,} then gluing these segments together at their endpoints in an appropriate way. This approach leads to a topology, called the {\it continuum topology,} on the resulting quotient space, called the {\it continuum model} of $M$. More precisely, define the {\bf relation set} $R(x)$ at an element $x$ in $M$ to be the subset of $R$ consisting of all relations $r$ such that either $i(r)=x$, or $t(r)=x$, or both. Identify every element of $R$ with a copy of the closed unit interval $[0,1]$ in $\mathbb{R}$, equipped with its Euclidean metric topology. Label the initial and terminal points of each interval with the initial and terminal elements of the corresponding relation. Define the {\bf continuum model} $M_\mathbb{R}$ of $M$ to be the quotient space given by identifying all endpoints of intervals with matching labels. In particular, the set of intervals with at least one endpoint labeled by $x$ corresponds to $R(x)$. The quotient topology $\ms{T}_\mathbb{R}$ on $M_\mathbb{R}$ is called the {\bf continuum topology}. Although the continuum topology involves a great deal of {\it extrinsic structure,} it has the advantage of providing natural families of neighborhoods that capture and isolate local multidirected structure near each element $x$ of $M$. In particular, for every $0< \rho<1$, define the {\bf (open) star at $x$ of radius $\rho$} to be the {\it star-shaped subset} $\tn{St}_\rho(x)$ of $M_\mathbb{R}$ with {\it limbs of length $\rho$} glued at $x$; i.e., the subset \[\tn{St}_\rho(x):=\{t\in M_\mathbb{R}\hspace*{.1cm}|\hspace*{.1cm} d(t,x)<\rho\},\] where $d$ is the Euclidean metric on each unit interval. Any neighborhood $U$ of $x$ in the continuum topology on $M_\mathbb{R}$ contains $\tn{St}_\rho(x)$ for sufficiently small values of $\rho$. Hence, given such a neighborhood $U$, one may ascertain the cardinality of the relation set $R(x)$ at $x$, and hence the local multidirected structure near $x$, by examining an appropriate star $\tn{St}_\rho(x)$ contained in $U$. Note that if $r$ is a reflexive relation $x\prec x$, and if $\rho\ge 1/2$, then $\tn{St}_\rho(x)$ contains the {\it entire interval corresponding to} $r$, so {\it``star-shaped"} is merely a suggestive term, carried over from the acyclic case. Figure \hyperref[topologies]{\ref{topologies}}c above illustrates the star $\tn{St}_{1/2}(x)$ at an element $x$ in the continuum topology on $M_\mathbb{R}$, with limbs represented by the dark segments. The figure is deliberately drawn to suggest a {\it one-dimensional subspace of three-dimensional Euclidean space.} This is intended to emphasize the difference between $M$ and $M_\mathbb{R}$, particularly the large number of extra elements of $M_\mathbb{R}$. \refstepcounter{textlabels}\label{startop} {\bf Star Topology.} The best features of the continuum topology may be retained without the burden of extrinsic structure, as I now explain. Define the {\bf star model} $M_\star$ of $M=(M,R,i,t)$ to be the union $M_\star:=M\cup R$ of (the underlying set of) $M$ with its relation set $R$. For each element $x$ in $M$, define the {\bf star at $x$ in $M_\star$} to be the subset \[\tn{St}(x):=\{x\}\cup R(x),\] of $M_\star$. Define the {\bf star topology} $\ms{T}_\star$ on $M_\star$ to be the topology with subbasis $\{\tn{St}(x)\hspace*{.1cm}|\hspace*{.1cm} x\in M\}$. It is useful to compare $\tn{St}(x)$ with the corresponding open stars $\tn{St}_\rho(x)$, defined above in the context of the continuum topology. In particular, note that while there exist uncountably many open stars $\tn{St}_\rho(x)$ of different radii, $\tn{St}(x)$ is unique. Unlike the extra elements of $M_\mathbb{R}$, which originate ultimately from the continuum interval $[0,1]$, an external source {\it a priori} unrelated to $M$, the extra elements $R$ of $M_\star$ are an intrinsic, uniquely-defined part of the structure of $M=(M,R,i,t)$. Variations of the star topology may be defined in terms of {\it irreducible stars,} {\it past stars,} {\it future stars,} and so on. Figure \hyperref[topologies]{\ref{topologies}}d above illustrates the star $\tn{St}(x)$ at an element $x$ in the star topology on $M_\star$. As in the case of the continuum topology, the appearance of this figure warrants some comment. The elements of the relation set $R$ of $M=(M,R,i,t)$ now have a {\it dual role:} they serve both as {\it relations} between pairs of elements of $M$ in the usual sense, and as {\it elements} of $M_\star$ in their own right. I have schematically highlighted this difference by drawing the edges in the ``Hasse diagram" for $M_\star$ slightly darker than the corresponding edges in figures \hyperref[topologies]{\ref{topologies}}a and \hyperref[topologies]{\ref{topologies}}b, and by slightly separating the edges from the nodes representing the elements of $M$. I represent $R$ in a different diagrammatic fashion in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below, where it serves as the underlying set of the {\it relation space} $\ms{R}(M)$ over $M$. \refstepcounter{textlabels}\label{cardvalscalar} {\bf Cardinal-Valued Scalar Fields in the Star Topology.} It is interesting to consider the question of how much information about a multidirected set $M=(M,R,i,t)$ may be recovered from the star topology on $M_\star$. The star topology detects the {\it valences} $v(x)$ at each element $x$ of $M$; i.e., the number of relations incident at $x$.\footnotemark\footnotetext{Reflexive relations $x\prec x$ are counted twice.} These valences form a {\bf cardinal-valued scalar field} $v_M$ on $M$, called the {\bf valence field} of $M$. An arbitrary cardinal-valued scalar field on an arbitrary set $S$ is generally {\it not} a valence field for any multidirected structure on $S$. For example, there is no multidirected structure on a singleton $S=\{x\}$ such that $v_S(x)=3$.\footnotemark\footnotetext{The inverse problem of recovering a ``multidirected set" from an arbitrary ``valence field" suggests consideration of generalized multidirected sets in which the initial and terminal element maps are only partially defined. This idea is actually somewhat interesting, due to an analogy involving the {\it causal path semicategories} studied in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}. However, I do not explore it in this paper.} The valence field $v_M$ of a multidirected set $M$ may be decomposed into past and future components $v_{M}^\pm$, where $v_M^-(x)$ is the cardinality of the set of relations terminating at $x$, and $v_M^+(x)$ is the cardinality of the set of relations beginning at $x$. Unlike the valence field itself, the past and future valence fields generally {\it cannot} be detected by the star topology. For each relation $r$ in the relation set $R$ of $M$, one may consider linear combinations of $v_M(i(r))$ and $v_M(t(r))$ or of $v_M^\pm(i(r))$ and $v_M^\pm(t(r))$; these are analogues of {\it timelike vector fields} in an obvious sense. Under the axiom of locally finiteness, introduced in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} below, all such fields take nonnegative integer values, and their arithmetic properties are of significant interest. \subsection{Interval Finiteness versus Local Finiteness}\label{subsectionintervalfiniteness} \refstepcounter{textlabels}\label{intcriticisms} {\bf Initial Criticisms of Interval Finiteness.} The causal set axiom of interval finiteness (\hyperref[if]{IF}) is utterly inadequate as a local finiteness condition for classical spacetime structure under Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). Physically, it permits the very type of pathological behavior an appropriate local finiteness condition should rule out. It is not {\it causally local} in the sense of definition \hyperref[definitioncausallocality]{\ref{definitioncausallocality}} above, since the direct causes and effects associated with a given element generally cannot be isolated in an open interval. Mathematically, it is not a local condition for the interval topology, since it involves arbitrarily large neighborhoods, nor is it a consequence of the topological local finiteness condition induced by the interval topology. Worse, open intervals do not even {\it capture} local multidirected structure in any suitable sense, let alone isolate such structure. These shortcomings, and others, are discussed in more detail below. \refstepcounter{textlabels}\label{locfinintro} {\bf Local Finiteness.} Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}) virtually demands {\it some} type of local finiteness condition, since the associated measure axiom (\hyperref[m]{M}) assigns an infinite spacetime volume to an infinite subset of a directed set. In this context, local infinities produce absurd behavior of the worst sort from a physical perspective. An example of such behavior occurs when an element $x$ in a directed set has an {\it infinite number of maximal predecessors or minimal successors,} corresponding to the instantaneous collapse or expansion of an infinite volume of spacetime. Certain other types of ``locally infinite" behavior may be interpreted as artifacts of transitivity, but this particular type is a fatal threat to any theory permitting it. One could try to formulate a different version of the causal metric hypothesis, in an effort to ameliorate this problem without imposing a local finiteness condition. For example, one might consider altering the measure axiom to admit discrete measures assigning {\it arbitrarily small volumes} to subsets of a directed set; i.e., measures without an {\it effective volume gap.} However, this strategy would lead far away from the basic insight that ``number" can serve as a proxy for volume in the context of Malament's metric recovery theorem. Axiomatic changes of such significance should be made for physical reasons, not for the purpose of rescuing other axioms.\footnotemark\footnotetext{As discussed at the end of this section, there are legitimate reasons to consider {\it relaxing} the measure axiom, to allow volume to depend on local causal structure. However, this requires a satisfactory treatment of local conditions, which open intervals cannot supply.} How, then, {\it should} local finiteness be axiomatized? The crucial {\it physical} requirement is that local finiteness should coincide with {\it causal local finiteness} in the context of classical spacetime structure. Modeling classical spacetime as a directed set $D$, and applying the independence convention (\hyperref[ic]{IC}), the definition of causal local finiteness translates to the condition that {\it the relation set $R(x)$ at each element $x$ of $D$ is finite.} Since the same condition is useful for general multidirected sets, I state the axiom in this context. \refstepcounter{textlabels}\label{lf} \hspace*{.3cm} LF. {\bf Local Finiteness}: {\it For every element $x$ of $M$, the relation set $R(x)$ at $x$ is finite.} In the special case of acyclic directed sets, the cardinality of $R(x)$ is the same as the number of {\it direct predecessors and successors} of $x$. Hence, local finiteness may be stated in terms of {\it elements} in this case. For directed sets including cycles, this correspondence is altered slightly, by the possibility of reflexive relations, and relations in both directions between a given pair of elements. Finally, for multidirected sets, one may imagine situations, such as the existence of an {\it infinite number of reflexive relations at a singleton,}\footnotemark\footnotetext{This is actually not so strange; it is a causal analogue of the category-theoretic representation of an infinite group.} in which local finiteness cannot be determined by counting predecessors and successors. For these reasons, it is best in the general case to express local finiteness in terms of relation sets. Local finiteness may be detected by either the continuum topology or the star topology. In fact, as mentioned above, local finiteness is the same condition as topological local finiteness for a multidirected set $M$ with respect to the star topology on the star model $M_\star$ of $M$. This is the content of lemma \hyperref[localfinitenesslemma]{\ref{localfinitenesslemma}}. \vspace*{.2cm} \refstepcounter{textlabels}\label{locfintoplocfinstar} \begin{lem}\label{localfinitenesslemma} Local finiteness \tn{(\hyperref[lf]{\tn{LF}})} for a multidirected set $M=(M,R,i,t)$ coincides with topological local finiteness of $M$ with respect to the pair $(M_\star,\ms{T}_\star)$. \end{lem} \begin{proof} Local finiteness means that the relation set $R(x)$ is finite for each element $x$ of $M$. This is true if and only if the star $\tn{St}(x)$ at $x$ is finite for each $x$ in $M$. Since $\tn{St}(x)$ is a neighborhood of $x$ in $M_\star$ with respect to $\ms{T}_\star$, this immediately implies topological local finiteness of $M$ with respect to $(M_\star,\ms{T}_\star)$. Conversely, the intersection of two stars $\tn{St}(x)\cap\tn{St}(y)$ is either a family of relations between $x$ and $y$, or empty. Hence, the smallest neighborhood of $x$ in the star topology on $M_\star$ is $\tn{St}(x)$. Therefore, topological local finiteness of $M$ with respect to $(M_\star,\ms{T}_\star)$, implies that $\tn{St}(x)$, and hence $R(x)$, is finite for every $x$ in $M$. \end{proof} Another equivalent condition is that the valence field $v_M$ of $M$ takes values in $\mathbb{N}$. \refstepcounter{textlabels}\label{varlocfin} {\bf Variations on Local Finiteness for Directed Sets.} The axiom of local finiteness (\hyperref[lf]{LF}) may be modified in a number of different ways. For example, in the general context of multidirected sets, one may first collapse all relations of the form $x\prec y$, for each pair of elements $x$ and $y$ in $M$, into a single relation, before testing for local finiteness. This leads to the condition of {\it local element finiteness} (\hyperref[lef]{LEF}), introduced in section \hyperref[subsectionhierarchyfiniteness]{\ref{subsectionhierarchyfiniteness}} below. In the special case of directed sets, it is sometimes useful to consider alternative finiteness conditions involving the addition or removal of {\it reducible relations.} Though inferior in some ways to local finiteness for describing classical spacetime structure, such conditions have a sufficiently important auxiliary role to warrant mention. The two extreme cases of such conditions admit convenient descriptions in terms of the transitive closure and the skeleton. \refstepcounter{textlabels}\label{lft} \hspace*{.3cm} LFT. {\bf Local Finiteness in the Transitive Closure}: $\tn{tr}(D)$ is locally finite. \refstepcounter{textlabels}\label{lfs} \hspace*{.3cm} LFS.\hspace*{.05cm} {\bf Local Finiteness in the Skeleton}: $\tn{sk}(D)$ is locally finite. Local finiteness in the transitive closure is not a causally local condition, since it requires the entire past and future of each element of $D$ to be finite. This does {\it not} imply that $D$ itself has finite ``temporal size;" for example, a disjoint union of chains of each nonnegative integer length is locally finite in the transitive closure. However, this condition is still too restrictive of global structure to serve as a suitable replacement for interval finiteness in the context of classical spacetime. Local finiteness in the skeleton, meanwhile, is far too weak a condition to restrict attention to physically plausible directed sets under Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). In particular, this condition permits any interpolative directed set, including the continuum, Minkowski spacetime, domains, etc., since the skeleta of such sets have no relations. It also permits ``pathologically complex cycles" in ``otherwise simple" directed sets, since the skeleton operation destroys cycles. Local finiteness in the skeleton is interesting primarily in concert with other axioms. For example, if one insists on retaining interval finiteness, local finiteness in the skeleton cures some of its deficiencies. \refstepcounter{textlabels}\label{intfinlocfinincomp} {\bf Incomparability of Interval Finiteness and Local Finiteness.} Interval finiteness (\hyperref[if]{IF}) and local finiteness (\hyperref[lf]{LF}) are incomparable conditions; i.e., neither condition implies the other. This is true even in the special case of acyclic directed sets. More comprehensive {\it pairwise} comparisons of various finiteness conditions, for both acyclic directed sets and multidirected sets, are given in theorems \hyperref[theoremhierarchyfiniteness]{\ref{theoremhierarchyfiniteness}} and \hyperref[theoremhierarchyfinitenessmulti]{\ref{theoremhierarchyfinitenessmulti}} of section \hyperref[subsectionhierarchyfiniteness]{\ref{subsectionhierarchyfiniteness}} below. In this section, I merely present a few illustrative counterexamples. To see that interval finiteness does not imply local finiteness, consider the causal set with elements $\{w\}\cup\{x_i\hspace*{.1cm}|\hspace*{.1cm}i\in\mathbb{Z}\}$, and relations $w\prec x_i$ for all $i$, illustrated in figure \hyperref[jacobs]{\ref{jacobs}}a below. For future reference, I call this causal set the {\it infinite bouquet}. The infinite bouquet possesses no nonempty open intervals, since its longest chain has length one, and is therefore interval finite. However, it is not locally finite, even in the skeleton, because the element $w$ has an infinite number of minimal successors. The example of the infinite bouquet illustrates how interval finiteness permits the worst possible type of locally infinite behavior under Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). This type of pathology is not limited to causal sets of small ``temporal size." For example, figure \hyperref[jacobs]{\ref{jacobs}}b illustrates part of the skeleton of a physically absurd causal set with more than two generations. This example demonstrates how interval finiteness permits all manner of locally infinite behavior, restricted only by the condition that such behavior is not {\it complemented nonlocally} in such a way as to create infinite intervals. As mentioned above, such behavior includes both instantaneous expansion, and instantaneous collapse, of an infinite volume of spacetime, under the measure axiom (\hyperref[m]{M}). Dismissing implausible speculation about the Big Bang, the Big Crunch, eternal inflation, vacuum collapse, and other universal cataclysms, such scenarios have no reasonable place in discrete causal theory. To see that local finiteness does not imply interval finiteness, consider the acyclic directed set with elements $\{x_0, x_1,...\}\cup\{...,y_{-1},y_0\}$, and relations \[x_i\prec x_{i+1}, \hspace*{.5cm} x_i\prec y_{-i}, \hspace*{.5cm} y_{-j}\prec y_{-j+1},\] illustrated in figure \hyperref[jacobs]{\ref{jacobs}}c. I refer to this set as {\it Jacob's ladder}.\footnotemark\footnotetext{This is a reference to Genesis 28:12, in which the patriarch Jacob dreams of a ladder reaching from earth to heaven.} Each element $x_i$, with the exception of $x_0$, has exactly one direct predecessor and two direct successors, and each element $y_{-j}$, with the exception of $y_0$, has exactly two direct predecessors and one direct successor. The elements $x_0$ and $y_0$ are unique global minimal and maximal elements of Jacob's ladder, with two direct successors, and two direct predecessors, respectively. Hence, Jacob's ladder is locally finite. However, for every choice of nonnegative integers $i$ and $j$, there exist arbitrarily long chains of the form \[x_i\prec x_{i+1}\prec...\prec x_{i+n}\prec y_{-i-n}\prec y_{-i-n+1}\prec ...\prec y_{-j}.\] Thus, every open interval of the form $\llangle x_i,y_{-j}\rrangle$ in Jacob's ladder is infinite. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5) \pgfxyline(5.25,-.1)(5.25,5) \pgfxyline(11.4,-.1)(11.4,5) \pgfxyline(0,5)(16.5,5) \pgfxyline(16.5,-.1)(16.5,5) \pgfputat{\pgfxy(.15,4.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.4,4.7)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.55,4.7)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{-2.2cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(2.35,2)}{0.025cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2.45,2)}{0.025cm} \pgfnodecircle{Node3}[fill]{\pgfxy(2.55,2)}{0.025cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.75,2)}{0.065cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3,2)}{0.072cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.3,2)}{0.078cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.65,2)}{0.084cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4.05,2)}{0.09cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.5,2)}{0.095cm} \pgfnodecircle{Node12}[fill]{\pgfxy(5,2)}{0.095cm} \pgfnodecircle{Node13}[fill]{\pgfxy(5.45,2)}{0.09cm} \pgfnodecircle{Node14}[fill]{\pgfxy(5.85,2)}{0.084cm} \pgfnodecircle{Node15}[fill]{\pgfxy(6.2,2)}{0.078cm} \pgfnodecircle{Node16}[fill]{\pgfxy(6.5,2)}{0.072cm} \pgfnodecircle{Node17}[fill]{\pgfxy(6.75,2)}{0.065cm} \pgfnodecircle{Node20}[fill]{\pgfxy(6.95,2)}{0.025cm} \pgfnodecircle{Node21}[fill]{\pgfxy(7.05,2)}{0.025cm} \pgfnodecircle{Node22}[fill]{\pgfxy(7.15,2)}{0.025cm} \pgfnodecircle{Node23}[fill]{\pgfxy(4.75,0)}{0.1cm} \pgfnodeconnline{Node23}{Node6} \pgfnodeconnline{Node23}{Node7} \pgfnodeconnline{Node23}{Node8} \pgfnodeconnline{Node23}{Node9} \pgfnodeconnline{Node23}{Node10} \pgfnodeconnline{Node23}{Node11} \pgfnodeconnline{Node23}{Node12} \pgfnodeconnline{Node23}{Node13} \pgfnodeconnline{Node23}{Node14} \pgfnodeconnline{Node23}{Node15} \pgfnodeconnline{Node23}{Node16} \pgfnodeconnline{Node23}{Node17} \pgfputat{\pgfxy(4.75,-.3)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(4.4,2.3)}{\pgfbox[center,center]{$x_i$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.4cm}{.25cm}} \begin{pgftranslate}{\pgfpoint{3cm}{0cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.75cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2cm}{2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.5cm}{2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,1)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,1)}{0.07cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.2,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.3,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.4,1)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.75cm}{3cm}} \pgfnodecircle{Node0}[fill]{\pgfxy(-.4,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.3,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.2,1)}{0.025cm} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,1)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,1)}{0.07cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.2,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.3,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.4,1)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1cm}{0cm}} \pgfnodecircle{Node0}[fill]{\pgfxy(-.4,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.3,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.2,1)}{0.025cm} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,1)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.5cm}{1cm}} \pgfnodecircle{Node0}[fill]{\pgfxy(-.4,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.3,1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(-.2,1)}{0.025cm} \pgfnodecircle{Node1}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,1)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,1)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,1)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,0)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,0)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,0)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,0)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,0)}{0.07cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.2,0)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.3,0)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.4,0)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.5cm}{2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(0,0)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.25,0)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.5,0)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,0)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1.75,0)}{0.07cm} \pgfnodecircle{Node10}[fill]{\pgfxy(2,0)}{0.07cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.2,0)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.3,0)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.4,0)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node9} \pgfnodeconnline{Node1}{Node10} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.5cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,0)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.5cm}{2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,0)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.25cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.75,0)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.75cm}{1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.07cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,0)}{0.07cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,0)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.5,0)}{0.07cm} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \end{pgftranslate} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.8cm}{-.8cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(0,1.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.7,1.8)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.35,2.05)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.95,2.25)}{0.085cm} \pgfputat{\pgfxy(.1,1.2)}{\pgfbox[center,center]{$x_0$}} \pgfputat{\pgfxy(1.55,1.75)}{\pgfbox[center,center]{$x_i$}} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,2.42)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.95,2.55)}{0.075cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3.35,2.65)}{0.07cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.7,2.73)}{0.065cm} \pgfnodecircle{Node9}[fill]{\pgfxy(4,2.77)}{0.06cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4.25,2.8)}{0.055cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node9} \pgfnodeconnline{Node9}{Node10} \pgfnodecircle{Node21}[fill]{\pgfxy(0,4.5)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(.7,4.2)}{0.095cm} \pgfnodecircle{Node23}[fill]{\pgfxy(1.35,3.95)}{0.09cm} \pgfnodecircle{Node24}[fill]{\pgfxy(1.95,3.75)}{0.085cm} \pgfputat{\pgfxy(.1,4.8)}{\pgfbox[center,center]{$y_0$}} \pgfputat{\pgfxy(2.7,3.85)}{\pgfbox[center,center]{$y_{-j}$}} \pgfnodecircle{Node25}[fill]{\pgfxy(2.5,3.58)}{0.08cm} \pgfnodecircle{Node26}[fill]{\pgfxy(2.95,3.45)}{0.075cm} \pgfnodecircle{Node27}[fill]{\pgfxy(3.35,3.35)}{0.07cm} \pgfnodecircle{Node28}[fill]{\pgfxy(3.7,3.27)}{0.065cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4,3.23)}{0.06cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.25,3.2)}{0.055cm} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node23}{Node24} \pgfnodeconnline{Node24}{Node25} \pgfnodeconnline{Node25}{Node26} \pgfnodeconnline{Node26}{Node27} \pgfnodeconnline{Node27}{Node28} \pgfnodeconnline{Node28}{Node29} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node1}{Node21} \pgfnodeconnline{Node2}{Node22} \pgfnodeconnline{Node3}{Node23} \pgfnodeconnline{Node4}{Node24} \pgfnodeconnline{Node5}{Node25} \pgfnodeconnline{Node6}{Node26} \pgfnodeconnline{Node7}{Node27} \pgfnodeconnline{Node8}{Node28} \pgfnodeconnline{Node9}{Node29} \pgfnodeconnline{Node10}{Node30} \pgfnodecircle{Node21}[fill]{\pgfxy(4.4,3)}{0.025cm} \pgfnodecircle{Node22}[fill]{\pgfxy(4.5,3)}{0.025cm} \pgfnodecircle{Node23}[fill]{\pgfxy(4.6,3)}{0.025cm} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) the infinite bouquet is interval finite but not locally finite, even in the skeleton; b) absurd interval-finite behavior; c) Jacob's ladder is locally finite but not interval finite.} \label{jacobs} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{interactiontrans} {\bf Interaction with Transitivity.} The foregoing criticisms of interval finiteness (\hyperref[if]{IF}) might lead the reader to question why this axiom was ever proposed in the context of discrete causal theory. A {\it plausible} explanation is that it is an artifact of continuum theory, and the associated transitive paradigm, bolstered by the prominence of Alexandrov subsets and the Alexandrov topology in general relativity. For a transitive directed set $D$, viewed as a model of classical spacetime structure, local finiteness (\hyperref[lf]{LF}) is not a reasonable alternative to interval finiteness, since the direct past $D_0^-(x)$ and direct future $D_0^+(x)$ of an element $x$ in $D$ coincide with its total past $D^-(x)$ and total future $D^+(x)$, respectively. This situation is illustrated in figure \hyperref[interactiontransitivity]{\ref{interactiontransitivity}}a below. In this context, the independence convention (\hyperref[ic]{IC}) does not make sense physically, since ``most" relations terminating at $x$ come from its ``distant past," and ``most" relations beginning at $x$ reach to its ``distant future." Hence, the relation set $R(x)$ at $x$ is dominated by redundant information, and fails to isolate local causal structure in any suitable sense. From a different point of view, the condition of local finiteness is generally {\it too restrictive} under the transitive paradigm, since transitivity creates ``innocuous" locally infinite behavior. For example, the set $\mathbb{N}$ of nonnegative integers is locally infinite at each element, but is locally finite in the skeleton, as illustrated in figure \hyperref[interactiontransitivity]{\ref{interactiontransitivity}}b. Since $\mathbb{N}$ may be recovered from its skeleton by taking the transitive closure, this locally infinite behavior is ``no worse," in an information-theoretic sense, than the simple structure of $\tn{sk}(\mathbb{N})$. Any transitive cosmological model in discrete causal theory admitting unbounded chains is locally infinite for essentially the same reason, regardless of its actual information-theoretic complexity. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.35cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,4.95)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,4.95)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(5.5,-.1)(5.5,5.25) \pgfxyline(11,-.1)(11,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \begin{pgftranslate}{\pgfpoint{.2cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3.3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.5,1.5)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node1}{Node8} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node1}{Node13} \pgfnodeconnline{Node1}{Node12} \pgfnodeconnline{Node1}{Node14} \pgfnodeconnline{Node1}{Node15} \pgfnodeconnline{Node1}{Node18} \pgfnodeconnline{Node1}{Node19} \pgfnodeconnline{Node1}{Node22} \pgfnodeconnline{Node1}{Node23} \pgfnodeconnline{Node1}{Node30} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node2}{Node8} \pgfnodeconnline{Node4}{Node2} \pgfnodeconnline{Node4}{Node3} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node6} \pgfnodeconnline{Node4}{Node7} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node4}{Node12} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node4}{Node15} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node3} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node12} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node5}{Node15} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node2} \pgfnodeconnline{Node9}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node5} \pgfnodeconnline{Node9}{Node6} \pgfnodeconnline{Node9}{Node7} \pgfnodeconnline{Node9}{Node8} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node12} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node14} \pgfnodeconnline{Node9}{Node15} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node9}{Node18} \pgfnodeconnline{Node9}{Node19} \pgfnodeconnline{Node9}{Node22} \pgfnodeconnline{Node9}{Node23} \pgfnodeconnline{Node9}{Node30} \pgfnodeconnline{Node10}{Node3} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node8} \pgfnodeconnline{Node10}{Node12} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node10}{Node14} \pgfnodeconnline{Node10}{Node15} \pgfnodeconnline{Node10}{Node18} \pgfnodeconnline{Node10}{Node19} \pgfnodeconnline{Node10}{Node22} \pgfnodeconnline{Node10}{Node23} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node11}{Node3} \pgfnodeconnline{Node11}{Node7} \pgfnodeconnline{Node11}{Node8} \pgfnodeconnline{Node11}{Node12} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node11}{Node14} \pgfnodeconnline{Node11}{Node15} \pgfnodeconnline{Node11}{Node18} \pgfnodeconnline{Node11}{Node19} \pgfnodeconnline{Node11}{Node22} \pgfnodeconnline{Node11}{Node23} \pgfnodeconnline{Node11}{Node30} \pgfnodeconnline{Node12}{Node3} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node3} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node8} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node15} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node13}{Node22} \pgfnodeconnline{Node13}{Node23} \pgfnodeconnline{Node13}{Node30} \pgfnodeconnline{Node14}{Node3} \pgfnodeconnline{Node14}{Node8} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node3} \pgfnodeconnline{Node16}{Node7} \pgfnodeconnline{Node16}{Node8} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node12} \pgfnodeconnline{Node16}{Node13} \pgfnodeconnline{Node16}{Node14} \pgfnodeconnline{Node16}{Node15} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node18} \pgfnodeconnline{Node16}{Node19} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node16}{Node21} \pgfnodeconnline{Node16}{Node22} \pgfnodeconnline{Node16}{Node23} \pgfnodeconnline{Node16}{Node26} \pgfnodeconnline{Node16}{Node27} \pgfnodeconnline{Node16}{Node29} \pgfnodeconnline{Node16}{Node30} \pgfnodeconnline{Node16}{Node35} \pgfnodeconnline{Node17}{Node3} \pgfnodeconnline{Node17}{Node8} \pgfnodeconnline{Node17}{Node12} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node17}{Node15} \pgfnodeconnline{Node18}{Node3} \pgfnodeconnline{Node18}{Node8} \pgfnodeconnline{Node18}{Node12} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node15} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node18}{Node30} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node3} \pgfnodeconnline{Node20}{Node7} \pgfnodeconnline{Node20}{Node8} \pgfnodeconnline{Node20}{Node12} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node14} \pgfnodeconnline{Node20}{Node15} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node19} \pgfnodeconnline{Node20}{Node21} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node23} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node20}{Node27} \pgfnodeconnline{Node20}{Node29} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node20}{Node35} \pgfnodeconnline{Node21}{Node19} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node21}{Node30} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node22}{Node30} \pgfnodeconnline{Node24}{Node3} \pgfnodeconnline{Node24}{Node7} \pgfnodeconnline{Node24}{Node8} \pgfnodeconnline{Node24}{Node12} \pgfnodeconnline{Node24}{Node13} \pgfnodeconnline{Node24}{Node14} \pgfnodeconnline{Node24}{Node15} \pgfnodeconnline{Node24}{Node17} \pgfnodeconnline{Node24}{Node18} \pgfnodeconnline{Node24}{Node19} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node24}{Node21} \pgfnodeconnline{Node24}{Node22} \pgfnodeconnline{Node24}{Node23} \pgfnodeconnline{Node24}{Node26} \pgfnodeconnline{Node24}{Node27} \pgfnodeconnline{Node24}{Node29} \pgfnodeconnline{Node24}{Node30} \pgfnodeconnline{Node24}{Node35} \pgfnodeconnline{Node25}{Node3} \pgfnodeconnline{Node25}{Node8} \pgfnodeconnline{Node25}{Node12} \pgfnodeconnline{Node25}{Node14} \pgfnodeconnline{Node25}{Node15} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node19} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node22} \pgfnodeconnline{Node25}{Node23} \pgfnodeconnline{Node25}{Node27} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node30} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node25}{Node34} \pgfnodeconnline{Node25}{Node35} \pgfnodeconnline{Node26}{Node19} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node22} \pgfnodeconnline{Node26}{Node23} \pgfnodeconnline{Node26}{Node27} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node26}{Node30} \pgfnodeconnline{Node26}{Node35} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node27} \pgfnodeconnline{Node28}{Node29} \pgfnodeconnline{Node28}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node28}{Node34} \pgfnodeconnline{Node28}{Node35} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node3} \pgfnodeconnline{Node31}{Node8} \pgfnodeconnline{Node31}{Node12} \pgfnodeconnline{Node31}{Node14} \pgfnodeconnline{Node31}{Node15} \pgfnodeconnline{Node31}{Node17} \pgfnodeconnline{Node31}{Node19} \pgfnodeconnline{Node31}{Node21} \pgfnodeconnline{Node31}{Node22} \pgfnodeconnline{Node31}{Node23} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node26} \pgfnodeconnline{Node31}{Node27} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node29} \pgfnodeconnline{Node31}{Node30} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node31}{Node33} \pgfnodeconnline{Node31}{Node34} \pgfnodeconnline{Node31}{Node35} \pgfnodeconnline{Node32}{Node19} \pgfnodeconnline{Node32}{Node21} \pgfnodeconnline{Node32}{Node22} \pgfnodeconnline{Node32}{Node23} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node27} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node30} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node32}{Node34} \pgfnodeconnline{Node32}{Node35} \pgfnodeconnline{Node33}{Node27} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node30} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node34}{Node30} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.5,1.5)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node1}{Node7} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node7} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node9}{Node7} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node11}{Node7} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node16}{Node7} \pgfnodeconnline{Node20}{Node7} \pgfnodeconnline{Node24}{Node7} \pgfputat{\pgfxy(1.8,3.5)}{\pgfbox[center,center]{\large{$x$}}} \pgfputat{\pgfxy(4.35,2.3)}{\pgfbox[center,center]{\large{$D_0^-(x)=D^-(x)$}}} \pgfputat{\pgfxy(3,4.2)}{\pgfbox[center,center]{\large{$D_0^+(x)=D^+(x)$}}} \comment{ \color{white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.5,1.5)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \color{black} \pgfnodecircle{Node1}[stroke]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[stroke]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[stroke]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[stroke]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[stroke]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node8}[stroke]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[stroke]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[stroke]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[stroke]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node16}[stroke]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node20}[stroke]{\pgfxy(3.5,1.5)}{0.11cm} \pgfnodecircle{Node24}[stroke]{\pgfxy(4,.5)}{0.11cm} } \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.2cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3.3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.5,1.5)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node13}{Node7} \pgfputat{\pgfxy(1.8,3.5)}{\pgfbox[center,center]{\large{$x$}}} \pgfputat{\pgfxy(1.6,2.4)}{\pgfbox[center,center]{\large{$D_{\tn{max}}^-(x)$}}} \pgfputat{\pgfxy(1.6,5.1)}{\pgfbox[center,center]{\large{$D_{\tn{min}}^+(x)$}}} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6cm}{0cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1,1.3)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1,2.1)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1,2.9)}{0.085cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1,3.7)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,4.5)}{0.075cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1,4.7)}{0.025cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1,4.8)}{0.025cm} \pgfnodecircle{Node9}[fill]{\pgfxy(1,4.9)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconncurve{Node1}{Node3}{0}{0}{.5cm}{.5cm} \pgfnodeconncurve{Node1}{Node4}{0}{0}{.7cm}{.7cm} \pgfnodeconncurve{Node1}{Node5}{0}{0}{.9cm}{.9cm} \pgfnodeconncurve{Node1}{Node6}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node2}{Node4}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node2}{Node5}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node2}{Node6}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node3}{Node5}{0}{0}{.35cm}{.35cm} \pgfnodeconncurve{Node3}{Node6}{0}{0}{.55cm}{.55cm} \pgfnodeconncurve{Node4}{Node6}{0}{0}{.3cm}{.3cm} \pgfputat{\pgfxy(.7,.5)}{\pgfbox[center,center]{$0$}} \pgfputat{\pgfxy(.7,1.3)}{\pgfbox[center,center]{$1$}} \pgfputat{\pgfxy(.7,2.1)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(.7,2.9)}{\pgfbox[center,center]{$3$}} \pgfputat{\pgfxy(.7,3.7)}{\pgfbox[center,center]{$4$}} \pgfputat{\pgfxy(.7,4.5)}{\pgfbox[center,center]{$5$}} \pgfnodecircle{Node1}[fill]{\pgfxy(3.5,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3.5,1.3)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(3.5,2.1)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(3.5,2.9)}{0.085cm} \pgfnodecircle{Node5}[fill]{\pgfxy(3.5,3.7)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(3.5,4.5)}{0.075cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3.5,4.7)}{0.025cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.5,4.8)}{0.025cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.5,4.9)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Transitivity obscures the advantages of local finiteness; b) $\mathbb{N}$ is locally infinite, but locally finite in the skeleton; c) local finiteness in the skeleton is a better condition than interval finiteness in many cases.} \label{interactiontransitivity} \end{figure} \vspace*{-.5cm} Even in the transitive case, however, better local finiteness conditions than interval finiteness are available. For example, restricting attention to transitive directed sets recoverable from their skeleta,\footnotemark\footnotetext{This condition rules out objects like $\mathbb{Q}$, which is locally finite in the skeleton because of the interpolative property. It also rules out cycles, but may be easily amended to permit them.} local finiteness in the skeleton (\hyperref[lfs]{LFS}) is a better condition than interval finiteness, since it is causally local, and since it rules out pathologies such as the infinite bouquet. Figure \hyperref[interactiontransitivity]{\ref{interactiontransitivity}}c above illustrates the set $D_{\tn{max}}^-(x)$ of maximal predecessors and the set $D_{\tn{min}}^+(x)$ of minimal successors of an element $x$ in a directed set $D$. The cardinalities of such sets, as $x$ ranges over $D$, determine whether or not $D$ is locally finite in the skeleton. The most obvious way to ensure that a directed set is recoverable from its skeleton is simply to impose the condition of {\it irreducibility;} i.e., to {\it begin} with ``skeletal objects." In particular, this approach makes information-theoretic sense for causal sets, since reducible relations are already regarded as redundant in this context. However, imposing irreducibility reduces the condition of local finiteness in the skeleton to local finiteness. Further, the independence convention (\hyperref[ic]{IC}) is automatic in the irreducible directed case. Hence, one recovers the same basic conditions favored in the nontransitive case, only applied to a smaller class of objects; namely, the class of irreducible locally finite directed sets. Despite these restrictions, this class is {\it still} much different, and vastly richer in general, than the class of causal sets. \newpage \refstepcounter{textlabels}\label{interactionmeasure} {\bf Interaction with the Measure Axiom.} The clear picture of local structure provided by local finiteness (\hyperref[lf]{LF}) and the star topology suggests reconsideration of the measure axiom (\hyperref[m]{M}) in the context of classical spacetime structure. Following this axiom, the volume measure $\mu$ on a causal set $C$ assigns ``approximately" one fundamental unit of spacetime volume to each element $x$ of $C$, {\it regardless of the details of local causal structure near $x$.} While it would certainly be simplest, and most attractive, if ``order plus number" really {\it did} ``equal geometry," an {\it exact relationship} of this nature already seems to be in tension with known bounds on Lorentz invariance violation. As discussed in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} above, this is the reason for the caveat {\it ``up to Poisson-type fluctuations,"} appearing in the later causal set literature. Sharpening the picture by dropping transitivity and adopting local finiteness raises the possibility that {\it relations,} as well as elements, might play some role in the emergent notion of volume. Any particular proposal along these lines would require solid motivation, to avoid the conceptual trap of postulating increasing arbitrary and complex structures to ``rescue" a preconceived result. However, it does seem reasonable, and even natural, to consider at least {\it local} causal structure in the interpretation of volume. \subsection{Relative Multdirected Sets over a Fixed Base}\label{relativeacyclicdirected} \refstepcounter{textlabels}\label{infintmeaning} {\bf Infinite Intervals.} What, if any, significance should be attached to the infinite intervals appearing in locally finite directed sets such as Jacob's ladder, and more generally, in multidirected sets? Should such intervals be viewed as large-scale pathologies that one would prefer to avoid in a physical theory, or should they be taken seriously as structures of possible physical import? In the context of classical spacetime structure, I know of no {\it positive physical evidence} for the latter view, but there seem to be good structural grounds for considering it. The theory of {\it relative multidirected sets over a fixed base,} introduced in this section, provides an excellent perspective on why censoring infinite intervals seems arbitrary and unjustified. The terminology comes from Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}), introduced in section \hyperref[settheoretic]{\ref{settheoretic}} above. \refstepcounter{textlabels}\label{grothrelview} \refstepcounter{textlabels}\label{causetrelz} {\bf Causal Sets as Relative Directed Sets over the Integers.} Every nonempty causal set $C=(C,\prec)$ admits a bijective morphism into a linear suborder of the integers, either $\mathbb{Z}$ itself, the nonnegative integers $\mathbb{N}$, the nonpositive integers $-\mathbb{N}$, or a {\it finite simplex} $[n]:=\{0,1,...,n\}$ for some nonnegative integer $n$. Since $\mathbb{N}$, $-\mathbb{N}$, and $[n]$ all embed into $\mathbb{Z}$, every causal set may be presented as a {\it relative directed set over} $\mathbb{Z}$; i.e., a directed set together with a distinguished morphism into $\mathbb{Z}$. Generally, there are many such morphisms for a given causal set. Such a morphism is equivalent to an {\it extension of the partial order $\prec$ on} $C$ to an order isomorphic to that of its linear image. The existence of such an extension does {\it not} follow immediately from the order extension principle (\hyperref[oep]{OEP}), since a linear extension of an interval finite partial order is generally not interval finite. However, such an extension may be explicitly constructed. The proof I give here is essentially due to Joel David Hamkins \cite{Hamkins13}, though expressed in different language. It uses a special case of a technique I refer to as {\it atomic accretion,} in which an order-theoretic analogue of a convex set, called a {\it causal atom} in the context of classical spacetime structure, is built up by a sequential process. I revisit the theory of causal atoms in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} below. It seems that this simple extension result was first recognized in the context of pure order theory, in the work of Marcel Ern\'{e} circa 1979.\footnotemark\footnotetext{The earliest published references I can find are secondary remarks citing the unpublished lecture notes \cite{Erne79}, which Ern\'{e} was kind enough to send me copies of.}\\ \refstepcounter{textlabels} \begin{theorem}\label{theoremcausalsetrelint} Any nonempty, countable, interval finite, acyclic directed set admits a morphism into the integers. \end{theorem} \begin{proof} The proof has two steps. The first step is an induction argument establishing that any nonempty, countable, interval finite, acyclic directed set $A=(A,\prec)$ admits a {\it sequential atomic accretion} $\{x_i\}_{i\in I}$, where $I$ is an interval in $\mathbb{Z}$ containing $0$. This means that every finite, two-sided truncation $\{x_m,...,x_n\}$ of $\{x_i\}_{i\in I}$ defines a {\it causal atom} in $A$; i.e., a subset of $A$ containing the intersection of its past and future. The second step is merely the observation that such an atomic accretion defines a morphism into $\mathbb{Z}$. \begin{enumerate} \item {\it $A$ admits a sequential atomic accretion.} First note that $A$ admits an {\it enumeration} $A=\{y_0,y_1,...\}$ by hypothesis, since $A$ is countable. A sequential atomic accretion $\{x_i\}_{i\in I}$ of $A$ may be constructed from this enumeration by induction. By acyclicity, any singleton in $A$ is a causal atom, so set $x_0=y_0$. If $A$ has only one element, the accretion process terminates at this stage, and the result is obvious. Otherwise, one must proceed to consider the succeeding elements $y_1,y_2,...$ in the enumeration. I explicitly describe the step involving $y_1$ for illustrative purposes. If $x_0$ precedes $y_1$ in $(A,\prec)$, then $\llangle x_0,y_1\rrangle$ is a finite interval in $(A,\prec)$ by interval finiteness. If this interval is empty, set $x_1=y_1$. If not, choose $x_1$ to be any minimal element in $\llangle x_0,y_1\rrangle$. In either case, the set $\{x_0,x_1\}$ is a causal atom in $A$. If the interval $\llangle x_0,y_1\rrangle$ contains no other elements, set $x_2=y_1$. Otherwise, choose $x_2$ to be a minimal element in $\llangle x_0,y_1\rrangle-\{x_1\}$. In either case, the set $\{x_0,x_1,x_2\}$ is a causal atom in $A$. Continuing in this fashion, the interval $\llangle x_0,y_1\rrangle$ is exhausted after a finite number of steps. A symmetric argument applies if $x_0$ succeeds $y_1$; in this case, the elements in the interval $\llangle y_1,x_0\rrangle$ are chosen to be $x_{-1}$, $x_{-2}$, and so on. If $y_1$ is unrelated to $x_0$, then $y_1$ may be chosen to be either $x_1$ or $x_{-1}$, since in this case the intersection of the past and future of $\{x_0,y_1\}$ is empty. All these cases lead to an accretion process that terminates after a finite number of steps, yielding a causal atom containing $y_0$ and $y_1$. Now assume, by induction, that $\alpha:=\{x_m,...,x_n\}$ is a causal atom containing $y_0,...,y_k$, and consider $y_{k+1}$. If $\alpha=A$, or if $y_{k+1}\in\alpha$, there is nothing to show. Suppose that $y_{k+1}$ belongs to the future of $\alpha$, but not to $\alpha$ itself. Since $\alpha$ is a causal atom, $y_{k+1}$ does not belong to the past of $\alpha$. Consider the generalized open interval $\llangle \alpha, y_{k+1}\rrangle$ consisting of elements in the future of $\alpha$ and the past of $y_{k+1}$, but not belonging to $\alpha$ or equal to $y_{k+1}$. If $\llangle \alpha, y_{k+1}\rrangle$ is empty, set $x_{n+1}=y_{k+1}$. Otherwise, choose $x_{n+1}$ to be a minimal element in $\llangle \alpha,y_1\rrangle$. In either case, $\alpha\cup\{x_{n+1}\}$ is a causal atom. Continuing in this fashion, the interval $\llangle \alpha,y_1\rrangle$ is exhausted after a finite number of steps. A symmetric argument applies if $y_{k+1}$ precedes $\alpha$. If $y_1$ is unrelated to any element of $\alpha$, then $y_1$ may be chosen to be either $x_{n+1}$ or $x_{m-1}$. All these cases lead to an accretion process that terminates after a finite number of steps, yielding a causal atom containing $y_0,...,y_{k+1}$. By induction, the accretion extends to all of $A$. \item {\it The set map $A\rightarrow\mathbb{Z}$ sending $x_i$ to $i$ is a morphism.} At each step in the accretion process, the image $i\in\mathbb{Z}$ of the ``next element" $x_i$ either precedes or succeeds all previously-defined images. The corresponding elements form a causal atom, so $x_i$ cannot precede some of these elements while succeeding others. Hence, the accretion process automatically respects the binary relation $\prec$ on $A$. \end{enumerate} \end{proof} Every causal set is a nonempty, countable, interval finite, acyclic directed set, and therefore admits a morphism into $\mathbb{Z}$ by the theorem. In fact, the proof of the theorem yields more: if $A$ is finite, with cardinality $n+1$, the construction described in the proof yields a family of bijective morphisms into $[n]$, determined by the choice of enumeration and the choices of maximal or minimal elements in generalized open intervals. If $A$ is infinite, bounded below,\footnotemark\footnotetext{There are many different notions of boundedness for acyclic directed sets, both local and global. In this case, {\it bounded below} means that every chain in $A$ has a minimal element. This is a relatively ``weak" boundedness condition.} and has a finite number of minimal elements, then the construction yields, after appropriate ``shifts," a family of bijective morphisms into $\mathbb{N}$. If $A$ is infinite, bounded above, and has a finite number of maximal elements, then the construction yields a family of bijective morphisms into $-\mathbb{N}$. If $A$ is infinite and unbounded above and below, the construction yields a family of bijective morphisms into $\mathbb{Z}$. The remaining cases may involve bijective morphisms into either $\mathbb{N},-\mathbb{N},$ or $\mathbb{Z}$, depending on the details. For example, a countably infinite set with no relations admits bijective morphisms into all three of these targets. \refstepcounter{textlabels}\label{relconseq} A directed set $(D,\prec)$, together with a {\it particular} morphism into $\mathbb{Z}$, bijective or otherwise, is called a {\bf relative directed set over} $\mathbb{Z}$. Hence, every causal set may be viewed as a relative directed set over $\mathbb{Z}$, generally in many different ways. Conversely, every transitive relative directed set $D$ of finite index\footnotemark\footnotetext{Here, {\it index} refers to the index of the morphism $D\rightarrow\mathbb{Z}$; i.e., the supremum of the cardinalities of its fibers.} over $\mathbb{Z}$ is a causal set, since every interval in $D$ is contained in the preimage of an interval in $\mathbb{Z}$, and since $(\mathbb{Z},<)$ itself is irreflexive and interval finite. The axioms of causal set theory may therefore be reduced from six to four, with countability (\hyperref[c]{C}), interval finiteness (\hyperref[if]{IF}), and irreflexivity (\hyperref[ir]{IR}) replaced by the existence of a morphism of finite index into $\mathbb{Z}$. This axiomatic streamlining is one advantage of the theory of relative multidirected sets over a fixed base. A more concrete application appears in Sorkin and Rideout's theory of sequential growth dynamics \cite{SorkinSequentialGrowthDynamics99}, in which a total labeling $\{x_0,x_1,x_2,...\}$ of a causal set $C$ exhibits $C$ as a relative directed set over $\mathbb{Z}$, via the inclusion $\mathbb{N}\rightarrow\mathbb{Z}$. \refstepcounter{textlabels}\label{relacdirarb} {\bf Relative Multidirected Sets over an Arbitrary Base.} More generally, let $M''=(M'',R'',i'',t'')$ be a fixed multidirected set. A {\bf relative multidirected set over} $M''$ is a multidirected set $M=(M,R,i,t)$, together with a morphism $M\rightarrow M''$. The {\bf category of relative multidirected sets over} $M''$ is the category whose objects are multidirected sets over $M''$, and whose morphisms are commutative triangles of the form \begin{pgfpicture}{0cm}{0cm}{15cm}{2cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(8.5, .25)}{\pgfbox[center,center]{$M''$}} \pgfputat{\pgfxy(7.5, 1.75)}{\pgfbox[center,center]{$M$}} \pgfputat{\pgfxy(9.5, 1.75)}{\pgfbox[center,center]{$M'$}} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(7.9,1.75)(9.1,1.75) \pgfxyline(7.6,1.4)(8.2,.5) \pgfxyline(9.4,1.4)(8.8,.5) \end{pgfmagnify}\end{pgfpicture} where the vertical arrows are the distinguished morphisms defining $M$ and $M'$ as multidirected sets over $M''$. In this context, $M''$ is called the {\bf base}, while $M$ and $M'$ are called {\bf sources}. Of particular interest in the study of classical spacetime structure is the case in which $M$, $M'$ and $M''$ are three directed sets $(D,\prec)$, $(D',\prec')$, and $(D'',\prec'')$, respectively, where the underlying sets coincide, where the morphisms are the identity set maps, and where the binary relations $\prec'$ and $\prec''$ are successive extensions of $\prec$. In this case, the morphisms represent {\it generalized frames of reference,} as mentioned in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above. \refstepcounter{textlabels}\label{democracy} {\bf Democracy of Bases.} Consider now the following question, which bears directly on the plausibility of the axiom of interval finiteness (\hyperref[if]{IF}) in discrete causal theory: \hspace*{.3cm}{\bf Question: }{\it Why should the causal structure of classical spacetime be modeled exclusively in terms \\\hspace*{2.25cm} of relative directed sets over $\mathbb{Z}$?} Even causal sets, and in particular, connected\footnotemark\footnotetext{Connected, that is, in a graph-theoretic sense. The point of highlighting this property is that the existence of bijective morphisms into ``strange" bases for sufficiently ``sparse" causal sets, such as a countable set with no relations, is pedestrian.} causal sets, generally admit bijective morphisms into linearly ordered bases {\it not contained in the integers,} and it is difficult to argue that such bases should be ignored. For example, consider the causal set $C=(C,\prec)$ whose skeleton is depicted in figure \hyperref[alternativeextensions]{\ref{alternativeextensions}}a below. $C$ consists of two infinite chains, one bounded below, with minimal element $x$; the other bounded above, with maximal element $y$; these chains are connected by a relation $x\prec y$. Approaching $C$ from the viewpoint of sequential growth, one wishes to provide a ``complete kinematic account of the evolution of $C$;" i.e., to exhibit $C$ bijectively as a relative directed set over a linearly ordered base. The addition of relations $v\prec x$ and $y\prec u$, as shown in figure \hyperref[alternativeextensions]{\ref{alternativeextensions}}b, uniquely determines a linear order on $C$, since every pair of elements is now connected by a chain. This linear order defines a bijective morphism into the integers. However, an alternative linear order on $C$, extending $\prec$, may be defined by taking the transitive closure of the acyclic directed set illustrated in figure \hyperref[alternativeextensions]{\ref{alternativeextensions}}c, which I introduced as {\it Jacob's ladder} in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} above. This linear order defines a bijective morphism from $C$ into a linearly ordered acyclic directed set, depicted in figure \hyperref[alternativeextensions]{\ref{alternativeextensions}}d, which is {\it not interval finite.} Rather, it is isomorphic to $\mathbb{N}\coprod-\mathbb{N}$, where every element of $\mathbb{N}$ is taken to precede every element of $-\mathbb{N}$. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5) \pgfxyline(4.125,-.1)(4.125,5) \pgfxyline(8.25,-.1)(8.25,5) \pgfxyline(12.375,-.1)(12.375,5) \pgfxyline(0,5)(16.5,5) \pgfxyline(16.5,-.1)(16.5,5) \pgfputat{\pgfxy(.15, 4.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(4.275, 4.7)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(8.4, 4.7)}{\pgfbox[left,center]{c)}} \pgfputat{\pgfxy(12.525, 4.7)}{\pgfbox[left,center]{d)}} \begin{pgftranslate}{\pgfpoint{0cm}{.2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.3,2)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1.3,2.4)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.3,2.8)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.3,3.2)}{0.085cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.3,3.6)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1.3,4)}{0.075cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2.8,2.4)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(2.8,2)}{0.095cm} \pgfnodecircle{Node23}[fill]{\pgfxy(2.8,1.6)}{0.09cm} \pgfnodecircle{Node24}[fill]{\pgfxy(2.8,1.2)}{0.085cm} \pgfnodecircle{Node25}[fill]{\pgfxy(2.8,.8)}{0.08cm} \pgfnodecircle{Node26}[fill]{\pgfxy(2.8,.4)}{0.075cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.2)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.3)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.4)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,.2)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,.1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,0)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node1}{Node21} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node23}{Node24} \pgfnodeconnline{Node24}{Node25} \pgfnodeconnline{Node25}{Node26} \pgfputat{\pgfxy(1,2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(3.1,2.4)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(1,2.4)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.1,2)}{\pgfbox[center,center]{$v$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.125cm}{.2cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.3,2)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1.3,2.9)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.3,3.3)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.3,3.6)}{0.085cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.3,3.82)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1.3,4)}{0.075cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2.8,2.4)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(2.8,1.5)}{0.095cm} \pgfnodecircle{Node23}[fill]{\pgfxy(2.8,1.1)}{0.09cm} \pgfnodecircle{Node24}[fill]{\pgfxy(2.8,.8)}{0.085cm} \pgfnodecircle{Node25}[fill]{\pgfxy(2.8,.58)}{0.08cm} \pgfnodecircle{Node26}[fill]{\pgfxy(2.8,.4)}{0.075cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.2)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.3)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1.3,4.4)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,.2)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,.1)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(2.8,0)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node1}{Node21} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node23}{Node24} \pgfnodeconnline{Node24}{Node25} \pgfnodeconnline{Node25}{Node26} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfnodeconnline{Node22}{Node1} \pgfnodeconnline{Node21}{Node2} \pgfputat{\pgfxy(1,2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(3.1,2.4)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(1,2.9)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.1,1.5)}{\pgfbox[center,center]{$v$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.8cm}{-.7cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(0,1.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.7,1.8)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.35,2.05)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.95,2.25)}{0.085cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,2.42)}{0.08cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.95,2.55)}{0.075cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodecircle{Node21}[fill]{\pgfxy(0,4.5)}{0.1cm} \pgfnodecircle{Node22}[fill]{\pgfxy(.7,4.2)}{0.095cm} \pgfnodecircle{Node23}[fill]{\pgfxy(1.35,3.95)}{0.09cm} \pgfnodecircle{Node24}[fill]{\pgfxy(1.95,3.75)}{0.085cm} \pgfnodecircle{Node25}[fill]{\pgfxy(2.5,3.58)}{0.08cm} \pgfnodecircle{Node26}[fill]{\pgfxy(2.95,3.45)}{0.075cm} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node23}{Node24} \pgfnodeconnline{Node24}{Node25} \pgfnodeconnline{Node25}{Node26} \pgfnodeconnline{Node1}{Node21} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfnodeconnline{Node2}{Node22} \pgfnodeconnline{Node3}{Node23} \pgfnodeconnline{Node4}{Node24} \pgfnodeconnline{Node5}{Node25} \pgfnodeconnline{Node6}{Node26} \pgfnodecircle{Node21}[fill]{\pgfxy(3.1,3)}{0.025cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.2,3)}{0.025cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,3)}{0.025cm} \pgfputat{\pgfxy(0,1.2)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(0,4.8)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(.7,1.5)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(.7,4.5)}{\pgfbox[center,center]{$v$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{13cm}{-.4cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.7)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1,1.1)}{0.095cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1,1.5)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1,1.85)}{0.085cm} \pgfnodecircle{Node50}[fill]{\pgfxy(1,2.2)}{0.08cm} \pgfnodecircle{Node60}[fill]{\pgfxy(1,2.5)}{0.075cm} \pgfnodecircle{Node61}[fill]{\pgfxy(1,3.2)}{0.075cm} \pgfnodecircle{Node51}[fill]{\pgfxy(1,3.5)}{0.08cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1,3.85)}{0.085cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,4.2)}{0.09cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1,4.6)}{0.095cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1,5)}{0.1cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1,2.75)}{0.02cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1,2.85)}{0.02cm} \pgfnodecircle{Node0}[fill]{\pgfxy(1,2.95)}{0.02cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node4}{Node50} \pgfnodeconnline{Node50}{Node60} \pgfnodeconnline{Node5}{Node51} \pgfnodeconnline{Node51}{Node61} \pgfnodeconncurve{Node1}{Node8}{0}{0}{1.5cm}{1.5cm} \pgfnodeconncurve{Node1}{Node7}{0}{0}{1.4cm}{1.4cm} \pgfnodeconncurve{Node2}{Node8}{0}{0}{1.4cm}{1.4cm} \pgfnodeconncurve{Node1}{Node6}{0}{0}{1.3cm}{1.3cm} \pgfnodeconncurve{Node2}{Node7}{0}{0}{1.3cm}{1.3cm} \pgfnodeconncurve{Node3}{Node8}{0}{0}{1.3cm}{1.3cm} \pgfnodeconncurve{Node1}{Node5}{0}{0}{1.2cm}{1.2cm} \pgfnodeconncurve{Node2}{Node6}{0}{0}{1.2cm}{1.2cm} \pgfnodeconncurve{Node3}{Node7}{0}{0}{1.2cm}{1.2cm} \pgfnodeconncurve{Node4}{Node8}{0}{0}{1.2cm}{1.2cm} \pgfnodeconncurve{Node1}{Node51}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node2}{Node5}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node3}{Node6}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node4}{Node7}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node50}{Node8}{0}{0}{1.1cm}{1.1cm} \pgfnodeconncurve{Node1}{Node61}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node2}{Node51}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node3}{Node5}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node4}{Node6}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node50}{Node7}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node60}{Node8}{0}{0}{1cm}{1cm} \pgfnodeconncurve{Node1}{Node60}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node2}{Node61}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node3}{Node51}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node4}{Node5}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node50}{Node6}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node60}{Node7}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node61}{Node8}{0}{0}{.8cm}{.8cm} \pgfnodeconncurve{Node1}{Node50}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node2}{Node60}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node3}{Node61}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node4}{Node51}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node50}{Node5}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node60}{Node6}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node61}{Node7}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node51}{Node8}{0}{0}{.6cm}{.6cm} \pgfnodeconncurve{Node1}{Node4}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node2}{Node50}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node3}{Node60}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node4}{Node61}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node50}{Node51}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node60}{Node5}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node61}{Node6}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node51}{Node7}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node5}{Node8}{0}{0}{.4cm}{.4cm} \pgfnodeconncurve{Node1}{Node3}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node2}{Node4}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node3}{Node50}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node4}{Node60}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node50}{Node61}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node60}{Node51}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node61}{Node5}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node51}{Node6}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node5}{Node7}{0}{0}{.2cm}{.2cm} \pgfnodeconncurve{Node6}{Node8}{0}{0}{.2cm}{.2cm} \pgfputat{\pgfxy(.75,.7)}{\pgfbox[center,center]{$0$}} \pgfputat{\pgfxy(.75,1.1)}{\pgfbox[center,center]{$1$}} \pgfputat{\pgfxy(.75,5)}{\pgfbox[center,center]{$0$}} \pgfputat{\pgfxy(.7,4.6)}{\pgfbox[center,center]{$-1$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Skeleton of a causal set $(C,\prec)$; b) adding relations $v\prec x$ and $y\prec u$, and taking the transitive closure, defines a bijective morphism $C\rightarrow\mathbb{Z}$; c) transitive closure of Jacob's ladder gives an alternative extension, which is not interval finite; d) this extension defines a bijective morphism $C\rightarrow\mathbb{N}\coprod-\mathbb{N}$.} \label{alternativeextensions} \end{figure} \vspace*{-.5cm} Both of the linearly ordered sets $\mathbb{Z}$ and $\mathbb{N}\coprod-\mathbb{N}$ are examples of {\it discrete linearly ordered sets,} already mentioned in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}} above, in the context of the partial order formulation of causal set theory. To review, a discrete linearly ordered set is a linearly ordered set in which every nonextremal element has a maximal predecessor and minimal successor. Hence, except at extremal elements, every discrete linearly ordered set is {\it locally isomorphic} to $\mathbb{Z}$, in either the interval topology, or the irreducible star topology. Discrete linearly ordered sets may be viewed as ``one-dimensional manifolds with boundary over the integers." The fact that discrete linearly ordered sets are defined in terms of {\it local properties} makes them particularly attractive as bases for relative directed sets in the context of classical spacetime structure, since it is better not to {\it prescribe global structure.} The relationship between $\mathbb{Z}$ and an arbitrary discrete linearly ordered set may be viewed as analogous to the relationship between $\mathbb{R}^4$, in Newtonian mechanics, and pseudo-Riemannian manifolds, in general relativity, which are only {\it locally} isomorphic to $\mathbb{R}^4$. \refstepcounter{textlabels}\label{infkleitroth} {\bf Countable Analogue of the Kleitman-Rothschild Pathology.} Order theory provides notions of ``size" more nuanced than cardinality; namely, the isomorphism classes of linearly ordered sets. In this sense, Cantor's first transfinite ordinal $\omega$ is ``larger" than $\mathbb{N}$, and the discrete linearly ordered set $\mathbb{N}\coprod-\mathbb{N}$ is ``larger" still. Restrictions on ``minimal linearly ordered bases" of relative acyclic directed sets, imposed by finiteness conditions such as interval finiteness (\hyperref[if]{IF}) or local finiteness (\hyperref[lf]{LF}), may be viewed as restrictions on {\it temporal size} in this refined sense. For example, theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} above demonstrates that interval finiteness limits acyclic directed sets to ``temporal size at most $\mathbb{Z}$," while local finiteness permits ``much larger temporal sizes." This is another way of expressing how severely interval finiteness restricts the global structure of causal sets. As mentioned in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}} above, large finite causal sets are known to be biased toward large spatial and small temporal size, with the three-generation Kleitman-Rothschild orders dominating asymptotically. Restriction to temporal size at most $\mathbb{Z}$ creates an analogous bias in the case of countably infinite interval finite acyclic directed sets, since ``spatial size is unrestricted" in this case, except by countability. Figure \hyperref[semiordinalspatial]{\ref{semiordinalspatial}} below illustrates the skeleton of a causal set of ``spatial size greater than $\mathbb{Z}$." A better justification of this vague statement may be given in terms of the theory of {\it relation space,} introduced in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below. In particular, the three connected subgraphs indicated by the braces in the figure may be viewed as ``antichains of size $\mathbb{Z}$ in relation space." \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{3.1cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,3.1)(16.5,3.1) \pgfxyline(0,-.1)(0,3.1) \pgfxyline(16.5,-.1)(16.5,3.1) \begin{pgftranslate}{\pgfpoint{.4cm}{0cm}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.3,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1.6,1.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.9,1)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(2.2,1.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,1)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.8,1.5)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3.1,1)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.4,1.5)}{0.10cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.7,1)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,1.5)}{0.10cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.3,1)}{0.10cm} \pgfnodecircle{Node00}[fill]{\pgfxy(.9,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1.1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.5,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.6,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.7,1.25)}{0.03cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node9} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node10}{Node11} \pgfmoveto{\pgfxy(.8,.8)} \pgfcurveto{\pgfxy(.7,.5)}{\pgfxy(2.8,.9)}{\pgfxy(2.8,.5)} \pgfcurveto{\pgfxy(2.8,.9)}{\pgfxy(4.9,.5)}{\pgfxy(4.8,.8)} \pgfstroke \pgfputat{\pgfxy(2.8,.2)}{\pgfbox[center,center]{$\mathbb{Z}$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5cm}{0cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.3,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1.6,1.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.9,1)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(2.2,1.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,1)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.8,1.5)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3.1,1)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.4,1.5)}{0.10cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.7,1)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,1.5)}{0.10cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.3,1)}{0.10cm} \pgfnodecircle{Node00}[fill]{\pgfxy(.9,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1.1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.5,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.6,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.7,1.25)}{0.03cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node9} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node10}{Node11} \pgfmoveto{\pgfxy(.8,.8)} \pgfcurveto{\pgfxy(.7,.5)}{\pgfxy(2.8,.9)}{\pgfxy(2.8,.5)} \pgfcurveto{\pgfxy(2.8,.9)}{\pgfxy(4.9,.5)}{\pgfxy(4.8,.8)} \pgfstroke \pgfputat{\pgfxy(2.8,.2)}{\pgfbox[center,center]{$\mathbb{Z}$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10cm}{0cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.3,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(1.6,1.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.9,1)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(2.2,1.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,1)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2.8,1.5)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(3.1,1)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.4,1.5)}{0.10cm} \pgfnodecircle{Node9}[fill]{\pgfxy(3.7,1)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(4,1.5)}{0.10cm} \pgfnodecircle{Node11}[fill]{\pgfxy(4.3,1)}{0.10cm} \pgfnodecircle{Node00}[fill]{\pgfxy(.9,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(1.1,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.5,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.6,1.25)}{0.03cm} \pgfnodecircle{Node00}[fill]{\pgfxy(4.7,1.25)}{0.03cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node9} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node10}{Node11} \pgfmoveto{\pgfxy(.8,.8)} \pgfcurveto{\pgfxy(.7,.5)}{\pgfxy(2.8,.9)}{\pgfxy(2.8,.5)} \pgfcurveto{\pgfxy(2.8,.9)}{\pgfxy(4.9,.5)}{\pgfxy(4.8,.8)} \pgfstroke \pgfputat{\pgfxy(2.8,.2)}{\pgfbox[center,center]{$\mathbb{Z}$}} \end{pgftranslate} \pgfnodecircle{Node100}[fill]{\pgfxy(5.3,2.5)}{0.10cm} \pgfnodecircle{Node100}[fill]{\pgfxy(10.3,2.5)}{0.10cm} \pgfxyline(2.8,1.5)(5.3,2.5) \pgfxyline(5.3,2.5)(7.8,1.5) \pgfxyline(7.8,1.5)(10.3,2.5) \pgfxyline(10.3,2.5)(12.8,1.5) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{Skeleton of a causal set of ``spatial size greater than $\mathbb{Z}$."} \label{semiordinalspatial} \end{figure} \vspace*{-.5cm} \subsection{Eight Arguments against Interval Finiteness and Similar Conditions}\label{subsectionintervalfinitenessdeficient} I have now assembled the necessary background to fully expose the shortcomings of interval finiteness (\hyperref[if]{IF}) as a local finiteness axiom for directed sets, viewed as models of classical spacetime structure under Sorkin's version of the causal metric hypothesis (\hyperref[cmh]{CMH}). Some of these shortcomings have already been mentioned in previous sections. The following arguments also show that interval finiteness cannot be repaired by working only with arbitrarily small open intervals, or intersections of open intervals. In particular, topological local finiteness for the interval topology is also an unsuitable axiom. The examples appearing below involve only acyclic directed sets. This makes it clear that the arguments apply even under conservative assumptions. \refstepcounter{textlabels}\label{intfinnonloc} {\bf 1.} {\bf Interval finiteness is not a causally local condition in the context of classical spacetime structure.} Referring to definition \hyperref[definitioncausallocality]{\ref{definitioncausallocality}} above, this means that interval finiteness generally cannot be checked by examining the direct causes and effects associated with each element in a directed set. For example, consider the element $x$ in the acyclic directed set illustrated in figure \hyperref[intervaldeficiency]{\ref{intervaldeficiency}}a below. The only open interval containing $x$ is the interval $\llangle w,y\rrangle$, which also contains all of the directed structure highlighted by the grey circle, which is causally unrelated to $x$. This is also the smallest open set containing $x$ in the interval topology. \refstepcounter{textlabels}\label{nonlocininttop} {\bf 2.} {\bf Interval finiteness is not a local condition for the interval topology.} In particular, interval finiteness is not the same as topological local finiteness for the interval topology, as already mentioned in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} above. This is intuitively obvious, since interval finiteness involves open intervals of arbitrary size. More generally, {\it no} condition stated solely in terms of open intervals, even ``small" open intervals, is local in the interval topology, because open intervals do not form a basis for the topology: a neighborhood containing an element need not contain an {\it open interval} containing the element. For example, consider the element $x$ in the acyclic directed set illustrated in figure \hyperref[intervaldeficiency]{\ref{intervaldeficiency}}b. The only nonempty open intervals in this set are $\llangle w,y\rrangle$ and $\llangle w,z\rrangle$, both of which contain $x$. Their intersection is the singleton $\{x\}$, which is therefore an open set in the interval topology, but {\it not} an open interval. Of course, interval finiteness is even worse in this regard, since it is defined in terms of {\it all} open intervals. \refstepcounter{textlabels}\label{toplocfinintnofinnbhd} {\bf 3.} {\bf Topological local finiteness for the interval topology does not imply that each element is contained in a finite open interval.} This illustrates the fact that interval finiteness is {\it not even close} to having a reasonable topological interpretation as a local finiteness condition. It is already clear from figure \hyperref[intervaldeficiency]{\ref{intervaldeficiency}}b that an element in a multidirected set may have neighborhoods in the interval topology which are smaller than any nonempty open interval. In the worst case scenario, a minimal nonempty open interval containing a given element may be {\it infinite} even though the element has a finite neighborhood in the interval topology! For example, consider the acyclic directed set $A=(A,\prec)$ with elements \[\{x\}\cup\{y_{p,q}\hspace*{.1cm}|\hspace*{.1cm}p,q \tn{ prime, } q\neq p\}\cup\{z_l\hspace*{.1cm}|\hspace*{.1cm}l \tn{ prime}\},\] and relations \[x\prec y_{p,q},\hspace*{.3cm} y_{p,q}\prec z_p,\hspace*{.3cm}\tn{and}\hspace*{.3cm}y_{p,q}\prec z_q.\] A portion of $A$ is illustrated in figure \hyperref[intervaldeficiency]{\ref{intervaldeficiency}}c, with the elements $y_{p,q}$ grouped together for each $p$, and with the relations $y_{p,q}\prec z_q$ indicated by dark segments. The only nonempty open intervals in $A$ are the infinite intervals $\llangle x,z_{l}\rrangle=\{y_{p,q}\hspace*{.1cm}|\hspace*{.1cm} p=l\tn{ or }q=l \}$ for each $l$. The intersection of two such intervals $\llangle x,z_{l}\rrangle$ and $\llangle x,z_{l'}\rrangle$ is the two-point set $\{y_{l,l'},y_{l',l}\}$. Meanwhile, the open set $D^+(y_{p,q})$ in $A$ is the two-point set $\{z_p\}\cup\{z_q\}$, and the open set $D^-(y_{p,q})$ is the singleton $\{x\}$ for all $p,q$. Hence, every element of $A$ is contained in a finite open subset of $A$ in the interval topology, but {\it no element of $A$ lies in a finite open interval.} \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{3.6cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,3.5) \pgfxyline(3.75,-.1)(3.75,3.5) \pgfxyline(7.5,-.1)(7.5,3.5) \pgfxyline(12.75,-.1)(12.75,3.5) \pgfxyline(0,3.5)(16.5,3.5) \pgfxyline(16.5,-.1)(16.5,3.5) \pgfputat{\pgfxy(.15,3.2)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(3.9,3.2)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(7.65,3.2)}{\pgfbox[left,center]{c)}} \pgfputat{\pgfxy(12.9,3.2)}{\pgfbox[left,center]{d)}} \begin{pgftranslate}{\pgfpoint{.5cm}{0cm}} \begin{colormixin}{30!white} \pgfnodecircle{Node10}[fill]{\pgfxy(2,1.5)}{.7cm} \end{colormixin} \pgfputat{\pgfxy(1.5,.2)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(.2,1.5)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.5,2.8)}{\pgfbox[center,center]{$y$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.5,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,1)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(0.5,1.5)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.5,1.5)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,1.5)}{0.1cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2,2)}{0.1cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,2.5)}{0.1cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node2}{Node5} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node4}{Node6} \pgfnodeconnline{Node4}{Node7} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node6}{Node7} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{0cm}} \pgfputat{\pgfxy(1.5,.2)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.2,1.5)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(2,2.8)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(1,2.8)}{\pgfbox[center,center]{$y$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1.5,.5)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,1.5)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.5,1.5)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.5,1.5)}{0.1cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,2.5)}{0.1cm} \pgfnodecircle{Node7}[fill]{\pgfxy(2,2.5)}{0.1cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node4}{Node6} \pgfnodeconnline{Node4}{Node7} \pgfnodeconnline{Node5}{Node7} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.75cm}{-.2cm}} \pgfputat{\pgfxy(1,.3)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(-.35,1.75)}{\pgfbox[center,center]{$y_{2,3}$}} \pgfputat{\pgfxy(.2,3.3)}{\pgfbox[center,center]{$z_2$}} \pgfputat{\pgfxy(1.4,3.3)}{\pgfbox[center,center]{$z_3$}} \pgfputat{\pgfxy(2.6,3.3)}{\pgfbox[center,center]{$z_5$}} \pgfnodecircle{Node0}[fill]{\pgfxy(1,.6)}{0.1cm} \pgfnodecircle{Node00}[fill]{\pgfxy(3.3,1.5)}{0.025cm} \pgfnodecircle{Node00}[fill]{\pgfxy(3.4,1.5)}{0.025cm} \pgfnodecircle{Node00}[fill]{\pgfxy(3.5,1.5)}{0.025cm} \pgfnodecircle{Node23}[fill]{\pgfxy(0,1.5)}{0.1cm} \pgfnodecircle{Node25}[fill]{\pgfxy(.2,1.4)}{0.1cm} \pgfnodecircle{Node27}[fill]{\pgfxy(.4,1.3)}{0.1cm} \pgfnodecircle{Node211}[fill]{\pgfxy(.6,1.2)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.2,3)}{0.1cm} \pgfnodecircle{Node32}[fill]{\pgfxy(1,1.6)}{0.1cm} \pgfnodecircle{Node35}[fill]{\pgfxy(1.4,1.4)}{0.1cm} \pgfnodecircle{Node37}[fill]{\pgfxy(1.6,1.3)}{0.1cm} \pgfnodecircle{Node311}[fill]{\pgfxy(1.8,1.2)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,3)}{0.1cm} \pgfnodecircle{Node52}[fill]{\pgfxy(2.2,1.6)}{0.1cm} \pgfnodecircle{Node53}[fill]{\pgfxy(2.4,1.5)}{0.1cm} \pgfnodecircle{Node57}[fill]{\pgfxy(2.8,1.3)}{0.1cm} \pgfnodecircle{Node511}[fill]{\pgfxy(3,1.2)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.6,3)}{0.1cm} \pgfnodeconnline{Node0}{Node23} \pgfnodeconnline{Node0}{Node25} \pgfnodeconnline{Node0}{Node27} \pgfnodeconnline{Node0}{Node211} \pgfnodeconnline{Node0}{Node32} \pgfnodeconnline{Node0}{Node35} \pgfnodeconnline{Node0}{Node37} \pgfnodeconnline{Node0}{Node311} \pgfnodeconnline{Node0}{Node52} \pgfnodeconnline{Node0}{Node53} \pgfnodeconnline{Node0}{Node57} \pgfnodeconnline{Node0}{Node511} \pgfnodeconnline{Node23}{Node2} \pgfnodeconnline{Node25}{Node2} \pgfnodeconnline{Node27}{Node2} \pgfnodeconnline{Node211}{Node2} \pgfnodeconnline{Node32}{Node3} \pgfnodeconnline{Node35}{Node3} \pgfnodeconnline{Node37}{Node3} \pgfnodeconnline{Node311}{Node3} \pgfnodeconnline{Node52}{Node5} \pgfnodeconnline{Node53}{Node5} \pgfnodeconnline{Node57}{Node5} \pgfnodeconnline{Node511}{Node5} \pgfsetlinewidth{1.5pt} \pgfnodeconnline{Node23}{Node3} \pgfnodeconnline{Node25}{Node5} \pgfnodeconnline{Node32}{Node2} \pgfnodeconnline{Node35}{Node5} \pgfnodeconnline{Node52}{Node2} \pgfnodeconnline{Node53}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{13cm}{0cm}} \pgfputat{\pgfxy(2,.2)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(1.2,1.5)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(2,2.8)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(1,.2)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(1,2.8)}{\pgfbox[center,center]{$v$}} \pgfputat{\pgfxy(1.5,2.8)}{\pgfbox[center,center]{$z$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,.5)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.5,1.5)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(2.5,1.5)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1,2.5)}{0.1cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1.5,2.5)}{0.1cm} \pgfnodecircle{Node7}[fill]{\pgfxy(2,2.5)}{0.1cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node4}{Node7} \pgfsetlinewidth{1.5pt} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) The only open interval containing $x$ also contains irrelevant directed structure; b) the singleton $\{x\}$ is open in the interval topology, but does not contain an open interval; c) every element lies in a finite open set in the interval topology, but not in a finite open interval; d) the open interval $\llangle w,y\rrangle$ contains $x$, but does not capture local directed structure near $x$.} \label{intervaldeficiency} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{intnotcapture} {\bf 4.} {\bf Open intervals fail to even capture local multidirected structure, let alone isolate it.} Heuristically, this means that open intervals are not only {\it imprecise,} but also {\it inaccurate.} Despite being ``too large" to {\it resolve} local multidirected structure, open intervals generally do not even contain adequate information about local multidirected structure near each of their elements. For example, consider the element $x$ lying in the open interval $\llangle w,y \rrangle$ in the acyclic directed set illustrated in figure \hyperref[intervaldeficiency]{\ref{intervaldeficiency}}d above. Three relations involving $x$; namely, $u\prec x$, $x\prec v$, and $x\prec z$, represented in the figure by dark segments, also involve elements outside of $\llangle w,y\rrangle$. Hence, $\llangle w,y \rrangle$ contains relatively little local information relevant to $x$. Open sets in the interval topology are even worse in general, since they are generally smaller than any open interval. \refstepcounter{textlabels}\label{intnotimplyloc} {\bf 5.} {\bf Interval finiteness does not imply local finiteness.} This was already demonstrated in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} above, with the infinite bouquet (figure \hyperref[jacobs]{\ref{jacobs}}a) as a counterexample. Since open intervals are generally ``too large" to resolve local multidirected structure, one might hope that requiring even such ``large" structures to be finite would at least constrain local causal structure, viewed as ``small," to be finite as well. The fact that this is not true illustrates how badly open intervals fail to capture local causal structure. Topological local finiteness in the interval topology is even worse in this regard, since it is a weaker condition. \refstepcounter{textlabels}\label{locnotimplyint} {\bf 6.} {\bf Interval finiteness does not follow from local finiteness.} Again, this was demonstrated in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} above, with Jacob's ladder (figure \hyperref[jacobs]{\ref{jacobs}}c) as a counterexample. This fact is not surprising, since conditions stated in terms of arbitrary open intervals are not local even in the interval topology. However, topological local finiteness in the interval topology does not follow from local finiteness either. To see this, ``augment" Jacob's ladder by replacing the relation $x_0\prec y_0$ with a $2$-chain $x_0\prec z_{0}\prec y_{0}$. Then the only open interval containing $z_0$ is the infinite interval $\llangle x_0,y_0\rrangle$. \refstepcounter{textlabels}\label{fatalsorkin} {\bf 7.} {\bf Interval finiteness permits physically fatal local behavior under Sorkin's version of the causal metric hypothesis.} In particular, it permits instantaneous collapse, or instantaneous expansion, of an infinite volume of spacetime, as explained in section \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} above. Mixing finite and infinite local behavior is dubious even without taking the measure axiom (\hyperref[m]{M}) into account, but is particularly problematic when the measure axiom is assumed to hold. Topological local finiteness in the interval topology is equally problematic in this regard. \refstepcounter{textlabels}\label{unjustglobal} {\bf 8.} {\bf Interval finiteness imposes unjustified restrictions on the global structure of classical spacetime.} This was made precise for causal sets by theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} above, which demonstrates that every causal set may be exhibited as a relative directed set over the integers. Nothing we know about cosmology justifies this restriction. Topological local finiteness in the interval topology does not impose such a restriction, since it is at least a local condition. However, the absence of this particular pathology is not enough to rescue this condition. \subsection{Six Arguments for Local Finiteness}\label{subsectionarglocfin} I now briefly review some of the principal arguments in favor of local finiteness as a replacement for interval finiteness (\hyperref[lf]{LF}). These arguments have all appeared, explicitly or implicitly, in the foregoing analysis. \refstepcounter{textlabels}\label{locfincausloc} {\bf 1.} {\bf Local finiteness is a causally local condition in the context of classical spacetime structure.} Referring to definition \hyperref[definitioncausallocality]{\ref{definitioncausallocality}} above, this means that local finiteness may be checked by examining the independent causes and effects associated with each element in a directed set $D$. These causes and effects are represented by the relation sets $R(x)$ at each element $x$ in $D$, whose cardinality determines local finiteness. \refstepcounter{textlabels}\label{locfintoploc} {\bf 2.} {\bf Local finiteness is a local condition in the topological sense.} More precisely, local finiteness coincides with the condition of topological local finiteness for the star topology on the star model $M_\star$ of a multidirected set $M$, as proven in lemma \hyperref[localfinitenesslemma]{\ref{localfinitenesslemma}} above. \refstepcounter{textlabels}\label{locfincaptures} {\bf 3.} {\bf Local finiteness both captures and isolates local multidirected structure.} This means that consideration of the entire relation set $R(x)$ at each element $x$ in a multidirected set $M$ is both necessary and sufficient to determine local finiteness. This follows immediately from the definition of local finiteness. \refstepcounter{textlabels}\label{compatsorkin} {\bf 4.} {\bf Local finiteness is compatible with Sorkin's version of the causal metric hypothesis.} In particular, local finiteness allows a minimum finite spacetime volume to be assigned to each element in a directed set, without the danger of fatal local behavior. Indeed, since the set $D_0^-(x)\cup D_0^+(x)$ of direct predecessors and successors of any element $x$ in {\it any} locally finite multidirected set $(M,\prec)$ is finite, this set has a finite volume for {\it any} measure $\mu:\ms{P}(M)\rightarrow\mathbb{R}^+$. \refstepcounter{textlabels}\label{notunjustglobal} {\bf 5.} {\bf Local finiteness does not impose unjustified restrictions on the global structure of classical spacetime.} For example, in the acyclic case, local finiteness admits relative directed sets over a wide variety of discrete linearly ordered sets, not only suborders of $\mathbb{Z}$. \refstepcounter{textlabels}\label{naturalnontrans} {\bf 6.} {\bf Local finiteness is a natural condition for nontransitive relations under the independence convention.} Under the independence convention (\hyperref[ic]{IC}), each element $r$ of the relation set $R(x)$ at an element $x$ of a multidirected set $M$ possesses independent information-theoretic significance. In this context, local finiteness is determined precisely by the information ``directly relevant to $x$," as $x$ ranges over $M$. \subsection{Hierarchy of Finiteness Conditions}\label{subsectionhierarchyfiniteness} To conclude this section, I briefly introduce a few more finiteness conditions for multidirected sets, and prove two elementary theorems relating these conditions for acyclic directed sets and multidirected sets, respectively. The corresponding conditions for directed sets may be easily inferred from the second of these two theorems; namely, theorem \hyperref[theoremhierarchyfinitenessmulti]{\ref{theoremhierarchyfinitenessmulti}}. \newpage Let $M=(M,R,i,t)$ be a multidirected set. Consider the following finiteness conditions on $M$: \refstepcounter{textlabels}\label{eltfin} \hspace*{.3cm} EF.\hspace*{.2cm} {\bf Element Finiteness:} {\it $M$ has finite cardinality.}\label{ef} \refstepcounter{textlabels}\label{loceltfin} \hspace*{.3cm} LEF. {\bf Local Element Finiteness:} {\it Every element of $M$ has a finite number of direct predecessors \\\hspace*{1.3cm} and successors.}\label{lef} \refstepcounter{textlabels}\label{relfin} \hspace*{.3cm} RF.\hspace*{.2cm} {\bf Relation Finiteness:} {\it $R$ has finite cardinality; i.e., $M$ has a finite number of \\ \hspace*{1.3cm} relations.}\label{rf} \refstepcounter{textlabels}\label{parrelfin} \hspace*{.3cm} PRF. {\bf Pairwise Relation Finiteness:} {\it Every pair of elements in $M$ has a finite number of \\\hspace*{1.3cm} relations between them.}\label{prf} \refstepcounter{textlabels}\label{chainfin} \hspace*{.3cm} CF.\hspace*{.2cm} {\bf Chain Finiteness:} {\it Every chain in $M$ has finite length.}\label{cf} \refstepcounter{textlabels}\label{antichainfin} \hspace*{.3cm} AF.\hspace*{.2cm} {\bf Antichain Finiteness:} {\it Every antichain in $M$ has finite cardinality.}\label{af} \refstepcounter{textlabels}\label{discussfincond} Local element finiteness coincides with local finiteness for directed sets, and pairwise relation finiteness follows directly from the definition in this case. The remaining four conditions are all too restrictive to serve as axioms for directed sets as models of classical spacetime structure under the causal metric hypothesis (\hyperref[cmh]{CMH}). Element finiteness is too restrictive because it prescribes a finite classical spacetime, and relation finiteness is too restrictive because it prescribes a classical spacetime with finite connected components. Chain finiteness is too restrictive because it prescribes a classical spacetime that is {\it locally temporally finite,} in the sense that every sequence of events has a first cause and final effect. Finally, antichain finiteness is too restrictive because it prescribes a classical spacetime that is spatially finite ``in any given frame of reference." Each of these prescriptions rules out interesting cosmological models. \\ \begin{theorem}\label{theoremhierarchyfiniteness} For acyclic directed sets, the conditions of element finiteness \tn{(\hyperref[ef]{\tn{EF}})}, relation finiteness \tn{(\hyperref[rf]{\tn{RF}})}, local finiteness in the transitive closure \tn{(\hyperref[lft]{\tn{LFT}})}, local finiteness \tn{(\hyperref[lf]{\tn{LF}})}, local finiteness in the skeleton \tn{(\hyperref[lfs]{\tn{LFS}})}, chain finiteness \tn{(\hyperref[cf]{\tn{CF}})}, interval finiteness \tn{(\hyperref[if]{\tn{IF}})}, and antichain finiteness \tn{(\hyperref[af]{\tn{AF}})}, are related as follows, where arrows indicate logical implications: \begin{pgfpicture}{0cm}{0cm}{15cm}{2.3cm} \begin{pgfmagnify}{1.03}{1.03} \begin{pgftranslate}{\pgfpoint{1cm}{-.5cm}} \pgfputat{\pgfxy(4.75,1.5)}{\pgfbox[center,center]{\tn{EF}}} \pgfputat{\pgfxy(5.5,2.5)}{\pgfbox[center,center]{\tn{RF}}} \pgfputat{\pgfxy(6.25,1.5)}{\pgfbox[center,center]{\tn{CF}}} \pgfputat{\pgfxy(7,2.5)}{\pgfbox[center,center]{\tn{LFT}}} \pgfputat{\pgfxy(8.5,2.5)}{\pgfbox[center,center]{\tn{LF}}} \pgfputat{\pgfxy(9.25,1.5)}{\pgfbox[center,center]{\tn{LFS}}} \pgfputat{\pgfxy(7,.5)}{\pgfbox[center,center]{\tn{AF}}} \pgfputat{\pgfxy(7.75,1.5)}{\pgfbox[center,center]{\tn{IF}}} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(4.85,1.8)(5.25,2.2) \pgfxyline(6.9,2.2)(6.5,1.8) \pgfxyline(5.9,2.5)(6.5,2.5) \pgfxyline(7.5,2.5)(8.1,2.5) \pgfxyline(8.6,2.2)(9,1.8) \pgfxyline(7.1,2.2)(7.5,1.8) \pgfxyline(5,1.3)(6.7,.7) \pgfxyline(7.3,.7)(8.9,1.2) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} Moreover, these are the only pairwise logical implications among these conditions. \end{theorem} \begin{proof} Let $A=(A,\prec)$ be an acyclic directed set. I break the proof into two parts, first showing that the claimed logical implications are valid, then showing that no other pairwise logical implications exist among these conditions. Some of the statements proven here are more general than actually necessary to establish the theorem. \begin{enumerate} \item {\it The claimed logical implications are valid.} EF $\Rightarrow$ RF: For {\it any} directed set, relations may be identified with pairs of elements; hence, the number of relations is bounded by the square of the number of elements. RF $\Rightarrow$ LFT: Suppose that $A$ has $n$ relations. The cardinality of the relation set $R(x)$ at any element $x$ in $\tn{tr}(A)$ is bounded by the total number of relations in $\tn{tr}(A)$, which is bounded by the total number of chains in $A$. Since chains may be identified with families of relations in the acyclic case, this number is bounded by the cardinality of the power set of the set of relations of $A$, which is $2^{n}$. LFT $\Rightarrow$ CF: I prove the contrapositive. Suppose that $A$ has an infinite chain $\gamma$, and let $x$ be an element of $A$ lying on $\gamma$. Then $\tn{tr}(A)$ includes a relation between $x$ and every other element of $\gamma$, so the relation set at $x$ in $\tn{tr}(A)$ is infinite. LFT $\Rightarrow$ IF: I prove the contrapositive. Suppose that $A$ is not interval finite, and let $\llangle w,y \rrangle $ be an infinite open interval in $A$. Every element in this interval is a direct successor of $w$ in $\tn{tr}(A)$, so $A$ is not locally finite in the transitive closure. LFT $\Rightarrow$ LF: Every relation in $A$ corresponds to a relation in $\tn{tr}(A)$, so any element with a finite relation set in $\tn{tr}(A)$ also has a finite relation set in $A$. LF $\Rightarrow$ LFS: Every relation in $\tn{sk}(A)$ corresponds to a relation in $A$, so any element with a finite relation set in $A$ also has a finite relation set in $\tn{sk}(A)$. EF $\Rightarrow$ AF: The number of elements in any antichain in {\it any} multidirected set is bounded by the total number of elements. If the latter is finite, then so is the former. AF $\Rightarrow$ LFS: I prove the contrapositive. Suppose that $A$ is not locally finite in the skeleton. Then there exists an element in $A$ with either an infinite number of maximal predecessors or an infinite number of minimal successors, which comprise an infinite antichain in $A$ by the definition of irreducibility. \item {\it No other pairwise logical implications exist among these conditions.} Note that disproving a particular implication also disproves every ``upstream implication." For example, since element finiteness (EF) implies relation finiteness (RF), disproving the implication LFT$\Rightarrow$ RF automatically disproves the implication LFT$\Rightarrow$ EF. RF $\nRightarrow$ AF: A countably infinite set with no relations is a counterexample. LFT $\nRightarrow$ RF, AF: A countably infinite number of disjoint $1$-chains is a counterexample. LF $\nRightarrow$ CF, IF, AF: Augment Jacob's ladder by replacing each relation $x_i\prec y_{-i}$ with a $2$-chain $x_i\prec z_{i}\prec y_{-i}$. The resulting acyclic directed set is a counterexample for all three conditions. LFS $\nRightarrow$ CF, IF, AF, LF: The transitive closure of the acyclic directed set in the previous step is a counterexample for all four conditions. AF $\nRightarrow$ CF, IF, LF: Cantor's first transfinite ordinal is a counterexample for all three conditions. IF $\nRightarrow$ CF, AF, LFS: The disjoint union of the infinite bouquet and $\mathbb{Z}$ is a counterexample for all three conditions. CF $\nRightarrow$ AF, IF, LFS: The {\it mirror bouquet,} with elements $x$, $\{y_i\}_{i\in\mathbb{Z}}$, and $z$, and relations $x\prec y_i$ and $y_i\prec z$ for all $i$, is a counterexample for all three conditions. \end{enumerate} \end{proof} The main purpose of theorem \hyperref[theoremhierarchyfiniteness]{\ref{theoremhierarchyfiniteness}} is to compare finiteness conditions for conservative models of classical spacetime structure in the discrete causal context. Theorem \hyperref[theoremhierarchyfinitenessmulti]{\ref{theoremhierarchyfinitenessmulti}}, meanwhile, is intended to apply primarily to the abstract structure of configuration spaces of directed sets, particularly kinematic schemes, arising in the histories approach to quantum causal theory developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below.\\ \newpage \begin{theorem}\label{theoremhierarchyfinitenessmulti} For multidirected sets, the conditions of element finiteness \tn{(\hyperref[ef]{\tn{EF}})}, local element finiteness \tn{(\hyperref[lef]{\tn{LEF}})}, relation finiteness \tn{(\hyperref[rf]{\tn{RF}})}, pairwise relation finiteness \tn{(\hyperref[prf]{\tn{PRF}})}, local finiteness \tn{(\hyperref[lf]{\tn{LF}})}, chain finiteness \tn{(\hyperref[cf]{\tn{CF}})}, interval finiteness \tn{(\hyperref[if]{\tn{IF}})}, and antichain finiteness \tn{(\hyperref[af]{\tn{AF}})}, are related as follows, where arrows indicate logical implications: \begin{pgfpicture}{0cm}{0cm}{15cm}{2.3cm} \begin{pgfmagnify}{1.03}{1.03} \begin{pgftranslate}{\pgfpoint{2cm}{0cm}} \pgfputat{\pgfxy(4.75,2)}{\pgfbox[center,center]{\tn{CF}}} \pgfputat{\pgfxy(6.25,2)}{\pgfbox[center,center]{\tn{EF}}} \pgfputat{\pgfxy(7.75,2)}{\pgfbox[center,center]{\tn{AF}}} \pgfputat{\pgfxy(4,1)}{\pgfbox[center,center]{\tn{LEF}}} \pgfputat{\pgfxy(5.5,1)}{\pgfbox[center,center]{\tn{LF}}} \pgfputat{\pgfxy(7,1)}{\pgfbox[center,center]{\tn{PRF}}} \pgfputat{\pgfxy(8.5,1)}{\pgfbox[center,center]{\tn{IF}}} \pgfputat{\pgfxy(6.25,0)}{\pgfbox[center,center]{\tn{RF}}} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(6.65,2)(7.3,2) \pgfxyline(5.9,1.8)(4.5,1.2) \pgfxyline(6.6,1.8)(8.15,1.2) \pgfxyline(6.6,.2)(8.15,.8) \pgfxyline(6,.3)(5.7,.7) \pgfxyline(5.2,1)(4.5,1) \pgfxyline(5.8,1)(6.5,1) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} Moreover, these are the only pairwise logical implications among these conditions. \end{theorem} \begin{proof} Let $M=(M,R,i,t)$ be a multidirected set. As in theorem \hyperref[theoremhierarchyfiniteness]{\ref{theoremhierarchyfiniteness}}, I first prove the claimed logical implications, then prove that no other pairwise logical implications exist among these conditions. \begin{enumerate} \item {\it The claimed logical implications are valid.} EF $\Rightarrow$ LEF: The number of direct predecessors and successors of each element in $M$ is bounded by the total number of elements of $M$. If the latter number is finite, then so is the former. EF $\Rightarrow$ AF: The proof of this implication in theorem \hyperref[theoremhierarchyfiniteness]{\ref{theoremhierarchyfiniteness}} is for general multidirected sets. EF $\Rightarrow$ IF: The number of elements in any open interval in $M$ is bounded by the total number of elements of $M$. If the latter number is finite, then so is the former. RF $\Rightarrow$ LF: The number of relations in the relation set $R(x)$ at any element $x$ in $M$ bounded by the cardinality of the relation set $R$ of $M$. If the latter number is finite, then so is the former. RF $\Rightarrow$ IF: I prove the contrapositive. Suppose that $M$ possesses an infinite interval $\llangle w,y\rrangle$, and let $\{x_i\}_{i\in\mathbb{Z}}$ be an infinite subfamily of $\llangle w,y\rrangle$. By the definition of an open interval, there exist chains $\gamma_i$ in $M$ beginning at $w$ and terminating at $x_i$ for each $i$. The terminal relations in these chains are all distinct, since they have distinct terminal elements. These relations form an infinite subset of the relation set $R$ of $M$. LF $\Rightarrow$ LEF: The number of elements directly related to an element $x$ of $M$ is bounded by the cardinality of the relation set $R(x)$ at $x$. If $M$ is locally finite, then the latter number is finite by definition, so the former number is also finite. LF $\Rightarrow$ PRF: The number of relations between two elements $x$ and $y$ in $M$ is bounded by the cardinalities of both the relation sets $R(x)$ and $R(y)$. If either of the latter numbers is finite, then so is the former. \item {\it No other pairwise logical implications exist among these conditions.} All the counterexamples from theorem \hyperref[theoremhierarchyfiniteness]{\ref{theoremhierarchyfiniteness}} remain valid, since multdirected sets are more general than acyclic directed sets. The remaining necessary counterexamples are as follows: EF $\nRightarrow$ CF, PRF: A multidirected set with two elements and a countably infinite number of relations between them in both directions is a counterexample for both conditions. LEF $\nRightarrow$ CF, AF, IF, PRF: The disjoint union of the previous counterexample with the augmented version of Jacob's ladder, in which each relation $x_i\prec y_{-i}$ is replaced with a chain $x_i\prec z_{i}\prec y_{-i}$, is a counterexample for all four conditions. PRF $\nRightarrow$ CF, LEF, AF, IF: The disjoint union of the infinite bouquet and the augmented version of Jacob's ladder from the previous counterexample, is a counterexample for all four conditions. CF $\nRightarrow$ LEF, PRF: Replace each relation in the infinite bouquet with a countably infinite number of relations. The resulting multidirected set is a counterexample for both conditions. AF $\nRightarrow$ LEF, PRF: The disjoint union of $\mathbb{Z}$ with a multidirected set consisting of a pair of elements with a countably infinite number of relations between them is a counterexample for both conditions. IF $\nRightarrow$ LEF, PRF: The previous counterexample suffices. RF $\nRightarrow$ CF: A reflexive relation $x\prec x$ is a counterexample. \end{enumerate} \end{proof} \newpage \section{The Binary Axiom: Events versus Elements}\label{sectionbinary} The motivation for isolating the physical interpretation of causal set elements as {\it spacetime events} in the binary axiom (\hyperref[b]{B}), is to distinguish the causal set viewpoint from interesting alternative viewpoints regarding similar generalized order-theoretic structures, in which elements may be assigned {\it different physical interpretations}. In this section, I describe a variety of such interpretations, and explore some of their applications. This section is not intended to advocate any actual change in the axioms of causal set theory, beyond those changes already suggested in sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}} above, although possible alterations of the binary axiom are briefly outlined in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}}, in the context of {\it classical holism.} Instead, this section provides an expanded structural perspective, introduces new technical methods, and begins the process of exploiting the conclusions of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}}. Mathematically, it has an {\it algebraic} flavor, with multiple appearances of Grothendieck's relative viewpoint (\hyperref[rv]{RV}), and the principle of hidden structure (\hyperref[hs]{HS}). In section \hyperref[subsectionrelation]{\ref{subsectionrelation}} below, I show how {\it relations} in a multidirected set $M=(M,R,i,t)$ may be interpreted as {\it elements} of an induced directed set $\ms{R}(M)=(R,\prec)$, called the {\it relation space} over $M$. The {\it induced (binary) relation} $\prec$ on $\ms{R}(M)$ is defined by setting $r\prec s$ if and only if the terminal element of $r$ coincides with the initial element of $s$ in $M$. Relation spaces are called {\it line digraphs} in the context of graph theory. They possess special information-theoretic properties, such as {\it impermeability of maximal antichains}, proven in theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}}. This property is crucial to the theory of {\it path summation} over a multidirected set; in particular, it implies the information-theoretic adequacy of the {\it causal Schr\"{o}dinger-type equations} derived via path summation in section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}}. The theoretical importance of relation space motivates the consideration of analogous spaces whose elements have ``more complicated internal structures" than relations. In section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}}, I discuss a special class of such spaces, called {\it power spaces}, which are power sets endowed with multidirected structures. I focus on the case of power spaces {\it induced} by the structure of a multidirected set $M=(M,R,i,t)$. From a conservative perspective, power spaces provide a convenient tool for organizing the information contained in a multidirected set. An example is the theory of {\it causal atoms,} already used in the proof of theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} above. More ambitiously, power spaces may be used to encode causal structure that {\it cannot be described} by relations between pairs of elements representing individual events. This prompts the consideration of {\it top-down causation} and {\it classical holism} in discrete causal theory. In section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}, I discuss the theory of {\it causal path spaces} over multidirected sets, which abstract and generalize different aspects of relation spaces. Paths have special significance in theoretical physics, representing the flow of information or causal influence. In the context of discrete causal theory, a {\it causal path} is simply a {\it directed path} in a multidirected set $M$; i.e., a morphism from a {\it linear directed set} into $M$. I usually drop the qualifier ``causal" when referring to individual paths. Causal path spaces are amenable to a wide array of modern algebraic tools, providing powerful technical methods for approaching basic physical problems. In particular, the theory of {\it causal path semicategories} and {\it causal path algebras} leads to a precise algebraic implementation of the histories approach to quantum causal theory, developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}. Relations between pairs of elements in a multidirected set $M$ generalize to yield a family of multidirected structures called {\it splice relations} on causal path spaces over $M$. Two of the simplest and most important splice relations are the {\it concatenation relation} and the {\it directed product relation,} which underlie product operations on special causal path semicategories and causal path algebras of particular interest in the quantum setting. In section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}}, I outline the theory of path summation over a multidirected set. Path sums combine the values of special maps called {\it path functionals,} leading ultimately to the generalized quantum amplitudes, wave functions, and Schr\"{o}dinger-type equations studied in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}. Path functionals over multidirected sets play a role similar to the role of the {\it classical action} and the {\it Lagrangian} in continuum theory. \subsection{Relation Space over a Multidirected Set}\label{subsectionrelation} \refstepcounter{textlabels}\label{spacesinphys} \refstepcounter{textlabels}\label{relspace} {\bf Relation Space Preliminaries.} Physicists are well-accustomed to working in terms of ``spaces" other than ordinary spacetime. Some of these spaces, such as {\it Hilbert spaces} in quantum theory, {\it configuration spaces} and {\it phase spaces} in mechanics and thermodynamics, {\it measure spaces} in physical applications of probability, and the various {\it Lie groups,} {\it bundles,} and {\it sheaves} arising in field theories, correspond to spacetime only in the sense that the same algebraic, topological, and geometric intuitions and techniques are useful for studying both. Other spaces, such as {\it frequency spaces} in electronics, {\it momentum spaces} in classical and quantum mechanics, {\it twistor spaces} in Roger Penrose's twistor theory, and the {\it conformal boundary} in the AdS/CFT correspondence, represent {\it the same or very similar information} as the ordinary space or spacetime to which they correspond.\footnotemark\footnotetext{Such spaces may often be studied in the context of an appropriate {\it duality} theory, such as {\it Pontryagin duality,} {\it Tannaka-Klein duality,} or {\it Maldacena duality.} The rudiments of such a duality theory for multidirected sets and their relation spaces appear in definition \hyperref[defabstractelementset]{\ref{defabstractelementset}} and theorem \hyperref[theoreminterior]{\ref{theoreminterior}} below.} The {\it relation space} $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$, introduced in definition \hyperref[defirelationspace]{\ref{defirelationspace}} below, falls approximately into this latter class. The relation space $\ms{R}(M)$ over $M$ is a directed set whose underlying set is the relation set $R$ of $M$, identified in section \hyperref[subsectionchains]{\ref{subsectionchains}} with the set $\tn{Ch}_1(M)$ of $1$-chains in $M$, and used in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} in the definition of the continuum and star topologies. The binary relation $\prec$ on $\ms{R}(M)$, called the {\it induced relation,} arises naturally from the multidirected structure of $M$, as described in definition \hyperref[defirelationspace]{\ref{defirelationspace}} below. $\ms{R}(M)$ represents the {\it same information} as $M$ itself, {\it except on the boundary} $\partial M$ of $M$; i.e., its subset of extremal elements. This statement is made precise in theorem \hyperref[theoreminterior]{\ref{theoreminterior}}. $\ms{R}(M)$ is therefore ``approximately equivalent" to $M$ in an information-theoretic sense, but provides a superior viewpoint in several important ways. For example, {\it passage to relation space reduces multidirected structure to directed structure,} since $\ms{R}(M)$ is merely a directed set. Further, $\ms{R}(M)$ does not suffer from the generic problem of {\it permeability of maximal antichains in multidirected sets,} discussed later in this section. Relation space methods seem to be best suited to the discrete, nontransitive context, which makes them particularly useful for studying the spacetime models of principal interest in this paper. Relation space represents another application of Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}), with the {\it ``objects"} involved being the elements of a multidirected set $M=(M,R,i,t)$, and the {\it ``relationships between them"} being the elements of its relation set $R$. In the special case of directed sets, viewed as models of classical spacetime, these ``objects" and ``relationships" represent spacetime events and independent causal relations between pairs of events, respectively. For a multidirected set arising from the abstract structure of a kinematic scheme, {\it ``objects"} are directed sets, viewed as {\it classical universes,} and {\it ``relationships"} are {\it co-relative histories,} discussed in detail in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} below. {\bf Definition and First Properties of Relation Space.} The induced binary relation $\prec$ on the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$ is defined by taking one relation (i.e., element of $R$) to precede another if and only if the terminal element (in $M$) of the first relation coincides with the initial element of the second. Since $\prec$ is defined in terms of {\it ordered pairs} of elements of $R$, the relation space $\ms{R}(M)$ is automatically a {\it directed set;} i.e., no pair of (induced) relations in $\ms{R}(M)$ share the same initial and terminal elements (in $R$). This is the precise meaning of the above statement that {\it passage to relation space reduces multidirected structure to directed structure.}\\ \begin{defi}\label{defirelationspace} Let $M=(M,R,i,t)$ be a multidirected set, and let $r$ and $s$ be elements of its relation set $R$. \begin{enumerate} \item The {\bf induced relation} $\prec$ on $R$ is the binary relation defined by setting $r\prec s$ if and only if $t(r)=i(s)$. \item The directed set $\ms{R}(M)=(R, \prec)$ is called the {\bf relation space over} $M$. \end{enumerate} \end{defi} In the parlance of graph theory, $\ms{R}(M)$ is called the {\it line digraph} of $M$, where $M$ is viewed as a small directed multigraph. Here, {\it line} means {\it relation,} and {\it digraph} means {\it directed graph.} The first appearance of line digraphs in the literature seems to be Frank Harary and Robert Norman's 1960 paper {\it Some Properties of Line Digraphs} \cite{HararyLineDigraphs60}. As far as I know, these objects have received little or no attention in the context of fundamental spacetime structure, although they do appear in recent articles on physically relevant topics, such as the {\it control theory of network dynamics} \cite{NepuszEdgeDynamics12}. Figure \hyperref[relationspace]{\ref{relationspace}} below illustrates the construction of the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$. Figure \hyperref[relationspace]{\ref{relationspace}}a shows the induced relation between two elements $r$ and $s$ of $R$, with initial and terminal elements $x$ and $y$, and $y$ and $z$, in $M$, respectively. Figure \hyperref[relationspace]{\ref{relationspace}}b shows the global structure of $\ms{R}(M)$ as a directed set. In this particular example, $M$ is an acyclic directed set, so $\ms{R}(M)$ is an {\it irreducible} acyclic directed set, as proven in theorem \hyperref[theoremrelationfunctor]{\ref{theoremrelationfunctor}} below. In this figure, I introduce the convention of representing elements of relation space by {\it square nodes,} while elements of ordinary element space are represented by circular nodes in the usual way. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.35cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(0,8.5)}{\pgfbox[left,center]{{\bf }}} \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(7.15,4.95)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(7,-.1)(7,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \begin{pgftranslate}{\pgfpoint{1.25cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node13}{Node19} \color{white} \pgfmoveto{\pgfxy(-1,4.4)} \pgflineto{\pgfxy(-1,1.45)} \pgflineto{\pgfxy(1.3,1.45)} \pgflineto{\pgfxy(1.3,4.4)} \pgflineto{\pgfxy(-1,4.4)} \pgffill \pgfmoveto{\pgfxy(3.75,5)} \pgflineto{\pgfxy(3.75,2.8)} \pgflineto{\pgfxy(5.65,2.8)} \pgflineto{\pgfxy(5.65,5)} \pgflineto{\pgfxy(3.75,5)} \pgffill \color{black} \pgfmoveto{\pgfxy(-1,4.4)} \pgflineto{\pgfxy(-1,1.45)} \pgflineto{\pgfxy(1.3,1.45)} \pgflineto{\pgfxy(1.3,4.4)} \pgflineto{\pgfxy(-1,4.4)} \pgfstroke \pgfmoveto{\pgfxy(3.75,5)} \pgflineto{\pgfxy(3.75,2.8)} \pgflineto{\pgfxy(5.65,2.8)} \pgflineto{\pgfxy(5.65,5)} \pgflineto{\pgfxy(3.75,5)} \pgfstroke \pgfputat{\pgfxy(3.4,1.1)}{\pgfbox[center,center]{\large{$x$}}} \pgfputat{\pgfxy(2.9,1.7)}{\pgfbox[center,center]{\large{$r$}}} \pgfputat{\pgfxy(2.25,2.6)}{\pgfbox[center,center]{\large{$y$}}} \pgfputat{\pgfxy(2.55,3.95)}{\pgfbox[center,center]{\large{$s$}}} \pgfputat{\pgfxy(2.9,5.1)}{\pgfbox[center,center]{\large{$z$}}} \pgfputat{\pgfxy(-.75,4)}{\pgfbox[left,center]{relations}} \pgfputat{\pgfxy(-.75,3.45)}{\pgfbox[left,center]{$r$ and $s$}} \pgfputat{\pgfxy(-.75,2.95)}{\pgfbox[left,center]{represented}} \pgfputat{\pgfxy(-.75,2.45)}{\pgfbox[left,center]{by square}} \pgfputat{\pgfxy(-.75,1.95)}{\pgfbox[left,center]{nodes}} \pgfputat{\pgfxy(4,4.7)}{\pgfbox[left,center]{induced}} \pgfputat{\pgfxy(4,4.2)}{\pgfbox[left,center]{relation}} \pgfputat{\pgfxy(4,3.7)}{\pgfbox[left,center]{between}} \pgfputat{\pgfxy(4,3.2)}{\pgfbox[left,center]{$r$ and $s$}} \pgfnoderect{Node2013}[fill]{\pgfxy(2.95,2)}{\pgfxy(0.2,.2)} \pgfnoderect{Node1319}[fill]{\pgfxy(2.7,3.7)}{\pgfxy(0.2,.2)} \pgfsetlinewidth{1.5pt} \pgfnodeconnline{Node2013}{Node1319} \pgfsetlinewidth{.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(1.3,2.4)(2.65,2.05) \pgfxyline(1.3,3.2)(2.4,3.55) \pgfxyline(3.75,3.5)(3,2.9) \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.75cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \comment{ \begin{colormixin}{50!white} \pgfxyline(0,0)(12,0) \pgfxyline(0,.5)(12,.5) \pgfxyline(0,1)(12,1) \pgfxyline(0,1.5)(12,1.5) \pgfxyline(0,2)(12,2) \pgfxyline(0,2.5)(12,2.5) \pgfxyline(0,3)(12,3) \pgfxyline(0,3.5)(12,3.5) \pgfxyline(0,4)(12,4) \pgfxyline(0,4.5)(12,4.5) \pgfxyline(0,5)(12,5) \pgfxyline(0,5.5)(12,5.5) \pgfxyline(0,6)(12,6) \pgfxyline(0,0)(0,6) \pgfxyline(1,0)(1,6) \pgfxyline(2,0)(2,6) \pgfxyline(3,0)(3,6) \pgfxyline(4,0)(4,6) \pgfxyline(5,0)(5,6) \pgfxyline(6,0)(6,6) \pgfxyline(7,0)(7,6) \pgfxyline(8,0)(8,6) \pgfxyline(9,0)(9,6) \pgfxyline(10,0)(10,6) \pgfxyline(11,0)(11,6) \pgfxyline(12,0)(12,6) \end{colormixin} \pgfputat{\pgfxy(0,.25)}{\pgfbox[center,center]{0}} \pgfputat{\pgfxy(1,.25)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(2,.25)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(3,.25)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(4,.25)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(5,.25)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(6,.25)}{\pgfbox[center,center]{6}} \pgfputat{\pgfxy(7,.25)}{\pgfbox[center,center]{7}} \pgfputat{\pgfxy(8,.25)}{\pgfbox[center,center]{8}} \pgfputat{\pgfxy(9,.25)}{\pgfbox[center,center]{9}} \pgfputat{\pgfxy(10,.25)}{\pgfbox[center,center]{10}} \pgfputat{\pgfxy(11,.25)}{\pgfbox[center,center]{11}} \pgfputat{\pgfxy(12,.25)}{\pgfbox[center,center]{12}} \pgfputat{\pgfxy(0,1)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(0,2)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(0,3)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(0,4)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(0,5)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(0,6)}{\pgfbox[center,center]{6}} } \begin{colormixin}{100!white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0104}[fill]{\pgfxy(2.5,.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0105}[fill]{\pgfxy(1.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0110}[fill]{\pgfxy(3.7,.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0203}[fill]{\pgfxy(.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0206}[fill]{\pgfxy(1.3,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0207}[fill]{\pgfxy(2.1,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0405}[fill]{\pgfxy(2.3,1.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0603}[fill]{\pgfxy(1.7,4.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0703}[fill]{\pgfxy(2.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0803}[fill]{\pgfxy(2.9,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0904}[fill]{\pgfxy(3.1,.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2026}[fill]{\pgfxy(7.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2122}[fill]{\pgfxy(7.4,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2127}[fill]{\pgfxy(7.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2420}[fill]{\pgfxy(7.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2533}[fill]{\pgfxy(9.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2730}[fill]{\pgfxy(8.5,4.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2833}[fill]{\pgfxy(9.5,1.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2927}[fill]{\pgfxy(8.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2930}[fill]{\pgfxy(9,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2935}[fill]{\pgfxy(9.4,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3125}[fill]{\pgfxy(8.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3128}[fill]{\pgfxy(9.4,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3132}[fill]{\pgfxy(10.2,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3226}[fill]{\pgfxy(8.3,2.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3233}[fill]{\pgfxy(10.2,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3435}[fill]{\pgfxy(10.2,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3530}[fill]{\pgfxy(9.6,4.8)}{\pgfxy(0.2,0.2)} \pgfnodeconnline{Node0102}{Node0203} \pgfnodeconnline{Node0102}{Node0206} \pgfnodeconnline{Node0102}{Node0207} \pgfnodeconnline{Node0104}{Node0405} \pgfnodeconnline{Node0104}{Node0414} \pgfnodeconnline{Node0105}{Node0502} \pgfnodeconnline{Node0105}{Node0506} \pgfnodeconnline{Node0105}{Node0507} \pgfnodeconnline{Node0105}{Node0508} \pgfnodeconnline{Node0105}{Node0514} \pgfnodeconnline{Node0110}{Node1007} \pgfnodeconnline{Node0110}{Node1013} \pgfnodeconnline{Node0206}{Node0603} \pgfnodeconnline{Node0207}{Node0703} \pgfnodeconnline{Node0207}{Node0708} \pgfnodeconnline{Node0405}{Node0502} \pgfnodeconnline{Node0405}{Node0506} \pgfnodeconnline{Node0405}{Node0507} \pgfnodeconnline{Node0405}{Node0508} \pgfnodeconnline{Node0405}{Node0514} \pgfnodeconnline{Node0414}{Node1412} \pgfnodeconnline{Node0414}{Node1415} \pgfnodeconnline{Node0502}{Node0203} \pgfnodeconnline{Node0502}{Node0206} \pgfnodeconnline{Node0502}{Node0207} \pgfnodeconnline{Node0506}{Node0603} \pgfnodeconnline{Node0507}{Node0703} \pgfnodeconnline{Node0507}{Node0708} \pgfnodeconnline{Node0508}{Node0803} \pgfnodeconnline{Node0514}{Node1412} \pgfnodeconnline{Node0514}{Node1415} \pgfnodeconnline{Node0708}{Node0803} \pgfnodeconnline{Node0904}{Node0405} \pgfnodeconnline{Node0904}{Node0414} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0703} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1208}{Node0803} \pgfnodeconnline{Node1307}{Node0703} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1620}{Node2026} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2026}{Node2621} \pgfnodeconnline{Node2026}{Node2629} \pgfnodeconnline{Node2122}{Node2219} \pgfnodeconnline{Node2122}{Node2223} \pgfnodeconnline{Node2127}{Node2730} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2420}{Node2013} \pgfnodeconnline{Node2420}{Node2017} \pgfnodeconnline{Node2420}{Node2018} \pgfnodeconnline{Node2420}{Node2022} \pgfnodeconnline{Node2420}{Node2026} \pgfnodeconnline{Node2517}{Node1714} \pgfnodeconnline{Node2521}{Node2122} \pgfnodeconnline{Node2521}{Node2127} \pgfnodeconnline{Node2529}{Node2927} \pgfnodeconnline{Node2529}{Node2930} \pgfnodeconnline{Node2529}{Node2935} \pgfnodeconnline{Node2533}{Node3329} \pgfnodeconnline{Node2533}{Node3334} \pgfnodeconnline{Node2621}{Node2122} \pgfnodeconnline{Node2621}{Node2127} \pgfnodeconnline{Node2629}{Node2927} \pgfnodeconnline{Node2629}{Node2930} \pgfnodeconnline{Node2629}{Node2935} \pgfnodeconnline{Node2833}{Node3329} \pgfnodeconnline{Node2833}{Node3334} \pgfnodeconnline{Node2927}{Node2730} \pgfnodeconnline{Node2935}{Node3530} \pgfnodeconnline{Node3125}{Node2517} \pgfnodeconnline{Node3125}{Node2521} \pgfnodeconnline{Node3125}{Node2529} \pgfnodeconnline{Node3125}{Node2533} \pgfnodeconnline{Node3128}{Node2833} \pgfnodeconnline{Node3132}{Node3226} \pgfnodeconnline{Node3132}{Node3229} \pgfnodeconnline{Node3132}{Node3233} \pgfnodeconnline{Node3226}{Node2621} \pgfnodeconnline{Node3226}{Node2629} \pgfnodeconnline{Node3229}{Node2927} \pgfnodeconnline{Node3229}{Node2930} \pgfnodeconnline{Node3229}{Node2935} \pgfnodeconnline{Node3233}{Node3329} \pgfnodeconnline{Node3233}{Node3334} \pgfnodeconnline{Node3329}{Node2927} \pgfnodeconnline{Node3329}{Node2930} \pgfnodeconnline{Node3329}{Node2935} \pgfnodeconnline{Node3334}{Node3435} \pgfnodeconnline{Node3435}{Node3530} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Induced relation between two relations $r$ and $s$ in a multidirected set $M=(M,R,i,t)$; b) relation space $\ms{R}(M)$ over $M$, with induced relation $\prec$.} \label{relationspace} \end{figure} \vspace*{-.5cm} Theorem \hyperref[theoremrelationfunctor]{\ref{theoremrelationfunctor}} establishes the basic properties of relation space. \refstepcounter{textlabels}\label{thmrelspacefunctor} \vspace*{.2cm} \begin{theorem}\label{theoremrelationfunctor} Passage to relation space defines a functor $\ms{R}$ from the category $\ms{M}$ of multidirected sets to the category $\ms{D}$ of directed sets. This functor sends acyclic multidirected sets to irreducible acyclic directed sets, and preserves local finiteness. Moreover, a cycle of length $n$ in a multidirected set induces a cycle of length $n$ in its relation space. \end{theorem} \begin{proof} Let $M=(M,R,i,t)$ be a multidirected set, and let $\ms{R}(M)=(R,\prec)$ be its relation space. The proof proceeds in five steps: \begin{enumerate} \item {\it A morphism $\phi:M\rightarrow M'$ in $\ms{M}$ induces a morphism $\ms{R}(\phi):\ms{R}(M)\rightarrow \ms{R}(M')$ in $\ms{D}$.} Let $M'=(M',R',i',t')$ be a second multidirected set, and let $\ms{R}(M')=(R',\prec')$ be its relation space. Recall that a morphism $\phi:M\rightarrow M'$ consists of a {\it pair of maps} $\phi_{\tn{\fsz elt}}:M\rightarrow M'$ and $\phi_{\tn{\fsz rel}}:R\rightarrow R'$, respecting $i,t,i'$ and $t'$. Define $\ms{R}(\phi)$ to be the map $\phi_{\tn{\fsz rel}}$, as a map of sets, and let $r\prec s$ be a relation in $\ms{R}(M)$. Then by definition, $t(r)=i(s)$. Since $\phi$ is a morphism, \[t'(\phi_{\tn{\fsz rel}}(r))=\phi_{\tn {\fsz elt}}(t(r))=\phi_{\tn {\fsz elt}}(i(s))=i'(\phi_{\tn{\fsz rel}}(s)).\] Hence, $\phi_{\tn{\fsz rel}}(r)\prec'\phi_{\tn{\fsz rel}}(s)$ in $\ms{R}(M')$, so $\ms{R}(\phi)$ is a morphism in $\ms{D}$. \item {\it $\ms{R}$ is a functor.} It remains to show that $\ms{R}$ preserves identity morphisms and compositions. The identity morphism $\tn{Id}_{(M,R,i,t)}$ consists of the identity maps $\tn{Id}_M$ and $\tn{Id}_R$ on the {\it sets} $M$ and $R$, respectively, the latter of which coincides with the induced morphism $\ms{R}(\tn{Id}_{(M,R,i,t)})$ on the relation space $\ms{R}(M)$, by definition. Hence, $\ms{R}$ preserves identities. Similarly, given two morphisms $\phi:(M,R,i,t)\rightarrow (M',R',i',t')$ and $\psi:(M',R',i',t')\rightarrow (M'',R'',i'',t'')$, the map $(\psi\circ\phi)_{\tn{\fsz rel}}$ is by definition equal to the composition $\psi_{\tn{\fsz rel}}\circ\phi_{\tn{\fsz rel}}$, so $\ms{R}$ preserves compositions. \item {\it If $M$ is acyclic, then so is $\ms{R}(M)$.} I prove the contrapositive. Suppose that $r_0\prec r_1\prec...\prec r_n\prec r_0$ is a cycle in $\ms{R}(M)$. Then $t(r_j)=i(r_{j+1})$ for $0\le j<n$, and $t(r_n)=i(r_0)$, by the definition of $\ms{R}(M)$. Therefore $i(r_0)\prec i(r_1)\prec...\prec i(r_n)\prec i(r_0)$ is a cycle in $M$. \item {\it If $M$ is acyclic, then $\ms{R}(M)$ is irreducible.} Again, I prove the contrapositive. Suppose that $r\prec s$ is a relation in $\ms{R}(M)$ and $r=r_0\prec r_1\prec...\prec r_n=s$ is a reducing chain. Then $t(r)$ coincides with the both $i(s)$ and $i(r_1)$. If $n=2$, then $t(r_1)$ coincides with $i(s)$, so $r_1$ is a reflexive relation. If $n>2$, then $t(r_1)$ precedes $i(s)$, so $i(r_1)\prec...\prec i(s)=i(r_1)$ is a cycle in $M$. \item {\it If $M$ is locally finite, then so is $\ms{R}(M)$.} I prove the contrapositive. Suppose that $\ms{R}(M)$ is not locally finite, and let $r$ be an element of $\ms{R}(M)$ with infinite relation set. Since $\ms{R}(M)$ is a directed set, either the set of direct predecessors of $r$, or the set of direct successors of $r$, is infinite. Assume the former condition, and let $\{r_j\}_{j\in\mathbb{Z}}$ be an infinite subfamily of direct predecessors of $r$. By the definition of $\ms{R}(M)$, the initial element $i(r)$ of $r$ coincides with each terminal element $t(r_j)$ for all $j$. Therefore, the relation set $R(i(r))$ in $M$ is infinite, so $M$ is not locally finite. A symmetric argument applies if the set of direct successors of $r$ is infinite. \item {\it A cycle of length $n$ in $M$ induces a cycle of length $n$ in $\ms{R}(M)$.} Let $x_0\prec x_1\prec...\prec x_n= x_0$ be a cycle of length $n$ in $M$, where the distinguished relation $x_j\prec x_{j+1}$ is labeled as $r_j$.\footnotemark\footnotetext{Of course, there are generally other relations between these two elements, since $M$ is multidirected, but $r_j$ is the relation belonging to the cycle under consideration.} Then $r_0\prec r_1\prec...\prec r_{n-1}\prec r_0$ is a cycle of length $n$ in $\ms{R}(M)$ by the definition of relation space. \end{enumerate} \end{proof} \refstepcounter{textlabels} \label{acyclicinterpolativediscussion} The relation space $\ms{R}(A)$ over an acyclic directed set $A=(A,\prec)$ satisfies a stronger condition than irreducibility. For example, no acyclic directed set with multiple $2$-chains between a given pair of elements is the relation space of an acyclic directed set $A$, since the existence of such chains implies multidirected structure on $A$. Harary and Norman \cite{HararyLineDigraphs60} give a precise criterion identifying when a digraph is the line digraph of a multidigraph, but this result is not needed for the purposes of this paper. As mentioned above, the relation space functor $\ms{R}$ is well-suited to the discrete context. Preservation of local finiteness in theorem \hyperref[theoremrelationfunctor]{\ref{theoremrelationfunctor}} is one indication of this. It is interesting to observe how strongly this contrasts with the behavior of $\ms{R}$ in the interpolative context. For example, the relation space over the real line is not only noninterpolative, but irreducible. To see this, observe that the relation space $\ms{R}(\mathbb{R})$ may be identified, as a set, with the half plane above the diagonal $\{(x,x)\hspace*{.1cm}\big|\hspace*{.1cm} x\in\mathbb{R}\}$ in $\mathbb{R}^2$. In the induced relation on $\ms{R}(\mathbb{R})$, each element $(x,y)$ directly precedes all elements of the form $(y,z)$, and these relations are irreducible. Similar considerations apply to Minkowski spacetime $X$; every relation directly precedes a ``future light cone" of relations in a section of $X\times X$, and these relations are irreducible. \refstepcounter{textlabels}\label{abeltspace} {\bf Abstract Element Space.} I turn now to explaining the sense in which the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$ represents the same information as $M$, except on the boundary $\partial M$ of $M$. A conceptually useful way to approach this topic is by constructing an {\it approximate inverse} to the relation space functor $\ms{R}$, called the {\it abstract element space functor,} denoted by $\ms{E}$. Heuristically, one imagines that the elements of some directed set $R=(R,\prec)$ are {\it actually relations,} then constructs in a canonical way a multidirected set $\ms{E}(R)=(M,R,i,t)$, called the {\it abstract element space over} $R$, whose relation space is isomorphic to $(R,\prec)$.\footnotemark\footnotetext{The functor $\ms{E}$ may be applied to {\it any} directed set; the letter ``$R$" is used for suggestive purposes. However, not every directed set is the relation space of a multidirected set, as proven by Harary and Norman \cite{HararyLineDigraphs60}.} As demonstrated in theorem \hyperref[theoreminterior]{\ref{theoreminterior}} below, $\ms{E}(R)$ has the property that {\it any multidirected set whose relation space is isomorphic to $R$ has interior isomorphic to} $\tn{Int}(\ms{E}(R))$. Abstract element space is constructed via the following procedure:\\ \begin{defi}\label{defabstractelementset} Let $R=(R,\prec)$ be a directed set. Construct a multidirected set $\ms{E}(R)=(M,R,i,t)$ by means of the following procedure: \begin{enumerate} \item Define abstract sets $M^-:=\{x_r\hspace*{.1cm}|\hspace*{.1cm}r\in R\}$ and $M^+:=\{y_r\hspace*{.1cm}|\hspace*{.1cm}r\in R\}$. \item Form the quotient set $M:=M^-\coprod M^+/(y_r\sim x_s \tn{ iff } r\prec s \tn{ in } R)$. Denote the equivalence classes of $x_r$ and $y_r$ by $\widetilde{x_r}$ and $\widetilde{y_r}$, respectively. \item Define maps $i$ and $t$ from $M$ to $R$ by setting $i(r)=\widetilde{x_r}$ and $t(r)=\widetilde{y_r}$ for each $r\in R$. \end{enumerate} The multidirected set $\ms{E}(R)=(M,R,i,t)$ is called the {\bf abstract element space} over $R$. \end{defi} Figure \hyperref[abstractelt]{\ref{abstractelt}} below illustrates the construction of the abstract element space $\ms{E}(R)$ over a directed set $R$. The elements of $R$ are represented by circular nodes in figure \hyperref[abstractelt]{\ref{abstractelt}}a, to emphasize that $R$ need not be a relation space, although applying the relation space functor $\ms{R}$ to $\ms{E}(R)$ recovers $R$ in this particular example. The same elements are represented by directed edges in figure \hyperref[abstractelt]{\ref{abstractelt}}c, where they are ``realized" as relations. The elements of $M^\pm$ and $\ms{E}(R)$ are represented by open circular nodes. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{16cm}{5.4cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5.3) \pgfxyline(5.5,-.1)(5.5,5.3) \pgfxyline(11,-.1)(11,5.3) \pgfxyline(16.5,-.1)(16.5,5.3) \pgfxyline(0,5.3)(16.5,5.3) \pgfputat{\pgfxy(.15,5)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,5)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,5)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{-1.5cm}{.25cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(4,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3,1.75)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(5,1.25)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(5,2.25)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(4,3)}{0.1cm} \pgfnodecircle{Node6}[fill]{\pgfxy(6,1.5)}{0.1cm} \pgfputat{\pgfxy(4,.2)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(2.75,1.75)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(5.25,1.25)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(5.25,2.25)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.75,3)}{\pgfbox[center,center]{$v$}} \pgfputat{\pgfxy(6.25,1.5)}{\pgfbox[center,center]{$w$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(4,.5)(3.5,1.125) \pgfxyline(4,.5)(4.5,.875) \pgfxyline(5,1.25)(5,1.75) \pgfxyline(3,1.75)(3.5,2.375) \pgfxyline(5,2.25)(4.5,2.625) \pgfmoveto{\pgfxy(4,3)} \pgfcurveto{\pgfxy(4.8,3.5)}{\pgfxy(4.8,4)}{\pgfxy(4,4)} \pgfstroke \pgfclearendarrow \pgfxyline(3.5,1.125)(3,1.75) \pgfxyline(4.5,.875)(5,1.25) \pgfxyline(5,1.75)(5,2.25) \pgfxyline(3.5,2.375)(4,3) \pgfxyline(4.5,2.625)(4,3) \pgfmoveto{\pgfxy(4,3)} \pgfcurveto{\pgfxy(3.2,3.5)}{\pgfxy(3.2,4)}{\pgfxy(4,4)} \pgfstroke \pgfputat{\pgfxy(4.25,4.5)}{\pgfbox[center,center]{$R$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{.55cm}} \pgfnodecircle{Node1}[stroke]{\pgfxy(2.5,.5)}{0.1cm} \pgfnodecircle{Node2}[stroke]{\pgfxy(3.2,.5)}{0.1cm} \pgfnodecircle{Node3}[stroke]{\pgfxy(3.9,.5)}{0.1cm} \pgfnodecircle{Node4}[stroke]{\pgfxy(4.6,.5)}{0.1cm} \pgfnodecircle{Node5}[stroke]{\pgfxy(5.3,.5)}{0.1cm} \pgfnodecircle{Node6}[stroke]{\pgfxy(6,.5)}{0.1cm} \pgfnodecircle{Node11}[stroke]{\pgfxy(2.5,2.5)}{0.1cm} \pgfnodecircle{Node21}[stroke]{\pgfxy(3.2,2.5)}{0.1cm} \pgfnodecircle{Node31}[stroke]{\pgfxy(3.9,2.5)}{0.1cm} \pgfnodecircle{Node41}[stroke]{\pgfxy(4.6,2.5)}{0.1cm} \pgfnodecircle{Node51}[stroke]{\pgfxy(5.3,2.5)}{0.1cm} \pgfnodecircle{Node61}[stroke]{\pgfxy(6,2.5)}{0.1cm} \pgfputat{\pgfxy(2.3,1.6)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(3,1.6)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(3.7,1.6)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(4.4,1.6)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(5.1,1.6)}{\pgfbox[center,center]{$v$}} \pgfputat{\pgfxy(5.8,1.6)}{\pgfbox[center,center]{$w$}} \pgfputat{\pgfxy(2.5,.2)}{\pgfbox[center,center]{$x_r$}} \pgfputat{\pgfxy(3.2,.2)}{\pgfbox[center,center]{$x_s$}} \pgfputat{\pgfxy(3.9,.2)}{\pgfbox[center,center]{$x_t$}} \pgfputat{\pgfxy(4.6,.2)}{\pgfbox[center,center]{$x_u$}} \pgfputat{\pgfxy(5.3,.2)}{\pgfbox[center,center]{$x_v$}} \pgfputat{\pgfxy(6,.2)}{\pgfbox[center,center]{$x_w$}} \pgfputat{\pgfxy(2.5,2.8)}{\pgfbox[center,center]{$y_r$}} \pgfputat{\pgfxy(3.2,2.8)}{\pgfbox[center,center]{$y_s$}} \pgfputat{\pgfxy(3.9,2.8)}{\pgfbox[center,center]{$y_t$}} \pgfputat{\pgfxy(4.6,2.8)}{\pgfbox[center,center]{$y_u$}} \pgfputat{\pgfxy(5.3,2.8)}{\pgfbox[center,center]{$y_v$}} \pgfputat{\pgfxy(6,2.8)}{\pgfbox[center,center]{$y_w$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(2.5,.6)(2.5,1.5) \pgfxyline(3.2,.6)(3.2,1.5) \pgfxyline(3.9,.6)(3.9,1.5) \pgfxyline(4.6,.6)(4.6,1.5) \pgfxyline(5.3,.6)(5.3,1.5) \pgfxyline(6,.6)(6,1.5) \pgfclearendarrow \pgfxyline(2.5,1.5)(2.5,2.4) \pgfxyline(3.2,1.5)(3.2,2.4) \pgfxyline(3.9,1.5)(3.9,2.4) \pgfxyline(4.6,1.5)(4.6,2.4) \pgfxyline(5.3,1.5)(5.3,2.4) \pgfxyline(6,1.5)(6,2.4) \pgfputat{\pgfxy(4.25,4.25)}{\pgfbox[center,center]{$M^\pm$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{9cm}{.25cm}} \pgfnodecircle{Node1}[stroke]{\pgfxy(4,.5)}{0.1cm} \pgfnodecircle{Node2}[stroke]{\pgfxy(4,1.5)}{0.1cm} \pgfnodecircle{Node3}[stroke]{\pgfxy(5,2.25)}{0.1cm} \pgfnodecircle{Node4}[stroke]{\pgfxy(4,3)}{0.1cm} \pgfnodecircle{Node5}[stroke]{\pgfxy(6.5,1)}{0.1cm} \pgfnodecircle{Node6}[stroke]{\pgfxy(6.5,3)}{0.1cm} \pgfputat{\pgfxy(3.8,1)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(3.8,2.25)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(4.65,1.8)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(4.5,2.85)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.25,3.6)}{\pgfbox[center,center]{$v$}} \pgfputat{\pgfxy(6.3,2.1)}{\pgfbox[center,center]{$w$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(4,.6)(4,1.1) \pgfxyline(4.05,1.55)(4.55,1.9) \pgfxyline(4,1.6)(4,2.4) \pgfxyline(4.95,2.3)(4.45,2.67) \pgfxyline(6.5,1.1)(6.5,2.1) \pgfmoveto{\pgfxy(4.06,3.06)} \pgfcurveto{\pgfxy(4.8,3.5)}{\pgfxy(4.8,4)}{\pgfxy(4,4)} \pgfstroke \pgfclearendarrow \pgfxyline(4,1.1)(4,1.4) \pgfxyline(4.55,1.9)(4.95,2.2) \pgfxyline(4,2.4)(4,2.9) \pgfxyline(4.45,2.67)(4.05,2.95) \pgfxyline(6.5,2.1)(6.5,2.9) \pgfmoveto{\pgfxy(3.94,3.06)} \pgfcurveto{\pgfxy(3.2,3.5)}{\pgfxy(3.2,4)}{\pgfxy(4,4)} \pgfstroke \pgfputat{\pgfxy(4.25,4.5)}{\pgfbox[center,center]{$\ms{E}(R)$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) A directed set $R$; b) abstract sets $M^\pm$ (open nodes); arrows anticipate relations in $\ms{E}(R)$; c) abstract element space $\ms{E}(R)$; note that appropriate pairs of elements of $M^\pm$ have been identified according to definition \hyperref[defabstractelementset]{\ref{defabstractelementset}}.2.} \label{abstractelt} \end{figure} \vspace*{-.2cm} In general, $R$ cannot be recovered from $\ms{E}(R)$ by application of $\ms{R}$. This is true even for small finite irreducible acyclic directed sets, as illustrated in figure \hyperref[einducescycles]{\ref{einducescycles}} below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{16cm}{5.4cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5.3) \pgfxyline(5.5,-.1)(5.5,5.3) \pgfxyline(11,-.1)(11,5.3) \pgfxyline(16.5,-.1)(16.5,5.3) \pgfxyline(0,5.3)(16.5,5.3) \pgfputat{\pgfxy(.15,5)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,5)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,5)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{-1.5cm}{.25cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(5,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3,.5)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(4,2)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(5,2)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(3.7,3.3)}{0.1cm} \pgfputat{\pgfxy(5,.2)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(3,.2)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(4,2.3)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(5.3,2)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.7,3.6)}{\pgfbox[center,center]{$v$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(5,.5)(5,1.25) \pgfxyline(5,.5)(4.5,1.2) \pgfxyline(3,.5)(3.5,1.2) \pgfxyline(3,.5)(3.35,1.9) \pgfxyline(5,2)(4.35,2.6) \pgfclearendarrow \pgfxyline(5,1.25)(5,2) \pgfxyline(4.5,1.2)(4,2) \pgfxyline(3.5,1.2)(4,2) \pgfxyline(3.35,1.9)(3.7,3.3) \pgfxyline(4.35,2.6)(3.7,3.3) \pgfputat{\pgfxy(4.25,4.5)}{\pgfbox[center,center]{$R$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{.25cm}} \pgfnodecircle{Node1}[stroke]{\pgfxy(5,0)}{0.1cm} \pgfnodecircle{Node2}[stroke]{\pgfxy(2.6,.5)}{0.1cm} \pgfnodecircle{Node3}[stroke]{\pgfxy(4,1.75)}{0.1cm} \pgfnodecircle{Node4}[stroke]{\pgfxy(4.8,3.3)}{0.1cm} \pgfnodecircle{Node5}[stroke]{\pgfxy(3.5,3.8)}{0.1cm} \pgfputat{\pgfxy(4.85,.75)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(3.3,.85)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(4.65,2.7)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(5.2,1.75)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.85,3.15)}{\pgfbox[center,center]{$v$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(4.95,0.06)(4.5,.875) \pgfxyline(4.05,1.82)(4.47,2.65) \pgfxyline(3.98,1.83)(3.7,2.96) \pgfxyline(2.66,.55)(3.35,1.2) \pgfmoveto{\pgfxy(4.06,1.68)} \pgfcurveto{\pgfxy(4.4,1)}{\pgfxy(5,1)}{\pgfxy(5,1.75)} \pgfstroke \pgfclearendarrow \pgfxyline(4.5,.875)(4.05,1.65) \pgfxyline(4.47,2.65)(4.74,3.24) \pgfxyline(3.7,2.96)(3.52,3.72) \pgfxyline(3.35,1.2)(3.94,1.7) \pgfmoveto{\pgfxy(4.06,1.82)} \pgfcurveto{\pgfxy(4.4,2.5)}{\pgfxy(5,2.5)}{\pgfxy(5,1.75)} \pgfstroke \pgfputat{\pgfxy(4.25,4.5)}{\pgfbox[center,center]{$\ms{E}(R)$}} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{9.5cm}{.25cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(5,.5)}{0.1cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3,.5)}{0.1cm} \pgfnodecircle{Node3}[fill]{\pgfxy(4,2)}{0.1cm} \pgfnodecircle{Node4}[fill]{\pgfxy(5,2)}{0.1cm} \pgfnodecircle{Node5}[fill]{\pgfxy(3.7,3.3)}{0.1cm} \pgfputat{\pgfxy(5,.2)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(3,.2)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(4,2.3)}{\pgfbox[center,center]{$t$}} \pgfputat{\pgfxy(5.3,2)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(3.7,3.6)}{\pgfbox[center,center]{$v$}} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(5,.5)(5,1.25) \pgfxyline(5,.5)(4.5,1.2) \pgfxyline(3,.5)(3.5,1.2) \pgfxyline(3,.5)(3.35,1.9) \pgfxyline(5,2)(4.35,2.6) \pgfclearendarrow \pgfxyline(5,1.25)(5,2) \pgfxyline(4.5,1.2)(4,2) \pgfxyline(3.5,1.2)(4,2) \pgfxyline(3.35,1.9)(3.7,3.3) \pgfxyline(4.35,2.6)(3.7,3.3) \pgfsetdash{{2pt}{2pt}}{0pt} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(5,.5)(4.35,1.9) \pgfxyline(3,.5)(4,1.25) \pgfxyline(5,2)(4.5,2) \pgfmoveto{\pgfxy(5,2)} \pgfcurveto{\pgfxy(5.4,1.2)}{\pgfxy(6,1.2)}{\pgfxy(6,2)} \pgfstroke \pgfclearendarrow \pgfxyline(4.35,1.9)(3.7,3.3) \pgfxyline(4,1.25)(5,2) \pgfxyline(4.5,2)(4,2) \pgfmoveto{\pgfxy(5,2)} \pgfcurveto{\pgfxy(5.4,2.8)}{\pgfxy(6,2.8)}{\pgfxy(6,2)} \pgfstroke \pgfputat{\pgfxy(4.25,4.5)}{\pgfbox[center,center]{$\ms{R}\ms{E}(R)$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Irreducible acyclic directed set $R$; b) abstract element space $\ms{E}(R)$; c) relation space of $\ms{E}(R)$; note ``extra relations."} \label{einducescycles} \end{figure} \vspace*{-.2cm} In certain important cases, the directed set $R=(R,\prec)$ in definition \hyperref[defabstractelementset]{\ref{defabstractelementset}} arises by application of the relation space functor $\ms{R}$ to some ``preexisting" multidirected set. In such cases, there are {\it two different notions} of initial and terminal elements for an element $r$ of $R$. The images of $r$ under the initial and terminal element maps of the preexisting multidirected set are called {\bf concrete initial and terminal elements} of $r$, while the equivalence classes $\widetilde{x_r}$ and $\widetilde{y_r}$, given by the construction in definition \hyperref[defabstractelementset]{\ref{defabstractelementset}}, are called the {\bf abstract initial and terminal elements} of $r$. Theorem \hyperref[theoreminterior]{\ref{theoreminterior}} below shows that {\it concrete interior elements correspond bijectively to abstract interior elements,} while {\it each concrete boundary element corresponds to a family of abstract boundary elements.} Hence, any multidirected set with relation space isomorphic to $R$ may be obtained from $\ms{E}(R)$, up to isomorphism, by identifying appropriate families of abstract boundary elements. It is interesting to note that the abstract element space over the continuum is a multidirected set consisting of $|\mathbb{R}|$ reflexive relations on a singleton. In particular, all notion of linear order is lost after applying $\ms{E}$. To see this, note that any pair $r$ and $s$ of real numbers have a common successor, whose abstract initial element coincides with the abstract terminal elements of $r$ and $s$, by the definition of $\ms{E}(\mathbb{R})$. Thus, all terminal elements in $\ms{E}(\mathbb{R})$ coincide. Similarly, all initial elements in $\ms{E}(\mathbb{R})$ coincide. An analogous argument applies to Minkowski spacetime. More generally, the same argument applies to any directed set satisfying the property that each pair of elements has a common predecessor and a common successor; the abstract element spaces over such sets depend only on their cardinality.\footnotemark\footnotetext{I thank Johnny Feng \cite{Feng13} for pointing out this obvious fact to me. Using conventional terminology, such sets are often called ``upward directed" and ``downward directed," respectively.} It is also useful to examine the interaction of $\ms{E}$ with the transitive closure functor $\tn{tr}$ and the skeleton operation $\tn{sk}$. The integers $\mathbb{Z}$ and rational numbers $\mathbb{Q}$ provide instructive examples. Recall that $\tn{tr}(\mathbb{Z})=\mathbb{Z}$ and $\tn{tr}(\mathbb{Q})=\mathbb{Q}$, since the binary relations on $\mathbb{Z}$ and $\mathbb{Q}$ are transitive, but $\tn{sk}(\mathbb{Z})$ is an irreducible unbounded chain, while $\tn{sk}(\mathbb{Q})$ is a set of cardinality $\aleph_0$ with no relations. The abstract element spaces $\ms{E}(\mathbb{Z})$ and $\ms{E}(\mathbb{Q})$ both consist of $\aleph_0$ reflexive relations on a singleton, but $\ms{E}(\tn{sk}(\mathbb{Z}))\cong\tn{sk}(\mathbb{Z})$, while $\ms{E}(\tn{sk}(\mathbb{Q}))$ consists of $\aleph_0$ disjoint $1$-chains. These examples illustrate another reason for abstaining from transitivity in modeling discrete spacetime structure, since redundant reducible relations introduce useless clutter under application of $\ms{E}$. Theorem \hyperref[theoremelementfunctor]{\ref{theoremelementfunctor}} establishes the basic properties of abstract element space. \refstepcounter{textlabels}\label{thmelspacefunctor} \vspace*{.2cm} \begin{theorem}\label{theoremelementfunctor} Passage to abstract element space defines a functor $\ms{E}$ from the category $\ms{D}$ of directed sets to the category $\ms{M}$ of multidirected sets. This functor preserves local finiteness, but fails to preserve acyclicity, even for irreducible directed sets. \end{theorem} \begin{proof} Let $R=(R,\prec)$ and $R'=(R',\prec')$ be directed sets, with abstract element spaces $\ms{E}(R)=(M,R,i,t)$ and $\ms{E}(R')=(M',R',i',t')$, respectively. Let $\phi:R\rightarrow R'$ be a morphism. Recall that for elements $r$ and $r'$ in $R$ and $R'$, respectively, \[i(r)=\widetilde{x_r}, \hspace*{.5cm} t(r)=\widetilde{y_r}, \hspace*{.5cm} i'(r')=\widetilde{x_{r'}}, \hspace*{.5cm}\tn{and}\hspace*{.5cm} t'(r')=\widetilde{y_{r'}},\] by part 3 of definition \hyperref[defabstractelementset]{\ref{defabstractelementset}}. The induced morphism $\ms{E}(\phi)$ consists of a map of elements $\phi_{\tn{\fsz elt}}:M\rightarrow M'$, and a map of relations $\phi_{\tn{\fsz rel}}:R\rightarrow R'$. Define $\phi_{\tn{\fsz elt}}$ by setting $\phi_{\tn{\fsz elt}}(\widetilde{x_r})=\widetilde{x_{\phi(r)}}$ and $\phi_{\tn{\fsz elt}}(\widetilde{y_r})=\widetilde{y_{\phi(r)}}$. Define $\phi_{\tn{\fsz rel}}$ to coincide with the given morphism $\phi:R\rightarrow R'$. The proof proceeds in four steps: \begin{enumerate} \item {\it The map of elements $\phi_{\tn{\fsz elt}}:M\rightarrow M'$ is well-defined.} A given element $z\in M$ may be the abstract initial element of many different relations in $R$, and the abstract terminal element of many others. It is necessary to show that the images of these elements under $\phi_{\tn{\fsz elt}}$ all coincide in $M'$. For example, suppose that $z=\widetilde{x_r}=\widetilde{x_{s}}$ for two different relations $r$ and $s$ in $R$, and consider the corresponding images $\widetilde{x_{\phi(r)}}$ and $\widetilde{x_{\phi(s)}}$ in $M'$ under $\phi_{\tn{\fsz elt}}$. In this case, $z$ is nonminimal in $M$, since extremal elements in abstract element space, viewed as equivalence classes, are singletons. Therefore, there exists an element $q$ in $R$ whose terminal element $t(q)$ in $M$ is $z$; i.e., $q$ directly precedes both $r$ and $s$ in $R$. Since $\phi$ is a morphism, $\phi_{\tn{\fsz rel}}(q)$ directly precedes both $\phi(r)$ and $\phi(s)$ in $R'$. Hence, the abstract terminal element $\widetilde{y_{\phi(q)}}$ of $q$ coincides with both the abstract initial elements $\widetilde{x_{\phi(r)}}$ and $\widetilde{x_{\phi(s)}}$, which therefore coincide with each other. A symmetric argument applies if $z=t(r)=t(s)$. Finally, if $z=t(r)=i(s)$, then $r$ directly precedes $s$ in $R$, so $\phi(r)$ directly precedes $\phi(s)$ in $R'$, and therefore $\widetilde{y_{\phi(r)}}=\widetilde{x_{\phi(s)}}$. \item {\it The pair $\ms{E}(\phi):=(\phi_{\tn{\fsz elt}},\phi_{\tn{\fsz rel}})$ is a morphism of multidirected sets.} It is necessary to show that $\phi_{\tn{\fsz elt}}$ and $\phi_{\tn{\fsz rel}}$ respect initial and terminal element maps in the sense that \[\phi_{\tn{\fsz elt}}\big(i(r)\big)=i'\big(\phi_{\tn{\fsz rel}}(r)\big)\hspace*{.5cm}\tn{and}\hspace*{.5cm} \phi_{\tn{\fsz elt}}\big(t(r)\big)=t'\big(\phi_{\tn{\fsz rel}}(r)\big),\] for every $r$ in $R$. Writing this in terms of equivalence classes, and recalling that $\phi_{\tn{\fsz rel}}$ is just the given morphism $\phi$, this becomes \[\phi_{\tn{\fsz elt}}(\widetilde{x_r})=\widetilde{x_{\phi(r)}}\hspace*{.5cm}\tn{and}\hspace*{.5cm} \phi_{\tn{\fsz elt}}(\widetilde{y_r})=\widetilde{y_{\phi(r)}},\] which is the {\it definition} of $\phi_{\tn{\fsz elt}}$. \item {\it $\ms{E}$ is a functor.} If $\phi:R\rightarrow R$ is the identity morphism on $R$, then the map of relations $\phi_{\tn{\fsz rel}}$ in $\ms{E}(\phi)=(\phi_{\tn{\fsz elt}},\phi_{\tn{\fsz rel}})$ is the identity morphism on $R$ by definition, and the map of elements $\phi_{\tn{\fsz elt}}$ sends $\widetilde{x_r}$ to $\widetilde{x_{\phi(r)}}=\widetilde{x_r}$. A symmetric statement applies to abstract terminal elements. Hence, $\ms{E}$ preserves identities. Similarly, given two morphisms $\phi:R\rightarrow R'$ and $\psi:R'\rightarrow R''$, where $R''=(R'',\prec'')$ is a third directed set, the map of relations $(\psi\circ\phi)_{\tn{\fsz rel}}$ in the induced morphism $\ms{E}(\psi\circ\phi)$ is the composition $\psi\circ\phi=\psi_{\tn{\fsz rel}}\circ\phi_{\tn{\fsz rel}}$ by definition, while the map of elements $(\psi\circ\phi)_{\tn{\fsz elt}}$ sends $\widetilde{x_r}$ to $\widetilde{x_{\psi\circ\phi(r)}}=\psi_{\tn{\fsz elt}}(\widetilde{x_{\phi(r)}})=\psi_{\tn{\fsz elt}}\circ\phi_{\tn{\fsz elt}}(\widetilde{x_r})$, and similarly for terminal elements. Hence, $\ms{E}$ preserves compositions. \item {\it If $R$ is locally finite, then so is $\ms{E}(R)$.} I prove the contrapositive. Suppose that $\ms{E}(R)$ is not locally finite, and let $x$ be an element of $\ms{E}(R)$ with infinite relation set. Then either $x$ is related to an infinite number of other elements of $\ms{E}(R)$, or there exist an infinite number of relations between $x$ and some other individual element of $\ms{E}(R)$. First assume that $x$ has an infinite number of direct successors $\{y_n\}_{n\in\mathbb{Z}}$ in $\ms{E}(R)$. Since $R$ is the relation set of $\ms{E}(R)$, there exist distinct $r_n$ in $R$ for each $n$, such that $x=\widetilde{x_{r_n}}$ and $y=\widetilde{y_{r_n}}$ for each $n$. Since the abstract initial elements $\widetilde{x_{r_n}}$ of the relations $r_n$ coincide in $\ms{E}(R)$, there exists a relation $r^-$ in $R$ such that $\widetilde{y_{r^-}}=\widetilde{x_{r_n}}$ in $\ms{E}(R)$ for each $n$. But this implies that $\{r_n\}_{n\in\mathbb{Z}}$ constitutes an infinite family of direct successors of $r^-$ in $R$, so $R$ is not locally finite. A symmetric argument applies if $x$ has an infinite number of directed predecessors. Next, assume that there exist an infinite number of relations of the form $x\prec y$ for some other element $y$ in $\ms{E}(R)$. Then there exists an infinite family $\{r_n\}_{n\in\mathbb{Z}}$ of elements of $R$ such that $x=\widetilde{x_{r_n}}$ and $y=\widetilde{y_{r_n}}$ for all $n$. Since the abstract initial elements of each $r_n$ coincide in $\ms{E}(R)$, and similarly for the abstract terminal elements, there exist elements $r^\pm$ of $R$ such that $\widetilde{y_{r^-}}=\widetilde{x_{r_n}}$ and $\widetilde{x_{r^+}}=\widetilde{y_{r_n}}$ in $\ms{E}(R)$ for each $n$. Hence, $\{r_n\}_{n\in\mathbb{Z}}$ simultaneously constitutes an infinite family of direct successors of $r^-$ and an infinite family of direct predecessors of $r^+$ in $R$, so again $R$ is not locally finite. A symmetric argument applies if there exist an infinite number of relations of the form $w\prec x$ for some other element $w$ in $\ms{E}(R)$. \item {\it $\ms{E}$ fails to preserve acyclicity, even in the irreducible case.} The irreducible acyclic directed set illustrated in figure \hyperref[einducescycles]{\ref{einducescycles}} above is a counterexample. \end{enumerate} \end{proof} {\bf Approximate Information Preservation by $\ms{R}$.} The relation space functor $\ms{R}$, and the abstract element space functor $\ms{E}$, interact in interesting and useful ways. Theorem \hyperref[theoreminterior]{\ref{theoreminterior}} below renders precise the assertion that the relation space $\ms{R}(M)$ of a multidirected set $M=(M,R,i,t)$ represents the same information as $M$, except on the boundary $\partial M$ of $M$. This theorem has the same substance as theorem 3 of Harary and Norman \cite{HararyLineDigraphs60}. This result has important physical applications, particularly in the quantum theoretic context. For example, the replacement of certain ``badly-behaved" multidirected sets with their ``well-behaved" relation spaces facilitates the development of a suitable version of the histories approach to quantum causal theory, expressed via path summation, as described in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper.\\ \refstepcounter{textlabels}\label{thmrelpreserves} \begin{theorem}\label{theoreminterior} Any pair of multidirected sets with isomorphic relation spaces have isomorphic interiors. \end{theorem} \begin{proof} The proof actually establishes the more specific statement that if $R=(R,\prec)$ is a directed set, and if $M$ is any multidirected set whose relation space $\ms{R}(M)$ is isomorphic to $R$, then $\tn{Int}(M)\cong\tn{Int}(\ms{E}(R))$. Since the relation space $\ms{R}(M)$ of $M$ is isomorphic to $(R,\prec)$, $M$ may be expressed, up to isomorphism, as a quadruple $(M,R,i,t)$, where the ``$R$" appearing in the quadruple coincides, as a set, with the given directed set $(R,\prec)$, and where the maps $i$ and $t$ identify initial and terminal elements in $M$; i.e., {\it concrete} initial and terminal elements. Throughout the proof, I use the ``tilde notation," appearing in part 2 of definition \hyperref[defabstractelementset]{\ref{defabstractelementset}} above, to denote elements of $\tn{Int}(\ms{E}(R))$; i.e., {\it abstract} initial and terminal elements. The interior $\tn{Int}(M)$ of $M$, viewed as a subobject of $M$, is the multidirected set whose elements are nonextremal elements of $M$, whose relations are elements of $R$ having nonextremal initial and terminal elements, and whose initial and terminal element maps are the restrictions of $i$ and $t$. In particular, any relation between two elements in $\tn{Int}(M)$ belongs to $\tn{Int}(\ms{R}(M))$, by the definition of the induced relation on $\tn{Int}(\ms{R}(M))$. Analogous statements apply to $\tn{Int}(\ms{E}(R))$. Therefore, an isomorphism $\phi$ between $\tn{Int}(M)$ and $\tn{Int}(\ms{E}(R))$, if it exists, consists of a map of elements $\phi_{\tn{\fsz elt}}:\tn{Int}(M)\rightarrow \tn{Int}(\ms{E}(R))$, and a map of relations $\phi_{\tn{\fsz rel}}:\tn{Int}(\ms{R}(M))\rightarrow \tn{Int}(R)$. I define the latter map $\phi_{\tn{\fsz rel}}$ to be the restriction of the given isomorphism $\phi:\ms{R}(M)\rightarrow R$ to the interior of $\ms{R}(M)$. Using the fact that any element of $\tn{Int}(M)$ may be written as either an initial element or a terminal element, I define the former map $\phi_{\tn{\fsz elt}}$ by setting $\phi_{\tn{\fsz elt}}(i(r))=\widetilde{x_{\phi_{\tn{\fsz rel}}(r)}}$ and $\phi_{\tn{\fsz elt}}(t(r))=\widetilde{y_{\phi_{\tn{\fsz rel}}(r)}}$. The proof then proceeds in four steps. \begin{enumerate} \item {\it The map $\phi_{\tn{\fsz elt}}$ is well-defined.} This is similar to part 1 of the proof of theorem \hyperref[theoremelementfunctor]{\ref{theoremelementfunctor}} above. Suppose that $z=i(r)=i(s)$ for two elements $r$ and $s$ of $R$. Then it is necessary to show that the images $\widetilde{x_{\phi_{\tn{\fsz rel}}(r)}}$ and $\widetilde{x_{\phi_{\tn{\fsz rel}}(s)}}$ belong to $\tn{Int}(\ms{E}(R))$ and are equal. Since $z$ belongs to $\tn{Int}(M)$, there exists an element $q$ of $R$ whose such that $t(q)=z$ in $M$. Therefore, $q$ directly precedes both $r$ and $s$ in $\ms{R}(M)$, so $\phi_{\tn{\fsz rel}}(q)$ directly precedes both $\phi_{\tn{\fsz rel}}(r)$ and $\phi_{\tn{\fsz rel}}(s)$ in $R$. Hence, the elements $\widetilde{x_{\phi_{\tn{\fsz rel}}(r)}}$ and $\widetilde{x_{\phi_{\tn{\fsz rel}}(s)}}$ coincide in $\ms{E}(R)$, by the definition of abstract element space. A symmetric argument applies if $z=t(r)=t(s)$. Finally, if $z=t(r)=i(s)$, then $\phi_{\tn{\fsz rel}}(r)$ directly precedes $\phi_{\tn{\fsz rel}}(s)$ in $R$, so $\widetilde{y_{\phi_{\tn{\fsz rel}}(r)}}=\widetilde{x_{\phi_{\tn{\fsz rel}}(s)}}$ in $\ms{E}(R)$, again by the definition of abstract element space. \item {\it The pair of maps $(\phi_{\tn{\fsz elt}},\phi_{\tn{\fsz rel}})$ define a morphism of multidirected sets.} The map $\phi_{\tn{\fsz rel}}$ simply identifies each relation in $\tn{Int}(\ms{R}(M))$ with the corresponding relation in $\tn{Int}(R)$, via the hypothesized isomorphism $\phi:\ms{R}(M)\cong R$ of directed sets. If $r$ belongs to $\tn{Int}(\ms{R}(M))$, then its concrete initial and terminal elements $i(r)$ and $t(r)$ in $M$ map to the abstract initial and terminal elements $\widetilde{x_{\phi_{\tn{\fsz rel}}(r)}}$ and $\widetilde{y_{\phi_{\tn{\fsz rel}}(r)}}$ of its image $\phi_{\tn{\fsz rel}}(r)$ in $\tn{Int}(\ms{E}(R))$, by the definition of $\phi_{\tn{\fsz elt}}$. Hence, $(\phi_{\tn{\fsz elt}},\phi_{\tn{\fsz rel}})$ respects initial and terminal element maps. \item {\it The map $\phi_{\tn{\fsz elt}}$ is bijective.} Suppose that $z=t(r)=i(s)$ and $w=t(\overline{r})=i(\overline{s})$ are two elements of $\tn{Int}(\ms{R}(M))$ whose images under $\phi_{\tn{\fsz elt}}$ coincide, where $r,s,\overline{r}$ and $\overline{s}$ are elements of $R$. Then by definition, $\widetilde{x_{\phi_{\tn{\fsz rel}}(s)}}=\widetilde{x_{\phi_{\tn{\fsz rel}}(\overline{s})}}=\widetilde{y_{\phi_{\tn{\fsz rel}}(r)}}=\widetilde{y_{\phi_{\tn{\fsz rel}}(\overline{r})}}$ in $\tn{Int}(M)$. Therefore, $\phi_{\tn{\fsz rel}}(r)$ and $\phi_{\tn{\fsz rel}}(\overline{r})$ both directly precede $\phi_{\tn{\fsz rel}}(s)$ and $\phi_{\tn{\fsz rel}}(\overline{s})$ in $\tn{Int}(R)$. Since $\phi_{\tn{\fsz rel}}$ is an isomorphism of directed sets, $r$ and $\overline{r}$ both directly precede $s$ and $\overline{s}$ in $\tn{Int}(\ms{R}(M))$. Hence, $t(r)=t(\overline{r})=i(s)=i(\overline{s})$, so $z=w$, and $\phi_{\tn{\fsz elt}}$ is injective. Next, observe that any element in $\tn{Int}(\ms{E}(R))$ may be written as $\widetilde{x_{\phi_{\tn{\fsz rel}}(s)}}$ for some relation $s$ in $\tn{Int}(\ms{R}(M))$. By definition of $\phi_{\tn{\fsz elt}}$, this element lifts to the element $i(s)$ in $\tn{Int}(M)$. Thus, $\phi_{\tn{\fsz elt}}$ is surjective. \item {\it The pair of inverse maps $(\phi_{\tn{\fsz elt}}^{-1},\phi_{\tn{\fsz rel}}^{-1})$ define a morphism of multidirected sets.} Since $\phi_{\tn{\fsz rel}}=\phi$ is an isomorphism, any element of $\tn{Int}(R)$ may be expressed as $\phi_{\tn{\fsz rel}}(r)$ for a unique element $r$ of $\ms{R}(M)$. The image of this element under the inverse relation map $\phi_{\tn{\fsz rel}}^{-1}$ is just $r$. By definition, the inverse element map $\phi_{\tn{\fsz elt}}^{-1}$ carries the abstract initial and terminal elements of $\phi_{\tn{\fsz rel}}(r)$ in $\tn{Int}(\ms{E}(R))$ to the concrete initial and terminal elements $i(r)$ and $t(r)$ in $\tn{Int}(M)$. Hence, the pair $(\phi_{\tn{\fsz elt}}^{-1},\phi_{\tn{\fsz rel}}^{-1})$ respects initial and terminal element maps. \end{enumerate} \end{proof} \refstepcounter{textlabels}\label{cauchyrel} {\bf Cauchy Surfaces; Impermeability.} I now discuss an information-theoretic device of great importance in continuum theory, both in the classical and quantum settings, whose adaptation to causal set theory has hitherto been hampered by a particular technical problem. A {\bf Cauchy surface} is a subset $\sigma$ of a region $X$ of continuum spacetime, admitting a {\it unique point of intersection with every causal path in $X$ between its past and future.} A Cauchy surface may be viewed as a {\it generalized spacelike section} of spacetime, representing a single ``instant in time." From an information-theoretic viewpoint, a Cauchy surface filters the flow of information from past to future through $X$. Foliation of spacetime via Cauchy surfaces is an important technique in relativistic cosmology, and also plays a central role in various approaches to quantum gravity, including loop quantum gravity, noncommutative geometry, and shape dynamics. Figure \hyperref[permeability]{\ref{permeability}}a below illustrates a Cauchy surface $\sigma$ in relativistic spacetime, with two representative causal paths passing through $\sigma$ from past to future. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.35cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(8.4,4.95)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(8.25,-.1)(8.25,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \begin{pgftranslate}{\pgfpoint{.5cm}{.2cm}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{20!white} \color{black} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgffill \end{colormixin} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgfstroke \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(1,0)(9.5,0) \pgfxyline(1,0)(1,6.8) \pgfclearendarrow \pgfsetlinewidth{1.5pt} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(4,3)}{\pgfxy(6,2.5)}{\pgfxy(8.5,4)} \pgfstroke \end{pgfscope} \begin{pgftranslate}{\pgfpoint{.8cm}{0cm}} \begin{colormixin}{100!white} \color{black} \pgfsetlinewidth{1pt} \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfmoveto{\pgfxy(3.5,.9)} \pgfcurveto{\pgfxy(2.5,1.5)}{\pgfxy(2.4,2)}{\pgfxy(2.3,2.4)} \pgfstroke \pgfmoveto{\pgfxy(2.3,2.4)} \pgfcurveto{\pgfxy(2.2,3.2)}{\pgfxy(3.2,3.6)}{\pgfxy(3,4.5)} \pgfstroke \pgfmoveto{\pgfxy(6.1,1.3)} \pgfcurveto{\pgfxy(5.5,1.7)}{\pgfxy(5.4,2.1)}{\pgfxy(5.2,2.4)} \pgfstroke \pgfmoveto{\pgfxy(5.2,2.4)} \pgfcurveto{\pgfxy(5,2.9)}{\pgfxy(5.2,3.2)}{\pgfxy(5.6,4.2)} \pgfstroke \end{pgfscope} \pgfmoveto{\pgfxy(3,4.5)} \pgfcurveto{\pgfxy(3,4.8)}{\pgfxy(2.7,5.5)}{\pgfxy(2.6,5.95)} \pgfstroke \pgfmoveto{\pgfxy(5.6,4.2)} \pgfcurveto{\pgfxy(5.7,4.5)}{\pgfxy(6,5.1)}{\pgfxy(6.5,5.92)} \pgfstroke \end{colormixin} \color{black} \pgfnodecircle{Node1}[fill]{\pgfxy(3.5,.9)}{0.14cm} \pgfnodecircle{Node3}[fill]{\pgfxy(6.1,1.3)}{0.14cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2.6,5.95)}{0.14cm} \pgfnodecircle{Node8}[fill]{\pgfxy(6.5,5.92)}{0.14cm} \end{pgftranslate} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(6.85,.45)}{\pgfbox[left,center]{$x$}} \pgfputat{\pgfxy(1.35,4.8)}{\pgfbox[left,center]{$t$}} \pgfputat{\pgfxy(4.1,3.7)}{\pgfbox[left,center]{$X$}} \pgfputat{\pgfxy(5.5,2.8)}{\pgfbox[left,center]{$\sigma$}} \begin{pgftranslate}{\pgfpoint{9.75cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node27}{Node30} \pgfsetlinewidth{.5pt} \color{white} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \color{black} \pgfnodecircle{Node2}[stroke]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node13}[stroke]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node17}[stroke]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node26}[stroke]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node33}[stroke]{\pgfxy(5,2.5)}{0.11cm} \pgfputat{\pgfxy(5.2,.35)}{\pgfbox[center,center]{\large{$x$}}} \pgfputat{\pgfxy(4.5,5.5)}{\pgfbox[center,center]{\large{$y$}}} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Cauchy surface in relativistic spacetime; b) maximal antichain in a multidirected set $M$ (open nodes); dark paths are permeating chains.} \label{permeability} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{eltspaceperm} The na\"{\i}ve analogue of a Cauchy surface in a multidirected set $M=(M,R,i,t)$ is a {\bf maximal antichain} in $M$; i.e., a subset $\sigma$ of $M$ such that no chain in $M$ connects any pair of elements of $\sigma$, but every element in the complement of $\sigma$ is connected to an element of $\sigma$ by a chain. This is illustrated in figure \hyperref[permeability]{\ref{permeability}}b above, which shows the acyclic directed case. The maximal antichain is represented by the five open nodes in the figure. The dark-colored chains, and the labels $x$ and $y$, are for future reference. A technical problem associated with maximal antichains in multidirected sets is that they are generally {\bf permeable}, meaning that there generally exist chains extending from the past of a maximal antichain $\sigma$ to its future {\it without intersecting} $\sigma$. Two such paths are illustrated in figure \hyperref[permeability]{\ref{permeability}}b. This problem is particularly severe in the transitive case, with numerous relations, mostly reducible, passing directly from the ``distant past" to the ``distant future" of $\sigma$. In causal set theory, reducible relations are taken to be dependent, so one might na\"{i}vely hope that the problem is only a formal one in this context, disappearing upon passage to the skeleton. This hope fails, as illustrated by the permeating chain from $x$ to $y$ in figure \hyperref[permeability]{\ref{permeability}}b, which involves only irreducible relations. The problem of permeability therefore {\it does not depend on the physical interpretation of reducible relations;} i.e., on whether or not one accepts the independence convention (\hyperref[ic]{IC}). \refstepcounter{textlabels}\label{rideoutperm} Seth Major, David Rideout, and Sumati Surya examine maximal antichains as analogues of Cauchy surfaces in causal set theory in their 2006 paper {\it Spatial Hypersurfaces in Causal Set Cosmology} \cite{RideoutSpatialHypersurface06}. Recognizing the permeability problem, they attempt to address it by {\it thickening} maximal antichains to ``capture" permeating chains.\footnotemark\footnotetext{Surya mentions the same problem in her 2011 review article \cite{Surya11}.} While such constructions carry some intrinsic interest, they are not particularly natural, and do not really solve the problem. A better approach is to pass to relation space, and work with {\it maximal antichains of relations.} Remarkably, no additional device is needed: the permeability problem disappears in relation space. This is made precise by theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} below, which states that {\it maximal antichains in relation space are impermeable.} Figure \hyperref[figimpermeability]{\ref{figimpermeability}}a below illustrates a typical maximal antichain in relation space. The underlying multidirected set from which this relation space is derived is the familiar example appearing in figure \hyperref[permeability]{\ref{permeability}}b above. Figure \hyperref[figimpermeability]{\ref{figimpermeability}}b is used in the proof of theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{16cm}{5.6cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5.5) \pgfxyline(10.15,-.1)(10.15,5.5) \pgfxyline(16.5,-.1)(16.5,5.5) \pgfxyline(0,5.5)(16.5,5.5) \pgfputat{\pgfxy(.15,5.2)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(10.3,5.2)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{0cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \comment{ \begin{colormixin}{50!white} \pgfxyline(0,0)(12,0) \pgfxyline(0,.5)(12,.5) \pgfxyline(0,1)(12,1) \pgfxyline(0,1.5)(12,1.5) \pgfxyline(0,2)(12,2) \pgfxyline(0,2.5)(12,2.5) \pgfxyline(0,3)(12,3) \pgfxyline(0,3.5)(12,3.5) \pgfxyline(0,4)(12,4) \pgfxyline(0,4.5)(12,4.5) \pgfxyline(0,5)(12,5) \pgfxyline(0,5.5)(12,5.5) \pgfxyline(0,6)(12,6) \pgfxyline(0,0)(0,6) \pgfxyline(1,0)(1,6) \pgfxyline(2,0)(2,6) \pgfxyline(3,0)(3,6) \pgfxyline(4,0)(4,6) \pgfxyline(5,0)(5,6) \pgfxyline(6,0)(6,6) \pgfxyline(7,0)(7,6) \pgfxyline(8,0)(8,6) \pgfxyline(9,0)(9,6) \pgfxyline(10,0)(10,6) \pgfxyline(11,0)(11,6) \pgfxyline(12,0)(12,6) \end{colormixin} \pgfputat{\pgfxy(0,.25)}{\pgfbox[center,center]{0}} \pgfputat{\pgfxy(1,.25)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(2,.25)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(3,.25)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(4,.25)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(5,.25)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(6,.25)}{\pgfbox[center,center]{6}} \pgfputat{\pgfxy(7,.25)}{\pgfbox[center,center]{7}} \pgfputat{\pgfxy(8,.25)}{\pgfbox[center,center]{8}} \pgfputat{\pgfxy(9,.25)}{\pgfbox[center,center]{9}} \pgfputat{\pgfxy(10,.25)}{\pgfbox[center,center]{10}} \pgfputat{\pgfxy(11,.25)}{\pgfbox[center,center]{11}} \pgfputat{\pgfxy(12,.25)}{\pgfbox[center,center]{12}} \pgfputat{\pgfxy(0,1)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(0,2)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(0,3)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(0,4)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(0,5)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(0,6)}{\pgfbox[center,center]{6}} } \begin{colormixin}{100!white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0104}[fill]{\pgfxy(2.5,.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0105}[fill]{\pgfxy(1.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0110}[fill]{\pgfxy(3.7,.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0203}[fill]{\pgfxy(.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0206}[fill]{\pgfxy(1.3,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0207}[fill]{\pgfxy(2.1,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0405}[fill]{\pgfxy(2.3,1.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0603}[fill]{\pgfxy(1.7,4.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0703}[fill]{\pgfxy(2.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0803}[fill]{\pgfxy(2.9,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0904}[fill]{\pgfxy(3.1,.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2026}[fill]{\pgfxy(7.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2122}[fill]{\pgfxy(7.4,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2127}[fill]{\pgfxy(7.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2420}[fill]{\pgfxy(7.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2533}[fill]{\pgfxy(9.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2730}[fill]{\pgfxy(8.5,4.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2833}[fill]{\pgfxy(9.5,1.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2927}[fill]{\pgfxy(8.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2930}[fill]{\pgfxy(9,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2935}[fill]{\pgfxy(9.4,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3125}[fill]{\pgfxy(8.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3128}[fill]{\pgfxy(9.4,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3132}[fill]{\pgfxy(10.2,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3226}[fill]{\pgfxy(8.3,2.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3233}[fill]{\pgfxy(10.2,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3435}[fill]{\pgfxy(10.2,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3530}[fill]{\pgfxy(9.6,4.8)}{\pgfxy(0.2,0.2)} \pgfnodeconnline{Node0102}{Node0203} \pgfnodeconnline{Node0102}{Node0206} \pgfnodeconnline{Node0102}{Node0207} \pgfnodeconnline{Node0104}{Node0405} \pgfnodeconnline{Node0104}{Node0414} \pgfnodeconnline{Node0105}{Node0502} \pgfnodeconnline{Node0105}{Node0506} \pgfnodeconnline{Node0105}{Node0507} \pgfnodeconnline{Node0105}{Node0508} \pgfnodeconnline{Node0105}{Node0514} \pgfnodeconnline{Node0110}{Node1007} \pgfnodeconnline{Node0110}{Node1013} \pgfnodeconnline{Node0206}{Node0603} \pgfnodeconnline{Node0207}{Node0703} \pgfnodeconnline{Node0207}{Node0708} \pgfnodeconnline{Node0405}{Node0502} \pgfnodeconnline{Node0405}{Node0506} \pgfnodeconnline{Node0405}{Node0507} \pgfnodeconnline{Node0405}{Node0508} \pgfnodeconnline{Node0405}{Node0514} \pgfnodeconnline{Node0414}{Node1412} \pgfnodeconnline{Node0414}{Node1415} \pgfnodeconnline{Node0502}{Node0203} \pgfnodeconnline{Node0502}{Node0206} \pgfnodeconnline{Node0502}{Node0207} \pgfnodeconnline{Node0506}{Node0603} \pgfnodeconnline{Node0507}{Node0703} \pgfnodeconnline{Node0507}{Node0708} \pgfnodeconnline{Node0508}{Node0803} \pgfnodeconnline{Node0514}{Node1412} \pgfnodeconnline{Node0514}{Node1415} \pgfnodeconnline{Node0708}{Node0803} \pgfnodeconnline{Node0904}{Node0405} \pgfnodeconnline{Node0904}{Node0414} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0703} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1208}{Node0803} \pgfnodeconnline{Node1307}{Node0703} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1620}{Node2026} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2026}{Node2621} \pgfnodeconnline{Node2026}{Node2629} \pgfnodeconnline{Node2122}{Node2219} \pgfnodeconnline{Node2122}{Node2223} \pgfnodeconnline{Node2127}{Node2730} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2420}{Node2013} \pgfnodeconnline{Node2420}{Node2017} \pgfnodeconnline{Node2420}{Node2018} \pgfnodeconnline{Node2420}{Node2022} \pgfnodeconnline{Node2420}{Node2026} \pgfnodeconnline{Node2517}{Node1714} \pgfnodeconnline{Node2521}{Node2122} \pgfnodeconnline{Node2521}{Node2127} \pgfnodeconnline{Node2529}{Node2927} \pgfnodeconnline{Node2529}{Node2930} \pgfnodeconnline{Node2529}{Node2935} \pgfnodeconnline{Node2533}{Node3329} \pgfnodeconnline{Node2533}{Node3334} \pgfnodeconnline{Node2621}{Node2122} \pgfnodeconnline{Node2621}{Node2127} \pgfnodeconnline{Node2629}{Node2927} \pgfnodeconnline{Node2629}{Node2930} \pgfnodeconnline{Node2629}{Node2935} \pgfnodeconnline{Node2833}{Node3329} \pgfnodeconnline{Node2833}{Node3334} \pgfnodeconnline{Node2927}{Node2730} \pgfnodeconnline{Node2935}{Node3530} \pgfnodeconnline{Node3125}{Node2517} \pgfnodeconnline{Node3125}{Node2521} \pgfnodeconnline{Node3125}{Node2529} \pgfnodeconnline{Node3125}{Node2533} \pgfnodeconnline{Node3128}{Node2833} \pgfnodeconnline{Node3132}{Node3226} \pgfnodeconnline{Node3132}{Node3229} \pgfnodeconnline{Node3132}{Node3233} \pgfnodeconnline{Node3226}{Node2621} \pgfnodeconnline{Node3226}{Node2629} \pgfnodeconnline{Node3229}{Node2927} \pgfnodeconnline{Node3229}{Node2930} \pgfnodeconnline{Node3229}{Node2935} \pgfnodeconnline{Node3233}{Node3329} \pgfnodeconnline{Node3233}{Node3334} \pgfnodeconnline{Node3329}{Node2927} \pgfnodeconnline{Node3329}{Node2930} \pgfnodeconnline{Node3329}{Node2935} \pgfnodeconnline{Node3334}{Node3435} \pgfnodeconnline{Node3435}{Node3530} \end{colormixin} \color{white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.2,0.2)} \color{black} \pgfnoderect{Node0102}[stroke]{\pgfxy(.9,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0414}[stroke]{\pgfxy(3.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0502}[stroke]{\pgfxy(1.4,2.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0506}[stroke]{\pgfxy(1.8,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0507}[stroke]{\pgfxy(2.3,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0508}[stroke]{\pgfxy(2.8,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0514}[stroke]{\pgfxy(3,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1007}[stroke]{\pgfxy(3.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1307}[stroke]{\pgfxy(3.9,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1312}[stroke]{\pgfxy(4.1,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1314}[stroke]{\pgfxy(4.5,3.4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1319}[stroke]{\pgfxy(5.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1714}[stroke]{\pgfxy(6,2.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1814}[stroke]{\pgfxy(4.8,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1819}[stroke]{\pgfxy(5.8,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1822}[stroke]{\pgfxy(6.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1823}[stroke]{\pgfxy(6.2,3.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2022}[stroke]{\pgfxy(7,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2521}[stroke]{\pgfxy(7.5,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2529}[stroke]{\pgfxy(8.7,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2621}[stroke]{\pgfxy(7.9,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2629}[stroke]{\pgfxy(8.4,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3229}[stroke]{\pgfxy(9.2,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3329}[stroke]{\pgfxy(9.7,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3334}[stroke]{\pgfxy(10.4,3.1)}{\pgfxy(0.2,0.2)} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.35cm}{-.35cm}} \pgfputat{\pgfxy(6.5,2)}{\pgfbox[center,center]{$t(s_m)=i(s)$}} \pgfputat{\pgfxy(6.5,1.5)}{\pgfbox[center,center]{$=t(r)=i(r_0)$}} \pgfputat{\pgfxy(4.75,2)}{\pgfbox[center,center]{\large{$r$}}} \pgfputat{\pgfxy(4.75,4)}{\pgfbox[center,center]{\large{$s$}}} \pgfputat{\pgfxy(5.45,3.5)}{\pgfbox[center,center]{\large{$r_0$}}} \pgfputat{\pgfxy(6.7,4.75)}{\pgfbox[center,center]{\large{$r_n$}}} \pgfputat{\pgfxy(4.3,3.1)}{\pgfbox[center,center]{\large{$s_m$}}} \pgfputat{\pgfxy(2.7,1.6)}{\pgfbox[center,center]{\large{$s_0$}}} \pgfnodecircle{Node1}[fill]{\pgfxy(5,1)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(5,3)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(6,3.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(6.75,4.25)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(7.25,5)}{0.11cm} \pgfnodecircle{Node0}[fill]{\pgfxy(6.275,3.775)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(6.375,3.875)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(6.475,3.975)}{0.025cm} \pgfnodecircle{Node7}[fill]{\pgfxy(2.75,1)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(3.25,1.75)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node0}[fill]{\pgfxy(3.525,2.025)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(3.625,2.125)}{0.025cm} \pgfnodecircle{Node0}[fill]{\pgfxy(3.725,2.225)}{0.025cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node9}{Node2} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(6.5,2.2)(5.2,2.85) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Maximal antichain in relation space; b) maximal antichains in relation space are impermeable.} \label{figimpermeability} \end{figure} \vspace*{-.2cm} \refstepcounter{textlabels}\label{thmrelspaceimperm} \addtocounter{theorem}{1} \begin{theorem}\label{theoremrelimpermeable} Maximal antichains in relation space are impermeable. \end{theorem} \begin{proof} Let $M=(M,R,i,t)$ be a multidirected set with relation space $\ms{R}(M)=(R,\prec)$, and let $\sigma$ be a maximal antichain in $\ms{R}(M)$. Note that by the definition of an antichain, the past and future of $\sigma$ are disjoint. Indeed, if $r\in \ms{R}(M)$ belongs to both the past and future of $\sigma$, then by definition, there exist elements $r^-$ and $r^+$ in $\sigma$, together with chains from $r^-$ to $r$ and from $r$ to $r^+$. These chains join together to form a chain from $r^-$ to $r^+$, contradicting the hypothesis that $\sigma$ is an antichain. The proof is by contradiction. Suppose that $\sigma$ is permeable, and let $\gamma$ be a chain in $\ms{R}(M)$ permeating $\sigma$. By definition, $\gamma$ may be expressed as a sequence of elements and relations in $\ms{R}(M)$ with initial element in the past of $\sigma$ and terminal element in the future of $\sigma$. Hence, there is at least one relation $r\prec s$ in $\gamma$ with initial element $r$ in the past of $\sigma$, and terminal element $s$ in the future of $\sigma$.\footnotemark\footnotetext{As I have defined it, $\gamma$ may ``wrap around cycles in $\ms{R}(M)$," and may permeate $\sigma$ multiple times. It is possible to choose $\gamma$ with tamer properties, but all that is necessary for the argument is to isolate {\it one} relation in $\gamma$ passing from the past of $\sigma$ to the future of $\sigma$.} Since $r$ is in the past of $\sigma$, there exists a chain $r\prec r_0\prec...\prec r_n$ in $\ms{R}(M)$, with $r_n\in\sigma$. Similarly, since $s$ is in the future of $\sigma$, there exists a chain $s_0\prec...\prec s_m\prec s$ in $\ms{R}(M)$, with $s_0\in\sigma$. Now \[t(s_m)=i(s)=t(r)=i(r_0)\hspace*{.3cm} \tn{ in } \hspace*{.1cm} M,\] since $s_m\prec s$, $r\prec s$, and $r\prec r_0$ in $\ms{R}(M)$. This sequence of equalities is illustrated in figure \hyperref[figimpermeability]{\ref{figimpermeability}}b above, which shows the {\it element space viewpoint.} Hence, $s_m\prec r_0$ in $\ms{R}(M)$, so the two chains $s_0\prec...\prec s_m$ and $r_0\prec...\prec r_n$ fit together to define a composite chain $s_0\prec...\prec s_m\prec r_0\prec...\prec r_n$, which is a chain between two elements $s_0$ and $r_n$ of $\sigma$. This contradicts the hypothesis that $\sigma$ is an antichain. \end{proof} \refstepcounter{textlabels}\label{preferrelspace} Impermeability of maximal antichains in relation space is of great importance in the {\it dynamics} of directed sets, especially in the context of quantum theory. In particular, the theory of path summation over a multidirected set, which is the causal analogue of {\it path integration} over continuum spacetime, relies on impermeability in a critical way, as explained in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} below. Viewing spacetime structure in terms of relations rather than elements is much more than a philosophical preference, since some of the most important associated methods {\it do not apply in element space}. Besides impermeability, there are other reasons to prefer relation space over element space for formulating {\it spatial notions} such as Cauchy surfaces. One such reason is {\it adjacency.} In the acyclic case, any pair of relations with the same initial or terminal element form a minimal nontrivial antichain in relation space, which may be viewed as a ``building block" of spatial structure. Such relations are adjacent, in the sense that they share a common element; i.e., they form a connected subgraph. While it is {\it not} true that a maximal antichain in relation space is always connected as a subgraph, a ``high degree of connectivity" is common. For example, the antichain of relations illustrated in in figure \hyperref[causalpathsum]{\ref{causalpathsum}}b of section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} below is connected as a subgraph. Similarly, the maximal antichain of relations consisting of the ``first generation" of relations in the causal set illustrated in figure \hyperref[semiordinalspatial]{\ref{semiordinalspatial}} above has just three connected components, each corresponding to a copy of $\mathbb{Z}$ in an obvious way. This supports the vague statement made in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}} that this causal set has ``spatial size greater than $\mathbb{Z}$." \refstepcounter{textlabels}\label{analogymorphism} {\bf Analogy between $\ms{R}(M)$ and Morphism Categories.} A {\it morphism category} is a category whose objects are the morphisms of another category, and whose morphisms are {\it morphisms between morphisms}, called {\it $2$-morphisms}. An {\it augmented category} possessing objects, morphisms, and $2$-morphisms, is called a {\it $2$-category.} Let $W,X,Y,$ and $Z$ be objects of some category. Then a $2$-morphism between morphisms $f:X\rightarrow Y$ and $g:W\rightarrow Z$ is defined to be a {\it commutative square} of the form illustrated in figure \hyperref[2cat]{\ref{2cat}}a below: \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{3cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,3)(16.5,3) \pgfxyline(0,-.1)(0,3) \pgfxyline(5,-.1)(5,3) \pgfxyline(11.5,-.1)(11.5,3) \pgfxyline(16.5,-.1)(16.5,3) \pgfputat{\pgfxy(.15,2.7)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.15,2.7)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.65,2.7)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{-6cm}{.15cm}} \pgfputat{\pgfxy(7.25,1.25)}{\pgfbox[center,center]{$Y$}} \pgfputat{\pgfxy(8.25,.25)}{\pgfbox[center,center]{$X$}} \pgfputat{\pgfxy(8.25,2.25)}{\pgfbox[center,center]{$Z$}} \pgfputat{\pgfxy(9.25,1.25)}{\pgfbox[center,center]{$W$}} \pgfputat{\pgfxy(7.5,.55)}{\pgfbox[center,center]{$f$}} \pgfputat{\pgfxy(9,.55)}{\pgfbox[center,center]{$b$}} \pgfputat{\pgfxy(7.5,1.85)}{\pgfbox[center,center]{$a$}} \pgfputat{\pgfxy(9,1.85)}{\pgfbox[center,center]{$g$}} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(8,.4)(7.4,1) \pgfxyline(9.1,1.5)(8.5,2.1) \pgfxyline(8.5,.4)(9.1,1) \pgfxyline(7.4,1.5)(8,2.1) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-1.25cm}{.15cm}} \pgfputat{\pgfxy(7.25,1.25)}{\pgfbox[center,center]{$Y$}} \pgfputat{\pgfxy(8.25,.25)}{\pgfbox[center,center]{$X$}} \pgfputat{\pgfxy(8.25,2.25)}{\pgfbox[center,center]{$Z$}} \pgfputat{\pgfxy(9.25,1.25)}{\pgfbox[center,center]{$Y$}} \pgfputat{\pgfxy(11.25,1.25)}{\pgfbox[center,center]{$Y$}} \pgfputat{\pgfxy(11.25,.25)}{\pgfbox[center,center]{$X$}} \pgfputat{\pgfxy(11.25,2.25)}{\pgfbox[center,center]{$Z$}} \pgfputat{\pgfxy(7.5,.55)}{\pgfbox[center,center]{$f$}} \pgfputat{\pgfxy(9,.55)}{\pgfbox[center,center]{$f$}} \pgfputat{\pgfxy(7.5,1.85)}{\pgfbox[center,center]{$g$}} \pgfputat{\pgfxy(9,1.85)}{\pgfbox[center,center]{$g$}} \pgfputat{\pgfxy(11.5,.75)}{\pgfbox[center,center]{$f$}} \pgfputat{\pgfxy(11.5,1.75)}{\pgfbox[center,center]{$g$}} \begin{pgfscope} \pgfsetlinewidth{1.5pt} \pgfsetstartarrow{\pgfarrowtriangle{4pt}} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(9.9,1.25)(10.5,1.25) \end{pgfscope} \pgfsetendarrow{\pgfarrowlargepointed{3pt}} \pgfxyline(8,.4)(7.4,1) \pgfxyline(9.1,1.5)(8.5,2.1) \pgfxyline(11.25,.5)(11.25,1) \pgfxyline(11.25,1.5)(11.25,2) \pgfxyline(8.5,.4)(9.1,1) \pgfxyline(7.4,1.5)(8,2.1) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{.15cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(10,.25)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(10,1.25)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(10,2.25)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfputat{\pgfxy(10.25,.25)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(9.75,.75)}{\pgfbox[center,center]{$r$}} \pgfputat{\pgfxy(10.25,1.25)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(9.75,1.75)}{\pgfbox[center,center]{$s$}} \pgfputat{\pgfxy(10.25,2.25)}{\pgfbox[center,center]{$z$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) A $2$-morphism; b) ``commutativity" without composition; c) $2$-chain analogue. } \label{2cat} \end{figure} \vspace*{-.5cm} A basic requirement for commutativity to make sense is the {\it composition axiom} of category theory, which guarantees that the ordered pairs of morphisms $(f,a)$ and $(b,g)$ compose to define morphisms $a\circ f:X\rightarrow Z$ and $g\circ b:X\rightarrow Z$. As discussed in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}} above, the composition axiom is roughly analogous to the causal set axiom of transitivity. However, there remains at least one circumstance under which the square may reasonably be regarded as ``commutative" even without the composition axiom: namely, when $Y=W$, $b=f$, and $a=g$. In this case, the square is equivalent to the sequence of morphisms $X\overset{f}{\rightarrow}Y\overset{g}{\rightarrow}Z$, as illustrated in figure \hyperref[2cat]{\ref{2cat}}b above. This sequence, in turn, is analogous to a $2$-chain in a multidirected set $M=(M,R,i,t)$, consisting of relations $r$ and $s$, with initial and terminal elements $i(r)=x$, $t(r)=i(s)=y$, and $t(s)=z$, as illustrated in figure \hyperref[2cat]{\ref{2cat}}c. By the definition of relation space, such a $2$-chain corresponds to a relation $r\prec s$ in $\ms{R}(M)$. One way to interpret this analogy is to view the relation space $\ms{R}(M)$ over a multidirected set $M$, together with its induced relation $\prec$, as a {\it generalized morphism category.} More holistically, the original multidirected set $M$ may be viewed as a {\it generalized $2$-category,} whose ``objects" are its elements, whose ``morphisms" are its relations, and whose ``$2$-morphisms" are its $2$-chains. This provides an example of a phenomenon mentioned in section \hyperref[settheoretic]{\ref{settheoretic}}, in the context of Isham's topos theoretic approach to physics: causal theory naturally presents structures that are {\it not} categories but that play similar roles. Unmodified appropriation of category-theoretic notions can do more harm than good in this context. For example, the na\"{\i}vest causal analogue of a $2$-morphism is a {\it causal diamond,} but this analogy leads in the wrong direction in the context of relation space. Many other such ``category-like" structures appear in succeeding sections. \refstepcounter{textlabels}\label{starrevisited} {\bf The Star Model Revisited.} The star model $M_\star$ of a multidirected set $M=(M,R,i,t)$, used in section \hyperref[sectioninterval]{\ref{sectioninterval}} above to facilitate the study of local finiteness, may be elevated to a multidirected set $(M_\star,R_\star,i_\star, t_\star)$ in its own right, by means of the induced relation $\prec$ on the relation space $\ms{R}(M)$ over $M$. As a set, the star model $M_\star$ is just the union $M\cup R$ of (the underlying set of) $M$ and its relation set $R$. This set may be endowed with multidirected structure by taking its relation set $R_\star$ to be the union $R\cup\prec$, where elements of $R$ relate elements of $M$, and elements of $\prec$ relate elements of $R$. Initial and terminal element maps $i_\star$ and $t_\star$ may then be defined in the obvious way, by combining the maps $i$ and $t$ with corresponding maps from $\prec$ to $R$, sending each induced relation $r\prec s$ to its initial and terminal elements $r$ and $s$ in $R$, respectively. This multidirected version of the star model combines the ``lowest-order graded pieces" of the {\it causal concatenation spaces} over $M$ and $\ms{R}(M)$, described in more detail in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. The star model of a directed set is directed, rather than multidirected, since both $R$ and $\prec$ define binary relations in this case. \refstepcounter{textlabels}\label{inducedtwoelt} {\bf An Induced Binary Relation on Two-Element Subsets.} Alternatively, the induced relation $\prec$ on the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$ may be used to construct a binary relation on the set $M_2$ of {\it two-element subsets} of $M$. This relation is defined by taking subsets $\{x,y\}$ and $\{u,v\}$ to be related if and only if there is a relation $x\prec y$ or $y\prec x$, a relation $u\prec v$ or $v\prec u$, and the terminal element of the first relation is the initial element of the second. This construction generally does not preserve information, even in the interior of $M$, since merely specifying a pair of elements of $M$ neither determines which element is initial and which is terminal, nor distinguishes among different relations having the same initial and terminal elements. Further, information about reflexive relations is lost entirely. Less information loss occurs in the special case of an acyclic directed set $A=(A,\prec)$, since each two-element subset $\{x,y\}$ of $A$ corresponds to at most one relation in $A$ involving $x$ and $y$, and since no reflexive relations exist in this case. For these reasons, the induced relation on two-element subsets may be viewed as a trivial extension of the induced relation on the relation space $\ms{R}(A)$. Most of the multidirected sets used as examples in this paper induce very {\it sparse} relations on their sets of two-element subsets, since most of these subsets do not consist of the initial and terminal elements of a relation. Figure \hyperref[twoelement]{\ref{twoelement}} below illustrates this sparseness, using the familiar $35$-element multidirected set $M$ appearing repeatedly in previous examples. Figure \hyperref[twoelement]{\ref{twoelement}}a shows that a ``typical" two-element subset of $M$ does not correspond to relation; indeed, $M$ has $\binom{35}{2}=595$ two-element subsets, but only $75$ relations. Hence, most of the elements of $M_2$ are isolated in the induced relation on $M_2$. Figure \hyperref[twoelement]{\ref{twoelement}}b shows $M_2$ with its induced relation; isolated elements of $M_2$ are represented by small square nodes, while non-isolated elements are represented by large square nodes. Since $M$ is in fact an acyclic directed set in this example, the induced relation on $M_2$ may be viewed as a trivial extension of the induced relation on $\ms{R}(M)$. This may be seen by comparing figure \hyperref[twoelement]{\ref{twoelement}}b to figure \hyperref[relationspace]{\ref{relationspace}}b above. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.35cm} \begin{pgfmagnify}{1.03}{1.03} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \begin{pgftranslate}{\pgfpoint{1cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.5,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \color{white} \pgfmoveto{\pgfxy(1,3.5)} \pgflineto{\pgfxy(3.5,3.5)} \pgflineto{\pgfxy(3.5,4.25)} \pgflineto{\pgfxy(1,4.25)} \pgflineto{\pgfxy(1,3.5)} \pgffill \color{black} \pgfmoveto{\pgfxy(1,3.5)} \pgflineto{\pgfxy(3.5,3.5)} \pgflineto{\pgfxy(3.5,4.25)} \pgflineto{\pgfxy(1,4.25)} \pgflineto{\pgfxy(1,3.5)} \pgfstroke \pgfputat{\pgfxy(1.25,3.875)}{\pgfbox[left,center]{not related}} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2,3.5)(1.2,1.7) \pgfxyline(2.5,3.5)(3.8,2.6) \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.75cm}{-.2cm}} \begin{pgfmagnify}{.9}{.9} \comment{ \begin{colormixin}{50!white} \pgfxyline(0,0)(12,0) \pgfxyline(0,.5)(12,.5) \pgfxyline(0,1)(12,1) \pgfxyline(0,1.5)(12,1.5) \pgfxyline(0,2)(12,2) \pgfxyline(0,2.5)(12,2.5) \pgfxyline(0,3)(12,3) \pgfxyline(0,3.5)(12,3.5) \pgfxyline(0,4)(12,4) \pgfxyline(0,4.5)(12,4.5) \pgfxyline(0,5)(12,5) \pgfxyline(0,5.5)(12,5.5) \pgfxyline(0,6)(12,6) \pgfxyline(0,0)(0,6) \pgfxyline(1,0)(1,6) \pgfxyline(2,0)(2,6) \pgfxyline(3,0)(3,6) \pgfxyline(4,0)(4,6) \pgfxyline(5,0)(5,6) \pgfxyline(6,0)(6,6) \pgfxyline(7,0)(7,6) \pgfxyline(8,0)(8,6) \pgfxyline(9,0)(9,6) \pgfxyline(10,0)(10,6) \pgfxyline(11,0)(11,6) \pgfxyline(12,0)(12,6) \end{colormixin} \pgfputat{\pgfxy(0,.25)}{\pgfbox[center,center]{0}} \pgfputat{\pgfxy(1,.25)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(2,.25)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(3,.25)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(4,.25)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(5,.25)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(6,.25)}{\pgfbox[center,center]{6}} \pgfputat{\pgfxy(7,.25)}{\pgfbox[center,center]{7}} \pgfputat{\pgfxy(8,.25)}{\pgfbox[center,center]{8}} \pgfputat{\pgfxy(9,.25)}{\pgfbox[center,center]{9}} \pgfputat{\pgfxy(10,.25)}{\pgfbox[center,center]{10}} \pgfputat{\pgfxy(11,.25)}{\pgfbox[center,center]{11}} \pgfputat{\pgfxy(12,.25)}{\pgfbox[center,center]{12}} \pgfputat{\pgfxy(0,1)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(0,2)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(0,3)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(0,4)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(0,5)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(0,6)}{\pgfbox[center,center]{6}} } \begin{colormixin}{100!white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0104}[fill]{\pgfxy(2.5,.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0105}[fill]{\pgfxy(1.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0110}[fill]{\pgfxy(3.7,.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0203}[fill]{\pgfxy(.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0206}[fill]{\pgfxy(1.3,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0207}[fill]{\pgfxy(2.1,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0405}[fill]{\pgfxy(2.3,1.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0603}[fill]{\pgfxy(1.7,4.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0703}[fill]{\pgfxy(2.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0803}[fill]{\pgfxy(2.9,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0904}[fill]{\pgfxy(3.1,.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2026}[fill]{\pgfxy(7.6,2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2122}[fill]{\pgfxy(7.4,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2127}[fill]{\pgfxy(7.8,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2420}[fill]{\pgfxy(7.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2533}[fill]{\pgfxy(9.1,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2730}[fill]{\pgfxy(8.5,4.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2833}[fill]{\pgfxy(9.5,1.9)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2927}[fill]{\pgfxy(8.5,3.7)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2930}[fill]{\pgfxy(9,4.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node2935}[fill]{\pgfxy(9.4,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3125}[fill]{\pgfxy(8.4,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3128}[fill]{\pgfxy(9.4,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3132}[fill]{\pgfxy(10.2,1.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3226}[fill]{\pgfxy(8.3,2.2)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3233}[fill]{\pgfxy(10.2,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3435}[fill]{\pgfxy(10.2,4)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node3530}[fill]{\pgfxy(9.6,4.8)}{\pgfxy(0.2,0.2)} \pgfnodeconnline{Node0102}{Node0203} \pgfnodeconnline{Node0102}{Node0206} \pgfnodeconnline{Node0102}{Node0207} \pgfnodeconnline{Node0104}{Node0405} \pgfnodeconnline{Node0104}{Node0414} \pgfnodeconnline{Node0105}{Node0502} \pgfnodeconnline{Node0105}{Node0506} \pgfnodeconnline{Node0105}{Node0507} \pgfnodeconnline{Node0105}{Node0508} \pgfnodeconnline{Node0105}{Node0514} \pgfnodeconnline{Node0110}{Node1007} \pgfnodeconnline{Node0110}{Node1013} \pgfnodeconnline{Node0206}{Node0603} \pgfnodeconnline{Node0207}{Node0703} \pgfnodeconnline{Node0207}{Node0708} \pgfnodeconnline{Node0405}{Node0502} \pgfnodeconnline{Node0405}{Node0506} \pgfnodeconnline{Node0405}{Node0507} \pgfnodeconnline{Node0405}{Node0508} \pgfnodeconnline{Node0405}{Node0514} \pgfnodeconnline{Node0414}{Node1412} \pgfnodeconnline{Node0414}{Node1415} \pgfnodeconnline{Node0502}{Node0203} \pgfnodeconnline{Node0502}{Node0206} \pgfnodeconnline{Node0502}{Node0207} \pgfnodeconnline{Node0506}{Node0603} \pgfnodeconnline{Node0507}{Node0703} \pgfnodeconnline{Node0507}{Node0708} \pgfnodeconnline{Node0508}{Node0803} \pgfnodeconnline{Node0514}{Node1412} \pgfnodeconnline{Node0514}{Node1415} \pgfnodeconnline{Node0708}{Node0803} \pgfnodeconnline{Node0904}{Node0405} \pgfnodeconnline{Node0904}{Node0414} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0703} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1208}{Node0803} \pgfnodeconnline{Node1307}{Node0703} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1620}{Node2026} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2026}{Node2621} \pgfnodeconnline{Node2026}{Node2629} \pgfnodeconnline{Node2122}{Node2219} \pgfnodeconnline{Node2122}{Node2223} \pgfnodeconnline{Node2127}{Node2730} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2420}{Node2013} \pgfnodeconnline{Node2420}{Node2017} \pgfnodeconnline{Node2420}{Node2018} \pgfnodeconnline{Node2420}{Node2022} \pgfnodeconnline{Node2420}{Node2026} \pgfnodeconnline{Node2517}{Node1714} \pgfnodeconnline{Node2521}{Node2122} \pgfnodeconnline{Node2521}{Node2127} \pgfnodeconnline{Node2529}{Node2927} \pgfnodeconnline{Node2529}{Node2930} \pgfnodeconnline{Node2529}{Node2935} \pgfnodeconnline{Node2533}{Node3329} \pgfnodeconnline{Node2533}{Node3334} \pgfnodeconnline{Node2621}{Node2122} \pgfnodeconnline{Node2621}{Node2127} \pgfnodeconnline{Node2629}{Node2927} \pgfnodeconnline{Node2629}{Node2930} \pgfnodeconnline{Node2629}{Node2935} \pgfnodeconnline{Node2833}{Node3329} \pgfnodeconnline{Node2833}{Node3334} \pgfnodeconnline{Node2927}{Node2730} \pgfnodeconnline{Node2935}{Node3530} \pgfnodeconnline{Node3125}{Node2517} \pgfnodeconnline{Node3125}{Node2521} \pgfnodeconnline{Node3125}{Node2529} \pgfnodeconnline{Node3125}{Node2533} \pgfnodeconnline{Node3128}{Node2833} \pgfnodeconnline{Node3132}{Node3226} \pgfnodeconnline{Node3132}{Node3229} \pgfnodeconnline{Node3132}{Node3233} \pgfnodeconnline{Node3226}{Node2621} \pgfnodeconnline{Node3226}{Node2629} \pgfnodeconnline{Node3229}{Node2927} \pgfnodeconnline{Node3229}{Node2930} \pgfnodeconnline{Node3229}{Node2935} \pgfnodeconnline{Node3233}{Node3329} \pgfnodeconnline{Node3233}{Node3334} \pgfnodeconnline{Node3329}{Node2927} \pgfnodeconnline{Node3329}{Node2930} \pgfnodeconnline{Node3329}{Node2935} \pgfnodeconnline{Node3334}{Node3435} \pgfnodeconnline{Node3435}{Node3530} \end{colormixin} \begin{colormixin}{100!white} \begin{pgftranslate}{\pgfpoint{1cm}{0cm}} \pgfnoderect{Node1}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node2}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node3}[fill]{\pgfxy(1.8,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node4}[fill]{\pgfxy(1.5,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node5}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node6}[fill]{\pgfxy(1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node7}[fill]{\pgfxy(1.2,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node8}[fill]{\pgfxy(1.25,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node9}[fill]{\pgfxy(1.75,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node10}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2cm}{0cm}} \pgfnoderect{Node11}[fill]{\pgfxy(1.3,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node12}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node13}[fill]{\pgfxy(1.7,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node14}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node15}[fill]{\pgfxy(1.4,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node16}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node17}[fill]{\pgfxy(1.35,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node18}[fill]{\pgfxy(1.35,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node19}[fill]{\pgfxy(1.75,1.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node20}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3cm}{0cm}} \pgfnoderect{Node21}[fill]{\pgfxy(1.1,.21)}{\pgfxy(.08,.08)} \pgfnoderect{Node22}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node23}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node24}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node25}[fill]{\pgfxy(1.8,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node26}[fill]{\pgfxy(1.9,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node27}[fill]{\pgfxy(1.5,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node28}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node29}[fill]{\pgfxy(1.3,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node30}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{0cm}} \pgfnoderect{Node31}[fill]{\pgfxy(1.7,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node32}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node33}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node34}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node35}[fill]{\pgfxy(1.75,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node36}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node37}[fill]{\pgfxy(1.35,.63)}{\pgfxy(.08,.08)} \pgfnoderect{Node38}[fill]{\pgfxy(1.35,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node39}[fill]{\pgfxy(1.75,1.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node40}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5cm}{0cm}} \pgfnoderect{Node41}[fill]{\pgfxy(1.1,.21)}{\pgfxy(.08,.08)} \pgfnoderect{Node42}[fill]{\pgfxy(1.3,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node43}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node44}[fill]{\pgfxy(1,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node45}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node46}[fill]{\pgfxy(1.85,.56)}{\pgfxy(.08,.08)} \pgfnoderect{Node47}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node48}[fill]{\pgfxy(1.35,.63)}{\pgfxy(.08,.08)} \pgfnoderect{Node49}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node50}[fill]{\pgfxy(1.55,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6cm}{0cm}} \pgfnoderect{Node51}[fill]{\pgfxy(1.1,.21)}{\pgfxy(.08,.08)} \pgfnoderect{Node52}[fill]{\pgfxy(1.7,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node53}[fill]{\pgfxy(1.85,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node54}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node55}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node56}[fill]{\pgfxy(1.8,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node57}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node58}[fill]{\pgfxy(1.1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node59}[fill]{\pgfxy(1.65,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node60}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7cm}{0cm}} \pgfnoderect{Node61}[fill]{\pgfxy(1.3,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node62}[fill]{\pgfxy(1.85,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node63}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node64}[fill]{\pgfxy(1.4,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node65}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node66}[fill]{\pgfxy(1.35,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node67}[fill]{\pgfxy(1.75,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node68}[fill]{\pgfxy(1.1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node69}[fill]{\pgfxy(1.65,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node70}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8cm}{0cm}} \pgfnoderect{Node71}[fill]{\pgfxy(1.1,.21)}{\pgfxy(.08,.08)} \pgfnoderect{Node72}[fill]{\pgfxy(1.3,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node73}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node74}[fill]{\pgfxy(1,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node75}[fill]{\pgfxy(1.4,.42)}{\pgfxy(.08,.08)} \pgfnoderect{Node76}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node77}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node78}[fill]{\pgfxy(1.35,.63)}{\pgfxy(.08,.08)} \pgfnoderect{Node79}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node80}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{9cm}{0cm}} \pgfnoderect{Node81}[fill]{\pgfxy(1.1,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node82}[fill]{\pgfxy(1.3,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node83}[fill]{\pgfxy(1.65,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node84}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node85}[fill]{\pgfxy(1.75,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node86}[fill]{\pgfxy(1.65,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node87}[fill]{\pgfxy(1.5,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node88}[fill]{\pgfxy(1.55,.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node89}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node90}[fill]{\pgfxy(1.65,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.65cm}{0cm}} \pgfnoderect{Node91}[fill]{\pgfxy(1.1,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node92}[fill]{\pgfxy(1.35,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node93}[fill]{\pgfxy(1.45,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node94}[fill]{\pgfxy(1.15,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node95}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node96}[fill]{\pgfxy(1.35,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node97}[fill]{\pgfxy(1.2,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node98}[fill]{\pgfxy(1.15,.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node99}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node100}[fill]{\pgfxy(1.15,1.25)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{.9cm}} \pgfnoderect{Node101}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node102}[fill]{\pgfxy(1.75,.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node103}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node104}[fill]{\pgfxy(1.4,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node105}[fill]{\pgfxy(1.9,.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node106}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node107}[fill]{\pgfxy(1.35,.62)}{\pgfxy(.08,.08)} \pgfnoderect{Node108}[fill]{\pgfxy(1.75,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node109}[fill]{\pgfxy(1.1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node110}[fill]{\pgfxy(1.75,.95)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.95cm}{.85cm}} \pgfnoderect{Node111}[fill]{\pgfxy(1.3,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node112}[fill]{\pgfxy(1.7,.32)}{\pgfxy(.08,.08)} \pgfnoderect{Node113}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node114}[fill]{\pgfxy(1,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node115}[fill]{\pgfxy(1.85,.56)}{\pgfxy(.08,.08)} \pgfnoderect{Node116}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node117}[fill]{\pgfxy(1.65,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node118}[fill]{\pgfxy(1.1,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node119}[fill]{\pgfxy(1.55,.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node120}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.3cm}{0cm}} \pgfnoderect{Node121}[fill]{\pgfxy(.1,1.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node122}[fill]{\pgfxy(.1,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node123}[fill]{\pgfxy(.1,2.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node124}[fill]{\pgfxy(.1,2.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node125}[fill]{\pgfxy(.1,3.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node126}[fill]{\pgfxy(.11,3.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node127}[fill]{\pgfxy(.1,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node128}[fill]{\pgfxy(.1,4.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node129}[fill]{\pgfxy(.1,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node130}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.5cm}{-.1cm}} \pgfnoderect{Node131}[fill]{\pgfxy(.1,1.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node132}[fill]{\pgfxy(.1,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node133}[fill]{\pgfxy(.1,2.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node134}[fill]{\pgfxy(0,2.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node135}[fill]{\pgfxy(.1,3.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node136}[fill]{\pgfxy(.1,3.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node137}[fill]{\pgfxy(.1,4.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node138}[fill]{\pgfxy(0,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node139}[fill]{\pgfxy(.1,4.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node140}[fill]{\pgfxy(.2,5.4)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.69cm}{.1cm}} \pgfnoderect{Node141}[fill]{\pgfxy(.12,1.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node142}[fill]{\pgfxy(.1,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node143}[fill]{\pgfxy(.1,2.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node144}[fill]{\pgfxy(0,2.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node145}[fill]{\pgfxy(.05,3.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node146}[fill]{\pgfxy(0,3.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node147}[fill]{\pgfxy(0,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node148}[fill]{\pgfxy(.2,5.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node149}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node150}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.86cm}{0cm}} \pgfnoderect{Node151}[fill]{\pgfxy(.1,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node152}[fill]{\pgfxy(.15,1.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node153}[fill]{\pgfxy(.15,2.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node154}[fill]{\pgfxy(.1,2.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node155}[fill]{\pgfxy(.14,3.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node156}[fill]{\pgfxy(.12,4.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node157}[fill]{\pgfxy(.2,4.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node158}[fill]{\pgfxy(.1,5.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node159}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node160}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node161}[fill]{\pgfxy(1.1,.21)}{\pgfxy(.08,.08)} \pgfnoderect{Node162}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node163}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node164}[fill]{\pgfxy(1.4,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node165}[fill]{\pgfxy(1.8,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node166}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node167}[fill]{\pgfxy(1.5,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node168}[fill]{\pgfxy(1.5,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node169}[fill]{\pgfxy(1.65,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node170}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.05cm}{0cm}} \pgfnoderect{Node171}[fill]{\pgfxy(.1,1.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node172}[fill]{\pgfxy(.1,2.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node173}[fill]{\pgfxy(0,3)}{\pgfxy(.08,.08)} \pgfnoderect{Node174}[fill]{\pgfxy(0,3.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node175}[fill]{\pgfxy(.1,4)}{\pgfxy(.08,.08)} \pgfnoderect{Node176}[fill]{\pgfxy(.1,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node177}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node178}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node179}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node180}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.25cm}{.06cm}} \pgfnoderect{Node181}[fill]{\pgfxy(.2,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node182}[fill]{\pgfxy(.1,2.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node183}[fill]{\pgfxy(.2,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node184}[fill]{\pgfxy(.0,3.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node185}[fill]{\pgfxy(.1,4)}{\pgfxy(.08,.08)} \pgfnoderect{Node186}[fill]{\pgfxy(.1,4.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node187}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node188}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node189}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node190}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.44cm}{-.04cm}} \pgfnoderect{Node191}[fill]{\pgfxy(.1,2.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node192}[fill]{\pgfxy(0,3)}{\pgfxy(.08,.08)} \pgfnoderect{Node193}[fill]{\pgfxy(.1,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node194}[fill]{\pgfxy(.05,3.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node195}[fill]{\pgfxy(.15,4)}{\pgfxy(.08,.08)} \pgfnoderect{Node196}[fill]{\pgfxy(.1,4.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node197}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node198}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node199}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node200}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.64cm}{0cm}} \pgfnoderect{Node201}[fill]{\pgfxy(.1,2.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node202}[fill]{\pgfxy(.2,4)}{\pgfxy(.08,.08)} \pgfnoderect{Node203}[fill]{\pgfxy(.25,4.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node204}[fill]{\pgfxy(.3,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node205}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node206}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node207}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node208}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node209}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node210}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.83cm}{.06cm}} \pgfnoderect{Node211}[fill]{\pgfxy(.2,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node212}[fill]{\pgfxy(0,2.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node213}[fill]{\pgfxy(.2,3.26)}{\pgfxy(.08,.08)} \pgfnoderect{Node214}[fill]{\pgfxy(.2,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node215}[fill]{\pgfxy(.3,4.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node216}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node217}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node218}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node219}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node220}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.09cm}{-.01cm}} \pgfnoderect{Node221}[fill]{\pgfxy(.1,2.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node222}[fill]{\pgfxy(0,2.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node223}[fill]{\pgfxy(.05,3.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node224}[fill]{\pgfxy(.25,4.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node225}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node226}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node227}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node228}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node229}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node230}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.3cm}{.06cm}} \pgfnoderect{Node231}[fill]{\pgfxy(.1,2.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node232}[fill]{\pgfxy(.2,3.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node233}[fill]{\pgfxy(.3,4.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node234}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node235}[fill]{\pgfxy(.1,5)}{\pgfxy(.08,.08)} \pgfnoderect{Node236}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node237}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node238}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node239}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node240}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.5cm}{0cm}} \pgfnoderect{Node241}[fill]{\pgfxy(.1,2.05)}{\pgfxy(.08,.08)} \pgfnoderect{Node242}[fill]{\pgfxy(0,2.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node243}[fill]{\pgfxy(.2,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node244}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node245}[fill]{\pgfxy(.1,5)}{\pgfxy(.08,.08)} \pgfnoderect{Node246}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node247}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node248}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node249}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node250}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.82cm}{.05cm}} \pgfnoderect{Node251}[fill]{\pgfxy(.15,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node252}[fill]{\pgfxy(.2,2.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node253}[fill]{\pgfxy(.25,3.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node254}[fill]{\pgfxy(.1,4.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node255}[fill]{\pgfxy(.15,4.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node256}[fill]{\pgfxy(.2,4.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node257}[fill]{\pgfxy(.25,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node258}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node259}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node260}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.15cm}{0cm}} \pgfnoderect{Node261}[fill]{\pgfxy(-.2,1.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node262}[fill]{\pgfxy(.1,1.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node263}[fill]{\pgfxy(.1,2.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node264}[fill]{\pgfxy(.05,3.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node265}[fill]{\pgfxy(.1,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node266}[fill]{\pgfxy(.1,4.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node267}[fill]{\pgfxy(.2,5.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node268}[fill]{\pgfxy(.2,5.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node269}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node270}[fill]{\pgfxy(.1,5.9)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node271}[fill]{\pgfxy(3.5,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node272}[fill]{\pgfxy(3.45,5.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node273}[fill]{\pgfxy(3.5,5.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node274}[fill]{\pgfxy(3.5,5)}{\pgfxy(.08,.08)} \pgfnoderect{Node275}[fill]{\pgfxy(3.5,4.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node276}[fill]{\pgfxy(3.6,4.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node277}[fill]{\pgfxy(3.65,4.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node278}[fill]{\pgfxy(3.5,4.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node279}[fill]{\pgfxy(3.5,3.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node280}[fill]{\pgfxy(3.6,3.7)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node281}[fill]{\pgfxy(3.5,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node282}[fill]{\pgfxy(3.55,2.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node283}[fill]{\pgfxy(3.65,2.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node284}[fill]{\pgfxy(3.9,2.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node285}[fill]{\pgfxy(3.5,1.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node286}[fill]{\pgfxy(3.5,1.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node287}[fill]{\pgfxy(3.5,1.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node288}[fill]{\pgfxy(3.3,1.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node289}[fill]{\pgfxy(3.4,1.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node290}[fill]{\pgfxy(4,1.1)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4cm}{5cm}} \pgfnoderect{Node291}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node292}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node293}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node294}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node295}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node296}[fill]{\pgfxy(1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node297}[fill]{\pgfxy(1.2,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node298}[fill]{\pgfxy(1.25,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node299}[fill]{\pgfxy(1.75,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node300}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.1cm}{5cm}} \pgfnoderect{Node301}[fill]{\pgfxy(1.2,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node302}[fill]{\pgfxy(1.45,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node303}[fill]{\pgfxy(1.15,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node304}[fill]{\pgfxy(1.35,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node305}[fill]{\pgfxy(1.15,.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node306}[fill]{\pgfxy(1,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node307}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node308}[fill]{\pgfxy(1.35,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node309}[fill]{\pgfxy(1.15,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node310}[fill]{\pgfxy(1.45,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.9cm}{5cm}} \pgfnoderect{Node311}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node312}[fill]{\pgfxy(1.7,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node313}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node314}[fill]{\pgfxy(1.8,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node315}[fill]{\pgfxy(1.35,.62)}{\pgfxy(.08,.08)} \pgfnoderect{Node316}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node317}[fill]{\pgfxy(1.2,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node318}[fill]{\pgfxy(1.25,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node319}[fill]{\pgfxy(1.75,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node320}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.9cm}{5.1cm}} \pgfnoderect{Node321}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node322}[fill]{\pgfxy(1.85,.26)}{\pgfxy(.08,.08)} \pgfnoderect{Node323}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node324}[fill]{\pgfxy(1.6,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node325}[fill]{\pgfxy(1.9,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node326}[fill]{\pgfxy(1.35,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node327}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node328}[fill]{\pgfxy(1.3,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node329}[fill]{\pgfxy(1.65,.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node330}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7.9cm}{4.9cm}} \pgfnoderect{Node331}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node332}[fill]{\pgfxy(1.85,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node333}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node334}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node335}[fill]{\pgfxy(1.9,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node336}[fill]{\pgfxy(1.35,.62)}{\pgfxy(.08,.08)} \pgfnoderect{Node337}[fill]{\pgfxy(1.75,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node338}[fill]{\pgfxy(1.2,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node339}[fill]{\pgfxy(1.75,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node340}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.85cm}{5cm}} \pgfnoderect{Node341}[fill]{\pgfxy(1.3,.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node342}[fill]{\pgfxy(1.85,.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node343}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node344}[fill]{\pgfxy(1.25,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node345}[fill]{\pgfxy(1.9,.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node346}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node347}[fill]{\pgfxy(1.35,.62)}{\pgfxy(.08,.08)} \pgfnoderect{Node348}[fill]{\pgfxy(1,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node349}[fill]{\pgfxy(1.2,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node350}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node351}[fill]{\pgfxy(3.75,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node352}[fill]{\pgfxy(4.05,5.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node353}[fill]{\pgfxy(4.2,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node354}[fill]{\pgfxy(4.7,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node355}[fill]{\pgfxy(4.85,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node356}[fill]{\pgfxy(5.5,5.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node357}[fill]{\pgfxy(5.95,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node358}[fill]{\pgfxy(6.7,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node359}[fill]{\pgfxy(6.75,5.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node360}[fill]{\pgfxy(7.35,5.75)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node361}[fill]{\pgfxy(7.9,5.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node362}[fill]{\pgfxy(9.5,5.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node363}[fill]{\pgfxy(10.4,5.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node364}[fill]{\pgfxy(3.85,5.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node365}[fill]{\pgfxy(3.9,5.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node366}[fill]{\pgfxy(3.85,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node367}[fill]{\pgfxy(3.95,4.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node368}[fill]{\pgfxy(3.85,3.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node369}[fill]{\pgfxy(3.7,3.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node370}[fill]{\pgfxy(4.05,3.7)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.3cm}{0cm}} \pgfnoderect{Node371}[fill]{\pgfxy(.2,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node372}[fill]{\pgfxy(.2,1.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node373}[fill]{\pgfxy(.3,2.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node374}[fill]{\pgfxy(.1,2.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node375}[fill]{\pgfxy(.15,2.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node376}[fill]{\pgfxy(.2,3.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node377}[fill]{\pgfxy(.4,3.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node378}[fill]{\pgfxy(.1,3.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node379}[fill]{\pgfxy(.15,3.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node380}[fill]{\pgfxy(.2,4)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.3cm}{0cm}} \pgfnoderect{Node381}[fill]{\pgfxy(.35,4.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node382}[fill]{\pgfxy(.2,4.25)}{\pgfxy(.08,.08)} \pgfnoderect{Node383}[fill]{\pgfxy(.5,4.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node384}[fill]{\pgfxy(.12,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node385}[fill]{\pgfxy(.1,5)}{\pgfxy(.08,.08)} \pgfnoderect{Node386}[fill]{\pgfxy(.2,5.15)}{\pgfxy(.08,.08)} \pgfnoderect{Node387}[fill]{\pgfxy(.2,5.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node388}[fill]{\pgfxy(.1,5.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node389}[fill]{\pgfxy(.2,5.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node390}[fill]{\pgfxy(.2,5.95)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node391}[fill]{\pgfxy(5,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node392}[fill]{\pgfxy(4.95,1.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node393}[fill]{\pgfxy(5,3)}{\pgfxy(.08,.08)} \pgfnoderect{Node394}[fill]{\pgfxy(4.9,3.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node395}[fill]{\pgfxy(5.8,3.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node396}[fill]{\pgfxy(5.9,3.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node397}[fill]{\pgfxy(6,4.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node398}[fill]{\pgfxy(6.1,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node399}[fill]{\pgfxy(6,4.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node400}[fill]{\pgfxy(6.1,4.2)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node401}[fill]{\pgfxy(6.3,4.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node402}[fill]{\pgfxy(6.4,4.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node403}[fill]{\pgfxy(4.9,4.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node404}[fill]{\pgfxy(6.6,3.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node405}[fill]{\pgfxy(6.7,3.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node406}[fill]{\pgfxy(6.7,3)}{\pgfxy(.08,.08)} \pgfnoderect{Node407}[fill]{\pgfxy(6.7,3.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node408}[fill]{\pgfxy(6.4,3.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node409}[fill]{\pgfxy(6.4,3.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node410}[fill]{\pgfxy(6.2,3.3)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node411}[fill]{\pgfxy(5.8,1.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node412}[fill]{\pgfxy(5.8,1.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node413}[fill]{\pgfxy(6.1,1.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node414}[fill]{\pgfxy(6.2,1.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node415}[fill]{\pgfxy(6.35,1.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node416}[fill]{\pgfxy(6.45,1.05)}{\pgfxy(.08,.08)} \pgfnoderect{Node417}[fill]{\pgfxy(6.45,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node418}[fill]{\pgfxy(6.45,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node419}[fill]{\pgfxy(6.6,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node420}[fill]{\pgfxy(6.7,2.1)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node421}[fill]{\pgfxy(6.7,2.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node422}[fill]{\pgfxy(6.8,2.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node423}[fill]{\pgfxy(6.65,2.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node424}[fill]{\pgfxy(6.75,2.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node425}[fill]{\pgfxy(6.85,2.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node426}[fill]{\pgfxy(7,2.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node427}[fill]{\pgfxy(7,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node428}[fill]{\pgfxy(6.9,1.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node429}[fill]{\pgfxy(7.1,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node430}[fill]{\pgfxy(7.3,2.6)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.1cm}{2cm}} \pgfnoderect{Node431}[fill]{\pgfxy(1.2,0)}{\pgfxy(.08,.08)} \pgfnoderect{Node432}[fill]{\pgfxy(1.7,.42)}{\pgfxy(.08,.08)} \pgfnoderect{Node433}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node434}[fill]{\pgfxy(1.11,.35)}{\pgfxy(.08,.08)} \pgfnoderect{Node435}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node436}[fill]{\pgfxy(1.75,.56)}{\pgfxy(.08,.08)} \pgfnoderect{Node437}[fill]{\pgfxy(1.9,.66)}{\pgfxy(.08,.08)} \pgfnoderect{Node438}[fill]{\pgfxy(1.45,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node439}[fill]{\pgfxy(1.65,.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node440}[fill]{\pgfxy(1.65,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.5cm}{3.5cm}} \pgfnoderect{Node441}[fill]{\pgfxy(1.55,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node442}[fill]{\pgfxy(1.7,.12)}{\pgfxy(.08,.08)} \pgfnoderect{Node443}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node444}[fill]{\pgfxy(1.4,.42)}{\pgfxy(.08,.08)} \pgfnoderect{Node445}[fill]{\pgfxy(1.6,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node446}[fill]{\pgfxy(1.15,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node447}[fill]{\pgfxy(1.35,.63)}{\pgfxy(.08,.08)} \pgfnoderect{Node448}[fill]{\pgfxy(1.5,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node449}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node450}[fill]{\pgfxy(1.55,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.8cm}{4.2cm}} \pgfnoderect{Node451}[fill]{\pgfxy(1.2,0)}{\pgfxy(.08,.08)} \pgfnoderect{Node452}[fill]{\pgfxy(1.7,.42)}{\pgfxy(.08,.08)} \pgfnoderect{Node453}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node454}[fill]{\pgfxy(1.11,.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node455}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node456}[fill]{\pgfxy(1.75,.56)}{\pgfxy(.08,.08)} \pgfnoderect{Node457}[fill]{\pgfxy(1.35,.83)}{\pgfxy(.08,.08)} \pgfnoderect{Node458}[fill]{\pgfxy(1.45,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node459}[fill]{\pgfxy(1.2,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node460}[fill]{\pgfxy(1.65,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.3cm}{4cm}} \pgfnoderect{Node461}[fill]{\pgfxy(1.2,0)}{\pgfxy(.08,.08)} \pgfnoderect{Node462}[fill]{\pgfxy(1.7,.42)}{\pgfxy(.08,.08)} \pgfnoderect{Node463}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node464}[fill]{\pgfxy(1.4,.52)}{\pgfxy(.08,.08)} \pgfnoderect{Node465}[fill]{\pgfxy(1.6,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node466}[fill]{\pgfxy(1.15,.65)}{\pgfxy(.08,.08)} \pgfnoderect{Node467}[fill]{\pgfxy(1.35,.83)}{\pgfxy(.08,.08)} \pgfnoderect{Node468}[fill]{\pgfxy(.8,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node469}[fill]{\pgfxy(1.2,.95)}{\pgfxy(.08,.08)} \pgfnoderect{Node470}[fill]{\pgfxy(1.85,.85)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node471}[fill]{\pgfxy(10.6,1.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node472}[fill]{\pgfxy(10.7,1.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node473}[fill]{\pgfxy(9,1)}{\pgfxy(.08,.08)} \pgfnoderect{Node474}[fill]{\pgfxy(9.1,1.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node475}[fill]{\pgfxy(8.9,1.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node476}[fill]{\pgfxy(8.8,1.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node477}[fill]{\pgfxy(8.7,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node478}[fill]{\pgfxy(8.8,1.85)}{\pgfxy(.08,.08)} \pgfnoderect{Node479}[fill]{\pgfxy(7.6,1.45)}{\pgfxy(.08,.08)} \pgfnoderect{Node480}[fill]{\pgfxy(7.8,1.7)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node481}[fill]{\pgfxy(10.7,3.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node482}[fill]{\pgfxy(10.6,3.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node483}[fill]{\pgfxy(10.6,3.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node484}[fill]{\pgfxy(10.7,3.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node485}[fill]{\pgfxy(10.5,3.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node486}[fill]{\pgfxy(10.7,4.1)}{\pgfxy(.08,.08)} \pgfnoderect{Node487}[fill]{\pgfxy(10.7,4.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node488}[fill]{\pgfxy(10.75,4.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node489}[fill]{\pgfxy(10.5,4.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node490}[fill]{\pgfxy(10.1,4.8)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node491}[fill]{\pgfxy(10.1,4.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node492}[fill]{\pgfxy(10.3,4.2)}{\pgfxy(.08,.08)} \pgfnoderect{Node493}[fill]{\pgfxy(10.5,4.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node494}[fill]{\pgfxy(10,1.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node495}[fill]{\pgfxy(9.6,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node496}[fill]{\pgfxy(9.8,1.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node497}[fill]{\pgfxy(8.4,3.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node498}[fill]{\pgfxy(8.5,2.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node499}[fill]{\pgfxy(8,2.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node500}[fill]{\pgfxy(7.3,3.3)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7.5cm}{4.1cm}} \pgfnoderect{Node501}[fill]{\pgfxy(1.55,.32)}{\pgfxy(.08,.08)} \pgfnoderect{Node502}[fill]{\pgfxy(1.7,.3)}{\pgfxy(.08,.08)} \pgfnoderect{Node503}[fill]{\pgfxy(1.85,.22)}{\pgfxy(.08,.08)} \pgfnoderect{Node504}[fill]{\pgfxy(1.4,.4)}{\pgfxy(.08,.08)} \pgfnoderect{Node505}[fill]{\pgfxy(1.6,.5)}{\pgfxy(.08,.08)} \pgfnoderect{Node506}[fill]{\pgfxy(1.25,.55)}{\pgfxy(.08,.08)} \pgfnoderect{Node507}[fill]{\pgfxy(1.35,.63)}{\pgfxy(.08,.08)} \pgfnoderect{Node508}[fill]{\pgfxy(1.2,.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node509}[fill]{\pgfxy(1.3,.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node510}[fill]{\pgfxy(1.35,.75)}{\pgfxy(.08,.08)} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfnoderect{Node511}[fill]{\pgfxy(6.9,4)}{\pgfxy(.08,.08)} \pgfnoderect{Node512}[fill]{\pgfxy(6.9,3.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node513}[fill]{\pgfxy(6.8,4.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node514}[fill]{\pgfxy(6.4,4.75)}{\pgfxy(.08,.08)} \pgfnoderect{Node515}[fill]{\pgfxy(10.7,5)}{\pgfxy(.08,.08)} \pgfnoderect{Node516}[fill]{\pgfxy(10.7,4.8)}{\pgfxy(.08,.08)} \pgfnoderect{Node517}[fill]{\pgfxy(10.3,4.9)}{\pgfxy(.08,.08)} \pgfnoderect{Node518}[fill]{\pgfxy(10,1.7)}{\pgfxy(.08,.08)} \pgfnoderect{Node519}[fill]{\pgfxy(8.2,1.6)}{\pgfxy(.08,.08)} \pgfnoderect{Node520}[fill]{\pgfxy(9.2,1.6)}{\pgfxy(.08,.08)} \end{pgftranslate} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \color{white} \pgfmoveto{\pgfxy(7,4.65)} \pgflineto{\pgfxy(7.7,4.65)} \pgflineto{\pgfxy(7.7,5.25)} \pgflineto{\pgfxy(7,5.25)} \pgflineto{\pgfxy(7,4.65)} \pgffill \color{black} \pgfmoveto{\pgfxy(7,4.65)} \pgflineto{\pgfxy(7.7,4.65)} \pgflineto{\pgfxy(7.7,5.25)} \pgfstroke \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(7,-.1)(7,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(7.15,4.95)}{\pgfbox[left,center]{b)}} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Most two-element subsets of $M$ do not consist of the initial and terminal elements of a relation; b) induced relation on $M_2$.} \label{twoelement} \end{figure} \subsection{Power Spaces}\label{subsectionpowerset} \refstepcounter{textlabels}\label{powerspaces} {\bf Power Spaces; Induced and Holistic Power Spaces.} The notion of relations between pairs of two-element subsets of a multidirected set $M=(M,R,i,t)$, discussed at the end of section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, naturally motivates consideration of relations between pairs of subsets of $M$ of arbitrary cardinality. Since subsets of $M$ correspond to {\it elements} of the power set $\ms{P}(M)$ over $M$, families of relations between pairs of subsets of $M$ correspond to multidirected structures on $\ms{P}(M)$, called {\it power spaces over} $M$. The following general set-theoretic definition makes this idea precise: \vspace*{.2cm} \begin{defi}\label{defpowerspaces} Let $M$ be an arbitrary set. A {\bf power space} over $M$ is a multidirected set whose underlying set is the power set $\ms{P}(M)$ of $M$. \end{defi} \refstepcounter{textlabels}\label{inducedpowerspaces} The reason for taking $M$ to be an arbitrary set in definition \hyperref[defpowerspaces]{\ref{defpowerspaces}} is that the ``general power space viewpoint" takes multidirected structure on $M$ to be {\it residual structure} left over after ignoring relations on $\ms{P}(M)$ involving nonsingleton subsets. As discussed below, it is generally not possible to recover a power space from this residual structure. Power spaces that {\it do} arise naturally from the structure of a ``preexisting" multidirected set $M=(M,R,i,t)$ are called {\bf induced power spaces}. The majority of power spaces considered in this paper are induced, but the reader should be aware that these are exceptional cases. A ``trivial" example of an induced power space over a multidirected set $M=(M,R,i,t)$ is given by taking the relation set over $\ms{P}(M)$ to be $R$, viewed as a set of relations between pairs of singleton subsets of $M$, with no relations involving subsets of higher cardinality. A slightly more interesting example may be constructed by combining the relation set $R$ of $M$ with the induced relation on two-element subsets of $M$, introduced at the end of section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above. This defines a power space over $M$ with relations between distinguished pairs of one and two-element subsets. If $A$ is an acyclic directed set, then the star model $A_\star$ of $A$, viewed as a directed set as described in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, is the ``nontrivial part" of this power space, from a generalized order-theoretic perspective. However, the star model of a general multidirected set $M$ does not embed naturally into a power space over $M$, since relations in $M$ generally cannot be identified with two-element subsets of $M$. A third induced power space over $M$ is the space with one relation between each pair of subsets $M'$ and $M''$ of $M$ for each relation in $R$ between an element of $M'$ and an element of $M''$. \refstepcounter{textlabels}\label{holisticpowerspaces} In contrast to these examples, most power spaces over a set $M$ are {\bf holistic,} meaning that their information content consists at least partly of {\it irreducibly complex relationships} between pairs of nonsingleton subsets of $M$. Such power spaces are {\it not} induced in any natural way by any preexisting multidrected structure on $M$.\footnotemark\footnotetext{The caveat {\it naturally} is included because one {\it can} define bijections between the set of {\it finite} multidirected structures on a {\it finite} set and the set of {\it finite} multidirected structures on its power set, since both sets are countably infinite. Under such a bijection, a typical power space could be encoded in an unnatural and arbitrary way by invoking large numbers of relations between typical elements of $M$.} It has recently become popular in certain scientific fields, not limited to physics, to take such holistic relationships seriously, even at the classical level. I briefly discuss a few instances of this notion below, including the general concept of {\it top-down causation,} and some specific holistic approaches to fundamental physics. \refstepcounter{textlabels}\label{higherinduced} {\bf Higher Induced Relations; Splice Relations.} Since relations in a multidirected set $M=(M,R,i,t)$ may be identified with $1$-chains in $M$, an obvious way to define induced relations between pairs of distinguished nonsingleton subsets of $M$ is to modify the induced binary relation $\prec$ on the relation space $\ms{R}(M)$ over $M$ to apply to longer chains. Complications immediately arise in the general multidirected context. For example, an $n$-chain in a multidirected set $M$ does not always correspond to an $n$-element subset of $M$, since chains may intersect themselves. Further, multiple chains may share the same sequence of elements. Hence, this approach generally involves a choice between accepting significant loss of information, as in the case of induced relations between pairs of two-element subsets of $M$, discussed at the end of section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, and leaving the realm of power spaces, as in the theory of {\it causal path spaces,} introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. However, in the special case where $M$ is an acyclic directed set, each $n$-chain in $M$ uniquely identifies an $n$-element subset of $M$, permitting relations between pairs of chains to be viewed as relations between elements of the power set $\ms{P}(M)$ of $M$. A straightforward way to define such relations between pairs of chains is to {\it iterate the process of forming relation spaces.} Let $A$ be an acyclic directed set, with relation space $\ms{R}(A)$, and set of $2$-chains $\tn{Ch}_2(A)$. Here, it is convenient to denote the binary relation on $A$ by $\prec_0$, and the induced binary relation on $\ms{R}(A)$ by $\prec_1$. Now let $w\prec_0 x\prec_0 y\prec_0 z$ be a $3$-chain in $A$, and denote the relations $w\prec_0 x$, $x\prec_0 y$, and $y\prec_0 z$, by $q$, $r$, and $s$, respectively. Denote the $2$-chains $w\prec_0 x\prec_0 y$ and $x\prec_0 y\prec_0 z$ by $\gamma$ and $\delta$, respectively. Then there exists an {\it induced $2$-chain} $q\prec_1 r\prec_1 s$ in $\ms{R}(A)$. The {\it second relation space} $\ms{R}^2(A):=\ms{R}(\ms{R}(A))$ of $A$ corresponds to $\tn{Ch}_2(A)$ in an obvious way, and the {\it second induced relation} $\prec_2$ on $\ms{R}^2(A)$, induced by $\prec_1$, may be naturally identified with the binary relation on $\tn{Ch}_2(A)$ defined by taking one $2$-chain to precede another if and only if these chains ``overlap" in a single relation in $A$. For example, in the case described above, $\gamma\prec_2 \delta$, since $\gamma$ and $\delta$ overlap in the relation $x \prec_0 y$. One may proceed iteratively to define {\bf higher induced relations} $\prec_n$ on the chain sets $\tn{Ch}_n(A)$ of $A$. As described so far, this construction applies to a general multidirected set. However, since $A$ is an acyclic directed set, the higher induced relations $\prec_n$ descend to binary relations on the corresponding sets of $n$-element subsets of $A$. Any subfamily of these relations may be used to define a power space over $A$. Whenever two overlapping $n$-chains $\gamma=\{x_0\prec x_1\prec...\prec x_n\}$ and $\delta=\{x_1\prec...\prec x_n\prec x_{n+1}\}$ in $A$ each extend to $n$-simplices $\overline{\gamma}$ and $\overline{\delta}$ in $A$; i.e., transitive linear suborders of $A$, the relation $\gamma\prec_n \delta$ may be viewed as {\it joining the maximal face of the ``initial" simplex $\overline{\gamma}$ to the minimal face of the ``terminal" simplex $\overline{\delta}$.} This viewpoint motivates the definition of alternative, ``sparser" power spaces over $A$ than those defined in terms of the higher induced relations described above. These {\bf simplicial power spaces} are defined by recognizing only relations between pairs of $n$-element subsets of $A$ that define simplices in $A$. The ``nontrivial parts" of simplicial power spaces are generalizations of the star model $A_\star$ of $A$, introduced in section \hyperref[subsectiontopology]{\ref{subsectiontopology}} above. Despite obvious complications, these ideas may be extended in interesting ways to apply to general multidirected sets.\footnotemark\footnotetext{This is one possible way of viewing causal dynamical triangulations.} Power spaces defined in terms of higher induced relations, and simplicial power spaces, are both examples of {\bf binary power spaces,} which involve only {\it directed,} rather than multidirected, structures on power sets. \refstepcounter{textlabels}\label{splicerel} Higher induced relations, and the binary relations defining simplicial power spaces, are special cases of {\bf splice relations}, in which pairs of chains, or more complicated ``generalized elements" associated with a multidirected set $M$, are taken to be related if they may be ``spliced together" in some specified way to form a ``generalized element of greater temporal size." These ``generalized elements" may often be formalized in terms of {\it spaces of morphisms into} $M$, such as the {\it causal path spaces} introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} below. These spaces are generally {\it not} power spaces, since morphisms into $M$ generally involve information that cannot be recovered from the images of their element maps. Moreover, since there is generally more than one way to ``splice together" a given pair of ``generalized elements," splice relations generally define multidirected, rather than merely directed, structures. The splice relations of greatest interest in this paper are {\it not} higher induced relations, nor relations defining simplicial power spaces, which relate pairs of chains or simplices overlapping in increasingly long subchains or subsimplices. More useful are splice relations relating pairs of ``generalized elements of small or empty intersection," such as the {\it concatenation relation} and the {\it directed product relation} for pairs of paths, introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}. These relations define causal path spaces called the {\it causal concatenation space} and the {\it causal directed product space,} respectively. \newpage \refstepcounter{textlabels}\label{causalatoms} {\bf Causal Atoms; Causal Atomic Resolution.} A useful application of induced power spaces is to {\it organize information in a multidirected set} into families of subobjects, for the purposes of approximation and computation. Here I present a method of using {\it nested families of subobjects} of a multidirected set to approximate structure at different scales. Such approximation methods are important for practical purposes in the study of fundamental spacetime structure, since the scale of individual elements may be many orders of magnitude too small to be directly accessible to experimental observation in the short term. The most interesting such approximation methods are those that {\it organize information in a manner compatible with multidirected structure.} Here, I briefly discuss one such method, in which elements of a multidirected set are grouped into subsets, called {\it causal atoms,} or more generally, {\it local causal atoms,} which may be viewed as elements of a ``coarser" multidirected set. The reader should compare this to the {\it decimation} approach to {\it coarse-graining} of causal sets, appearing, for example, in \cite{SorkinEvidenceofContLimVer208}. The computational utility of causal atoms has already been illustrated in the technique of {\it atomic accretion} used in the proof of theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} above. A {\bf causal atom} $\alpha$ in a multidirected set $M=(M,R,i,t)$ is a {\it convex subset} of $M$ in a generalized order-theoretic sense, satisfying the property that every element lying on a chain in $M$ between a pair of elements in $\alpha$ also belongs to $\alpha$. This means that every element in the complement $M-\alpha$ is either exclusively in the past of $\alpha$, exclusively in the future of $\alpha$, or unrelated to any element of $\alpha$. In particular, $\alpha$ {\it absorbs} any cycle it intersects. Hence, the possible relationships between a causal atom and an element in its complement are analogous to the possible relationships between a pair of elements in an acyclic directed set. Causal atoms may therefore be viewed as ``generalized elements," which explains the choice of terminology. To temper the absorption property in multidirected sets involving large cycles, one may define {\bf local causal atoms,} which include all elements lying on {\it sufficiently short} chains between pairs of their elements. I do not explore these details here, however. Causal atoms may be endowed with internal multidirected structure by elevating them to full subobjects of $M$. \refstepcounter{textlabels}\label{atomictop} Relationships between pairs of causal atoms in a multidirected set $M=(M,R,i,t)$ are generally more complicated than relationships between pairs of elements of $M$, or between a causal atom in $M$ and an element of $M$. If $M$ is acyclic, its individual elements are causal atoms, but elements lying on cycles are ``smaller than any causal atom," due to the absorption property. The open interval $\llangle x,z\rrangle$ between two elements $x$ and $z$ in $M$ is a causal atom, since every chain between two elements in $\llangle x,z\rrangle$ extends to a chain from $x$ to $z$. Antichains are causal atoms, since they admit no chains between pairs of their elements. Maximal cycles are causal atoms, since any element belonging to both the past and future of a cycle lies on an intersecting cycle. The family of causal atoms in $M$ is closed under intersection, so appropriate subfamilies can serve as {\it bases,} rather than merely {\it subbases,} for topologies on the underlying set of $M$, called {\bf atomic topologies}. Intersections of open intervals are causal atoms, so the interval topology is an atomic topology. The discrete topology is atomic only in the acyclic case. Many other examples of atomic topologies exist. The star topology on the star model $A_\star$ of an acyclic directed set $A$ is an atomic topology for the directed structure on $A_\star$ described in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above. In the general case, however, the star $\tn{St}(x)$ at an element $x$ in $M$ may be ``smaller than any causal atom containing $x$." {\bf Local atomic topologies} may be defined in terms of local causal atoms. \refstepcounter{textlabels} \label{causalatomicdec} A {\bf causal atomic decomposition} of a multidirected set $M=(M,R,i,t)$ is a partitioning of the underlying set of $M$ into causal atoms. In this context, it is sometimes useful to relabel $M$ as $M^0$, while the causal atoms in the decomposition are viewed as elements of a ``coarser" set $M^1$. This set may be endowed with multidirected structure in a variety of different ways. In particular, $M^1$ may have ``higher-level cycles," even though every cycle in $M$ is absorbed into a unique causal atom. For example, if a pair of causal atoms $\alpha$ and $\alpha'$ in $M$ each have elements in the past of the other, relations may be defined in both directions between $\alpha$ and $\alpha'$ in $M^1$. {\bf Local causal atomic decompositions} may also be defined. Multidirected structures arising from either type of decomposition define power spaces over $M$, since causal atoms in $M$ correspond uniquely to subsets of $M$. A causal atomic decomposition of a multidirected set $M=M^0$ is illustrated in figure \hyperref[atomic]{\ref{atomic}}a below, with the elements represented by the nodes in each grey region grouped together as causal atoms. Figure \hyperref[atomic]{\ref{atomic}}b illustrates one possible method of assigning a multidirected structure to the set $M^1$ of causal atoms in this decomposition. In this case, the chosen structure is a binary relation, defined as follows: $\alpha\prec\alpha'$ in $M^1$ if and only if 1) every element of $\alpha$ precedes some element of $\alpha'$ in $M$; 2) no element of $\alpha$ succeeds any element of $\alpha'$ in $M$; and 3) some element of $\alpha$ {\it directly} precedes some element of $\alpha'$ in $M$. This is a very ``sparse" choice of structure for $M^1$. The further grouping of elements of $M^1$ indicated by the grey regions in figure \hyperref[atomic]{\ref{atomic}}b represents a second causal atomic decomposition, as discussed below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.25cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,4.95)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,4.95)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(5.5,-.1)(5.5,5.25) \pgfxyline(11,-.1)(11,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \begin{pgftranslate}{\pgfpoint{.2cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{30!white} \pgfmoveto{\pgfxy(.7,.3)} \pgfcurveto{\pgfxy(1.5,.3)}{\pgfxy(1.5,3)}{\pgfxy(.7,2.8)} \pgfcurveto{\pgfxy(0,2.8)}{\pgfxy(.2,.2)}{\pgfxy(.7,.3)} \pgffill \pgfmoveto{\pgfxy(2,.7)} \pgfcurveto{\pgfxy(2.4,.7)}{\pgfxy(2.5,1.7)}{\pgfxy(1.9,1.7)} \pgfcurveto{\pgfxy(1.4,1.7)}{\pgfxy(1.5,.6)}{\pgfxy(2,.7)} \pgffill \pgfmoveto{\pgfxy(1.7,3.5)} \pgfcurveto{\pgfxy(2.2,5.5)}{\pgfxy(0,5.5)}{\pgfxy(.3,5)} \pgfcurveto{\pgfxy(1,3)}{\pgfxy(1.5,3)}{\pgfxy(1.7,3.5)} \pgffill \pgfmoveto{\pgfxy(1.9,4.2)} \pgfcurveto{\pgfxy(1.9,4.5)}{\pgfxy(2.1,5.7)}{\pgfxy(2.5,5.1)} \pgfcurveto{\pgfxy(2.6,4.5)}{\pgfxy(2.9,3)}{\pgfxy(1.9,4.2)} \pgffill \pgfmoveto{\pgfxy(2.2,2)} \pgfcurveto{\pgfxy(2.2,3)}{\pgfxy(3,3)}{\pgfxy(3.2,2.5)} \pgfcurveto{\pgfxy(3.2,2.2)}{\pgfxy(2.4,1.4)}{\pgfxy(2.2,2)} \pgffill \pgfmoveto{\pgfxy(3,3.1)} \pgfcurveto{\pgfxy(4.3,3.2)}{\pgfxy(4,3.7)}{\pgfxy(3.5,4.2)} \pgfcurveto{\pgfxy(3.1,4.3)}{\pgfxy(2.6,3)}{\pgfxy(3,3.1)} \pgffill \pgfmoveto{\pgfxy(3,4.5)} \pgfcurveto{\pgfxy(3.7,4.1)}{\pgfxy(3.7,5.1)}{\pgfxy(2.8,5)} \pgfcurveto{\pgfxy(2.6,4.8)}{\pgfxy(2.8,4.5)}{\pgfxy(3,4.5)} \pgffill \pgfmoveto{\pgfxy(3,.5)} \pgfcurveto{\pgfxy(5.2,-.3)}{\pgfxy(4.1,1.8)}{\pgfxy(3.4,1.7)} \pgfcurveto{\pgfxy(3.1,1.8)}{\pgfxy(2.5,.7)}{\pgfxy(3,.5)} \pgffill \pgfmoveto{\pgfxy(4.2,4.3)} \pgfcurveto{\pgfxy(5.2,5.7)}{\pgfxy(4.3,5.6)}{\pgfxy(4,4.8)} \pgfcurveto{\pgfxy(3.6,4)}{\pgfxy(4,4)}{\pgfxy(4.2,4.3)} \pgffill \pgfmoveto{\pgfxy(5.2,1)} \pgfcurveto{\pgfxy(5.6,2.2)}{\pgfxy(4.2,2)}{\pgfxy(4.2,1.6)} \pgfcurveto{\pgfxy(5,.1)}{\pgfxy(5.2,.1)}{\pgfxy(5.2,1)} \pgffill \pgfmoveto{\pgfxy(4.5,2.2)} \pgfcurveto{\pgfxy(6,2.2)}{\pgfxy(4.7,3.6)}{\pgfxy(4.5,3.5)} \pgfcurveto{\pgfxy(4,3.1)}{\pgfxy(3.3,2.2)}{\pgfxy(4.5,2.2)} \pgffill \pgfmoveto{\pgfxy(5.3,3.7)} \pgfcurveto{\pgfxy(5.2,5.4)}{\pgfxy(4.6,4.6)}{\pgfxy(4.8,4.1)} \pgfcurveto{\pgfxy(4.8,3.5)}{\pgfxy(5.4,2.8)}{\pgfxy(5.3,3.7)} \pgffill \end{colormixin} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfputat{\pgfxy(3.85,3.55)}{\pgfbox[left,center]{\large{$x$}}} \pgfputat{\pgfxy(4.3,2.8)}{\pgfbox[left,center]{\huge{$\alpha_1$}}} \pgfputat{\pgfxy(4,4.8)}{\pgfbox[left,center]{\huge{$\alpha_3$}}} \pgfputat{\pgfxy(4.9,4)}{\pgfbox[left,center]{\huge{$\alpha_2$}}} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.7cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{30!white} \pgfmoveto{\pgfxy(1,.7)} \pgfcurveto{\pgfxy(3,0)}{\pgfxy(3.5,3.5)}{\pgfxy(2.5,3)} \pgfcurveto{\pgfxy(1,2.5)}{\pgfxy(-1,1.5)}{\pgfxy(1,.7)} \pgffill \pgfmoveto{\pgfxy(1.5,3.9)} \pgfcurveto{\pgfxy(3.1,3.9)}{\pgfxy(2.9,5.2)}{\pgfxy(2,5.2)} \pgfcurveto{\pgfxy(.5,5.2)}{\pgfxy(.5,3.9)}{\pgfxy(1.5,3.9)} \pgffill \pgfmoveto{\pgfxy(4.5,2.2)} \pgfcurveto{\pgfxy(5.5,2.5)}{\pgfxy(6,5.5)}{\pgfxy(4,5.2)} \pgfcurveto{\pgfxy(3.5,5.2)}{\pgfxy(4,2.2)}{\pgfxy(4.5,2.2)} \pgffill \pgfmoveto{\pgfxy(3.5,3.1)} \pgfcurveto{\pgfxy(4,3.1)}{\pgfxy(3.7,5.5)}{\pgfxy(3,5)} \pgfcurveto{\pgfxy(2.5,5)}{\pgfxy(3,3.1)}{\pgfxy(3.5,3.1)} \pgffill \pgfmoveto{\pgfxy(3.5,.5)} \pgfcurveto{\pgfxy(6,1)}{\pgfxy(6,2.5)}{\pgfxy(4,1.7)} \pgfcurveto{\pgfxy(2.5,1.5)}{\pgfxy(2.5,.5)}{\pgfxy(3.5,.5)} \pgffill \end{colormixin} \pgfnodecircle{Node100}[fill]{\pgfxy(.7,1.5)}{0.17cm} \pgfnodecircle{Node200}[fill]{\pgfxy(1.3,4.5)}{0.17cm} \pgfnodecircle{Node300}[fill]{\pgfxy(2,1.3)}{0.17cm} \pgfnodecircle{Node400}[fill]{\pgfxy(2.3,4.5)}{0.17cm} \pgfnodecircle{Node500}[fill]{\pgfxy(2.6,2.4)}{0.17cm} \pgfnodecircle{Node600}[fill]{\pgfxy(3.5,1)}{0.17cm} \pgfnodecircle{Node700}[fill]{\pgfxy(3.4,3.6)}{0.17cm} \pgfnodecircle{Node800}[fill]{\pgfxy(3.1,4.6)}{0.17cm} \pgfnodecircle{Node900}[fill]{\pgfxy(4.8,1.4)}{0.17cm} \pgfnodecircle{Node1000}[fill]{\pgfxy(4.5,2.7)}{0.17cm} \pgfnodecircle{Node1100}[fill]{\pgfxy(4.2,4.7)}{0.17cm} \pgfnodecircle{Node1200}[fill]{\pgfxy(5,4)}{0.17cm} \pgfsetlinewidth{2pt} \pgfnodeconnline{Node100}{Node200} \pgfnodeconnline{Node300}{Node200} \pgfnodeconnline{Node300}{Node500} \pgfnodeconnline{Node500}{Node200} \pgfnodeconnline{Node500}{Node400} \pgfnodeconnline{Node600}{Node500} \pgfnodeconnline{Node600}{Node700} \pgfnodeconnline{Node600}{Node1000} \pgfnodeconnline{Node700}{Node800} \pgfnodeconnline{Node900}{Node1000} \pgfnodeconnline{Node1000}{Node1100} \pgfnodeconnline{Node1000}{Node1200} \pgfnodeconnline{Node1100}{Node1200} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \pgfnodecircle{Node101}[fill]{\pgfxy(1.5,1.5)}{0.23cm} \pgfnodecircle{Node201}[fill]{\pgfxy(1.8,4.5)}{0.23cm} \pgfnodecircle{Node301}[fill]{\pgfxy(3,4.2)}{0.23cm} \pgfnodecircle{Node401}[fill]{\pgfxy(4,1)}{0.23cm} \pgfnodecircle{Node501}[fill]{\pgfxy(4.5,4)}{0.23cm} \pgfsetlinewidth{3pt} \pgfnodeconnline{Node101}{Node201} \pgfnodeconnline{Node401}{Node501} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Grouping elements of a multidirected set $M=M^0$ into causal atoms; b) causal atomic decomposition $M^1$, with elements grouped for further decomposition; c) second decomposition $M^2$.} \label{atomic} \end{figure} \vspace*{-.5cm} An important classical case of causal atomic decomposition is the {\it foliation of a spacetime region by Cauchy surfaces,} as mentioned in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above. The analogue of such foliation in the discrete causal context may be called {\bf generational dynamics}, in contrast to the sequential growth dynamics introduced by Sorkin and Rideout \cite{SorkinSequentialGrowthDynamics99}. I briefly return to the topic of generational dynamics in the context of kinematic schemes in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. A simple example of {\it local} causal atomic decomposition in the cyclic interpolative case is the partitioning of a directed circle into a pair of half-open arcs. \refstepcounter{textlabels}\label{causalatomicres} A {\bf causal atomic resolution} of a multidirected set $M=(M,R,i,t)$ is a sequence \[\mbf{M}:=\{M=M^0,M^1,M^2,...\}\] of multidirected sets, where $M^{n+1}$ is a causal atomic decomposition of $M^n$ for all $n\ge0$. The multidirected set $M^n$ is called the $n$th {\it level} of the causal atomic resolution $\mbf{M}$. Note that causal atomic decomposition is generally {\it not} a transitive operation, meaning that $M^{n+2}$ is generally not a causal atomic decomposition of $M^n$. More precisely, if $\alpha_1,...,\alpha_k$ are causal atoms in $M^n$, viewed as elements of $M^{n+1}$, such that $\{\alpha_1,...,\alpha_k\}$ is a causal atom in $M^{n+1}$, the union of the elements of $\alpha_1,...,\alpha_k$ is generally not a causal atom in $M^n$. For example, the element $x$ in figure \hyperref[atomic]{\ref{atomic}}a above is in the intersection of the past and future of the set $\alpha_1\cup \alpha_2\cup\alpha_3$, which is therefore not a causal atom in $M=M^0$. However, $\{\alpha_1,\alpha_2,\alpha_3\}$ {\it is} a causal atom in $M^1$. This is yet another example in which a notion analogous to category-theoretic composition fails to apply in discrete causal theory.\footnotemark\footnotetext{More generally, failure of ``composition" is not unusual in "information-filtering constructions." A familiar example is given by {\it composition series} of finite groups in the context of the Jordan-H\"{o}lder theorem, since formation of normal subgroups is nontransitive.} There are a number of different ways to characterize a causal atomic resolution $\mbf{M}$ in terms of the properties of its levels $M^n$, and the relationships among them. For example, the $n$th {\it growth factor} $\lambda_n$ of $\mbf{M}$ is the average number of elements of $M^{n-1}$ making up an element of $M^{n}$. If the sequence $\Lambda(\mbf{M}):=\{\lambda_1,\lambda_2,...\}$ of growth factors of $\mbf{M}$ is constant over a certain range of indices, then $\mbf{M}$ is called {\it exponential} over this range, since the sizes of typical atoms at different levels within this range are related by a power of the common growth factor. Similarly, if the sequence $\Lambda(\mbf{M})$ is decreasing or increasing over a certain range of indices, then $\mbf{M}$ is called {\it subexponential} or {\it superexponential} over this range, respectively. The growth factors of the causal atomic resolution illustrated in figure \hyperref[atomic]{\ref{atomic}} above are $\lambda_1=35/12\approx 2.92$ and $\lambda_2=12/5=2.4$, so this particular resolution is subexponential. One may also characterize the {\it uniformity} of a given level $M^n$ in a number of obvious ways. The method of causal atomic resolution is similar in spirit to {\it multiresolution analysis} in the study of {\it wavelets} in the field of {\it harmonic analysis.} The general aim of such methods is to provide a unified way of dealing with aspects of structure existing at different scales. This viewpoint is of particularly interest in physical theories where emergent phenomena are prominent, since emergence is often scale-dependent. \refstepcounter{textlabels}\label{classicalhol} \refstepcounter{textlabels}\label{ellis} {\bf Top-Down Causation; Classical Holism.} A number of physicists, computer scientists, chemists, biologists, neuroscientists, and philosophers have recently devoted attention to the concept of classical {\bf top-down causation}, which challenges the philosophy of {\bf classical reductionism}. Classical reductionism assumes that classical causal structure may be completely described in terms of relations between pairs of individual events, as implied by the binary axiom (\hyperref[b]{B}) of causal set theory. Representative samplings of the top-down viewpoint appear in the recent collections {\it Top-Down Causation} \cite{EllisNobleOConnor12}, and {\it Downward Causation and the Neurobiology of Free Will} \cite{MurphyEllisOConnor09}. As George Ellis \cite{EllisNature05} puts it, {\it ``the higher levels in the hierarchy of complexity have autonomous causal powers that are functionally independent of lower-level processes."} Depending on its specific version, top-down causation either approaches or crosses the threshold of {\bf classical holism}, the antithesis of the reductionist philosophy. To evaluate the legitimacy and implications of top-down causation, it is necessary to formulate the idea in precise mathematical terms. Since top-down causation involves influences exerted by entire families of events, the theory of power spaces is a natural tool to use.\footnotemark\footnotetext{Other alternatives exist, even in the context of discrete causal theory. In particular, even the most general power spaces preserve a certain {\it binary} character, by focusing on relationships {\it between pairs} of families of elements, rather than relationships {\it among families} of families of elements. There are excellent physical motivations for this focus, but one may choose to abstain from it. I briefly mention some alternatives in section \hyperref[subsectionomitted]{\ref{subsectionomitted}} below.} Different implementations are possible, some conservative, others more radical. \refstepcounter{textlabels}\label{degreeshol} A conservative example of the top-down viewpoint involving power spaces appears in the method of causal atomic resolution, where influences are ascribed to families of events purely for convenience. The power spaces arising in this context merely represent ways of organizing and filtering information; they do not involve any profound hypotheses about the nature of the universe. In particular, the ``approximate" multidirected structures given by causal atomic decompositions may {\it innocuously} exhibit properties that would arouse controversy in more fundamental contexts. For example, two causal atoms may be ``related to each other," in the sense that each contains events in the other's past, yet this is irrelevant to the controversial issue of classical causal cycles. Similarly, one causal atom may be related to another, even though some of the corresponding events are {\it unrelated,} yet this is irrelevant to the controversial issue of classical nonlocal interaction. More interesting are cases in which relationships involving complex families of events are considered ``exact," rather than mere approximations. The corresponding power spaces may still be induced by relations between pairs of individual elements, but there may be alternative families of relations that better represent the information contained in these power spaces. Choosing such a family is analogous to choosing a ``natural basis" from among a linearly dependent set of vectors. For example, the relation space $\ms{R}(M)$ over a multidirected set $M$, together with its induced relation $\prec$, gives an alternative representation of the information contained in $M$, except at the boundary $\partial M$ of $M$. The resolution of the impermeability problem for maximal antichains upon passage to $\ms{R}(M)$, proven in theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} above, illustrates that advantages may sometimes be gained by working in terms of one representation rather than another. This example demonstrates that the reductionist viewpoint is not always the most useful one, even when reductionism is information-theoretically adequate. The argument that ``emergence" and reductionism are not mutually exclusive has been recognized for some time; see, for example, the classic paper by Anderson \cite{AndersonMoreisDifferent72}, or the recent paper by Butterfield \cite{ButterfieldReductionism11}. \refstepcounter{textlabels}\label{twistor} A classical alternative representation of ``spacetime information" is given by the {\it twistor space} $T$ over Minkowski spacetime $X$ in Roger Penrose's {\it twistor theory}, introduced in his 1967 paper {\it Twistor Algebra} \cite{PenroseTwistor67}. A useful later reference for twistor theory is Ward and Wells' standard text {\it Twistor Geometry and Field Theory} \cite{WardWellsTwistor90}. Twistor space $T$ may be identified with a five-dimensional real hypersurface in the complex projective space $\mathbb{C}\mathbb{P}^3$, which is six-dimensional as a real manifold. Elements of $T$ correspond to {\it null lines} in $X$, physically interpreted as the spacetime paths of rays of light. Interesting analogies exist between twistor space and the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$. For example, at a given element $x$ in $M$, the relations incident at $x$ are may be viewed as {\it minimal null segments}\footnotemark\footnotetext{In causal set theory, {\it irreducible} relations; i.e., ``links," are treated as ``null segments;" see, for example, Sorkin's paper \cite{SorkinLightLinksCausalSets09}. In the present case, this structural analogy is not intended to fix a rigid ``geometric" interpretation of either irreducible or reducible relations.} at $x$. One of the most attractive features of twistor theory is that it permits the conversion of certain important but intractable equations over classical spacetime into more natural and manageable forms. In an analogous way, passage to relation space allows a natural and straightforward derivation of {\it causal Schr\"{o}dinger-type equations} in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper. \refstepcounter{textlabels}\label{shapedynamics} The most radical versions of top-down causation involve holistic power spaces, which contain more information than can be naturally encoded in any multidirected structure on their underlying sets. Such {\it irreducible holism} at the classical level is broadly agreed to be at least {\it strongly constrained} by well-supported principles such as classical locality, but this view is not universal. A modern theory incorporating a radical degree of irreducible holism is Julian Barbour's {\it shape dynamics} \cite{BarbourShape12}. Shape dynamics does not fall under the umbrella of causal theory, taking a nearly opposite, ``space-first" approach to fundamental spacetime structure, in which {\it ``one \tn{[spatial]} state of the universe succeeds another along a unique evolution curve"} in {\it shape space,} the {\it conformal superspace of all conformal $3$-geometries.}\footnotemark\footnotetext{This quotation is taken from Barbour, Koslowsi, and Mercati \cite{BarbourShape13}. Conformal superspace is only one possible choice of a shape space; for example, one may define {\it discrete shape spaces.}} Underlying shape dynamics is the philosophy of Ernst Mach (1838-1916), that {\it global structure determines local effects.}\footnotemark\footnotetext{Mach's ideas influenced Einstein in the development of general relativity. ``Mach's principle," under various interpretations, has since become a ``strange attractor" for dubious theorizing. Carlo Rovelli demonstrates the disunity of modern Machian ideas in an amusing way by listing {\it eight} different versions of ``Mach's principle" on page 76 of his book {\it Quantum Gravity} \cite{Rovelli04}.} Shape dynamics represents the utmost extreme in classical holism: every aspect of each ``state of the universe" is relevant to all subsequent behavior. Among other ``eyebrow-raising" implications, this entails a {\it universal time parameter} and {\it absolute simultaneity.} Shape dynamics also favors a {\it spatially closed universe,} disfavored by current cosmological observations. Despite this, the theory succeeds in reproducing certain results of the {\it ADM formulation} of general relativity. \subsection{Causal Path Spaces}\label{subsectionpathspaces} \refstepcounter{textlabels}\label{causalpaths} {\bf Causal Paths.} A {\it directed path} in a multidirected set $M=(M,R,i,t)$ is a {\it morphism from a linear directed set into} $M$. A relation in $M$ may be regarded as a directed path of length one, and an $n$-chain in $M$ may be regarded as a directed path of length $n$, whose source is irreducible. In the special case where $M$ is a directed set, viewed as a model of classical spacetime structure, directed paths are {\it causal paths,} in the sense that an event represented by an element $x$ in $M$ influences an event represented by an element $y$ in $M$ if and only if there is a directed path from $x$ to $y$ in $M$. Such causal paths differ from causal curves in general relativity in two important ways. First, following the causal metric hypothesis (\hyperref[cmh]{CMH}), causal paths represent {\it actual influence,} not merely {\it potential influence.} In the language of section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} above, causal paths are {\it descriptive,} rather than {\it prescriptive.} Second, whereas causal curves in general relativity are equivalence classes of paths with the same continuum source, there exist an immense variety of distinct types of linear directed sets that can serve as sources for causal paths in discrete causal theory. In this paper, I focus almost entirely on the case of chains; i.e., causal paths with irreducible sources. The physical intuition associated with causal paths in directed sets is useful for multidirected sets in general. For this reason, I refer to a number of general constructions involving directed paths in multidirected sets as {\it causal.} For directed paths themselves, I usually abbreviate and simply use the term {\it paths,} since ``undirected paths" play no role in this paper. {\it Causal path sets} associated with a multidirected set $M=(M,R,i,t)$ are natural generalizations of the relation set $R$ of $M$. The recognition of natural directed structures on causal path sets, supplied by {\it splice relations,} such as the {\it concatenation relation} and the {\it directed product relation,} elevates these sets to {\it causal path spaces,} analogous to the relation space $\ms{R}(M)$ over $M$. This directed structure may be defined algebraically, in terms of {\it splice products,} and the corresponding causal path spaces may be viewed algebraically, as {\it causal path semicategories.} These, in turn, supply the multiplicative structure of {\it causal path algebras,} which are crucial for converting the physical information represented by a multidirected set into a mathematically tractable form. Multidirected, rather than merely directed, structures, may also be defined on causal path sets, but these are not explored in this paper. More generally, it is not feasible here to replicate, for the various spaces and algebraic objects introduced in this section, the careful examination of functorial, information-theoretic, and structural properties carried out for the relation space $\ms{R}(M)$ in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above. \refstepcounter{textlabels}\label{pathspacephys} {\bf General Contextual Remarks.} The use of path spaces in mathematics and theoretical physics goes back too far, and originates from too many independent sources, for a historical synopsis to be of any use to the reader. I therefore limit explicit reference to modern sources of more direct relevance to discrete causal theory. However, it is worthwhile, for contextual purposes, to briefly mention some of the conventional uses of path spaces. In mathematics, path spaces form the basis for {\it homotopy theory}\footnotemark\footnotetext{That is, {\it ordinary} homotopy theory. Other variants of the same idea are possible, such as Voevodsky's theory of $\mathbb{A}^1$-{\it homotopy} in algebraic geometry. {\it Causal homotopy} is an interesting topic, but I do not pursue it here.} in algebraic topology, via the {\it fundamental groupoid} and its higher-dimensional analogues. Path spaces also feature prominently in the theory of {\it topological groups,} and particularly {\it Lie groups,} due to the importance of {\it one-parameter subgroups.} In virtually every area of theoretical physics based on continuum geometry, path spaces are central. At the simplest level, they represent trajectories of point particles. For more complex systems, they appear in the theory of {\it phase space,} with particularly iconic status in the field of {\it nonlinear dynamics.} Perhaps most importantly, they underlie the {\it Lagrangian} and {\it Hamiltonian formulations} of classical mechanics, extend to analogous methods in quantum theory, and culminate in the {\it path integral approach} to quantum mechanics and quantum field theory. Path spaces find additional applications in string theory, due to the one-dimensional nature of strings, and in loop quantum gravity, due to the use of {\it holonomy} in the construction of the {\it loop representation.} However, the {\it directedness} inherent in causal structure makes path spaces especially important in the context of causal theory. \refstepcounter{textlabels}\label{pathmorph} \refstepcounter{textlabels}\label{linsource} {\bf Causal Path Sets.} Before giving formal definitions of paths and causal path sets, I briefly discuss the source objects of paths in causal theory, which are {\it linear directed sets.} Linear directed sets are generally nontransitive analogues of the linearly ordered sets discussed in section \hyperref[settheoretic]{\ref{settheoretic}} above. In this section, I often suppress binary relations, relation sets, etc., to avoid notational clutter. A directed set $L$ is {\bf linear} if it is acyclic, and if every pair of elements of $L$ is connected by a chain. The {\bf length} of a finite linear directed set $L$ is defined to be the chain length of its skeleton $\tn{sk}(L)$ in the sense of part 2 of definition \hyperref[definitionchains]{\ref{definitionchains}} above; i.e., its maximal number of consecutive {\it relations}. Infinite linear directed sets may be ``longer" and ``more complicated" than any acyclic chain. In particular, every acyclic chain may be represented as a relative directed set over $\mathbb{Z}$, but a linear directed set may a ``too large" to be represented in this way. For example, the directed set I refer to as {\it Jacob's ladder,} illustrated in figure \hyperref[jacobs]{\ref{jacobs}}c above, is linear, but does not admit a bijective morphism into any linear suborder of $\mathbb{Z}$. The analogue of length for an infinite linear directed set $L$ is the {\it isomorphism class of its transitive closure} $\tn{tr}(L)$ as a linearly ordered set. \\ \refstepcounter{textlabels}\label{pathsets} \begin{defi}\label{defipaths} Let $M=(M,R,i,t)$ be a multidirected set. \begin{enumerate} \item A {\bf (directed) path} in $M$ is a morphism $\gamma:L\rightarrow M$ for some linear directed set $L$, called the {\bf source} of $\gamma$, considered up to isomorphism. A path may also be called a {\bf causal path}. \item If the source $L$ of a path $\gamma$ is finite, the {\bf length} of $\gamma$ is defined to be the length of $L$. If $L$ is infinite, the isomorphism class of $\tn{tr}(L)$ supersedes the usual notion of length. \item The {\bf causal path set} $\Gamma(M)$ of $M$ is the set whose elements are paths in $M$. \item A path $\gamma:L\rightarrow M$ is called {\bf irreducible} if its source $L$ is irreducible. The {\bf irreducible path set} $\Gamma_{\tn{\fsz{irr}}}(M)$ of $M$ is the subset of $\Gamma(M)$ whose elements are irreducible paths. \end{enumerate} \end{defi} The image of an irreducible path $\gamma:L\rightarrow M$ is a chain in $M$, {\it not} necessarily irreducible, since ``pairs of elements may acquire new relations between them in the image." Every chain in $M$ may be identified with an appropriate irreducible path in $M$, so the chain set $\tn{Ch}(M)$ of $M$ may be identified with the irreducible path set $\Gamma_{\tn{\fsz{irr}}}(M)$ of $M$. Supplying appropriate directed structure to $\Gamma(M)$ and $\tn{Ch}(M)$, respectively, as described below, elevates these sets to {\it causal path spaces} and {\it chain spaces over} $M$, respectively. It is convenient in many cases to define this structure {\it algebraically,} in terms of {\it products of paths.} One of many ways to generalize the above definitions is to consider sets whose elements are paths {\it together with} distinguished subpaths of length zero, physically representing choices of {\it present} in linear sequences of events. Such sets are structurally analogous to {\it flag varieties} in algebraic geometry. Extending the analogy between causal theory and twistor theory mentioned in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} above, such ``enriched causal path sets" may be used to define {\it correspondences,} analogous to those described in \cite{WardWellsTwistor90}. However, besides a footnote in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} mentioning an analogous ``higher-level" construction in the context of kinematic schemes, I do not explore enriched causal path sets further in this paper. \refstepcounter{textlabels}\label{pathprod} {\bf Products of Paths; Causal Path Spaces.} A wide variety of algebraic operations may be defined on causal path sets over directed sets and multidirected sets. Two of the most important are special cases of {\it splice products,} which are {\it partially defined binary operations} formalizing different notions of ``joining morphisms together end-to-end in a temporal sense." Russling \cite{RusslingConcatenation95} discusses essentially equivalent notions in the special case of chains, and applies these constructions to problems in algorithmic graph theory. These operations define natural directed structures on causal path sets, elevating them to {\it causal path spaces.} The first such operation, called the {\it concatenation product,} applies to pairs of paths in a directed set $(D,\prec)$, for which the terminal element in the image of the first path {\it directly precedes} the initial element in the image of the second path. The images of such a pair of paths $\alpha$ and $\beta$ are illustrated in figures \hyperref[pathalgebras]{\ref{pathalgebras}}a and \hyperref[pathalgebras]{\ref{pathalgebras}}b below. \refstepcounter{textlabels}\label{concatprod} \vspace*{.2cm} \begin{defi}\label{deficoncatprod} Let $\alpha:L\rightarrow D$ and $\beta:L'\rightarrow D$ be paths in a directed set $D=(D,\prec)$. Suppose that $L$ has a maximal element $\ell_+$, and $L'$ has a minimal element $\ell_-'$, such that $\alpha(\ell_+)\prec\beta(\ell_-')$ in $D$. \begin{enumerate} \item Denote by $L\sqcup L'$ the linear directed set whose underlying set is the disjoint union $L\coprod L'$, and whose binary relation is given by augmenting the union of the binary relations on $L$ and $L'$ with the single relation $\ell_+\prec\ell_-'$, called the {\bf connecting relation}. Then the {\bf concatenation product} $\alpha\sqcup\beta: L\sqcup L' \rightarrow D$ of $\alpha$ and $\beta$ is defined by setting \[\alpha\sqcup\beta(\ell)=\begin{cases} \alpha(\ell) & \tn{if } \ell\in L\\ \beta(\ell) &\tn{if } \ell\in L' \end{cases}.\] \item The {\bf concatenation relation} $\prec_\sqcup$ on the causal path set $\Gamma(D)$ of $D$ is the binary relation given by setting $\alpha\prec_\sqcup\beta$ if and only if $\alpha\sqcup\beta$ exists. The {\bf causal concatenation space} over $(D,\prec)$ is the directed set $\big(\Gamma(D),\prec_\sqcup\big)$. \end{enumerate} \end{defi} \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.35cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,4.95)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(5.65,4.95)}{\pgfbox[left,center]{b)}} \pgfputat{\pgfxy(11.15,4.95)}{\pgfbox[left,center]{c)}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.25)(16.5,5.25) \pgfxyline(0,-.1)(0,5.25) \pgfxyline(5.5,-.1)(5.5,5.25) \pgfxyline(11,-.1)(11,5.25) \pgfxyline(16.5,-.1)(16.5,5.25) \begin{pgftranslate}{\pgfpoint{.2cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node20}{Node26} \end{pgfmagnify} \begin{pgfmagnify}{1.1}{1.1} \pgfputat{\pgfxy(.5,1)}{\pgfbox[left,center]{\large{$\alpha$}}} \pgfputat{\pgfxy(3.1,1.5)}{\pgfbox[left,center]{\large{$\gamma$}}} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.7cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node27}{Node30} \end{pgfmagnify} \begin{pgfmagnify}{1.1}{1.1} \pgfputat{\pgfxy(1.2,3.3)}{\pgfbox[left,center]{\large{$\beta$}}} \pgfputat{\pgfxy(3.5,3.3)}{\pgfbox[left,center]{\large{$\delta$}}} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node20}{Node26} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node27}{Node30} \color{white} \pgfmoveto{\pgfxy(1.8,2)} \pgflineto{\pgfxy(1.8,3.15)} \pgflineto{\pgfxy(3.7,3.15)} \pgflineto{\pgfxy(3.7,2)} \pgflineto{\pgfxy(1.8,2)} \pgffill \color{black} \pgfmoveto{\pgfxy(1.8,2)} \pgflineto{\pgfxy(1.8,3.15)} \pgflineto{\pgfxy(3.7,3.15)} \pgflineto{\pgfxy(3.7,2)} \pgflineto{\pgfxy(1.8,2)} \pgfstroke \pgfputat{\pgfxy(2.75,2.95)}{\pgfbox[center,center]{image of}} \pgfputat{\pgfxy(2.75,2.6)}{\pgfbox[center,center]{connecting}} \pgfputat{\pgfxy(2.75,2.25)}{\pgfbox[center,center]{relation}} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(1.8,2.6)(1.45,2.6) \pgfclearendarrow \pgfsetdash{{2pt}{2pt}}{0pt} \pgfnodeconnline{Node5}{Node14} \end{pgfmagnify} \begin{pgfmagnify}{1.1}{1.1} \pgfputat{\pgfxy(.3,2.4)}{\pgfbox[left,center]{\large{$\alpha\sqcup\beta$}}} \pgfputat{\pgfxy(3.8,2.6)}{\pgfbox[left,center]{\large{$\gamma\vee\delta$}}} \end{pgfmagnify} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Paths $\alpha$ and $\gamma$ in a directed or multidirected set; b) paths $\beta$ and $\delta$; c) concatenation product $\alpha\sqcup\beta$ and directed product $\gamma\vee\delta$.} \label{pathalgebras} \end{figure} \vspace*{-.5cm} The concatenation product is uniquely defined only for directed sets. In a multidirected set, there may be multiple connecting relations between $\alpha(\ell_+)$ and $\beta(\ell_-')$. Unlike the concatenation product, the concatenation {\it relation} $\prec_\sqcup$ may be generalized in an obvious way to yield a multidirected structure on the causal path set over a multidirected set, but I do not pursue this generalization here. The notation $\sqcup$ for the concatenation product is deliberately chosen to evoke the disjoint union $\coprod$ of sets. The concatenation relation $\prec_\sqcup$ on $\Gamma(D)$ extends the binary relation $\prec$ on $D$. Hence, $(D,\prec)$ embeds as a full subobject into its causal concatenation space. The second operation on causal path sets, called {\it directed product,} applies to pairs of paths in a multidirected set $M=(M,R,i,t)$, for which the terminal element in the image of the first path {\it coincides with} the initial element in the image of the second path. The images of such a pair of paths $\gamma$ and $\delta$ are illustrated in figures \hyperref[pathalgebras]{\ref{pathalgebras}}a and \hyperref[pathalgebras]{\ref{pathalgebras}}b above.\\ \refstepcounter{textlabels}\label{dirprod} \begin{defi}\label{defidirprod} Let $\alpha:L\rightarrow M$ and $\beta:L'\rightarrow M$ be paths in a multidirected set $M=(M,R,i,t)$. Suppose that $L$ has a terminal element $\ell_+$, and $L'$ has an initial element $\ell_-'$, such that $\alpha(\ell_+)=\beta(\ell_-')$ in $M$. \begin{enumerate} \item Denote by $L\vee L'$ the linear directed set whose underlying set is the quotient space of the disjoint union $L\coprod L'$, given by identifying $\ell_+$ and $\ell_-'$, and whose binary relation is the union of the binary relations on $L$ and $L'$. Then the {\bf directed product} $\alpha\vee\beta: L\vee L' \rightarrow M$ of $\alpha$ and $\beta$ is defined by setting \[\alpha\vee\beta(\ell)=\begin{cases} \alpha(\ell) & \tn{if } \ell\in L \\ \beta(\ell) &\tn{if } \ell\in L' \end{cases}.\] \item The {\bf directed product relation} $\prec_\vee$ on the causal path set $\Gamma(M)$ of $M$ is the binary relation given by setting $\alpha\prec_\vee\beta$ if and only if $\alpha\vee\beta$ exists. The {\bf causal directed product space} over $M$ is the directed set $\big(\Gamma(M),\prec_\vee\big)$. \end{enumerate} \end{defi} The notation $\vee$ for the directed product is deliberately chosen to evoke the {\it wedge sum} $\bigvee$, familiar from topology, which joins topological spaces at distinguished basepoints. The directed product relation $\prec_\vee$ on $\Gamma(M)$ extends the induced relation on the relation space $\ms{R}(M)$ over $M$. Hence, $\ms{R}(M)$ embeds as a full subobject into the causal directed product space over $M$. \refstepcounter{textlabels}\label{spliceprod} The causal concatenation space and the causal directed product space are prototypical examples of {\bf causal path spaces}, which are causal path sets over a directed or multidirected set equipped with natural directed or multidirected structures. The concatenation product $\sqcup$ and the directed product $\vee$ are the simplest examples of {\bf splice products}, defined by ``splicing together morphisms." The corresponding relations $\prec_\sqcup$ and $\prec_\vee$ are the simplest examples of {\it splice relations,} already introduced in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} above. \refstepcounter{textlabels}\label{catsemicat} {\bf Categories and Semicategories.} I now introduce another analogy between causal theory and category theory, similar to those already cited in the fifth argument against transitivity in section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}}, and the discussion of morphism categories near the end of section \hyperref[subsectionrelation]{\ref{subsectionrelation}}. These analogies arise essentially because both causal theory and category theory involve notions of {\it multidirectedness}, {\it preservation of structure,} and {\it hierarchy.} Category theory was first developed in the 1940's by Eilenberg and MacLane, to study relationships among different {\it cohomology theories}, which are special families of functors, often from a category of topological spaces to a category of groups. In this context, additional notions such as {\it composition,} {\it associativity,} and {\it existence of identities} are natural, but these notions are perhaps less relevant to applications of directed and multidirected structure in theoretical physics.\footnotemark\footnotetext{Note the contrast with Chris Isham's viewpoint; Isham {\it adds additional properties} to categories to obtain {\it topoi,} whereas I emphasize certain category-theoretic properties, and de-emphasize others. Of course, Isham works with topoi of many different kinds, not only those whose objects are intended to model causal structure.} For example, elements and relations in a multidirected set may be compared to objects and morphisms of a category, respectively, but there are important qualitative differences in their formal properties: the ``composition" of relations is generally not defined, at least, not as a relation, and associativity makes no sense without composition. Identity morphisms, and more generally, endomorphisms, correspond to reflexive relations, which are not guaranteed to exist in the general case, and are forbidden in the acyclic case. {\it Semicategories,} or {\it non-unital categories,} are similar to categories, but without the guaranteed existence of identity morphisms. Causal path spaces over directed or multidirected sets, such as the causal concatenation space and the causal directed product space, may be viewed as {\it generalized semicategories,} admitting ``morphisms without sources and/or targets."\footnotemark\footnotetext{Heuristically, one may think of a field in which field lines either terminate on charges or extend to infinity.} More precisely, a {\bf semicategory} consists of a class of objects $\ms{S}$, and a class of morphisms $\Gamma(\ms{S})$, such that every morphism $\gamma$ has an initial object $\mbf{i}(\gamma)$ and a terminal object $\mbf{t}(\gamma)$. In the context of causal theory, $\ms{S}$ may be viewed as an abstraction of the underlying set of elements of a multidirected set $M=(M,R,i,t)$, while $\Gamma(\ms{S})$ may be viewed as an abstraction of the class of {\it finite chains} in $M$. However, generalizations of these analogies, involving the entire causal path set $\Gamma(M)$ of $M$, are needed below. The maps $\mbf{i}:\Gamma(\ms{S})\rightarrow \ms{S}$ and $\mbf{t}:\Gamma(\ms{S})\rightarrow \ms{S}$ may be viewed as abstractions of the initial and terminal element maps $i$ and $t$ of $M$, extended to apply to finite chains of arbitrary length, rather than merely relations. A semicategory $\ms{S}$ also admits {\it associative compositions} of morphisms, defined whenever the terminal object of the first morphism coincides with the initial object of the second. This composition may be viewed as an abstraction of the directed product of chains. Note that although the ``composition" of relations is not a {\it relation} in this context, it {\it is} a chain, more specifically, a $2$-chain. The analogy between complex chains and morphisms should not be carried too far. In section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, I compared $2$-chains to $2$-morphisms, rather than morphisms, and this is the {\it preferred structural analogy for causal theory,} at least in my opinion. The semicategory viewpoint regarding causal path spaces is closer to the category-theoretic approach to {\it groups}, mentioned in sections \hyperref[settheoretic]{\ref{settheoretic}} and \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}} above, in which the elements of a group are viewed as morphisms in a category with a single object. In this sense, a semicategory is a {\it grouplike structure} whose morphisms are its {\it elements,} and composition is viewed as a {\it partially defined algebraic operation} on these elements. From this perspective, a semicategory is simply a class together with an associative partially defined operation, generally noncommutative, and without identity or inverses. Familiar examples of such structures are more common than one might expect. These include the space of continuum paths in a topological space under the operation of directed product, the class of {\it all} matrices over a given ring under the operation of matrix multiplication, and the class of algebraic cycles on an algebraic variety under the operation of {\it proper} intersection.\footnotemark\footnotetext{In all three of these examples, information may be discarded to obtain a structure that is better-behaved mathematically, but less complete and natural; namely, fundamental groupoids or groups, matrix groups, and Chow rings, respectively.} \refstepcounter{textlabels}\label{causalpathsemicat} {\bf Causal Path Semicategories.} The causal path set $\Gamma(D)$ of a directed set $D=(D\prec)$, equipped with the partially defined operation of concatenation product $\sqcup$, is referred to in this paper as the {\bf concatenation semicategory} over $D$. Similarly, the causal path set $\Gamma(M)$ of a multidirected set $M=(M,R,i,t)$, equipped with the partially defined operation of directed product $\vee$, is referred to as the {\bf directed product semicategory} over $M$. Both of these structures, viewed as ``group-like objects," are examples of what I refer to as {\bf causal path semicategories}, which represent the {\it algebraic viewpoint} regarding directed structure on causal path sets, in contrast to the {\it generalized order-theoretic viewpoint} represented by causal path spaces. The ``if and only if" statements relating $\sqcup$ and $\prec_\sqcup$ in part 2 of definition \hyperref[deficoncatprod]{\ref{deficoncatprod}}, and $\vee$ and $\prec_\vee$ in part 2 of definition \hyperref[defidirprod]{\ref{defidirprod}} above, ensure that the two viewpoints are equivalent in the present context.\footnotemark\footnotetext{Of course, the concatenation relation $\prec_\sqcup$ generalizes to yield a multidirected structure on causal path sets over multidirected sets, while the concatenation product $\sqcup$ is uniquely defined only for directed sets, as explained above.} Causal path semicategories are {\it generalized semicategories,} whose morphisms, in this case paths, do not necessarily possess initial or terminal objects, and whose composition law, for morphisms $\gamma$ and $\delta$, is not necessarily defined in terms of the condition $\mbf{t}(\gamma)=\mbf{i}(\delta)$. In particular, different splice products on the same underlying causal path set define different causal path semicategories, with different composition laws. Interesting relationships exist among semicategory operations on causal path sets, and functors such as the relation space functor $\ms{R}$ and abstract element space functor $\ms{E}$. For example, the concatenation semicategory over the relation space $\ms{R}(M)$ over a multidirected set $M=(M,R,i,t)$ corresponds to a subsemicategory of the directed product semicategory over $M$ itself, as illustrated by the following example: if $r$ and $s$ are elements of $\ms{R}(M)$, with initial and terminal elements $i(r)=x$, $t(r)=i(s)=y,$ and $t(s)=z$, in $M$, then the directed product in $\Gamma(M)$ of $r$ and $s$, viewed as paths of length $1$, corresponds to the concatenation product in $\Gamma(\ms{R}(M))$ of $r$ and $s$, viewed as paths of length zero. This illustrates once again how the relation space functor $\ms{R}$ {\it reduces multidirected structure to directed structure,} since concatenation products are uniquely defined only in the directed case. The physical importance of {\it both} element space and relation space is the reason for introducing both types of semicategories. Since the operations $\sqcup$ and $\vee$ preserve irreducibility of paths, useful {\it subsemicategories} of $(\Gamma(D),\sqcup)$ and $(\Gamma(M),\vee)$ may be defined by restricting attention to the irreducible path sets of $M$ and $D$, which may be identified with the chain sets $\tn{Ch}(D)$ and $\tn{Ch}(M)$, respectively. I refer to these subsemicategories as the {\bf chain concatenation semicategory} over $D$, and the {\bf chain directed product semicategory} over $M$, respectively. These {\bf chain semicategories} $\big(\tn{Ch}(D),\sqcup\big)$ and $\big(\tn{Ch}(M),\vee\big)$ belong to a special class of causal path semicategories which may be studied without explicit reference to morphisms.\footnotemark\footnotetext{Another case in which explicit reference to morphisms can be avoided is the case in which all paths under consideration are {\it isomorphisms onto their images,} viewed as subobjects $M$. However, this case does not behave well algebraically. In particular, splice products of such paths are generally {\it not} isomorphisms onto their images.} They sometimes appear in the literature as implicit structural substrata for multiplicative operations in special types of {\it path algebras,} discussed in detail below. Though I focus mostly on the special case of chains in this paper, there are good reasons to develop the theory in terms of general paths: it makes the relative viewpoint (\hyperref[rv]{RV}) explicit, enables more refined treatment of target objects, and includes paths of transfinite length. Special {\it quotient semicategories} also play a significant role in the theory of causal path semicategories. In this context, quotient semicategories may be viewed concretely as families of {\it conguence classes of paths.} A {\it congruence relation} $\sim$ on an arbitrary semicategory $(\Gamma,\sqcup)$ is a reflexive, symmetric, transitive binary relation on $\Gamma$, preserving the semicategory operation $\sqcup$, in the sense that for any four elements $\alpha,\beta,\gamma,\delta$ of $\Gamma$, if $\alpha\sim\beta$ and $\gamma\sim\delta$, then $\alpha\sqcup\gamma\sim\beta\sqcup\delta$. Important congruence relations on $(\Gamma(D),\sqcup)$ and $(\Gamma(M),\vee)$ are given by taking two paths to be congruent if and only if their images share the same initial and terminal elements. Such congruence relations appear in the theory of {\it co-relative histories,} developed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} below, where they identify ``congruent families of kinematic accounts of evolutionary processes." \refstepcounter{textlabels}\label{causalpathalg} {\bf Causal Path Algebras.} Causal path semicategories are ``primitive" algebraic objects, in the sense that they involve partially defined operations that fail to satisfy a number of familiar algebraic properties. Richer objects, called {\it causal path algebras,} may be constructed from causal path semicategories, by incorporating the structure of a ring. An analogy from elementary algebra is useful to motivate this construction. Multiplication of monomials with real coefficients involves {\it two} distinct operations: multiplication of coefficients in $\mathbb{R}$, and addition of exponents in $\mathbb{N}$. Thus, the additive structure of $\mathbb{N}$ serves as a {\it substratum for multiplication} in the polynomial algebra $\mathbb{R}[x]$. More generally, {\it group} and {\it semigroup algebras} combine the structure of a ring, analogous to $\mathbb{R}$, with the structure of a group or semigroup, analogous to $\mathbb{N}$, which participates only in the multiplicative operation of the resulting algebra. Still more general are {\it semicategory algebras}, which derive a portion of their multiplicative structure from a semicategory operation. Two important types of (generalized) semicategory algebras arising naturally in discrete causal theory are {\it concatenation algebras} and {\it directed product algebras}, constructed by combining the structure of a ring $T$ with appropriate semicategory operations on causal path sets. I refer to both types of algebras as {\bf causal path algebras}, since they are generated by causal path semicategories. The notation $T$ in this context is intended to evoke the word {\it ``target,"} since $T$ serves as a {\it target object for path functionals,} as discussed in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}}. $T$ may be replaced with an object having less structure than a ring, such as an abelian group, or by a structure at a different level of algebraic hierarchy, such as a category. Correspondingly, causal path algebras may be replaced by {\it causal path modules,} or by higher-level structures such as {\it generalized enriched functor categories.} Generalizing beyond the context of paths, any splice product for a class of morphisms into a multidirected set $M$ may be used to define a {\it causal splice module} or {\it causal splice algebra} over $M$. Alternatively, one may simultaneously consider many distinct splice products on the same class of morphisms.\footnotemark\footnotetext{{\it Vertex operator algebras,} used extensively in string theory, provide familiar examples of algebraic objects involving many distinct operations.} Concatenation algebras are defined in the context of {\it directed sets,} since their multiplicative structures arise from concatenation products:\\ \begin{defi}\label{deficoncatalg} Let $D=(D,\prec)$ be a directed set, with causal path set $\Gamma(D)$. Let $T$ be a ring, usually assumed to be commutative with unit. \begin{enumerate} \item The {\bf concatenation algebra} $T^\sqcup(D)$ over $D$, with coefficients in $T$, is the algebra whose elements are formal sums $\sum_\gamma t_\gamma\gamma$, where $\gamma\in\Gamma(D)$ and $t_\gamma\in T$, and whose multiplication is defined by multiplying coefficients in $T$, and forming concatenation products of paths in $\Gamma(D)$, whenever these products exist: \[\Big(\sum_\gamma t_\gamma\gamma\Big)\sqcup\Big(\sum_{\gamma'} t_{\gamma'}\gamma'\Big):=\sum_{\gamma\prec_\sqcup\gamma'}t_\gamma t_{\gamma'}\gamma\sqcup\gamma'.\] \item The {\bf chain concatenation algebra} $T^\sqcup[D]$ over $D$, with coefficients in $T$, is the subalgebra of $T^\sqcup(D)$ given by restricting $\gamma$ to $\Gamma_{\tn{\fsz irr}}(D)=\tn{Ch}(D)$. \end{enumerate} The empty sum is identified with the additive identity $0$ in both algebras. \end{defi} Directed product algebras are defined in the context of multidirected sets, since their multiplicative structures arise from directed products:\\ \begin{defi}\label{defidirprodalg} Let $M=(M,R,i,t)$ be a multidirected set, with path set $\Gamma(M)$. Let $T$ be a ring, usually assumed to be commutative with unit. \begin{enumerate} \item The {\bf directed product algebra} $T^\vee(M)$ over $M$, with coefficients in $T$, is the algebra whose elements are formal sums $\sum_\gamma t_\gamma\gamma$, where $\gamma\in\Gamma(M)$ and $t_\gamma\in T$, and whose multiplication is defined by multiplying coefficients in $T$, and forming directed products of paths in $\Gamma(M)$, whenever these products exist: \[\Big(\sum_\gamma t_\gamma\gamma\Big)\vee\Big(\sum_{\gamma' }t_{\gamma'}\gamma'\Big):=\sum_{\gamma\prec_\vee\gamma'}t_\gamma t_{\gamma'}\gamma\vee\gamma'.\] \item The {\bf chain directed product algebra} $T^\vee[M]$ over $M$, with coefficients in $T$, is the subalgebra of $T^\vee(M)$ given by restricting $\gamma$ to $\Gamma_{\tn{\fsz irr}}(M)=\tn{Ch}(M)$. \end{enumerate} The empty sum is identified with the additive identity $0$ in both algebras. \end{defi} The causal path algebras introduced in definitions \hyperref[deficoncatalg]{\ref{deficoncatalg}} and \hyperref[defidirprodalg]{\ref{defidirprodalg}} above are noncommutative {\it generalized graded algebras,} since the ``length functionals" encoding the lengths of paths in $D$ and $M$ are ``additive"\footnotemark\footnotetext{This vague description is due to obvious caveats, such as the fact that the length of $\gamma\sqcup\delta$, when this product exists, is not the sum of the lengths of $\gamma$ and $\delta$, but the sum of these lengths {\it plus one.} Further, additivity has a generalized meaning, in terms of isomorphism classes, for paths of infinite length.} in a generalized sense. In particular, for each nonnegative integer $n$, the additive subgroups of these algebras generated by paths or chains of length $n$ are their ``$n$th graded pieces." From a purely algebraic viewpoint, this generalized graded structure is ultimately responsible for many of the most useful properties of causal path algebras, but I do not emphasize this perspective in this paper. It is sometimes useful to consider subalgebras of {\it finite} formal sums in causal path algebras. In this paper, I usually avoid this subtlety by working with {\it finite} subsets of the causal path sets $\Gamma(D)$ and $\Gamma(M)$. This restriction also avoids the use of cumbersome notation, such as $T^\sqcup\llbracket D\rrbracket$, to denote algebras of formal sums without finiteness assumptions. \refstepcounter{textlabels}\label{conventionalpath} {\bf Conventional Applications of Path Algebras.} Path algebras similar or equivalent to causal path algebras arise naturally in graph theory, information theory, computer science, network theory, category theory, noncommutative geometry, algebraic $K$-theory, and of course causal set theory. For example, in the special case where the coefficient ring $T$ is a field, the subalgebra of finite formal sums in the chain directed product algebra $T^\vee[M]$ is the {\it standard path algebra} appearing in \cite{AbramsPathAlgebra05}. The chain concatenation algebra $T^\sqcup[D]$ also appears in the literature as a special type of {\it quiver algebra}. Similar quiver algebras play important roles in recent and ongoing research in noncommutative algebra and noncommutative geometry, particularly in the study of so-called {\it quantum groups}, which are {\it deformations of Hopf algebras.} Path algebras possess a number of attractive mathematical properties, besides the generalized graded structure already noted above. The chain directed product algebra $T^\vee[A]$ over a finite acyclic directed set $A=(A,\prec)$ will serve for illustrative purposes. $T^\vee[A]$ has a multiplicative identity given by the formal sum of the elements of $A$. Each element $x$ in $A$ defines an {\it idempotent operator} on $T^\vee[A]$, whose image is an {\it indecomposable projective} $T^\vee[A]$-{\it module} $P(x)$, called the {\it punctual module} of $x$. Every other chain in $A$ has square zero. $T^\vee[A]$ is a {\it hereditary algebra}, which means that any submodule of a projective module over $T^\vee[A]$ is itself projective; this follows from the fact that every projective module over $T^\vee[A]$ is built up from the punctual modules $P(x)$. These properties play an important role in the study of related {\it module categories,} {\it derived categories,} and {\it cluster categories.} For further details, I suggest Bernhard Keller's recent paper {\it Cluster Algebras, Quiver Representations and Triangulated Categories} \cite{KellerClusterAlgebras10}. Some of these properties, or suitable analogues thereof, still apply for general multidirected sets, but others break down. For example, the formal sum of elements of a multidirected set $M$ is not a multiplicative identity for $T^\vee[M]$ if $M$ has unbounded chains, and a reflexive relation in $M$ does not have square zero, but generates an infinite cyclic multiplicative subgroup of $T^\vee[M]$. \refstepcounter{textlabels}\label{raptisincidence} {\bf Raptis' Incidence Algebra.} In his 2000 paper {\it Algebraic Quantization of Causal Sets} \cite{RaptisAlgebraicQuantization00}, Ioannis Raptis employs a noncommutative algebra called an {\it incidence algebra,} which is a close transitive analogue of the directed product algebra, with coefficients in $\mbb{C}$, over a locally finite acyclic directed set. As described by Raptis, a suggestive `ket-bra' notation may be used to represent elements of this algebra. In particular, each relation $x\prec y$ in a causal set $C$ may be viewed as an {\it operator} $|x\rangle\langle y|$, and relations may be ``multiplied" according to the formula \begin{equation} \label{shortcircuit}|x\rangle\langle y|u\rangle\langle v|=\langle y|u\rangle|x\rangle\langle v|, \hspace*{.5cm}\tn{where}\hspace*{.5cm} \langle y|u\rangle:=\delta_{yu}=\begin{cases} 1& \tn{if } y=u \\ 0&\tn{if } y\ne u,\end{cases} \end{equation} where the expression $\delta_{yu}$ denotes the {\it Kronecker delta function}, defined by the cases on the far right of equation \hyperref[shortcircuit]{\ref{shortcircuit}}. This definition requires the causal set axiom of transitivity (\hyperref[tr]{TR}), since the relation $x\prec v$ may not exist in the nontransitive case. Equation \hyperref[shortcircuit]{\ref{shortcircuit}} defines a semicategory operation on the relation space over a causal set, with Raptis' incidence algebra arising as the corresponding semicategory algebra with complex coefficients. The operation described by equation \hyperref[shortcircuit]{\ref{shortcircuit}} involves {\it loss of information} compared with the product operation in the corresponding directed product algebra, since there may be many distinct $2$-chains between a given pair of elements $x$ and $v$ in a causal set. Hence, Raptis' incidence algebra may be viewed as a special type of ``short-circuit algebra," a descriptive term for a ``degenerate splice algebra," which fails to preserve information. Raptis describes addition in the incidence algebra over a causal set $C$ as representing {\it quantum-theoretic superposition.} While this is a reasonable physical interpretation, it can only apply in the background-dependent context, since it involves combining algebraic data associated with substructures of a {\it particular} causal set. More relevant to the approach developed in this paper are the {\it monoids of arrow fields} used by Chris Isham in his {\it quantization on a category} \cite{IshamQuantisingI05}, described in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} below, which may be viewed as ``higher-level multiplicative analogues" of Raptis' incidence algebras. \subsection{Path Summation over a Multidirected Set}\label{subsectionpathsummation} \refstepcounter{textlabels}\label{pathfunctcont} \refstepcounter{textlabels}\label{lagrangehamilton} {\bf Path Functionals; Motivation from Continuum Theory.} I begin this section by briefly reviewing some standard notions from continuum mechanics, in preparation for analogous constructions in the context of multidirected sets. The reader should be aware that some of the notation appearing in this section, such as the standard notation for the {\it classical action} and the {\it Lagrangian,} is similar to notation I have used for completely different purposes elsewhere in the paper.\footnotemark\footnotetext{Actually, the mathscript symbol $\ms{L}$, which I use for the Lagrangian, is often used to denote a Lagrangian {\it density} in the context of field theory. In this paper, however, mathscript symbols generally appear at the ``category-functor level of hierarchy," which is appropriate for the Lagrangian. In any case, densities do not arise in the locally finite context.} {\bf Path functionals} are maps from a space of paths into a target object, such as an abelian group, a ring, or a field. In the study of particle motion in classical mechanics, the {\bf classical action} $\ms{S}$ is a real-valued path functional, given by integrating the {\bf Lagrangian} $\ms{L}$, which describes the difference between a particle's kinetic and potential energies, with respect to time, over a spacetime path $\gamma$, viewed as a possible particle trajectory: \begin{equation} \ms{S}(\gamma)=\int_\gamma \ms{L}dt. \end{equation} {\bf Hamilton's principle} states that {\it the classical trajectory of the particle renders the classical action stationary.} Heuristically, this means that the Lagrangian $\ms{L}$ ``chooses" a trajectory, determined by how $\ms{S}$ varies with $\gamma$. This is a more {\it holistic} interpretation of classical mechanics than the interpretation provided by Newtonian theory, since the value of the action $\ms{S}(\gamma)$ depends on the entire trajectory $\gamma$. Thus, Hamilton's principle provides a classical example of the general rule, mentioned in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} above, that a holistic viewpoint may be {\it useful}, even when the reductionist approach is information-theoretically adequate. The Lagrangian $\ms{L}$ is an {\bf infinitesimal path functional}, meaning that it depends only on the position and motion of the particle at a given time, not on the details of its trajectory. In fact, the Lagrangian is usually taken to depend only on {\it information up to first order}; i.e., on position, time, and velocity. As such, it may be viewed as a function on the {\it real tangent bundle} over continuum spacetime.\footnotemark\footnotetext{Similarly, important causal analogues of Lagrangians may be viewed as {\it relation functions,} as described later in this section. The reader may recall here the comparison between the relation set $R$ of a multidirected set $M=(M,R,i,t)$ and the {\it tangent sheaf} in geometry, mentioned in the last footnote in section \hyperref[subsectionacyclicdirected]{\ref{subsectionacyclicdirected}} above.} This restriction makes certain computations easier; for example, it simplifies the calculation of Feynman's path integral. Generalizations of $\ms{L}$ and $\ms{S}$ are many and varied, but the primitive versions cited here will serve for the purposes of illustration and analogy. In Feynman's original version of the histories approach to quantum theory, described in more detail in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} below, Feynman's {\bf phase map} $\Theta$ is a path functional assigning to each particle trajectory $\gamma$ the {\it complex exponential of its classical action, in units of Planck's reduced constant:} \begin{equation}\label{feynmanphase} \Theta(\gamma)=e^{\frac{i}{\hbar}\ms{S}(\gamma)}. \end{equation} The classical action $\ms{S}$ is additive for directed products of paths\footnotemark\footnotetext{This means essentially the same thing in the continuum context as in the context of multidirected sets: a pair of continuum paths has a directed product ``joining the two paths together" if and only if the terminal element in the image of the first path coincides with the initial element in the image of the second. This product is particularly familiar in homotopy theory, where it descends to multiplication in the fundamental groupoid.} by the definition of integration. Hence, Feynman's phase map is multiplicative, since exponentiation converts addition to multiplication. In more sophisticated terms, $\ms{S}$ is a ``semicategory morphism" from an appropriate causal path semicategory over spacetime to the underlying additive group of $\mathbb{R}$, and $\Theta$ is a ``semicategory morphism" from the same causal path semicategory into the unit circle, viewed as a multiplicative subgroup of the complex numbers $\mathbb{C}$. \refstepcounter{textlabels}\label{pathfunctac} {\bf Path Functionals for Multidirected Sets. } The analogue of an infinitesimal path functional, such as a Lagrangian, on the causal path set $\Gamma(M)$ over a multidirected set $M=(M,R,i,t)$, is a {\bf relation function}; i.e., a map $\theta:R\rightarrow T$, for an appropriate target object $T$. The analogue of an ordinary path functional is a map $\Theta:\Gamma(M)\rightarrow T$, which I refer to as a {\bf causal path functional}, or simply a {\bf phase map.} Using the latter term, the individual value $\Theta(\gamma)$ of $\Theta$ on a path $\gamma$ may be referred to as the {\bf phase} of $\gamma$ with respect to $\Theta$, which is both concise and physically suggestive. In the cases of principal interest in this paper, $\Gamma(M)$ is equipped with a semicategory operation, such as the directed product, and $\Theta$ is a ``semicategory morphism," analogous to Feynman's phase map. Depending on the context, the phases $\Theta(\gamma)$ of paths $\gamma$ with respect to a phase map $\Theta$ may reside in a group, a ring, a field, or perhaps an algebraic object physically suggestive in some specific way, such as an operator algebra.\footnotemark\footnotetext{For example, Isham uses functionals whose values are Hilbert space operators in his {\it quantization on a category} \cite{IshamQuantisingI05}.} At higher levels of algebraic hierarchy, such as in the theory of kinematic schemes, path functionals may be viewed as {\it assignments,} analogous to functors. For example, in the language of section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below, there exist ``higher-level path functionals" assigning to each co-relative history in a kinematic scheme its {\it causal Galois groups.}\footnotemark\footnotetext{It is actually preferable in many ways to think of phase maps as generalized functors even when working over an individual directed set, rather than a kinematic scheme. However, it is not feasible to undertake a formal exploration of hierarchical considerations in this paper.} A phase map $\Theta$ on the causal path space $\Gamma(M)$ over a multidirected set $M$ may be regarded as a {\it formal sum} $\sum_{\gamma\in\Gamma(M)}\Theta(\gamma)\gamma$, since this expression uniquely pairs each path with its corresponding phase. This viewpoint immediately evokes the causal path algebras described in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} above. Indeed, in the case where $T$ is a ring, a phase map $\Theta$ may be regarded as {\it supplying the coefficients of a causal path algebra element.} Conversely, each element of a causal path algebra specifies a unique phase map. Hence, the causal path algebras $T^\sqcup(D)$ and $T^\vee(M)$ supply multiplicative structures on the {\bf mapping spaces} $T^{\Gamma(D)}$ and $T^{\Gamma(M)}$ of maps $\Gamma(D)\rightarrow T$ and $\Gamma(M)\rightarrow T$, respectively. For example, if $\Theta\leftrightarrow\sum\Theta(\gamma)\gamma$ and $\Psi\leftrightarrow\sum\Psi(\gamma)\gamma$ are elements of $T^{\Gamma(D)}$ or $T^{\Gamma(M)}$; i.e., phase maps, then the formulae \begin{equation}\label{convolutions} \Phi\sqcup\Psi=\sum_{\gamma\prec_\sqcup\gamma'}\Phi(\gamma)\Psi(\gamma')\gamma\sqcup\gamma'\hspace*{.7cm}\tn{and}\hspace*{.7cm}\Phi\vee\Psi=\sum_{\gamma\prec_\vee\gamma'}\Phi(\gamma)\Psi(\gamma')\gamma\vee\gamma', \end{equation} given by multiplication in $T^\vee(D)$ and $T^\sqcup(M)$, respectively, define analogues of {\it convolution}\footnotemark\footnotetext{For similar reasons, certain types of path algebras are sometimes called {\it convolution algebras} in the literature.} on $T^{\Gamma(D)}$ and $T^{\Gamma(M)}$, respectively. Using the concatenation algebra as an example, the similarity of such a product to ordinary convolution of real valued functions, given by the formula \[f*g(t)=\int_\tau f(\tau)g(t-\tau)d\tau,\] may be seen by expressing the value $\Phi\sqcup\Psi(\beta)$, for a particular path $\beta$, as \[\Phi\sqcup\Psi(\beta)=\sum_{\alpha}\Phi(\alpha)\Psi(\beta-\alpha),\] where $\alpha$ ranges over all paths that ``may be extended in an appropriate way\footnotemark\footnotetext{This may be formalized via the theory of {\it transitions,} introduced in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} below, since the image of $\alpha$ embeds as a {\it full originary subobject} of the image of $\beta$. } to produce $\beta$," and where the expression $\beta-\alpha$ means ``the path such that $\alpha\sqcup(\beta-\alpha)=\beta$." Similarly, the chain algebras $T^\sqcup[D]$ and $T^\vee[M]$ supply multiplicative structures on the mapping spaces $T^{\tn{Ch}(D)}$ and $T^{\tn{Ch}(M)}$, while the corresponding causal path algebras of {\it finite} formal sums supply multiplicative structures on the spaces of {\it finitely supported} maps $\Gamma(D)\rightarrow T$ and $\Gamma(M)\rightarrow T$, where the {\it support} of a phase map $\Theta$ is the set of all paths $\gamma$ in $\Gamma(D)$ or $\Gamma(M)$ {\it not} mapping to $0_T$ under $\Theta$. \refstepcounter{textlabels}\label{pathsumac} {\bf Path Summation over a Multidirected Set.} Fortunately, there is no need to discuss the general theory of path integration in the continuum context before introducing the analogous theory of path summation over a multidirected set. Indeed, many of the difficulties of continuum path integration are irrelevant to the latter theory, at least in the locally finite case. In section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} below, I discuss in some detail the specific methods of continuum path integration used by Richard Feynman to formulate his original version of the histories approach to quantum theory, since these constructions are directly relevant to the corresponding quantum causal theory. Here, I merely describe the rudiments of path summation over a multidirected set, in the case where the target object for phase maps is a commutative ring with unit. I treat phase maps and path algebra elements interchangeably, as needed. Let $M=(M,R,i,t)$ be an a multidirected set, with causal path set $\Gamma(M)$, and let $T$ be a commutative ring with unit. The {\bf global evaluation map} $e_M:T^{\Gamma(M)}\dashedrightarrow T$ is the {\it partially defined} map given by ``collapsing" the formal sum associated with a phase map $\Theta$, to yield the corresponding {\it concrete sum} of elements of $T$, whenever it exists: \begin{equation}\label{globalevaluation} e_M(\Theta)=e_M\Bigg(\sum_{\gamma\in\Gamma(M)}\Theta(\gamma)\gamma\Bigg)= \sum_{\gamma\in\Gamma(M)}\Theta(\gamma). \end{equation} The global evaluation map $e_M$ may be expressed in terms of the {\bf indicator functions} $e_\gamma:\Gamma(M)\rightarrow T$, defined by the formulae \[e_\gamma(\gamma')=\delta_{\gamma\gamma'}=\begin{cases} 1_T & \tn{if } \gamma'=\gamma \\ 0_T &\tn{if } \gamma'\ne\gamma.\end{cases}\] The indicator functions $e_\gamma$ extend by $T$-linearity to yield totally defined functions $T^{\Gamma(M)}\rightarrow T$, given by the formulae $e_\gamma(\Theta)=e_\gamma\big(\sum_{\gamma'\in\Gamma(M)}\Theta(\gamma')\gamma'\big)=\Theta(\gamma)$. This particular application of $e_\gamma$ is roughly analogous to integrating a continuum function against a Dirac delta function, to obtain a function value at a particular point. Restricting to the case of finitely-supported phase maps, the formal sum of indicator functions gives a totally defined evaluation map. In the general case, however, the right-hand side of equation \hyperref[globalevaluation]{\ref{globalevaluation}} may not converge to an element of $T$. Let $\Theta:\Gamma(M)\rightarrow T$ a phase map. The $T$-valued {\bf path sum} $\Theta(M)$ over $M$ with respect to $\Theta$, {\it if it exists,} is the image of the global evaluation map: \begin{equation}\label{pathsum} \Theta(M):=e_{M}(\Theta)=\sum_{\gamma\in\Gamma(M)}\Theta(\gamma). \end{equation} If $M'$ is a subset of $M$, viewed as a full subobject, then $\Theta$ also defines an $T$-valued path sum $\Theta(M')$ over $M'$, given by restricting $\gamma$ to belong to $\Gamma(M')$. This sum may also be expressed in terms of indicator functions in the obvious way. The assignment $M'\mapsto \Theta(M')$ is a partially-defined {\it power set function} $\ms{P}(M)\dashedrightarrow T$, which may be viewed as a $T$-valued {\it generalized measure} on $M$. More generally, one may restrict summation to other subsets of the causal path set $\Gamma(M)$, not necessarily those consisting of {\it all} paths into a particular subobject of $M$. In particular, sets of {\it maximal chains} in finite subsets of directed sets, particularly acyclic directed sets, are important index sets for summation in quantum causal theory. \newpage \section{Quantum Causal Theory}\label{subsectionquantumcausal} In this section, I introduce a new version of quantum causal theory, using locally finite directed sets as underlying models of classical spacetime. Passage to the quantum regime is accomplished via an appropriate adaptation of the {\it histories approach to quantum theory,} expressed in terms of path summation over multidirected sets. The presentation in this section is not intended to be comprehensive, and involves a number of simplifying assumptions. Despite this, the approach introduced here is very general at the conceptual level. In particular, it may be used to construct {\it either} background-dependent theories of fields and particles existing {\it on} directed sets, {\it or} background-independent theories involving {\it configuration spaces of directed sets.} The former theories are analogous to quantum field theories in curved spacetime, while the latter theories provide suitable tools for studying the problems of fundamental spacetime structure, quantum gravity, and unification in the discrete quantum causal context. Applicability of the same abstract approach to both types of theories is due to the principle of {\it iteration of structure} (\hyperref[is]{IS}), a quantum-theoretic expression of the fact that configuration spaces of directed sets possess {\it higher-level abstract multidirected structures,} collectively induced by their member sets, and by relationships between pairs of these sets. Section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} below consists of preliminary material on quantum theory, beginning with a brief explanation of how path integration in conventional quantum physics relates to the general quantum-theoretic principle of {\it superposition.} Paths in distinguished continuum manifolds represent {\it histories,} whose superposition may be quantified, via path integration, to define {\it quantum amplitudes,} conventionally interpreted in terms of probabilities. Background independence elevates each history to a complete {\it universe.} In particular, under the classical causal metric hypothesis (\hyperref[ccmh]{CCMH}), histories are represented by directed sets. The na\"{i}ve analogues of path integrals in this context are therefore {\it sums over configuration spaces of directed sets.} Important examples of such sums appear in existing approaches to background-independent quantum theory, such as Isham's {\it quantization on a category} \cite{IshamQuantisingI05}, and Sorkin's {\it quantum measure theory} \cite{SorkinQuantalMeasure12}, \cite{SorkinQuantumMeasure94}. These approaches provide motivation for the principle of iteration of structure (\hyperref[is]{IS}). For both conceptual and technical reasons, however, it is advantageous to adopt a modified perspective, emphasizing {\it relationships between pairs of histories,} again utilizing Grothendieck's relative viewpoint (\hyperref[rv]{RV}). I refer to such relationships as {\it co-relative histories.} In the discrete causal context, they are represented by special morphisms of directed sets, called {\it transitions.} In section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}, I introduce an abstract causal analogue of Feynman's theory of path integration, expressed in terms of path summation over multidirected sets. As explained above, the same general theory applies in both the background-dependent and background-independent contexts. I review the construction of Feynman's path integral over Euclidean spacetime, and define an analogous path sum on the relation space over a multidirected set. Impermeability of maximal antichains in relation space, proven in theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} above, plays a crucial role here. Feynman's path integral yields quantum amplitudes associated with the motion of particles through Euclidean spacetime, while the analogous causal construction yields generalized quantum amplitudes, associated either with ``particles moving through relation space," or with ``evolution of families of initial directed sets into families of terminal directed sets." In section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}}, I review Feynman's re-derivation, via path integrals, of Schr\"{o}dinger's equation, which serves as a dynamical law in ordinary quantum theory. I then proceed to derive analogous {\it causal Schr\"{o}dinger-type equations,} which play the same role in the discrete quantum causal context. In section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}}, I introduce the theory of {\it kinematic schemes,} which are configuration spaces of directed sets, endowed with ``higher-level multidirected structures," supplied by families of co-relative histories, according to the principle of iteration of structure (\hyperref[is]{IS}). I discuss in some detail the {\it positive sequential kinematic scheme} $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, which is the analogue, in the nontransitive case, of Sorkin and Rideout's kinematic scheme describing the sequential growth of causal sets. I briefly mention analogies between kinematic schemes and number systems, such as $\mathbb{Q}$ and $\mathbb{R}$. I also introduce {\it kinematic space}, which represents ``all possible frames of reference for all possible histories." \subsection{Quantum Preliminaries; Iteration of Structure; Co-Relative Histories}\label{subsectionquantumprelim} \refstepcounter{textlabels}\label{superposition} \refstepcounter{textlabels}\label{pathintegration} \refstepcounter{textlabels}\label{histories} {\bf Superposition; Path Integration; Histories Approach to Quantum Theory.} A generic feature of quantum theory is the notion of {\it superposition}, in which multiple independent physical scenarios coexist. In ordinary quantum mechanics, each physical scenario is a solution of Schr\"{o}dinger's equation, and superposition is a consequence of its {\it linearity.} In Feynman's theory of path integration, each physical scenario may be understood as a possible {\it history} of a physical system, traced out as a {\it path} on a background spacetime manifold or configuration space. Predictions are made in this context by summing over these histories to obtain {\it quantum amplitudes,} which may be translated into probabilities. All ``possible" histories, even those that seem classically absurd, contribute to these amplitudes. Incorporating fields and special relativity, path integration becomes one of the principal tools, and one of the principal technical difficulties, of quantum field theory. The {\it histories approach to quantum theory} is a very general approach that grew out of path integration. The histories approach weighs contributions from members of a configuration space of histories. In the background-dependent context, these histories may be represented by particle trajectories, or field configurations, or some other auxiliary structure {\it inhabiting} spacetime. In the background-independent context, each history is an entire {\it classical universe,} which may be represented by a geometry, or a spinfoam, or a dynamical triangulation, or a causal set, or some other model. The flexibility and generality of the histories approach allows it to be applied in contexts where most other versions of quantum theory do not even make sense. \refstepcounter{textlabels}\label{historiesquantumcausal} \refstepcounter{textlabels}\label{backgrounddependentqct} \refstepcounter{textlabels}\label{backgroundindependentqct} {\bf Histories Approach to Quantum Causal Theory.} In the discrete causal context, it is straightforward, at least in a formal sense, to construct a {\bf background-dependent quantum causal theory}, by simply replacing the spacetime background in Feynman's theory of path integration with a locally finite directed set, and applying the theory of path summation introduced in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}} above. In this context, a {\it history} is represented by a ``particle trajectory," or some more complex auxiliary structure, inhabiting a {\it fixed} locally finite directed set. This approach supplies flexible discrete causal alternatives to quantum field theories in curved spacetime and lattice field theories. It is conceptually similar to existing treatments of particles and fields on causal set backgrounds, but may be considerably strengthened by the systematic use of relation space, and by the other improvements to the causal set viewpoint described in sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} of this paper.\footnotemark\footnotetext{Obvious complications arise in the cyclic case, where one must decide how to handle ``trajectories" wrapping around cycles arbitrary numbers of times. In this section, however, I work mostly with the acyclic case.} Much deeper, more difficult, and more interesting is {\bf background-independent quantum causal theory}, in which each history is represented by a {\it different} directed set. This requires consideration of entire configuration spaces of directed sets, in a manner analogous to Isham's {\it quantization on a category} and Sorkin's {\it quantum measure theory.} Of particular interest are {\it kinematic schemes,} which are special configuration spaces adapted to an {\it evolutionary} viewpoint. The abstract structure of kinematic schemes may be understood in terms of {\it relationships between pairs of histories,} which I refer to as {\it co-relative histories.} Along with the theory of relation space, the theory of co-relative histories is the most important physical application of Grothendieck's now-familiar relative viewpoint (\hyperref[rv]{RV}) appearing in this paper. \refstepcounter{textlabels}\label{ishamquantcat} {\bf Isham's Quantization on a Category.} Chris Isham has explored a number of ideas relevant to quantum causal theory in his groundbreaking work on physical applications of category theory. For example, in his 2005 paper {\it Quantising on a Category} \cite{IshamQuantisingI05}, Isham introduces a category-theoretic approach to ``quantizing" a broad variety of physical models, including causal sets. This represents only a small facet of Isham's ongoing research program to develop a general {\it topos-theoretic} framework for physics. A {\it topos} is an ``enriched" category with properties rendering it suitable to serve as a ``mathematical universe." For example, there is a {\it topos of sets,} which serves as Isham's topos for classical physics. Topos theory is yet another crucial structural idea arising from Alexander Grothendieck's approach to commutative algebraic geometry, where it was conceived as a way of organizing information about {\it presheaves} and {\it sheaves} on topological spaces and generalized topological spaces called {\it sites}. For the purposes of this paper, it suffices to highlight only a few specific aspects of Isham's approach. Isham \cite{IshamQuantisingI05} writes that his \begin{quotation}\noindent{\it ``...main topic of interest \tn{[is]} to construct a quantum framework for a system whose ``history configuration space"... ...is a collection of possible space-times, such as... ...causal sets... ...or... topological spaces."} (page 277) \end{quotation} and later, \begin{quotation}\noindent{\it ``...when quantising a system whose configuration space... ... is a collection... ...of sets with structure, the starting point will be an association to each pair \tn{[of such sets]} a set... ...of transformations \tn{[between them]} that, in some suitable sense, `respect,' or `reflect,' \tn{[their]} internal structures..."} (page 278) \end{quotation} Isham then proceeds to explain that such a family of ``structured sets" may be organized into a category $\ms{Q}$, whose morphisms are the aforementioned transformations. Abstracting this, he proposes to quantize general categories by \begin{quotation}\noindent{\it ``...representing \tn{[morphisms]} with operators on a Hilbert space in some way."} (page 280) \end{quotation} Without going into details, Isham's method involves a {\it presheaf of Hilbert spaces} and a {\it monoid of arrow fields} over $\ms{Q}$. An {\it arrow field} assigns to each object $Q$ of $\ms{Q}$ a morphism $Q\rightarrow Q'$ to some other object $Q'$. A natural monoid structure on the class of arrow fields over $\ms{Q}$ is given by composition of morphisms. As mentioned at the end of section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} above, Isham's monoid of arrow fields may be viewed as a higher-level, purely multiplicative analogue of Raptis' incidence algebras \cite{RaptisAlgebraicQuantization00}. All the information in either object is contained in the corresponding directed product algebra.\footnotemark\footnotetext{Isham mentions Raptis' incidence algebras in the preprint version of \cite{IshamQuantisingI05}.} Isham gives a simple example of this construction, using a category consisting of two causal sets, on pages 293-296 of \cite{IshamQuantisingI05}. However, for the purposes of this paper, it is Isham's {\it general reasoning,} represented by the above quotations, that is important. The technical details of the approach presented here are much different. In particular, I do not {\it a priori} assign Hilbert spaces, or any other ``ready-made" quantum-theoretic devices, to directed sets. I also prefer to work with path algebras than with incidence algebras, arrow fields, or similar constructions, which ``lose information" under composition. \refstepcounter{textlabels}\label{sorkinquantal} {\bf Sorkin's Quantum Measure Theory.} Rafael Sorkin has done very general work on the histories approach to background-independent quantum theory, and has applied these ideas in specific ways to causal set theory. In particular, in his 2012 paper {\it Toward a Fundamental Theorem of Quantal Measure Theory} \cite{SorkinQuantalMeasure12}, Sorkin describes quantum-theoretic measures in the context of the Sorkin-Rideout kinematic scheme $\ms{K}$ describing the sequential growth of causal sets, illustrated in figure \hyperref[sequential]{\ref{sequential}} of section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above. Sorkin works mostly in terms of {\it labeled} causal sets in this paper, which leads to the replacement of $\ms{K}$ with an appropriate {\it tree.} Sorkin writes, \begin{quotation}\noindent{\it ``...the collection of naturally labeled finite \tn{[causal sets]}... ...has itself the structure of a \tn{[tree]}, because its elements are labeled... ...Clearly, a particular realization of the growth process... ...can be conceived of as an upward path through this tree." } (page 7) \end{quotation} Here, a ``realization of the growth process" means a history, represented by a causal set $C$, together with a generalized frame of reference for $C$, represented by a specific labeling of $C$. In section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below, I revisit such ``upward paths" in the more general context of {\it co-relative kinematics,} and {\it completions of kinematic schemes.} \refstepcounter{textlabels}\label{iteration} {\bf Iteration of Structure.} As mentioned in the introduction to section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}, the same theory of path summation over directed sets and multidirected sets, outlined in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}} above, applies to both the background-dependent quantum theory of fields and particles on a directed set, and the background-independent quantum theory of co-relative histories and kinematic schemes. This fact is by no means obvious, since the path sums in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}} involve the causal path set over a {\it particular} directed or multidirected set, while background-independent quantum causal theory requires simultaneous consideration of entire configuration spaces of directed sets. The link between the two approaches is the principle of {\it iteration of structure}, which is based on the observation that configuration spaces of directed sets possess {\it abstract multidirected structures,} collectively induced by their member sets, and by relationships between pairs of these sets. The following statement of this principle is sufficiently general for the purposes of this paper, although other examples of the same basic idea may be recognized at other levels of hierarchy in causal theory.\footnotemark\footnotetext{For example, the relationship between kinematic schemes and {\it kinematic spaces,} briefly outlined at the end of section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}}, is an example of iteration of structure beyond the quantum level.} \refstepcounter{textlabels}\label{is} \hspace*{.3cm} IS. {\bf Iteration of Structure:} {\it Background independent quantum causal theory may be described \\ \hspace*{.9cm} in terms of ``multidirected sets whose elements are directed sets."} Iteration of structure is perhaps the most important generalization of the binary axiom (\hyperref[b]{B}). A prototypical example of iteration of structure is given by Sorkin and Rideout's kinematic scheme describing the sequential growth of causal sets \cite{SorkinSequentialGrowthDynamics99}, illustrated in figure \hyperref[sequential]{\ref{sequential}} of section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above. A closely related example, more relevant to the approach developed in this paper, is given by the {\it positive sequential kinematic scheme,} illustrated in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} of section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. Both of these kinematic schemes possess the abstract structures of countably infinite, locally finite, acyclic multidirected sets. Iteration of structure is a striking property, conceptually unifying classical and quantum theory as {\it different levels of algebraic hierarchy} in a single universal construction. This is a distinguishing characteristic of discrete causal theory, compared with other approaches to the study of fundamental spacetime structure, quantum gravity, and unification. For comparison, the abstract structures of configuration spaces of relativistic spacetime geometries, up to suitable equivalence, are generally nothing like individual spacetime geometries.\footnotemark\footnotetext{For example, the moduli space of {\it Einstein metrics} on the single diffeomorphism class of K3 surfaces admits a local diffeomorphism into a $57$-dimensional orbifold \cite{AndersonEinsteinMetrics09}.} \refstepcounter{textlabels}\label{transitions} {\bf Transitions.} Iteration of structure arises naturally in the search for suitable notions of {\it initial} and {\it terminal conditions} in causal theory. It is useful to express this idea in terms of {\it relationships between histories,} which I refer to as {\it co-relative histories.} In this paper, I focus on the special case of co-relative histories that formalize the ``evolution of one directed set into another."\footnotemark\footnotetext{As usual, only {\it isomorphism classes} of directed sets are significant in this context, but the necessary information may be described in terms of representatives.} Such co-relative histories are represented by {\it transitions,} which are special morphisms of directed sets. The term ``co-relative" is explained below.\\ \begin{defi}\label{defitransition} A {\bf transition} in the category $\ms{D}$ of directed sets is a monomorphism $\tau:D_i\rightarrow D_t$, embedding its {\bf initial set}, or {\bf source}, $D_i$, into its {\bf terminal set}, or {\bf target}, $D_t$, as a proper, full, originary subobject. This means that $\tau(D_i)$ has nontrivial complement in $D_t$ (proper), that $\tau(x)\prec\tau(y)$ in $D_t$ if and only if $x\prec y$ in $D_i$ (full), and that the isomorphic image $\tau(D_i)$ of $D_i$ in $D_t$ contains its own past (originary). \end{defi} Heuristically, a transition $\tau$ {\it augments} its source $D_i$, by adding directed structure in a manner compatible with its existing structure, thereby ``producing" its target $D_t$. There is no difficulty in extending the definition of a transition to multidirected sets, but such a generalization is not needed in this paper. Figure \hyperref[transitionfig]{\ref{transitionfig}} below illustrates a transition. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{14cm}{5.7cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(0,5.8) \pgfxyline(16.5,-.1)(16.5,5.8) \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.8)(16.5,5.8) \pgfmoveto{\pgfxy(11,5.6)} \pgfstroke \begin{pgftranslate}{\pgfpoint{1.25cm}{.25cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10.25cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{40!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \begin{colormixin}{100!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.11cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.11cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.11cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.11cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.11cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.11cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.11cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.11cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.11cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.11cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.11cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.11cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.11cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \end{colormixin} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(8.25,3.3)}{\pgfbox[center,center]{transition $\tau$}} \pgfputat{\pgfxy(3.75,5.2)}{\pgfbox[center,center]{source $D_i$}} \pgfputat{\pgfxy(12.75,5.2)}{\pgfbox[center,center]{target $D_t$}} \pgfputat{\pgfxy(8.75,1.4)}{\pgfbox[center,center]{\small{$D_i$ embedded as}}} \pgfputat{\pgfxy(8.75,1)}{\pgfbox[center,center]{\small{proper, full, originary}}} \pgfputat{\pgfxy(8.75,.6)}{\pgfbox[center,center]{\small{subobject of $D_t$}}} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(10.1,1.5)(10.6,1.8) \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(7,2.75)(9.5,2.75) \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{A transition $\tau:D_i\rightarrow D_t$.} \label{transitionfig} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{benderrobinson} Transitions are {\it a priori} too specific to be regarded as physically fundamental, since they generally contain extra, nonphysical information, due to symmetries of their sources and targets. For example, the composition of a transition $\tau$ with an automorphism of its source defines a ``physically equivalent" transition. The automorphism group of the target of $\tau$ also plays a role, whenever this group ``mixes the image of $\tau$ with its complement." Given the ``irregularity" of an arbitrary directed set, one might expect most such sets to lack ``large-scale symmetries;" i.e., automorphisms mapping large subsets to their complements. On the other hand, one might expect ``small-scale symmetries;" i.e., automorphisms permuting a few elements in a small symmetric subset, to be fairly common. An interesting result, in the acyclic case, is that {\it a typical finite acyclic directed set is rigid;} i.e., it has no nontrivial automorphisms at all. This result was proven, in an asymptotic sense, by Edward Bender and Robert Robinson, in their 1988 paper {\it The Asymptotic Number of Acyclic Digraphs II} \cite{BenderRobinsonAuto88}. The physical implications of Bender and Robinson's rigidity result in the context of discrete causal theory depend on a variety of factors. For example, a {\it typical} finite directed set has a ``large number of relations," due to elementary counting considerations. This tends to favor rigidity, since any given relation can break a symmetry. However, dynamical considerations may ultimately prove to favor ``sparser" directed sets, allowing nontrivial automorphism groups to play a significant role. This is an issue of considerable physical importance, since automorphisms of directed sets are analogous to ``external" or ``spacetime" symmetries in conventional physics, such as those encoded by the Poincar\'{e} symmetry group of Minkowski spacetime. Such external symmetries impose basic constraints on the properties of elementary particles in quantum field theory, via Lie representation theory. \refstepcounter{textlabels}\label{symmetry} {\bf Symmetry Preservation, Extension, Breaking, Generation.} Assuming, despite Bender and Robinson's rigidity result, that (``external") symmetries are relevant in discrete causal theory, it is worthwhile to briefly consider how such symmetries interact with transitions. In this context, a {\it symmetry} means simply an automorphism of a particular directed set. A transition may {\it preserve,} {\it extend,} or {\it break} symmetries of its source, or it may {\it generate} symmetries of its target, or some combination of the four. Care is required in the physical interpretation of these notions, particularly because the too-specific nature of individual transitions adds arbitrary information to the picture. For example, {\it different but equivalent} transitions mapping a common source to different subobjects of a common target generally preserve, extend, break, and/or generate {\it different but equivalent} families of symmetries. Let $\tau:D_i\rightarrow D_t$ be a transition of directed sets, and let $\tn{Aut}(D_i)$ and $\tn{Aut}(D_t)$ be the automorphism groups of the source $D_i$ and target $D_t$ of $\tau$, respectively. Let $F_\tau:=\tn{Aut}(D_t;\tau(D_i))$ be the subgroup of $\tn{Aut}(D_t)$ whose elements {\it permute $\tau(D_i)$ and its complement separately.} Each element of $F_\tau$ defines an automorphism of $D_i$, via restriction and conjugation by $\tau$. Hence, I refer to $F_\tau$ as the {\bf group of extensions} of symmetries of $D_i$. More precisely, there is a group homomorphism \[\rho_\tau:F_\tau\longrightarrow\tn{Aut}(D_i),\] \[\hspace*{.8cm}\beta\hspace*{.2cm}\mapsto\hspace*{.2cm}\tau^{-1}\circ\beta\circ\tau,\] which I refer to as the {\bf restriction homomorphism}, where the factors $\tau^{-1}$ and $\beta$ in the composition are understood to be restricted to the image $\tau(D_i)$ of $\tau$. If $\rho_\tau$ is surjective, then $\tau$ {\bf preserves the symmetries} of the source $D_i$ of $\tau$. The kernel $K_\tau$ of $\rho_\tau$ is the normal subgroup of $F_\tau$ whose elements consist of automorphisms of $D_t$ fixing $\tau(D_i)$. Hence, if $\tau$ preserves the symmetries of $D_i$, there is a short exact sequence \[1\rightarrow K_\tau\overset{\iota}{\rightarrow} F_\tau\overset{\rho_\tau}{\rightarrow} \tn{Aut}(D_i)\rightarrow 1,\] where $\iota$ is the inclusion of the kernel $K_\tau$ into $\tn{Aut}(D_i)$. Hence, $F_\tau$ is a {\it group extension} of $\tn{Aut}(D_i)$ by $K_\tau$ if and only if $\tau$ preserves the symmetries of $D_i$. The cokernel $F_\tau/K_\tau$ of $\rho_\tau$ may be identified with the normal subgroup of $F_\tau$ fixing the {\it complement} of $\tau(D_i)$ in $D_t$. The group $F_\tau$ is isomorphic to the direct product of $K_\tau$ and $G_\tau$: \[F_\tau\cong K_\tau\times G_\tau.\] \refstepcounter{textlabels}\label{galois} Whether or not $\rho_\tau$ is surjective, it induces a quotient monomorphism $\overline{\rho_\tau}:G_\tau\rightarrow \tn{Aut}(D_i)$, identifying $G_\tau$ with the {\it subgroup of automorphisms of $D_i$ preserved by} $\tau$. In the finite case, the subgroup index $[F_\tau:G_\tau]$ of $G_\tau$ in $F_\tau$ measures the number of different ways automorphisms of $D_i$ preserved by $\tau$ are extended by $\tau$. If $\rho_\tau$ is surjective, then $\overline{\rho_\tau}$ is an isomorphism. If $\rho_\tau$ is not surjective, then every element of the complement $\tn{Aut}(D_i)-\overline{\rho_\tau}(G_\tau)$ represents a {\bf symmetry broken by} $\tau$. In the finite case, the index $[\tn{Aut}(D_i):G_\tau]$ measures the extent of this symmetry breaking. The groups $K_\tau$ and $G_\tau$ may be called {\bf causal Galois groups}, since they are groups of automorphisms fixing distinguished subobjects of directed sets in the context of discrete causal theory.\footnotemark\footnotetext{Many interesting analogies exist between causal theory and Galois theory, which cannot be explored in this paper. For example, I have not mentioned the obviously relevant theory of {\it Galois connections.}} Figure \hyperref[transitiongalois]{\ref{transitiongalois}} below illustrates simple examples of symmetry preservation, extension, and breaking. Symmetry generation, also depicted in the figure, is discussed later in this section. Let $\tau_0:D_0\rightarrow D_1$ be the transition sending the elements $t_0$, $u_0,$ and $v_0$ of $D_0$ to the elements $t_1$, $u_1,$ and $v_1$ of $D_1$, respectively. The automorphism group $\tn{Aut}(D_0)$ of the source $D_0$ of $\tau_0$ is isomorphic to $\mathbb{Z}_2$, generated by the unique automorphism $\alpha$ of $D_0$ interchanging $u_0$ and $v_0$. The automorphism group $\tn{Aut}(D_1)$ of the target $D_1$ of $\tau_0$ is trivial. Hence, the subgroups $F_{\tau_0}$, $K_{\tau_0},$ and $G_{\tau_0}$ of $\tn{Aut}(D_1)$ are trivial. The index $[F_{\tau_0}:G_{\tau_0}]$ is therefore equal to $1$, which means that $\tau_0$ extends the trivial automorphism of $D_0$ to the trivial automorphism of $D_1$, with no other extensions. The index $[\tn{Aut}(D_0):G_{\tau_0}]$ is equal to $2$, which means that $\tau_0$ breaks the unique nontrivial symmetry $\alpha$ of $D_0$. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{4.2cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,4.1)(16.5,4.1) \pgfxyline(0,-.1)(0,4.1) \pgfxyline(16.5,-.1)(16.5,4.1) \begin{pgftranslate}{\pgfpoint{.4cm}{0cm}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm}} \pgfputat{\pgfxy(1,3.5)}{\pgfbox[center,center]{$D_0$}} \pgfputat{\pgfxy(2.5,1.8)}{\pgfbox[center,center]{$\tau_0$}} \pgfputat{\pgfxy(1,.7)}{\pgfbox[center,center]{$t_0$}} \pgfputat{\pgfxy(.25,1.8)}{\pgfbox[center,center]{$u_0$}} \pgfputat{\pgfxy(1.75,1.8)}{\pgfbox[center,center]{$v_0$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,1.8)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,1.8)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.2,1.5)(2.8,1.5) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.35cm}{0cm}} \pgfputat{\pgfxy(1,3.5)}{\pgfbox[center,center]{$D_1$}} \pgfputat{\pgfxy(2.5,1.8)}{\pgfbox[center,center]{$\tau_1$}} \pgfputat{\pgfxy(1,.7)}{\pgfbox[center,center]{$t_1$}} \pgfputat{\pgfxy(.25,1.8)}{\pgfbox[center,center]{$u_1$}} \pgfputat{\pgfxy(1.75,1.8)}{\pgfbox[center,center]{$v_1$}} \pgfputat{\pgfxy(0,2.9)}{\pgfbox[center,center]{$w_1$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,1.8)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,1.8)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.2,2.6)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.2,1.5)(2.8,1.5) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.7cm}{0cm}} \pgfputat{\pgfxy(1,3.5)}{\pgfbox[center,center]{$D_2$}} \pgfputat{\pgfxy(2.5,1.8)}{\pgfbox[center,center]{$\tau_2$}} \pgfputat{\pgfxy(1,.7)}{\pgfbox[center,center]{$t_2$}} \pgfputat{\pgfxy(.25,1.8)}{\pgfbox[center,center]{$u_2$}} \pgfputat{\pgfxy(1.75,1.8)}{\pgfbox[center,center]{$v_2$}} \pgfputat{\pgfxy(0,2.9)}{\pgfbox[center,center]{$w_2$}} \pgfputat{\pgfxy(2,2.9)}{\pgfbox[center,center]{$z_2$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,1.8)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,1.8)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.2,2.6)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.8,2.6)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.2,1.5)(2.8,1.5) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10.05cm}{0cm}} \pgfputat{\pgfxy(1,3.5)}{\pgfbox[center,center]{$D_3$}} \pgfputat{\pgfxy(2.5,1.8)}{\pgfbox[center,center]{$\tau_3$}} \pgfputat{\pgfxy(1,.7)}{\pgfbox[center,center]{$t_3$}} \pgfputat{\pgfxy(.25,1.8)}{\pgfbox[center,center]{$u_3$}} \pgfputat{\pgfxy(1.75,1.8)}{\pgfbox[center,center]{$v_3$}} \pgfputat{\pgfxy(0,2.9)}{\pgfbox[center,center]{$w_3$}} \pgfputat{\pgfxy(.75,2.9)}{\pgfbox[center,center]{$x_3$}} \pgfputat{\pgfxy(2,2.9)}{\pgfbox[center,center]{$z_3$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,1.8)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,1.8)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.2,2.6)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.8,2.6)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(.75,2.6)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node2}{Node6} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.2,1.5)(2.8,1.5) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{13.4cm}{0cm}} \pgfputat{\pgfxy(1,3.5)}{\pgfbox[center,center]{$D_4$}} \pgfputat{\pgfxy(1,.7)}{\pgfbox[center,center]{$t_4$}} \pgfputat{\pgfxy(.25,1.8)}{\pgfbox[center,center]{$u_4$}} \pgfputat{\pgfxy(1.75,1.8)}{\pgfbox[center,center]{$v_4$}} \pgfputat{\pgfxy(0,2.9)}{\pgfbox[center,center]{$w_4$}} \pgfputat{\pgfxy(2,2.9)}{\pgfbox[center,center]{$z_4$}} \pgfputat{\pgfxy(.75,2.9)}{\pgfbox[center,center]{$x_4$}} \pgfputat{\pgfxy(1.25,2.9)}{\pgfbox[center,center]{$y_4$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,1.8)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.4,1.8)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(.2,2.6)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1.8,2.6)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(.75,2.6)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.25,2.6)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node3}{Node7} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{Symmetry preservation, extension, breaking, and generation, via transitions.} \label{transitiongalois} \end{figure} \vspace*{-.5cm} It is instructive to examine the significance of this broken $\mathbb{Z}_2$ symmetry in an informal sense, first ``from the perspective of the source $D_0$," then ``from the perspective of the target $D_1$." There are obvious physical motivations for considering such viewpoints. In particular, one might begin with $D_0$, and examine the possible ways in which it could ``evolve into $D_1$;" conversely, one might begin with $D_1$, and investigate how it could have ```evolved from $D_0$." From the perspective of the source $D_0$ of $\tau_0$, the broken $\mathbb{Z}_2$ symmetry reflects the existence of two ``equivalent" ways to obtain a directed set isomorphic to $D_1$ by adding a single element to $D_0$. This ``new element" may be added either in the future of $u_0$, or in the future of $v_0$. These two constructions yield transitions from $D_0$ into two {\it different, but isomorphic,} targets. It is important to note, however, that the existence of multiple isomorphic targets is not always associated with a broken symmetry of the source. In particular, there may be multiple ``inequivalent" ways to construct the isomorphism class of a specified target from a specified source {\it even if both source and target have trivial automorphism groups.} This curious phenomenon, explained in more detail below, can occur even in cases where only a single new element is added. From the perspective of the target $D_1$ of $\tau_0$, there is a unique subobject of $D_1$ that may be identified with the source $D_0$. From this perspective, the broken $\mathbb{Z}_2$ symmetry reflects the fact that the unique nontrivial symmetry of this subobject does not extend to a symmetry of $D_1$. Returning for the moment to the case of an arbitrary transition $\tau:D_i\rightarrow D_t$, the complement of the subgroup $F_\tau$ in $\tn{Aut}(D_t)$ consists of all automorphisms of $D_t$ that {\it mix} the image $\tau(D_i)$ of $\tau$ with its complement in $D_t$. These automorphisms obviously do not form a subgroup of $\tn{Aut}(D_t)$, since they do not include the identity. Further, $F_\tau$ is generally not a normal subgroup of $\tn{Aut}(D_t)$,\footnotemark\footnotetext{For example, consider the automorphism $\beta$ of $D_4$ in figure \hyperref[transitiongalois]{\ref{transitiongalois}} interchanging $u_4,w_4$, and $x_4$ with $v_4, y_4,$ and $z_4,$ respectively. It is easy to see that $\beta^{-1}F_{\tau_3}\beta\ne F_{\tau_3}$.} so there is generally no quotient group to study. These automorphisms are new ``from the perspective of the transition $\tau$," since they do not restrict to automorphisms of the image $\tau(D_i)$. Hence, I refer to them as symmetries {\bf generated by} $\tau$. However, these automorphisms are not necessarily new ``from the perspective of the source $D_i$ of $\tau$." The meaning of this distinction is that there may exist other transitions $\tau':D_i\rightarrow D_t$, corresponding to different, but isomorphic, subgroups $F_{\tau'}$ of $\tn{Aut}(D_t)$, whose elements extend symmetries of $D_i$ broken by $\tau$, and which are {\it equivalent to} $\tau$, in the sense that their images are mapped isomorphically onto the image of $\tau$ by automorphisms of $D_t$. This notion of equivalence is made precise in definition \hyperref[deficorelative]{\ref{deficorelative}} below. From a physical perspective, this distinction reflects the fact that transitions are too specific to be considered physically fundamental. Referring again to figure \hyperref[transitiongalois]{\ref{transitiongalois}} above, let $\tau_1:D_1\rightarrow D_2$ be the transition sending the elements $t_1$, $u_1$, $v_1$, and $w_1$ of $D_1$ to the elements $t_2$, $u_2$, $v_2$, and $w_2$ of $D_2$, respectively. The automorphism group $\tn{Aut}(D_1)$ of the source $D_1$ of $\tau_1$ is trivial, while the automorphism group $\tn{Aut}(D_2)$ of the target $D_2$ of $\tau_1$ is isomorphic to $\mathbb{Z}_2$, generated by the unique automorphism $\beta$ of $D_2$ interchanging $u_2$ and $v_2$, and $w_2$ and $z_2$, respectively. The extension group $F_{\tau_1}$ is trivial, since $\beta$ mixes $\tau_1(D_1)$ with its complement. Hence, the causal Galois groups $K_{\tau_1}$ and $G_{\tau_1}$ are also trivial. The complement of $F_{\tau_1}$ in $\tn{Aut}(D_2)$ is the singleton $\{\beta\}$; therefore, $\beta$ is a symmetry generated by $\tau_1$. From the perspective of the source $D_1$ of $\tau_1$, there is only one way to add an extra element to $D_1$ to obtain a directed set isomorphic to the target $D_2$. However, from the perspective of the target $D_2$ of $\tau_1$, there are {\it two} different subobjects that may be identified with the source $D_1$. This reflects the fact that there is an equivalent transition ${\tau_1}':D_1\rightarrow D_2$, sending the elements $t_1$, $u_1$, $v_1,$ and $w_1$ of $D_1$ to the elements $t_2$, $v_2$, $u_2,$ and $z_2$ of $D_2$, respectively. The transition ${\tau_1}'$ {\it also} generates the unique nontrivial symmetry $\beta$ of $D_2$. The remaining transitions $\tau_2:D_2\rightarrow D_3$ and $\tau_3:D_3\rightarrow D_4$ appearing in figure \hyperref[transitiongalois]{\ref{transitiongalois}} involve more complex phenomena. The transition $\tau_2$ breaks the $\mathbb{Z}_2$-symmetry of $D_2$, but generates a {\it new} $\mathbb{Z}_2$ symmetry in $D_3$. Hence, the symmetry groups of $D_2$ and $D_3$ are ``accidentally" isomorphic. The transition $\tau_3$, meanwhile, generates multiple new symmetries. One may also consider compositions of the transitions in the figure; for example, the composition $\tau_1\circ\tau_0:D_0\rightarrow D_2$ preserves the symmetries of $D_0$, without generating any new symmetries. In this case, the fact that the symmetry groups of the source and target are isomorphic is {\it not} ``accidental." \refstepcounter{textlabels}\label{labelcorelative} {\bf Co-Relative Histories.} In defining co-relative histories, the objective is to precisely capture the {\it physical essence of particular relationships between pairs causal structures,} represented by directed sets, following Grothendieck's relative viewpoint (\hyperref[rv]{RV}). In this context, a na\"{\i}ve attempt to represent such relationships by individual morphisms in the category $\ms{D}$ of directed sets leads in the wrong direction, at least in the case in which the directed sets under consideration possess nontrivial automorphism groups. A better choice is to represent such relationships by {\it equivalence classes of morphisms whose images are related by automorphisms of their target sets.} Here, I focus on equivalence classes of {\it transitions,} since the goal is to formalize the evolution of causal structure. Such equivalence classes define {\it proper, full, originary co-relative histories} relating their common sources and targets. One may imagine more general relationships between pairs of histories, but maximal generality is not the goal in the present context. \refstepcounter{textlabels}\label{corelmulti} Examples of equivalent and inequivalent transitions between pairs of directed sets are illustrated in figure \hyperref[corelativeexamples]{\ref{corelativeexamples}}a below, with the labels on some of the elements suppressed to avoid clutter. The transitions $\tau$ and $\tau'$ between $D_i$ and $D_t$ are equivalent, since their images are related by the unique automorphism $\alpha$ of $D_t$ interchanging the elements $x$ and $y$. The transition $\tau''$ is {\it not} equivalent to $\tau$ or $\tau'$, since $\alpha$ is the only nontrivial automorphism of $D_t$, and $\alpha$ fixes the element $z$. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{6.2cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,6.1)(16.5,6.1) \pgfxyline(0,-.1)(0,6.1) \pgfxyline(16.5,-.1)(16.5,6.1) \pgfxyline(6.6,-.1)(6.6,6.1) \pgfputat{\pgfxy(.15,5.8)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(6.75,5.8)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{.4cm}{0cm}} \begin{pgftranslate}{\pgfpoint{-.5cm}{0cm}} \pgfputat{\pgfxy(3.5,.5)}{\pgfbox[center,center]{$D_i$}} \pgfputat{\pgfxy(1.5,5.3)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,4.8)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.5,4.8)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2,4.8)}{\pgfbox[center,center]{$z$}} \pgfputat{\pgfxy(2.3,1.95)}{\pgfbox[center,center]{$\tau$}} \pgfputat{\pgfxy(3.8,2.8)}{\pgfbox[center,center]{$\tau'$}} \pgfputat{\pgfxy(4.7,1.95)}{\pgfbox[center,center]{$\tau''$}} \pgfnodecircle{Node1}[fill]{\pgfxy(3,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4,.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(3.5,1.25)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(3.5,2)}{0.10cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1,4.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,4.5)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(1.5,4.5)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node20}{Node50} \pgfnodeconnline{Node30}{Node50} \pgfnodeconnline{Node30}{Node60} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(1,4.5)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.8,1.95)(2,2.4) \pgfxyline(3.5,2.4)(3.5,3.2) \pgfxyline(4.2,1.95)(5,2.4) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.5cm}{.5cm}} \pgfputat{\pgfxy(1.5,5.3)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,4.8)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.5,4.8)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2,4.8)}{\pgfbox[center,center]{$z$}} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1,4.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,4.5)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(1.5,4.5)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node20}{Node50} \pgfnodeconnline{Node30}{Node50} \pgfnodeconnline{Node30}{Node60} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(1.5,4.5)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.5cm}{0cm}} \pgfputat{\pgfxy(1.5,5.3)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,4.8)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.5,4.8)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2,4.8)}{\pgfbox[center,center]{$z$}} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1,4.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,4.5)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(1.5,4.5)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node20}{Node50} \pgfnodeconnline{Node30}{Node50} \pgfnodeconnline{Node30}{Node60} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(1,3)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(1.5,3.75)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(2,4.5)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10cm}{0cm}} \pgfputat{\pgfxy(1.75,.75)}{\pgfbox[center,center]{$D_i$}} \pgfputat{\pgfxy(-.2,2.1)}{\pgfbox[center,center]{$\tau$}} \pgfputat{\pgfxy(2.2,2.3)}{\pgfbox[center,center]{$\tau'$}} \pgfputat{\pgfxy(.8,.3)}{\pgfbox[center,center]{$u$}} \pgfputat{\pgfxy(.3,.8)}{\pgfbox[center,center]{$v$}} \pgfputat{\pgfxy(-.2,1.3)}{\pgfbox[center,center]{$w$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.5,1)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1,1)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(0,1.5)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(1,1.5)}{0.10cm} \pgfnodecircle{Node6}[fill]{\pgfxy(2,1.5)}{0.10cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node8}[fill]{\pgfxy(2.5,2)}{0.10cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node6} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node6}{Node7} \pgfnodeconnline{Node6}{Node8} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(.3,1.95)(-.4,2.6) \pgfxyline(1.8,2.2)(2.3,2.8) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7cm}{2.75cm}} \pgfputat{\pgfxy(1.5,2.8)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,2.3)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2.5,2.8)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(2.1,.5)}{\pgfbox[center,center]{$\tau(u)$}} \pgfputat{\pgfxy(3.6,2)}{\pgfbox[center,center]{$\tau(v)$}} \pgfputat{\pgfxy(4.1,2.5)}{\pgfbox[center,center]{$\tau(w)$}} \begin{colormixin}{50!white} \pgfnodecircle{Node00}[fill]{\pgfxy(1.5,.5)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(0,1)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(.5,1.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1.5,1.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(2.5,2.5)}{0.10cm} \pgfnodecircle{Node70}[fill]{\pgfxy(3,2)}{0.10cm} \pgfnodecircle{Node80}[fill]{\pgfxy(3.5,2.5)}{0.10cm} \pgfnodeconnline{Node10}{Node20} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node40}{Node30} \pgfnodeconnline{Node40}{Node50} \pgfnodeconnline{Node50}{Node60} \pgfnodeconnline{Node70}{Node60} \pgfnodeconnline{Node70}{Node80} \pgfnodeconnline{Node00}{Node10} \pgfnodeconnline{Node00}{Node40} \pgfnodeconnline{Node00}{Node70} \end{colormixin} \pgfnodecircle{Node00}[fill]{\pgfxy(1.5,.5)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(0,1)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(.5,1.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1.5,1.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodecircle{Node70}[fill]{\pgfxy(3,2)}{0.10cm} \pgfnodecircle{Node80}[fill]{\pgfxy(3.5,2.5)}{0.10cm} \pgfnodeconnline{Node10}{Node20} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node40}{Node30} \pgfnodeconnline{Node40}{Node50} \pgfnodeconnline{Node70}{Node80} \pgfnodeconnline{Node00}{Node10} \pgfnodeconnline{Node00}{Node40} \pgfnodeconnline{Node00}{Node70} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{12cm}{2.75cm}} \pgfputat{\pgfxy(1.5,2.8)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,2.3)}{\pgfbox[center,center]{$y$}} \pgfputat{\pgfxy(2.5,2.8)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(1.5,.2)}{\pgfbox[center,center]{$\tau'(u)$}} \pgfputat{\pgfxy(-.6,1)}{\pgfbox[center,center]{$\tau'(v)$}} \pgfputat{\pgfxy(-.1,1.6)}{\pgfbox[center,center]{$\tau'(w)$}} \begin{colormixin}{50!white} \pgfnodecircle{Node00}[fill]{\pgfxy(1.5,.5)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(0,1)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(.5,1.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1.5,1.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(2.5,2.5)}{0.10cm} \pgfnodecircle{Node70}[fill]{\pgfxy(3,2)}{0.10cm} \pgfnodecircle{Node80}[fill]{\pgfxy(3.5,2.5)}{0.10cm} \pgfnodeconnline{Node10}{Node20} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node40}{Node30} \pgfnodeconnline{Node40}{Node50} \pgfnodeconnline{Node50}{Node60} \pgfnodeconnline{Node70}{Node60} \pgfnodeconnline{Node70}{Node80} \pgfnodeconnline{Node00}{Node10} \pgfnodeconnline{Node00}{Node40} \pgfnodeconnline{Node00}{Node70} \end{colormixin} \pgfnodecircle{Node00}[fill]{\pgfxy(1.5,.5)}{0.10cm} \pgfnodecircle{Node10}[fill]{\pgfxy(0,1)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(.5,1.5)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(1.5,1.5)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(2.5,2.5)}{0.10cm} \pgfnodecircle{Node70}[fill]{\pgfxy(3,2)}{0.10cm} \pgfnodecircle{Node80}[fill]{\pgfxy(3.5,2.5)}{0.10cm} \pgfnodeconnline{Node10}{Node20} \pgfnodeconnline{Node40}{Node50} \pgfnodeconnline{Node50}{Node60} \pgfnodeconnline{Node70}{Node60} \pgfnodeconnline{Node70}{Node80} \pgfnodeconnline{Node00}{Node10} \pgfnodeconnline{Node00}{Node40} \pgfnodeconnline{Node00}{Node70} \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Equivalent and inequivalent transitions; b) McKay's example of inequivalent single-element transitions between the same pair of directed sets.} \label{corelativeexamples} \end{figure} \vspace*{-.5cm} An interesting fact, already mentioned above in the context of symmetry breaking, is that there may exist multiple ``inequivalent" ways to construct the isomorphism class of a specified target $D_t$ from a specified source $D_i$, even if both $\tn{Aut}(D_i)$ and $\tn{Aut}(D_t)$ are trivial, and even if $D_i$ and $D_t$ differ by only a single element. Equivalently, viewing $D_i$ and $D_t$ as ``given," rather than approaching the problem constructively, there may exist multiple inequivalent transitions between $D_i$ and $D_t$. Figure \hyperref[corelativeexamples]{\ref{corelativeexamples}}b above illustrates an example of this phenomenon, supplied by Brendan McKay \cite{McKayPrivate13}. The element labels $u$, $v$, $w$, etc., are included to clarify the existence of the transitions $\tau$ and $\tau'$. The elements $x$ and $y$ in the complements of $\tau$ and $\tau'$, respectively, are called {\it pseudosimilar vertices,} in the language of graph theory.\footnotemark\footnotetext{The theory of pseudosimilarity is of interest in the {\it graph reconstruction problem.} McKay tells me \cite{McKayPrivate13} that little work has been done on pseudosimilarity in the case of {\it directed} graphs.} Viewing $D_i$ and $D_t$ as histories, it is reasonably obvious that the relationships between them represented by $\tau$ and $\tau'$ differ in a physically significant way. The presence of this type of behavior has significant structural consequences in discrete causal theory. For example, as discussed below, it implies that the abstract structures of kinematic schemes are generally multidirected, rather than merely directed. In particular, Sorkin and Rideout's kinematic scheme describing the sequential growth of causal sets, illustrated in figure \hyperref[sequential]{\ref{sequential}} of section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, has the abstract structure of a multidirected set, {\it not} the skeleton of a partial order. The foregoing discussion motivates the following formal definition of co-relative histories, suitably general for the purposes of this paper. The terminology is chosen to suggest a {\it fixed source,} reflecting the intended evolutionary viewpoint. However, there are other equally valid perspectives. \vspace*{.2cm} \begin{defi}\label{deficorelative} A {\bf proper, full, originary co-relative history} $h:D_i\Rightarrow D_t$ between directed sets $D_i$ and $D_t$ is an equivalence class of transitions from $D_i$ to $D_t$, where two transitions $\tau$ and $\tau'$ are equivalent if and only if there exists an automorphism $\beta$ of $D_t$ mapping $\tau(D_i)$ onto $\tau'(D_i)$. The common initial set $D_i$ of the transitions making up $h$ is called the {\bf cobase} of $h$, and the common terminal set $D_t$ of these transitions is called the {\bf target} of $h$. \end{defi} The {\it double-arrow notation} $\Rightarrow$ in definition \hyperref[deficorelative]{\ref{deficorelative}} is intended to convey the idea that a co-relative history consists of a {\it family} of morphisms, ``bundled together" to encode a {\it single physical relationship.} More generally, one may specify a {\it cobase family} $\mbf{D}_i$ of directed sets, and consider {\it histories co-relative to} $\mbf{D}_i$. The reason for using the terms {\it co-relative} and {\it cobase,} rather than {\it relative} and {\it base,} is that $D_i$ is taken to represent ``initial," or ``known" information, while $D_t$ is allowed to vary over ``possible evolutionary outcomes." This viewpoint contrasts with the viewpoint accompanying the theory of relative multidirected sets over a fixed base, introduced in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}} above, in which the base is a fixed target set, and the source is allowed to vary.\footnotemark\footnotetext{The terminology in both cases originates from the category-theoretic notions of {\it covariant} and {\it contravariant} constructions, related to one another by ``reversing the arrows." Unfortunately, these terms carry many different meanings in both mathematics and physics.} The hyphen in {\it co-relative} is added to avoid confusion with the too-similar word {\it correlative}. The physical sense of definition \hyperref[deficorelative]{\ref{deficorelative}} is reasonably clear when the cobase $D_i$ and target $D_t$ are both finite. In the infinite case, {\it cycles of co-relative histories} are possible under definition \hyperref[deficorelative]{\ref{deficorelative}}, even when the directed sets involved are acyclic. This raises interesting interpretive challenges. Note that definition \hyperref[deficorelative]{\ref{deficorelative}} does {\it not} require that $\tau'=\beta\circ\tau$ as {\it morphisms;} it is only the {\it images} of $\tau$ and $\tau'$ that must be related by an automorphism of the target. For example, if $\tau$ breaks a symmetry $\alpha$ of its cobase $D_i$, then the transition $\tau'=\tau\circ\alpha$ cannot be recovered by applying an automorphism to its target $D_t$ {\it after} application of $\tau$. The simplest nontrivial example of this is illustrated in figure \hyperref[transitionnotcongruence]{\ref{transitionnotcongruence}}a below, in which $D_i$ consists of a pair of unrelated elements $x$ and $y$, and $D_t$ consists of three elements $x'$, $y'$, and $z'$, with one relation $y'\prec z'$. In this case, $\tn{Aut}(D_i)$ is isomorphic to $\mathbb{Z}_2$, generated by the automorphism interchanging $x$ and $y$, while $\tn{Aut}(D_t)$ is trivial. There is a unique co-relative history $h:D_i\Rightarrow D_t$, represented by two equivalent transitions $\tau$ and $\tau'$, where $\tau$ sends the elements $x$ and $y$ of $D_i$ to the elements $x'$ and $y'$ of $D_t$, respectively, and where $\tau'$ reverses these images. Both transitions break the unique nontrivial symmetry of $D_i$. These transitions are equivalent, via the identity automorphism of $D_t$, since they have the same image, but there is no automorphism of $D_t$ relating them as morphisms. \refstepcounter{textlabels}\label{corelativesubtle} Co-relative histories are subtle from a category-theoretic perspective. For example, equivalence of transitions is {\it not} a congruence relation under composition in the category $\ms{D}$ of directed sets. This means that, given equivalent transitions from a common source $D_0$ to a common target $D_1$, and from $D_1$ to a third directed set $D_2$, it is generally {\it not} true that the compositions of these transitions are equivalent. This is illustrated in figure \hyperref[transitionnotcongruence]{\ref{transitionnotcongruence}}b. The transitions $\tau_{01}$ and $\overline{\tau}_{01}$ from $D_0$ to $D_1$ are related by any automorphism of $D_1$ interchanging its two terminal elements, but the compositions $\tau_{12}\circ\tau_{01}$ and $\tau_{12}\circ\overline{\tau}_{01}$ represent different co-relative histories between $D_0$ and $D_2$. This example demonstrates that equivalence of transitions is not even {\it cancellative,} since the ``left-hand factor" $\tau_{12}$ is the same in both cases. More generally, pairs of ``consecutive" co-relative histories generally do not compose to give unique ``products." Instead, the ``composition" of a pair of ``consecutive" co-relative histories is a {\it family of co-relative histories.} \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{15cm}{5.5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.4)(16.5,5.4) \pgfxyline(0,-.1)(0,5.4) \pgfxyline(16.5,-.1)(16.5,5.4) \pgfxyline(6,-.1)(6,5.4) \pgfputat{\pgfxy(.15,5.1)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(6.15,5.1)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{.4cm}{0cm}} \begin{pgftranslate}{\pgfpoint{-.5cm}{0cm}} \pgfputat{\pgfxy(1.5,4)}{\pgfbox[center,center]{$D_i$}} \pgfputat{\pgfxy(2.85,2.8)}{\pgfbox[center,center]{$\tau$}} \pgfputat{\pgfxy(2.85,1.9)}{\pgfbox[center,center]{$\tau'$}} \pgfputat{\pgfxy(1,1.7)}{\pgfbox[center,center]{$x$}} \pgfputat{\pgfxy(2,1.7)}{\pgfbox[center,center]{$y$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,2)}{0.10cm} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.5,2.5)(3.25,2.5) \pgfxyline(2.5,2.2)(3.25,2.2) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.5cm}{0cm}} \pgfputat{\pgfxy(1.5,4)}{\pgfbox[center,center]{$D_t$}} \pgfputat{\pgfxy(1,1.7)}{\pgfbox[center,center]{$x'$}} \pgfputat{\pgfxy(2,1.7)}{\pgfbox[center,center]{$y'$}} \pgfputat{\pgfxy(2,3.3)}{\pgfbox[center,center]{$z'$}} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(2,3)}{0.10cm} \pgfnodeconnline{Node2}{Node3} \end{colormixin} \pgfnodecircle{Node1}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,2)}{0.10cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.2cm}{1.5cm}} \pgfputat{\pgfxy(1.5,2.5)}{\pgfbox[center,center]{$D_0$}} \pgfputat{\pgfxy(2.7,2.05)}{\pgfbox[center,center]{$\tau_{01}$}} \pgfputat{\pgfxy(2.7,.1)}{\pgfbox[center,center]{$\overline{\tau}_{01}$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.5,1.25)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.5,2)}{0.10cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.4,1.5)(3.4,2) \pgfxyline(2.4,.7)(3.4,0) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.55cm}{2.5cm}} \pgfputat{\pgfxy(3.15,2.25)}{\pgfbox[center,center]{$D_1$}} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(2.65,.5)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.65,.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(3.15,1.25)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(2.65,2)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(3.65,2)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node30}{Node50} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(2.65,.5)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.65,.5)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(3.15,1.25)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(2.65,2)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.55cm}{0cm}} \pgfputat{\pgfxy(3.15,2.25)}{\pgfbox[center,center]{$D_1$}} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(2.65,.5)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.65,.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(3.15,1.25)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(2.65,2)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(3.65,2)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node30}{Node50} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(2.65,.5)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.65,.5)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(3.15,1.25)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(3.65,2)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10.8cm}{1.5cm}} \pgfputat{\pgfxy(1.5,2.5)}{\pgfbox[center,center]{$D_1$}} \pgfputat{\pgfxy(3.7,2.5)}{\pgfbox[center,center]{$D_2$}} \pgfputat{\pgfxy(2.5,1.6)}{\pgfbox[center,center]{$\tau_{12}$}} \pgfnodecircle{Node1}[fill]{\pgfxy(1,.5)}{0.10cm} \pgfnodecircle{Node2}[fill]{\pgfxy(2,.5)}{0.10cm} \pgfnodecircle{Node3}[fill]{\pgfxy(1.5,1.25)}{0.10cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1,2)}{0.10cm} \pgfnodecircle{Node5}[fill]{\pgfxy(2,2)}{0.10cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node3}{Node5} \begin{colormixin}{50!white} \pgfnodecircle{Node10}[fill]{\pgfxy(3.2,.5)}{0.10cm} \pgfnodecircle{Node20}[fill]{\pgfxy(4.2,.5)}{0.10cm} \pgfnodecircle{Node30}[fill]{\pgfxy(3.7,1.25)}{0.10cm} \pgfnodecircle{Node40}[fill]{\pgfxy(3.2,2)}{0.10cm} \pgfnodecircle{Node50}[fill]{\pgfxy(4.2,2)}{0.10cm} \pgfnodecircle{Node60}[fill]{\pgfxy(4.7,2.75)}{0.10cm} \pgfnodeconnline{Node10}{Node30} \pgfnodeconnline{Node20}{Node30} \pgfnodeconnline{Node30}{Node40} \pgfnodeconnline{Node30}{Node50} \pgfnodeconnline{Node50}{Node60} \end{colormixin} \pgfnodecircle{Node11}[fill]{\pgfxy(3.2,.5)}{0.10cm} \pgfnodecircle{Node21}[fill]{\pgfxy(4.2,.5)}{0.10cm} \pgfnodecircle{Node31}[fill]{\pgfxy(3.7,1.25)}{0.10cm} \pgfnodecircle{Node41}[fill]{\pgfxy(3.2,2)}{0.10cm} \pgfnodecircle{Node51}[fill]{\pgfxy(4.2,2)}{0.10cm} \pgfnodeconnline{Node11}{Node31} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node31}{Node41} \pgfnodeconnline{Node31}{Node51} \pgfsetlinewidth{1.5pt} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfxyline(2.1,1.25)(2.9,1.25) \end{pgftranslate} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Equivalent transitions may be unrelated as {\it morphisms} due to symmetry breaking; b) equivalence of transitions is not a congruence relation under composition in $\ms{D}$.} \label{transitionnotcongruence} \end{figure} \vspace*{-.5cm} Congruence relations are so convenient that it is tempting to try to work with congruence classes of transitions, rather than the equivalence classes specified in definition \hyperref[deficorelative]{\ref{deficorelative}} above. One way to do this is to define a stricter equivalence relation, in which two transitions $\tau$ and $\tau'$ from $D_i$ to $D_t$ are taken to be equivalent if and only if they {\it share the same image.} This equivalence is obviously a congruence; the corresponding congruence classes may be called {\bf image-fixed co-relative histories}. A co-relative history may then be viewed as an equivalence class of image-fixed co-relative histories, with members indexed by an appropriate subgroup of the target $\tn{Aut}(D_t)$, moving the image of a representative transition around its orbit in $D_t$. Heuristically, transitions {\it freeze the actions} of both the automorphism groups $\tn{Aut}(D_i)$ and $\tn{Aut}(D_t)$, while image-fixed co-relative histories allow $\tn{Aut}(D_i)$ to act, and co-relative histories allow both groups to act. Co-relative histories embody the principle of iteration of structure (\hyperref[is]{IS}), by supplying ``higher-level relations" between pairs of directed sets in a configuration space of directed sets. Specifying an appropriate class of co-relative histories, ``suitable for the evolutionary viewpoint," elevates such a configuration space to a {\it kinematic scheme,} which is a ``category-like structure encoding evolutionary pathways for histories" in discrete causal theory, as described in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. Figure \hyperref[corelative]{\ref{corelative}} below illustrates how a particular family of four co-relative histories sharing a common cobase fits into a kinematic scheme called the {\it positive sequential kinematic scheme,} which is a close analogue of Sorkin and Rideout's kinematic scheme describing the sequential growth of causal sets. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{14cm}{6.45cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,5.9)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(8.4,5.9)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{-6.6cm}{-1cm}} \begin{pgfmagnify}{1.3}{1.3} \begin{pgftranslate}{\pgfpoint{1.25cm}{-.6cm}} \begin{colormixin}{20!white} \pgfnodecircle{Node3}[fill]{\pgfxy(7.25,2.75)}{.51cm}% \pgfnodecircle{Node5}[fill]{\pgfxy(5,3.5)}{.51cm}% \pgfnodecircle{Node6}[fill]{\pgfxy(6.25,4.1)}{.51cm}% \pgfnodecircle{Node7}[fill]{\pgfxy(7.5,4.5)}{.51cm}% \pgfnodecircle{Node8}[fill]{\pgfxy(9,4.5)}{.51cm}% \end{colormixin} \pgfsetlinewidth{1.2pt} \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.51cm}% \pgfnodecircle{Node5}[stroke]{\pgfxy(5,3.5)}{.51cm}% \pgfnodecircle{Node6}[stroke]{\pgfxy(6.25,4.1)}{.51cm}% \pgfnodecircle{Node7}[stroke]{\pgfxy(7.5,4.5)}{.51cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.51cm}% \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.25cm}{-1.6cm}} \begin{pgftranslate}{\pgfpoint{-1cm}{2.75cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.25cm}{3.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.15,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.15,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-2cm}{4.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.75cm}{4.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.2)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.2)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.75cm}{4.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \end{pgftranslate} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6cm}{-1.3cm}} \begin{pgfmagnify}{.85}{.85} \begin{pgftranslate}{\pgfpoint{0cm}{.7cm}} \begin{colormixin}{20!white} \pgfnodecircle{Node3}[fill]{\pgfxy(7.25,2.75)}{.51cm}% \pgfnodecircle{Node5}[fill]{\pgfxy(5,3.5)}{.51cm}% \pgfnodecircle{Node6}[fill]{\pgfxy(6.25,4.1)}{.51cm}% \pgfnodecircle{Node7}[fill]{\pgfxy(7.5,4.5)}{.51cm}% \pgfnodecircle{Node8}[fill]{\pgfxy(9,4.5)}{.51cm}% \end{colormixin} \pgfnodecircle{Node1}[virtual]{\pgfxy(8.25,0)}{.53cm}% \pgfnodecircle{Node2}[stroke]{\pgfxy(8.25,1.5)}{.53cm}% \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.53cm}% \pgfnodecircle{Node4}[stroke]{\pgfxy(9.25,2.75)}{.53cm}% \pgfnodecircle{Node5}[stroke]{\pgfxy(5,3.5)}{.53cm}% \pgfnodecircle{Node6}[stroke]{\pgfxy(6.25,4.1)}{.53cm}% \pgfnodecircle{Node7}[stroke]{\pgfxy(7.5,4.5)}{.53cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.53cm}% \pgfnodecircle{Node9}[stroke]{\pgfxy(10.25,4.1)}{.53cm}% \pgfnodecircle{Node10}[stroke]{\pgfxy(11.5,3.5)}{.53cm}% \pgfnodecircle{Node11}[virtual]{\pgfxy(1.5,4)}{.53cm}% \pgfnodecircle{Node12}[virtual]{\pgfxy(1.5,4.9)}{.53cm}% \pgfnodecircle{Node13}[virtual]{\pgfxy(1.8,5.2)}{.53cm}% \pgfnodecircle{Node14}[virtual]{\pgfxy(1.5,6.4)}{.53cm}% \pgfnodecircle{Node15}[virtual]{\pgfxy(2,6.4)}{.53cm}% \pgfnodecircle{Node16}[virtual]{\pgfxy(1.5,8.1)}{.53cm}% \pgfnodecircle{Node17}[virtual]{\pgfxy(2.2,7.8)}{.53cm}% \pgfnodecircle{Node18}[virtual]{\pgfxy(1.8,9)}{.53cm}% \pgfnodecircle{Node19}[stroke]{\pgfxy(5.1,6.4)}{.53cm}% \pgfnodecircle{Node20}[stroke]{\pgfxy(6.8,7.5)}{.53cm}% \pgfnodecircle{Node21}[stroke]{\pgfxy(4,8)}{.53cm}% \pgfnodecircle{Node22}[virtual]{\pgfxy(2.9,9.2)}{.53cm} \pgfnodecircle{Node23}[virtual]{\pgfxy(8.2,9.2)}{.53cm} \pgfnodecircle{Node24}[stroke]{\pgfxy(8,7)}{.53cm}% \pgfnodecircle{Node25}[virtual]{\pgfxy(6.1,8.7)}{.53cm}% \pgfnodecircle{Node26}[virtual]{\pgfxy(4.9,9)}{.53cm}% \pgfnodecircle{Node27}[virtual]{\pgfxy(3.7,9)}{.53cm}% \pgfnodecircle{Node28}[virtual]{\pgfxy(7.3,9)}{.53cm}% \pgfnodecircle{Node29}[stroke]{\pgfxy(9.2,8.2)}{.53cm}% \pgfnodecircle{Node30}[virtual]{\pgfxy(13.8,9)}{.53cm}% \pgfnodecircle{Node31}[virtual]{\pgfxy(13.4,7.9)}{.53cm}% \pgfnodecircle{Node32}[virtual]{\pgfxy(12.5,9)}{.53cm}% \pgfnodecircle{Node33}[virtual]{\pgfxy(11.6,9)}{.53cm}% \pgfnodecircle{Node34}[virtual]{\pgfxy(10.4,9)}{.53cm}% \pgfnodecircle{Node35}[virtual]{\pgfxy(9.8,9)}{.53cm}% \pgfnodecircle{Node36}[virtual]{\pgfxy(12.8,6.7)}{.53cm}% \pgfnodecircle{Node37}[virtual]{\pgfxy(14.6,7.1)}{.53cm}% \pgfnodecircle{Node38}[virtual]{\pgfxy(14.2,6)}{.53cm}% \pgfnodecircle{Node39}[virtual]{\pgfxy(15.4,5.8)}{.53cm}% \pgfnodecircle{Node40}[virtual]{\pgfxy(14.7,4.7)}{.53cm}% \pgfnodecircle{Node41}[virtual]{\pgfxy(15.8,4.3)}{.53cm} \pgfnodecircle{Node1901}[virtual]{\pgfxy(4.4,8.7)}{.1cm}% \pgfnodecircle{Node1902}[virtual]{\pgfxy(4.6,8.7)}{.1cm}% \pgfnodecircle{Node1903}[virtual]{\pgfxy(4.8,8.7)}{.1cm}% \pgfnodecircle{Node1904}[virtual]{\pgfxy(5,8.7)}{.1cm}% \pgfnodecircle{Node1905}[virtual]{\pgfxy(5.2,8.7)}{.1cm}% \pgfnodecircle{Node1906}[virtual]{\pgfxy(5.4,8.7)}{.1cm}% \pgfnodecircle{Node1907}[virtual]{\pgfxy(5.6,8.7)}{.1cm}% \pgfnodecircle{Node1908}[virtual]{\pgfxy(2.3,7.1)}{.1cm}% \pgfnodecircle{Node1909}[virtual]{\pgfxy(2.3,7.3)}{.1cm}% \pgfnodecircle{Node1910}[virtual]{\pgfxy(2.3,7.5)}{.1cm}% \pgfnodecircle{Node1911}[virtual]{\pgfxy(2.3,7.7)}{.1cm}% \pgfnodecircle{Node1912}[virtual]{\pgfxy(2.3,7.9)}{.1cm}% \pgfnodecircle{Node1913}[virtual]{\pgfxy(2.3,8.1)}{.1cm}% \pgfnodecircle{Node1914}[virtual]{\pgfxy(2.3,8.3)}{.1cm}% \pgfnodecircle{Node1915}[virtual]{\pgfxy(2.3,8.5)}{.1cm}% \pgfnodecircle{Node1916}[virtual]{\pgfxy(2.3,8.7)}{.1cm}% \pgfnodecircle{Node2001}[virtual]{\pgfxy(6.7,8.7)}{.1cm}% \pgfnodecircle{Node2002}[virtual]{\pgfxy(6.85,8.7)}{.1cm}% \pgfnodecircle{Node2003}[virtual]{\pgfxy(7,8.7)}{.1cm}% \pgfnodecircle{Node2004}[virtual]{\pgfxy(7.15,8.7)}{.1cm}% \pgfnodecircle{Node2005}[virtual]{\pgfxy(7.3,8.7)}{.1cm}% \pgfnodecircle{Node2006}[virtual]{\pgfxy(7.45,8.7)}{.1cm}% \pgfnodecircle{Node2007}[virtual]{\pgfxy(7.6,8.7)}{.1cm}% \pgfnodecircle{Node2008}[virtual]{\pgfxy(7.75,8.7)}{.1cm}% \pgfnodecircle{Node2009}[virtual]{\pgfxy(7.9,8.7)}{.1cm}% \pgfnodecircle{Node2010}[virtual]{\pgfxy(8.05,8.7)}{.1cm}% \pgfnodecircle{Node2011}[virtual]{\pgfxy(8.2,8.7)}{.1cm}% \pgfnodecircle{Node2012}[virtual]{\pgfxy(8.35,8.7)}{.1cm}% \pgfnodecircle{Node2013}[virtual]{\pgfxy(8.5,8.7)}{.1cm}% \pgfnodecircle{Node2014}[virtual]{\pgfxy(8.65,8.7)}{.1cm}% \pgfnodecircle{Node2015}[virtual]{\pgfxy(8.8,8.7)}{.1cm}% \pgfnodecircle{Node2016}[virtual]{\pgfxy(8.95,8.7)}{.1cm}% \pgfnodecircle{Node2401}[virtual]{\pgfxy(7.1,8.7)}{.1cm}% \pgfnodecircle{Node2402}[virtual]{\pgfxy(7.25,8.7)}{.1cm}% \pgfnodecircle{Node2403}[virtual]{\pgfxy(7.4,8.7)}{.1cm}% \pgfnodecircle{Node2404}[virtual]{\pgfxy(7.55,8.7)}{.1cm}% \pgfnodecircle{Node2405}[virtual]{\pgfxy(7.7,8.7)}{.1cm}% \pgfnodecircle{Node2406}[virtual]{\pgfxy(7.85,8.7)}{.1cm}% \pgfnodecircle{Node2407}[virtual]{\pgfxy(8,8.7)}{.1cm}% \pgfnodecircle{Node2408}[virtual]{\pgfxy(8.15,8.7)}{.1cm}% \pgfnodecircle{Node2409}[virtual]{\pgfxy(8.3,8.7)}{.1cm}% \pgfnodecircle{Node2410}[virtual]{\pgfxy(8.45,8.7)}{.1cm}% \pgfnodecircle{Node2411}[virtual]{\pgfxy(8.6,8.7)}{.1cm}% \pgfnodecircle{Node2412}[virtual]{\pgfxy(8.75,8.7)}{.1cm}% \pgfnodecircle{Node2413}[virtual]{\pgfxy(6.8,8.7)}{.1cm}% \pgfnodecircle{Node2414}[virtual]{\pgfxy(6.95,8.7)}{.1cm}% \pgfnodecircle{Node2415}[virtual]{\pgfxy(8.9,8.7)}{.1cm}% \pgfnodecircle{Node2416}[virtual]{\pgfxy(6.65,8.7)}{.1cm}% \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node4}{Node9} \pgfnodeconnline{Node4}{Node10} \pgfnodeconnline{Node5}{Node21} \pgfnodeconnline{Node8}{Node21} \pgfnodeconnline{Node5}{Node20} \pgfnodeconnline{Node5}{Node22} \pgfnodeconnline{Node7}{Node20} \pgfnodeconnline{Node5}{Node19} \pgfnodeconnline{Node6}{Node19} \pgfnodeconnline{Node5}{Node11} \pgfnodeconnline{Node5}{Node12} \pgfnodeconnline{Node5}{Node13} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node15} \pgfnodeconnline{Node6}{Node24} \pgfnodeconnline{Node7}{Node24} \pgfnodeconnline{Node6}{Node25} \pgfnodeconnline{Node8}{Node23} \pgfnodeconnline{Node6}{Node16} \pgfnodeconnline{Node6}{Node17} \pgfnodeconnline{Node6}{Node18} \pgfnodeconnline{Node6}{Node23} \pgfnodeconnline{Node7}{Node26} \pgfnodeconnline{Node7}{Node27} \pgfnodeconnline{Node7}{Node28} \pgfnodeconnline{Node7}{Node29} \pgfnodeconnline{Node8}{Node29} \pgfnodeconnline{Node8}{Node30} \pgfnodeconnline{Node8}{Node31} \pgfnodeconnline{Node8}{Node32} \pgfnodeconnline{Node8}{Node33} \pgfnodeconnline{Node9}{Node33} \pgfnodeconnline{Node8}{Node34} \pgfnodeconnline{Node10}{Node34} \pgfnodeconnline{Node9}{Node35} \pgfnodeconnline{Node9}{Node36} \pgfnodeconnline{Node9}{Node37} \pgfnodeconnline{Node9}{Node38} \pgfnodeconnline{Node9}{Node39} \pgfnodeconnline{Node10}{Node39} \pgfnodeconnline{Node10}{Node40} \pgfnodeconnline{Node10}{Node41} \pgfnodeconnline{Node19}{Node1901} \pgfnodeconnline{Node19}{Node1902} \pgfnodeconnline{Node19}{Node1903} \pgfnodeconnline{Node19}{Node1904} \pgfnodeconnline{Node19}{Node1905} \pgfnodeconnline{Node19}{Node1906} \pgfnodeconnline{Node19}{Node1907} \pgfnodeconnline{Node19}{Node1908} \pgfnodeconnline{Node19}{Node1909} \pgfnodeconnline{Node19}{Node1910} \pgfnodeconnline{Node19}{Node1911} \pgfnodeconnline{Node19}{Node1912} \pgfnodeconnline{Node19}{Node1913} \pgfnodeconnline{Node19}{Node1914} \pgfnodeconnline{Node19}{Node1915} \pgfnodeconnline{Node19}{Node1916} \pgfnodeconnline{Node20}{Node2001} \pgfnodeconnline{Node20}{Node2002} \pgfnodeconnline{Node20}{Node2003} \pgfnodeconnline{Node20}{Node2004} \pgfnodeconnline{Node20}{Node2005} \pgfnodeconnline{Node20}{Node2006} \pgfnodeconnline{Node20}{Node2007} \pgfnodeconnline{Node20}{Node2008} \pgfnodeconnline{Node20}{Node2009} \pgfnodeconnline{Node20}{Node2010} \pgfnodeconnline{Node20}{Node2011} \pgfnodeconnline{Node20}{Node2012} \pgfnodeconnline{Node20}{Node2013} \pgfnodeconnline{Node20}{Node2014} \pgfnodeconnline{Node20}{Node2015} \pgfnodeconnline{Node20}{Node2016} \pgfnodeconnline{Node24}{Node2401} \pgfnodeconnline{Node24}{Node2402} \pgfnodeconnline{Node24}{Node2403} \pgfnodeconnline{Node24}{Node2404} \pgfnodeconnline{Node24}{Node2405} \pgfnodeconnline{Node24}{Node2406} \pgfnodeconnline{Node24}{Node2407} \pgfnodeconnline{Node24}{Node2408} \pgfnodeconnline{Node24}{Node2409} \pgfnodeconnline{Node24}{Node2410} \pgfnodeconnline{Node24}{Node2411} \pgfnodeconnline{Node24}{Node2412} \pgfnodeconnline{Node24}{Node2413} \pgfnodeconnline{Node24}{Node2414} \pgfnodeconnline{Node24}{Node2415} \pgfnodeconnline{Node24}{Node2416} \begin{pgfscope} \pgfsetlinewidth{2pt} \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.51cm}% \pgfnodecircle{Node5}[stroke]{\pgfxy(5,3.5)}{.51cm}% \pgfnodecircle{Node6}[stroke]{\pgfxy(6.25,4.1)}{.51cm}% \pgfnodecircle{Node7}[stroke]{\pgfxy(7.5,4.5)}{.51cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.51cm}% \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \end{pgfscope} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{-.3cm}} \begin{pgftranslate}{\pgfpoint{0cm}{1.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-1cm}{2.75cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1cm}{2.75cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.25cm}{3.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.15,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.15,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-2cm}{4.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.75cm}{4.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.2)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.2)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.75cm}{4.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2cm}{4.1cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,.8)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.8)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.25cm}{3.5cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(7.9,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.6,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.15cm}{6.4cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.95)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-1.45cm}{7.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-4.25cm}{8cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \comment{ \begin{pgftranslate}{\pgfpoint{-2.15cm}{8.8cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.55)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} } \begin{pgftranslate}{\pgfpoint{-.25cm}{7cm}} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.1,1.3)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.95cm}{8.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.2,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.2)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.4,1.2)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.6,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \comment{ \begin{pgftranslate}{\pgfpoint{2.15cm}{9cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} } \end{pgftranslate} \end{pgfmagnify} \end{pgftranslate} \begin{pgfscope} \color{white} \pgfmoveto{\pgfxy(8.25,-.1)} \pgflineto{\pgfxy(8.25,6.65)} \pgflineto{\pgfxy(7.6,6.65)} \pgflineto{\pgfxy(7.6,-.1)} \pgflineto{\pgfxy(8.25,-.1)} \pgffill \pgfmoveto{\pgfxy(8.25,6.2)} \pgflineto{\pgfxy(8.25,6.65)} \pgflineto{\pgfxy(12.7,6.65)} \pgflineto{\pgfxy(12.7,6.2)} \pgflineto{\pgfxy(8.25,6.2)} \pgffill \pgfmoveto{\pgfxy(12.7,6.2)} \pgflineto{\pgfxy(12.7,6.75)} \pgflineto{\pgfxy(20,6.75)} \pgflineto{\pgfxy(20,6.2)} \pgflineto{\pgfxy(8.25,6.2)} \pgffill \pgfmoveto{\pgfxy(16.5,-.1)} \pgflineto{\pgfxy(16.5,6.65)} \pgflineto{\pgfxy(21,6.65)} \pgflineto{\pgfxy(21,-.1)} \pgflineto{\pgfxy(16.5,-.1)} \pgffill \pgfmoveto{\pgfxy(16.5,-.1)} \pgflineto{\pgfxy(16.5,-.3)} \pgflineto{\pgfxy(8.25,-.3)} \pgflineto{\pgfxy(8.25,-.1)} \pgflineto{\pgfxy(16.5,-.1)} \pgffill \end{pgfscope} \pgfxyline(0,-.1)(0,6.2) \pgfxyline(16.5,-.1)(16.5,6.2) \pgfxyline(8.25,-.1)(8.25,6.2) \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,6.2)(16.5,6.2) \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) a) Four irreducible co-relative histories sharing a common cobase; b) co-relative histories embedded in a kinematic scheme.} \label{corelative} \end{figure} \vspace*{-.5cm} Due to the nonuniqueness of co-relative histories between a given cobase and target, illustrated by McKay's example in figure \hyperref[corelativeexamples]{\ref{corelativeexamples}}b above, the abstract structure of a kinematic scheme is generally multidirected, rather than merely directed. ``Higher-level analogues" of irreducibility and independence are of interest in this context. For example, the four co-relative histories illustrated in figure \hyperref[corelative]{\ref{corelative}}a above are {\it irreducible,} since each target differs from the common cobase by a ``minimal amount;" i.e., by a single element. McKay's example implies that {\it irreducible co-relative histories are generally nonunique,} which is just another way of restating the fact that multidirected sets are necessary to encode the abstract structure of kinematic schemes. Turning to the subject of independence, a basic requirement built into the definition of a kinematic scheme is that it must include ``enough independent co-relative histories" to admit ``evolutionary pathways for every possible history." This requirement is formalized by the condition of {\it weak accessibility} (\hyperref[wa]{WA}), defined in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. \subsection{Abstract Quantum Causal Theory via Path Summation}\label{subsectionquantumpathsummation} \refstepcounter{textlabels}\label{adaptinghistories} {\bf Adapting the Histories Approach.} The histories approach to quantum theory, first implemented in the continuum setting via Feynman's path integral, may be adapted without {\it formal} difficulty to the discrete causal context, by means of the theory of path summation over a multidirected set, outlined in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}} above. In this section, I carry out this adaptation in an important special case, involving a {\it finite region of a locally finite acyclic multidirected set.} Following such an adaptation, the problem ceases to be a purely formal one, and the actual suitability of the resulting theory becomes the primary concern. Conceptually, one must consider the range of physical interpretations that may be assigned to the mathematical structures appearing in the theory, and must choose those interpretations most appropriate and useful for the intended applications. Technically, one must examine whether or not the chosen approach treats the information involved in the theory in an adequate fashion, and produces intelligible results. \refstepcounter{textlabels}\label{backgrounddepadapt} The most na\"{i}ve adaptation of the histories approach to discrete causal theory involves simply replacing continuum spacetime with a directed set $D$, and performing path summation over $D$. Examples of this approach, in the special case where $D$ is a causal set, already appear in the literature. These efforts suffer from multiple problems, including the structural deficiencies associated with transitivity and interval finiteness, the shortcomings of element space, the inadequacy of algebraic devices such as incidence algebras, and the problem of background dependence. With the exception of this latter problem, the results of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} of this paper offer the prospect of significant improvements, without requiring any radical conceptual changes. These results point the way toward flexible new alternatives to quantum field theories in curved spacetime and lattice field theories, involving {\it path summation on relation spaces over fixed locally finite directed sets.} \refstepcounter{textlabels}\label{backgroundindepadapt} The desire for a {\it background-independent} theory is the final catalyst for the radical change in viewpoint represented by the principle of iteration of structure (\hyperref[is]{IS}), and the theory of co-relative histories, introduced in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, and further explored in the context of kinematic schemes in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. While there is existing precedent for these ideas in the work of Sorkin, Isham, and others, it seems that a truly clear structural picture emerges only by combining the causal metric hypothesis (\hyperref[cmh]{CMH}), discreteness, the independence convention (\hyperref[ic]{IC}), local finiteness (\hyperref[lf]{LF}), hidden structure (\hyperref[hs]{HS}), the relative viewpoint (\hyperref[rv]{RV}), the histories approach, and background independence. This leads to the following conclusion, which I call the {\it path summation principle:} \refstepcounter{textlabels}\label{appliestoboth} \refstepcounter{textlabels}\label{ps} \hspace*{.3cm} PS. {\bf Path Summation:} {\it A background-independent version of the histories approach to discrete \\\hspace*{1cm} quantum causal theory may be constructed using the same abstract methods that apply in the \\\hspace*{1cm} background-dependent context, expressed in terms of path summation over multidirected sets. \\\hspace*{1cm} } The proper object over which to perform path summation in the background-independent context is the {\it relation space over the underlying multidirected set of a kinematic scheme of directed sets.} This relation space represents the abstract directed structure of a distinguished class of co-relative histories. The physical interpretation of this conclusion is that the quantum theory of fundamental spacetime structure ultimately involves not a space of elements, nor a space of relations, nor even a space of classical universes, but a space of {\it relationships between pairs of classical universes.} This viewpoint is further developed in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. \refstepcounter{textlabels}\label{contpathintegral} {\bf Feynman's Continuum Path Integral.} For the convenience of the reader, I briefly review Feynman's construction before discussing its discrete causal analogues. Following Feynman's 1948 paper \cite{FeynmanSOH48}, path integrals over continuum manifolds have found broad use, particularly in quantum field theory, in determining complex-valued {\it quantum amplitudes,} interpreted as encoding relative probabilities for the values of physical measurements. Although the credit for this approach rightly belongs to Feynman, it has important antecedents. Feynman explicitly credits the 1935 edition of Dirac's classic book {\it The Principles of Quantum Mechanics} \cite{Dirac35}, while Norbert Wiener's earlier work on Brownian motion, published in the early 1920's just before the introduction of ``modern quantum theory" by Schr\"{o}dinger and Heisenberg, seems to represent the first appearance of path integrals in the literature. Here, I give a sketch of continuum path integration using the case originally considered by Feynman: the {\it quantum-theoretic behavior of a single particle in a finite region of Euclidean spacetime}. This purpose of this sketch is to motivate the corresponding abstract quantum causal theory to follow. Perhaps surprisingly, there is little conceptual advantage to be gained, for this specific purpose, by considering more ``realistic" or elaborate scenarios involving, for instance, special relativity or field theory. In particular, most of the familiar technical machinery for evaluating path integrals in quantum field theory is wholly irrelevant in this context. Let $X$ be a finite subset of Euclidean spacetime, bounded between two spatial sections $t=t^\pm$, where $t$ is the universal Newtonian time parameter, and $t^\pm$ are specific time values. The {\it lower boundary} $\sigma^-$ of $X$ is the set of all lower endpoints of temporal intervals in $X$ at fixed spatial points, and the {\it upper boundary} $\sigma^+$ of $X$ is the set of all upper endpoints of such intervals.\footnotemark\footnotetext{Mathematically, one may imagine all manner of perverse spacetime regions. For the present illustrative purpose, it does no harm to assume that $X$ is bounded by a smooth hypersurface, or even that $X$ is convex.} Let $\Gamma=\Gamma_{\tn{\fsz max}}(X)$ be the set of all {\it maximal directed paths} in $X$; i.e., continuous future-directed paths in $X$ beginning somewhere on $\sigma^-$ and terminating somewhere on $\sigma^+$. Consider the motion of a single particle, with Lagrangian $\ms{L}$ and classical action $\ms{S}=\int_\gamma \ms{L}dt$, along a path $\gamma$ in $\Gamma$. Feynman's approach to studying the quantum behavior of such a particle is based on the two following ``postulates:" \refstepcounter{textlabels}\label{feynpost} \refstepcounter{textlabels}\label{f1} \hspace*{.3cm} F1. {\bf Quantum Amplitude}: {\it If an ``ideal measurement" is performed, to determine whether \\\hspace*{1.05cm} the trajectory of the particle belongs to $\Gamma=\Gamma_{\tn{\fsz max}}(X)$, then the probability of an affirmative \\\hspace*{1.05cm} result is the square modulus of a complex-valued {\bf quantum amplitude} $\psi(X;\ms{L})$, given \\\hspace*{1.05cm} by summing complex contributions from each path $\gamma\in\Gamma$.}(Adapted from \cite{FeynmanSOH48}, page 8.) \refstepcounter{textlabels}\label{f2} \hspace*{.3cm} F2. {\bf Phase Map}: {\it Each path $\gamma\in\Gamma$ contributes equally in magnitude to the amplitude $\psi(X;\ms{L})$, \\\hspace*{1.1cm} with complex {\bf phase} $\Theta(\gamma)=e^{\frac{i}{\hbar}\ms{S}(\gamma)}$, where $\ms{S}$ is the classical action corresponding to $\ms{L}$, \\\hspace*{1.1cm} and where $\hbar=h/2\pi$ is Planck's reduced constant.}(Adapted from \cite{FeynmanSOH48}, page 9.) \refstepcounter{textlabels}\label{feynamp} What these ``postulates" really describe is the {\it heuristic intent of a limiting procedure,} in which the interval $[t^-,t^+]$ is subjected to a sequence of increasingly fine {\it partitions} $\Delta=\{t^-=t_0,t_1,...,t_{N}=t^+\}$, and a ``path" $\gamma$ is specified by a sequence $\mbf{x}:=\{x_0,x_1,...,x_{N}\}$ of position values at each time in $\Delta$. A suitable interpolating convention must be chosen for the ``path increments" $\gamma_i$ of $\gamma$ between each pair of spacetime points $(x_i,t_i)$ and $(x_{i+1},t_{i+1})$. For example, $\gamma_i$ might be taken to be the {\it classical trajectory} between its endpoints, as determined by the Lagrangian $\ms{L}$, according to Hamilton's principle. Denote by $\sigma_i=\sigma_i(\Delta)$ the spatial section of $X$ at time $t=t_i$ in the partition $\Delta$, and let $|\Delta|$ be the norm of $\Delta$; i.e., the supremum of the values $t_{i+1}-t_i$. Then the amplitude $\psi(X;\ms{L})$ is {\it defined} to be the limit \[\psi(X;\ms{L}):=\lim_{|\Delta|\to0}\int_{\mathbf{x}}\psi(\Delta;\mathbf{x};\ms{L})d\mathbf{x}\] \begin{equation}\label{continuumpathintegral} :=\lim_{|\Delta|\to0}\int_{\sigma_N}... \int_{\sigma_1}\int_{\sigma_0} C\cdot\tn{exp}\Bigg(\sum_{i=1}^{N}\frac{i}{\hbar}\ms{S}(\gamma_i)\Bigg)dx_0dx_1...dx_{N}, \end{equation} if this limit exists. The first (i.e., top) expression in equation \hyperref[continuumpathintegral]{\ref{continuumpathintegral}} is purely symbolic, meaning ``integrate $N+1$ times, over all values of $x_0,...,x_N$, a complex quantity $\psi(\Delta;\mbf{x};\ms{L})$, depending on a particular choice of $x_0,...,x_N$ for the partition $\Delta$, and on the Lagrangian $\ms{L}$, then take the limit as $|\Delta|\rightarrow 0$." The second expression makes this more explicit at a particular level of approximation. The sum in the exponential comes from {\it multiplying} the phases of each ``increment" $\gamma_i$ of a typical ``path" $\gamma$, determined by $\Delta$ and $\mbf{x}$. The number $C$ appearing in the integrand is a proportionality factor which may be ignored in the present context. Figure \hyperref[feynman]{\ref{feynman}}a below illustrates the setting for Feynman's path integral. Comparing this with figure \hyperref[permeability]{\ref{permeability}}a of section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, one may observe that the spatial sections $\sigma_i$ serve as {\it Cauchy surfaces} for $X$, since every future-directed path $\gamma$ in $X$ from the past of $\sigma_i$ to the future of $\sigma_i$ intersects $\sigma_i$ uniquely. The physical significance of these Cauchy surfaces is that Feynman's path integral is {\it information-theoretically adequate;} i.e., it captures all information flowing from the lower boundary $\sigma^-$ of $X$ to the upper boundary $\sigma^+$ of $X$. In particular, there is no {\it permeability problem} of the type studied by Major, Rideout, and Surya \cite{RideoutSpatialHypersurface06} in the context of causal set theory. Figure \hyperref[feynman]{\ref{feynman}}b illustrates the derivation of Schr\"{o}dinger's equation via Feynman's path integral, which is described in section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{6.1cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,5.8)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(8.4,5.8)}{\pgfbox[left,center]{b)}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,6.1)(16.5,6.1) \pgfxyline(0,-.1)(0,6.1) \pgfxyline(8.25,-.1)(8.25,6.1) \pgfxyline(16.5,-.1)(16.5,6.1) \begin{pgftranslate}{\pgfpoint{.2cm}{.3cm}} \begin{pgftranslate}{\pgfpoint{.7cm}{.2cm}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{20!white} \color{black} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgffill \end{colormixin} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgfstroke \pgfmoveto{\pgfxy(4,2.7)} \pgfcurveto{\pgfxy(3.7,2.6)}{\pgfxy(4.1,3.4)}{\pgfxy(3.7,3.4)} \pgfmoveto{\pgfxy(3.7,3.4)} \pgfcurveto{\pgfxy(4.1,3.4)}{\pgfxy(3.7,4.2)}{\pgfxy(4,4.1)} \pgfstroke \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(1,0)(9.5,0) \pgfxyline(1,0)(1,7.5) \pgfclearendarrow \pgfsetlinewidth{1.5pt} \pgfxyline(2.03,2.6)(8.17,2.6) \pgfxyline(2.15,4.2)(8.5,4.2) \end{pgfscope} \begin{pgftranslate}{\pgfpoint{.8cm}{0cm}} \begin{colormixin}{100!white} \color{black} \pgfsetlinewidth{1pt} \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfmoveto{\pgfxy(3.3,2.6)} \pgfcurveto{\pgfxy(3.55,3)}{\pgfxy(3.7,3.2)}{\pgfxy(3.85,3.4)} \pgfstroke \end{pgfscope} \pgfmoveto{\pgfxy(2.7,1.1)} \pgfcurveto{\pgfxy(2.7,1.4)}{\pgfxy(2.9,1.9)}{\pgfxy(3.3,2.6)} \pgfstroke \pgfmoveto{\pgfxy(3.85,3.4)} \pgfcurveto{\pgfxy(4.5,4.3)}{\pgfxy(4.8,5.8)}{\pgfxy(4.8,6.45)} \pgfstroke \end{colormixin} \color{black} \pgfnodecircle{Node1}[fill]{\pgfxy(2.7,1.1)}{0.14cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4.8,6.45)}{0.14cm} \end{pgftranslate} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(7.05,.45)}{\pgfbox[left,center]{$x$}} \pgfputat{\pgfxy(1.55,5.3)}{\pgfbox[left,center]{$t$}} \pgfputat{\pgfxy(3.1,3.8)}{\pgfbox[left,center]{$X$}} \pgfputat{\pgfxy(5.2,3.4)}{\pgfbox[left,center]{$\sigma_{i+1}$}} \pgfputat{\pgfxy(5.2,2.25)}{\pgfbox[left,center]{$\sigma_i$}} \pgfputat{\pgfxy(2.9,2.6)}{\pgfbox[left,center]{$\gamma_i$}} \pgfputat{\pgfxy(4.15,3.7)}{\pgfbox[left,center]{$\gamma$}} \pgfputat{\pgfxy(.65,.8)}{\pgfbox[left,center]{$t_1$}} \pgfputat{\pgfxy(.75,2.02)}{\pgfbox[left,center]{$t_i$}} \pgfputat{\pgfxy(.65,3.14)}{\pgfbox[left,center]{$t_{i+1}$}} \pgfputat{\pgfxy(3.58,-.1)}{\pgfbox[center,center]{$x_i$}} \pgfputat{\pgfxy(4.33,-.1)}{\pgfbox[center,center]{$x_{i+1}$}} \pgfputat{\pgfxy(.55,4.75)}{\pgfbox[left,center]{$t_{N+1}$}} \pgfxyline(1.3,.8)(1.5,.8) \pgfxyline(1.3,2.02)(1.5,2.02) \pgfxyline(1.3,3.14)(1.5,3.14) \pgfxyline(1.3,4.75)(1.5,4.75) \pgfxyline(3.58,.1)(3.58,.3) \pgfxyline(4.27,.1)(4.27,.3) \pgfsetdash{{2pt}{2pt}}{0pt} \pgfxyline(1.4,.8)(7.4,.8) \pgfxyline(1.4,2.02)(7.4,2.02) \pgfxyline(1.4,3.14)(7.4,3.14) \pgfxyline(1.4,4.75)(7.4,4.75) \pgfxyline(3.58,.4)(3.58,2.02) \pgfxyline(4.27,.4)(4.27,3.14) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.5cm}{.3cm}} \begin{pgftranslate}{\pgfpoint{.7cm}{.2cm}} \begin{pgfmagnify}{.7}{.7} \begin{colormixin}{20!white} \color{black} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgffill \end{colormixin} \pgfmoveto{\pgfxy(2,3)} \pgfcurveto{\pgfxy(2,0)}{\pgfxy(8.5,0)}{\pgfxy(8.5,4)} \pgfcurveto{\pgfxy(8.5,7)}{\pgfxy(2,8)}{\pgfxy(2,3)} \pgfstroke \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{5pt}} \pgfxyline(1,0)(9.5,0) \pgfxyline(1,0)(1,7.5) \pgfclearendarrow \pgfsetlinewidth{1.5pt} \pgfxyline(2.02,3.5)(8.46,3.5) \pgfxyline(2.2,4.3)(8.48,4.3) \end{pgfscope} \begin{pgftranslate}{\pgfpoint{.8cm}{0cm}} \begin{colormixin}{100!white} \color{black} \pgfsetlinewidth{1pt} \begin{pgfscope} \pgfsetendarrow{\pgfarrowtriangle{4pt}} \pgfmoveto{\pgfxy(2.7,1.1)} \pgfcurveto{\pgfxy(2.8,1.5)}{\pgfxy(2.9,1.9)}{\pgfxy(3.1,2.4)} \pgfstroke \pgfmoveto{\pgfxy(3.1,2.4)} \pgfcurveto{\pgfxy(3.3,2.9)}{\pgfxy(3.6,3.5)}{\pgfxy(3.9,3.95)} \pgfstroke \end{pgfscope} \pgfmoveto{\pgfxy(3.9,3.95)} \pgfcurveto{\pgfxy(4,4.1)}{\pgfxy(4.05,4.2)}{\pgfxy(4.13,4.3)} \pgfstroke \end{colormixin} \color{black} \pgfnodecircle{Node1}[fill]{\pgfxy(2.7,1.1)}{0.14cm} \pgfnodecircle{Node2}[fill]{\pgfxy(4.13,4.3)}{0.14cm} \pgfnodecircle{Node2}[fill]{\pgfxy(3.62,3.51)}{0.14cm} \end{pgftranslate} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(7.05,.45)}{\pgfbox[left,center]{$x$}} \pgfputat{\pgfxy(1.55,5.3)}{\pgfbox[left,center]{$t$}} \pgfputat{\pgfxy(4.3,1.8)}{\pgfbox[left,center]{$X^-$}} \pgfputat{\pgfxy(4.4,4)}{\pgfbox[left,center]{$X^+$}} \pgfputat{\pgfxy(5.25,2.9)}{\pgfbox[left,center]{$\sigma'$}} \pgfputat{\pgfxy(5.25,3.5)}{\pgfbox[left,center]{$\sigma''$}} \pgfputat{\pgfxy(3.35,2.95)}{\pgfbox[left,center]{$\delta\gamma$}} \pgfputat{\pgfxy(3.15,2)}{\pgfbox[left,center]{$\gamma$}} \pgfputat{\pgfxy(.8,2.65)}{\pgfbox[left,center]{$t'$}} \pgfputat{\pgfxy(.75,3.21)}{\pgfbox[left,center]{$t''$}} \pgfxyline(1.3,2.65)(1.5,2.65) \pgfxyline(1.3,3.21)(1.5,3.21) \pgfsetdash{{2pt}{2pt}}{0pt} \pgfxyline(1.4,2.65)(7.4,2.65) \pgfxyline(1.4,3.21)(7.4,3.21) \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Setting for Feynman's path integral; b) deriving Schr\"{o}dinger's equation via Feynman's path integral.} \label{feynman} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{causalpathint} \refstepcounter{textlabels}\label{dependsimperm} {\bf Abstract Analogues of Feynman's Path Integral in the Discrete Causal Context.} Using the theory of path summation over a multidirected set, introduced in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}} above, it is straightforward, at least in a formal sense, to define abstract analogues of Feynman's path integral, with maximal antichains playing the role of spatial sections of spacetime. However, new issues, as well as new choices, arise in the discrete causal context. First, one must select a suitable class of directed sets to serve as models of classical spacetime. The results of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}} of this paper suggest that certain classes of nontransitive locally finite directed sets are potentially better choices than the class of causal sets for this purpose. Second, the generic problem of permeability of maximal antichains in directed sets, discussed in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, renders na\"{i}ve path summation over an arbitrary directed set information-theoretically inadequate. The obvious solution is {\it to consider path sums over relation space}, where the permeability problem disappears, by theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}}. Third, one must decide what kinds of paths to consider. For example, is it necessary to include {\it all} maximal paths in path sums, or does it suffice to work in terms of maximal chains? Assuming that chains do suffice, do they provide the most natural and insightful point of view? For elementary information-theoretic reasons, I work exclusively with chains in this section, but do not attempt to argue that more complicated paths are always irrelevant. A remaining issue, absent both in Feynman's construction and in the context of causal sets, is the question of whether or not to allow directed sets including cycles, and if so, how to treat path summation when cycles are involved. Even a small finite region of a directed set containing a cycle admits an infinite number of distinct chains wrapping around this cycle different numbers of times, potentially dominating the corresponding path sum. Moreover, as already mentioned in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, ``higher-level cycles" arise in the theory of co-relative histories and kinematic schemes, even when the individual directed sets involved are acyclic, unless one also restricts to the finite case. Hence, the issue of cycles remains relevant in the {\it background-independent context} even for relatively conservative models of classical spacetime. A number of possible approaches to this problem may be devised. First, one may simply apply the same formal methods as in the acyclic case, and try to show that the resulting path sums converge in some suitable sense. Second, one might attempt to treat cycles ``holistically" in some way. A simple method would be to define an equivalence relation $\sim$ on paths, in which $\gamma\sim\gamma'$ if and only if $\gamma$ coincides with $\gamma'$ ``outside of cycles," then assign a single phase to each equivalence class. A more elaborate version of the same idea is to perform a causal atomic decomposition, in which the causal atoms are maximal cycles, and the resulting directed set is a ``maximal acyclic approximation" of the original directed set. Paths in this acyclic approximation then correspond to the above equivalence classes, and the phases assigned to its paths are partially determined by the internal cyclic structures of the ``elements" in their images. A mathematically attractive holistic approach to cycles is to adopt the relative viewpoint (\hyperref[rv]{RV}) and {\it generalize cycles to morphisms whose sources are cyclic directed sets,}\footnotemark\footnotetext{That is, {\it totally} cyclic directed sets; i.e., sets in which there are chains in both directions between any pair of elements.} just as paths generalize chains to morphisms whose sources are linear directed sets. Finally, if all cycles in a directed set happen to be ``large," like many of the closed timelike curves appearing in solutions to general relativity, one could investigate local problems without considering cycles. In this section, I sidestep these issues by focusing on the abstract acyclic case. Applying these simplifying assumptions, let $R$ be a finite subspace of the relation space $\ms{R}(M')$ over a locally finite acyclic multidirected set $M'=(M',R',i',t')$.\footnotemark\footnotetext{Primes are used to denote the underlying multidirected set $M'=(M',R',i',t')$, in order to leave the ``unprimed" symbol $R$ free to denote the object of principal interest; namely, a finite subspace of $R'$.} Let $\big(\tn{Ch}(R),\sqcup\big)$ be the chain concatenation semicategory over $R$, and let $\Gamma=\tn{Ch}_{\tn{\fsz max}}(R)$ be the subset\footnotemark\footnotetext{From the algebraic viewpoint, the concatenation product of any pair of elements of $\Gamma$ is zero, since these elements are maximal chains. Similarly, from the order-theoretic viewpoint, no element of $\Gamma$ directly precedes any other element of $\Gamma$ under the causal concatenation relation. Hence, it makes no difference whether $\Gamma$ is called a ``subset," a ``subsemicategory," or a ``subspace" in this context. I choose the first option.} of $\big(\tn{Ch}(R),\sqcup\big)$ consisting of all maximal chains in $R$. Let $\sigma$ be a maximal antichain in $R$, separating $R$ into the disjoint union $R^-\coprod\sigma\coprod R^+$, and let $\gamma\in \Gamma$ be a typical maximal chain in $R$, intersecting $\sigma$ in a unique element $r$. Comparing these designations to Feynman's continuum construction, $\ms{R}(M')$ corresponds to Euclidean spacetime, $R$ to the finite spacetime region $X$, $\Gamma$ to the set of maximal directed paths in $X$, $\sigma$ to a spatial section of $X$, and $\gamma$ to a maximal directed path in $X$. This setup is illustrated in figure \hyperref[causalpathsum]{\ref{causalpathsum}}a below, with the corresponding picture in the original element space $M'$ shown in figure \hyperref[causalpathsum]{\ref{causalpathsum}}b. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{16cm}{5.5cm} \begin{pgfmagnify}{1.03}{1.03} \pgfputat{\pgfxy(.15,5.2)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(10.85,5.2)}{\pgfbox[left,center]{b)}} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5.5) \pgfxyline(10.7,-.1)(10.7,5.5) \pgfxyline(16.5,-.1)(16.5,5.5) \pgfxyline(0,5.5)(16.5,5.5) \begin{pgftranslate}{\pgfpoint{-.3cm}{-.2cm}} \begin{pgfmagnify}{1}{1} \comment{ \begin{colormixin}{50!white} \pgfxyline(0,0)(12,0) \pgfxyline(0,.5)(12,.5) \pgfxyline(0,1)(12,1) \pgfxyline(0,1.5)(12,1.5) \pgfxyline(0,2)(12,2) \pgfxyline(0,2.5)(12,2.5) \pgfxyline(0,3)(12,3) \pgfxyline(0,3.5)(12,3.5) \pgfxyline(0,4)(12,4) \pgfxyline(0,4.5)(12,4.5) \pgfxyline(0,5)(12,5) \pgfxyline(0,5.5)(12,5.5) \pgfxyline(0,6)(12,6) \pgfxyline(0,0)(0,6) \pgfxyline(1,0)(1,6) \pgfxyline(2,0)(2,6) \pgfxyline(3,0)(3,6) \pgfxyline(4,0)(4,6) \pgfxyline(5,0)(5,6) \pgfxyline(6,0)(6,6) \pgfxyline(7,0)(7,6) \pgfxyline(8,0)(8,6) \pgfxyline(9,0)(9,6) \pgfxyline(10,0)(10,6) \pgfxyline(11,0)(11,6) \pgfxyline(12,0)(12,6) \end{colormixin} \pgfputat{\pgfxy(0,.25)}{\pgfbox[center,center]{0}} \pgfputat{\pgfxy(1,.25)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(2,.25)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(3,.25)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(4,.25)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(5,.25)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(6,.25)}{\pgfbox[center,center]{6}} \pgfputat{\pgfxy(7,.25)}{\pgfbox[center,center]{7}} \pgfputat{\pgfxy(8,.25)}{\pgfbox[center,center]{8}} \pgfputat{\pgfxy(9,.25)}{\pgfbox[center,center]{9}} \pgfputat{\pgfxy(10,.25)}{\pgfbox[center,center]{10}} \pgfputat{\pgfxy(11,.25)}{\pgfbox[center,center]{11}} \pgfputat{\pgfxy(12,.25)}{\pgfbox[center,center]{12}} \pgfputat{\pgfxy(0,1)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(0,2)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(0,3)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(0,4)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(0,5)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(0,6)}{\pgfbox[center,center]{6}} } \begin{colormixin}{25!white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0104}[fill]{\pgfxy(2.5,.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0105}[fill]{\pgfxy(1.6,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0110}[fill]{\pgfxy(3.7,.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0203}[fill]{\pgfxy(.8,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0206}[fill]{\pgfxy(1.3,3.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0207}[fill]{\pgfxy(2.1,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0405}[fill]{\pgfxy(2.3,1.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0603}[fill]{\pgfxy(1.7,4.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0703}[fill]{\pgfxy(2.4,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0803}[fill]{\pgfxy(2.9,5.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0904}[fill]{\pgfxy(3.1,.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2026}[fill]{\pgfxy(7.6,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2122}[fill]{\pgfxy(7.4,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2127}[fill]{\pgfxy(7.8,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2420}[fill]{\pgfxy(7.4,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2533}[fill]{\pgfxy(9.1,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2730}[fill]{\pgfxy(8.5,4.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2833}[fill]{\pgfxy(9.5,1.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2927}[fill]{\pgfxy(8.5,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2930}[fill]{\pgfxy(9,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2935}[fill]{\pgfxy(9.4,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3125}[fill]{\pgfxy(8.4,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3128}[fill]{\pgfxy(9.4,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3132}[fill]{\pgfxy(10.2,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3226}[fill]{\pgfxy(8.3,2.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3233}[fill]{\pgfxy(10.2,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3435}[fill]{\pgfxy(10.2,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3530}[fill]{\pgfxy(9.6,4.8)}{\pgfxy(0.17,0.17)} \pgfnodeconnline{Node0102}{Node0203} \pgfnodeconnline{Node0102}{Node0206} \pgfnodeconnline{Node0102}{Node0207} \pgfnodeconnline{Node0104}{Node0405} \pgfnodeconnline{Node0104}{Node0414} \pgfnodeconnline{Node0105}{Node0502} \pgfnodeconnline{Node0105}{Node0506} \pgfnodeconnline{Node0105}{Node0507} \pgfnodeconnline{Node0105}{Node0508} \pgfnodeconnline{Node0105}{Node0514} \pgfnodeconnline{Node0110}{Node1007} \pgfnodeconnline{Node0110}{Node1013} \pgfnodeconnline{Node0206}{Node0603} \pgfnodeconnline{Node0207}{Node0703} \pgfnodeconnline{Node0207}{Node0708} \pgfnodeconnline{Node0405}{Node0502} \pgfnodeconnline{Node0405}{Node0506} \pgfnodeconnline{Node0405}{Node0507} \pgfnodeconnline{Node0405}{Node0508} \pgfnodeconnline{Node0405}{Node0514} \pgfnodeconnline{Node0414}{Node1412} \pgfnodeconnline{Node0414}{Node1415} \pgfnodeconnline{Node0502}{Node0203} \pgfnodeconnline{Node0502}{Node0206} \pgfnodeconnline{Node0502}{Node0207} \pgfnodeconnline{Node0506}{Node0603} \pgfnodeconnline{Node0507}{Node0703} \pgfnodeconnline{Node0507}{Node0708} \pgfnodeconnline{Node0508}{Node0803} \pgfnodeconnline{Node0514}{Node1412} \pgfnodeconnline{Node0514}{Node1415} \pgfnodeconnline{Node0708}{Node0803} \pgfnodeconnline{Node0904}{Node0405} \pgfnodeconnline{Node0904}{Node0414} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0703} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1208}{Node0803} \pgfnodeconnline{Node1307}{Node0703} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1620}{Node2026} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2026}{Node2621} \pgfnodeconnline{Node2026}{Node2629} \pgfnodeconnline{Node2122}{Node2219} \pgfnodeconnline{Node2122}{Node2223} \pgfnodeconnline{Node2127}{Node2730} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2420}{Node2013} \pgfnodeconnline{Node2420}{Node2017} \pgfnodeconnline{Node2420}{Node2018} \pgfnodeconnline{Node2420}{Node2022} \pgfnodeconnline{Node2420}{Node2026} \pgfnodeconnline{Node2517}{Node1714} \pgfnodeconnline{Node2521}{Node2122} \pgfnodeconnline{Node2521}{Node2127} \pgfnodeconnline{Node2529}{Node2927} \pgfnodeconnline{Node2529}{Node2930} \pgfnodeconnline{Node2529}{Node2935} \pgfnodeconnline{Node2533}{Node3329} \pgfnodeconnline{Node2533}{Node3334} \pgfnodeconnline{Node2621}{Node2122} \pgfnodeconnline{Node2621}{Node2127} \pgfnodeconnline{Node2629}{Node2927} \pgfnodeconnline{Node2629}{Node2930} \pgfnodeconnline{Node2629}{Node2935} \pgfnodeconnline{Node2833}{Node3329} \pgfnodeconnline{Node2833}{Node3334} \pgfnodeconnline{Node2927}{Node2730} \pgfnodeconnline{Node2935}{Node3530} \pgfnodeconnline{Node3125}{Node2517} \pgfnodeconnline{Node3125}{Node2521} \pgfnodeconnline{Node3125}{Node2529} \pgfnodeconnline{Node3125}{Node2533} \pgfnodeconnline{Node3128}{Node2833} \pgfnodeconnline{Node3132}{Node3226} \pgfnodeconnline{Node3132}{Node3229} \pgfnodeconnline{Node3132}{Node3233} \pgfnodeconnline{Node3226}{Node2621} \pgfnodeconnline{Node3226}{Node2629} \pgfnodeconnline{Node3229}{Node2927} \pgfnodeconnline{Node3229}{Node2930} \pgfnodeconnline{Node3229}{Node2935} \pgfnodeconnline{Node3233}{Node3329} \pgfnodeconnline{Node3233}{Node3334} \pgfnodeconnline{Node3329}{Node2927} \pgfnodeconnline{Node3329}{Node2930} \pgfnodeconnline{Node3329}{Node2935} \pgfnodeconnline{Node3334}{Node3435} \pgfnodeconnline{Node3435}{Node3530} \comment{ \color{white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} \color{black} \pgfnoderect{Node0102}[stroke]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[stroke]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[stroke]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[stroke]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[stroke]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[stroke]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[stroke]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[stroke]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[stroke]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[stroke]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[stroke]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[stroke]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[stroke]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[stroke]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} } \end{colormixin} \begin{colormixin}{40!white} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.22,0.22)} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2517}{Node1714} \end{colormixin} \color{white} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \color{black} \pgfnoderect{Node1007}[stroke]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[stroke]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[stroke]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[stroke]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[stroke]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[stroke]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[stroke]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[stroke]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[stroke]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[stroke]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[stroke]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.22,0.22)} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1819}{Node1930} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11cm}{0cm}} \begin{pgfmagnify}{.9}{.9} \begin{colormixin}{25!white} \pgfnodecircle{Node1}[fill]{\pgfxy(.7,.5)}{0.09cm} \pgfnodecircle{Node2}[fill]{\pgfxy(.6,2.5)}{0.09cm} \pgfnodecircle{Node3}[fill]{\pgfxy(.5,5)}{0.09cm} \pgfnodecircle{Node4}[fill]{\pgfxy(1.1,1.5)}{0.09cm} \pgfnodecircle{Node5}[fill]{\pgfxy(.9,2)}{0.09cm} \pgfnodecircle{Node6}[fill]{\pgfxy(1,3.7)}{0.09cm} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.09cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.09cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.09cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.09cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.09cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.09cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.09cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.09cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.09cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.09cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.09cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.09cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.09cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.09cm} \pgfnodecircle{Node21}[fill]{\pgfxy(3.7,3.5)}{0.09cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.09cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.09cm} \pgfnodecircle{Node24}[fill]{\pgfxy(4,.5)}{0.09cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.09cm} \pgfnodecircle{Node26}[fill]{\pgfxy(4,2.5)}{0.09cm} \pgfnodecircle{Node27}[fill]{\pgfxy(4,4.3)}{0.09cm} \pgfnodecircle{Node28}[fill]{\pgfxy(4.4,1.5)}{0.09cm} \pgfnodecircle{Node29}[fill]{\pgfxy(4.5,3.3)}{0.09cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.09cm} \pgfnodecircle{Node31}[fill]{\pgfxy(5,.6)}{0.09cm} \pgfnodecircle{Node32}[fill]{\pgfxy(5.1,1.5)}{0.09cm} \pgfnodecircle{Node33}[fill]{\pgfxy(5,2.5)}{0.09cm} \pgfnodecircle{Node34}[fill]{\pgfxy(5.2,3.5)}{0.09cm} \pgfnodecircle{Node35}[fill]{\pgfxy(5,4.5)}{0.09cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node1}{Node5} \pgfnodeconnline{Node1}{Node10} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node6} \pgfnodeconnline{Node2}{Node7} \pgfnodeconnline{Node4}{Node5} \pgfnodeconnline{Node4}{Node14} \pgfnodeconnline{Node5}{Node2} \pgfnodeconnline{Node5}{Node6} \pgfnodeconnline{Node5}{Node7} \pgfnodeconnline{Node5}{Node8} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node6}{Node3} \pgfnodeconnline{Node7}{Node3} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node8}{Node3} \pgfnodeconnline{Node9}{Node4} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node20}{Node26} \pgfnodeconnline{Node21}{Node22} \pgfnodeconnline{Node21}{Node27} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node24}{Node20} \pgfnodeconnline{Node25}{Node17} \pgfnodeconnline{Node25}{Node21} \pgfnodeconnline{Node25}{Node29} \pgfnodeconnline{Node25}{Node33} \pgfnodeconnline{Node26}{Node21} \pgfnodeconnline{Node26}{Node29} \pgfnodeconnline{Node27}{Node30} \pgfnodeconnline{Node28}{Node33} \pgfnodeconnline{Node29}{Node27} \pgfnodeconnline{Node29}{Node30} \pgfnodeconnline{Node29}{Node35} \pgfnodeconnline{Node31}{Node25} \pgfnodeconnline{Node31}{Node28} \pgfnodeconnline{Node31}{Node32} \pgfnodeconnline{Node32}{Node26} \pgfnodeconnline{Node32}{Node29} \pgfnodeconnline{Node32}{Node33} \pgfnodeconnline{Node33}{Node29} \pgfnodeconnline{Node33}{Node34} \pgfnodeconnline{Node34}{Node35} \pgfnodeconnline{Node35}{Node30} \end{colormixin} \begin{colormixin}{50!white} \pgfnodecircle{Node7}[fill]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node8}[fill]{\pgfxy(1.6,4.6)}{0.11cm} \pgfnodecircle{Node9}[fill]{\pgfxy(2,.9)}{0.11cm} \pgfnodecircle{Node10}[fill]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node12}[fill]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[fill]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node15}[fill]{\pgfxy(2.3,5.1)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node17}[fill]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[fill]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node22}[fill]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[fill]{\pgfxy(3.3,4.7)}{0.11cm} \pgfnodecircle{Node25}[fill]{\pgfxy(3.9,1.3)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfnodeconnline{Node7}{Node8} \pgfnodeconnline{Node9}{Node10} \pgfnodeconnline{Node9}{Node13} \pgfnodeconnline{Node9}{Node17} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node10}{Node13} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node12}{Node8} \pgfnodeconnline{Node12}{Node15} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node14}{Node12} \pgfnodeconnline{Node14}{Node15} \pgfnodeconnline{Node16}{Node10} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node16}{Node17} \pgfnodeconnline{Node16}{Node20} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node19}{Node30} \pgfnodeconnline{Node20}{Node13} \pgfnodeconnline{Node20}{Node17} \pgfnodeconnline{Node20}{Node18} \pgfnodeconnline{Node20}{Node22} \pgfnodeconnline{Node22}{Node19} \pgfnodeconnline{Node22}{Node23} \pgfnodeconnline{Node25}{Node17} \end{colormixin} \comment{ \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} } \pgfnodecircle{Node7}[virtual]{\pgfxy(1.5,3.4)}{0.11cm} \pgfnodecircle{Node10}[virtual]{\pgfxy(1.9,1.5)}{0.11cm} \pgfnodecircle{Node12}[virtual]{\pgfxy(2,4.2)}{0.11cm} \pgfnodecircle{Node13}[virtual]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node14}[virtual]{\pgfxy(2.4,4)}{0.11cm} \pgfnodecircle{Node17}[virtual]{\pgfxy(3,2.4)}{0.11cm} \pgfnodecircle{Node18}[virtual]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[virtual]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node20}[virtual]{\pgfxy(3.4,1.4)}{0.11cm} \pgfnodecircle{Node22}[virtual]{\pgfxy(3.4,4)}{0.11cm} \pgfnodecircle{Node23}[virtual]{\pgfxy(3.3,4.7)}{0.11cm} \pgfsetlinewidth{2.5pt} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node20}{Node22} \color{white} \pgfsetlinewidth{1.5pt} \pgfnodeconnline{Node10}{Node7} \pgfnodeconnline{Node13}{Node7} \pgfnodeconnline{Node13}{Node12} \pgfnodeconnline{Node13}{Node14} \pgfnodeconnline{Node13}{Node19} \pgfnodeconnline{Node17}{Node14} \pgfnodeconnline{Node18}{Node14} \pgfnodeconnline{Node18}{Node19} \pgfnodeconnline{Node18}{Node22} \pgfnodeconnline{Node18}{Node23} \pgfnodeconnline{Node20}{Node22} \color{black} \pgfnodecircle{Node11}[fill]{\pgfxy(2.5,2)}{0.11cm} \pgfnodecircle{Node13}[fill]{\pgfxy(2.5,2.6)}{0.11cm} \pgfnodecircle{Node16}[fill]{\pgfxy(3,.7)}{0.11cm} \pgfnodecircle{Node18}[fill]{\pgfxy(3,3.3)}{0.11cm} \pgfnodecircle{Node19}[fill]{\pgfxy(2.9,4.8)}{0.11cm} \pgfnodecircle{Node30}[fill]{\pgfxy(4.5,5.2)}{0.11cm} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node16}{Node11} \pgfnodeconnline{Node11}{Node13} \pgfnodeconnline{Node13}{Node18} \pgfnodeconnline{Node19}{Node30} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(8,3)}{\pgfbox[left,center]{\large{$\ms{R}(M')$}}} \pgfputat{\pgfxy(4.2,2.7)}{\pgfbox[left,center]{\large{$R$}}} \pgfputat{\pgfxy(5.65,4.05)}{\pgfbox[left,center]{\large{$r$}}} \pgfputat{\pgfxy(13.6,3.7)}{\pgfbox[left,center]{\large{$r$}}} \pgfputat{\pgfxy(14.5,2.7)}{\pgfbox[left,center]{\large{$M'$}}} \pgfputat{\pgfxy(5,3.2)}{\pgfbox[left,center]{\large{$\gamma$}}} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Path summation over a finite subset $R$ of relation space, represented by large nodes; maximal antichain $\sigma\subset R$ represented by large open nodes; maximal chain $\gamma$ in $R$ represented by large dark nodes and edges; b) corresponding view in element space; thick white lines represent $\sigma$.} \label{causalpathsum} \end{figure} \vspace*{-.5cm} Referring to Feynman's path integral formula for the quantum amplitude $\psi(X;\ms{L})$ in the continuum context, appearing in equation \hyperref[continuumpathintegral]{\ref{continuumpathintegral}} above, note that the ``path increments" $\gamma_i$, appearing in the second (i.e., bottom) expression, are defined in terms of {\it pairs} of consecutive spatial sections $\sigma_i$ and $\sigma_{i+1}$, since a single section of continuum spacetime has ``no thickness." In the discrete causal context, however, the {\it single} maximal antichain of relations $\sigma$ accomplishes an analogous purpose. Each relation $r$ belonging to $\sigma$ constitutes an ``increment" of any chain $...\prec r^-\prec r\prec r^+\prec...$ passing through $r$.\footnotemark\footnotetext{For general paths, the situation is more complicated, since the portion of such a chain containing $r$ may be ``short-circuited" by a reducible relation.} Having specified the necessary data for path summation {\it intrinsic to} $R$, I now consider the target object $T$ in which the values of path sums over $R$ are taken to lie, and the phase map $\Theta:\tn{Ch}(R)\rightarrow T$ that supplies the terms of the particular path sum analogous to Feynman's path integral. Both $T$ and $\Theta$ are {\it a priori} independent of $R$, but physical and philosophical considerations narrow the field of possibilities considerably. For example, the target object $T$ is strongly constrained if one wishes to interpret path summation in terms of probabilities.\footnotemark\footnotetext{As noted in section \hyperref[settheoretic]{\ref{settheoretic}} above, Chris Isham advocates replacing probabilities with ``generalized truth values." See, for example, \cite{Isham11}, page 6.} The phase map $\Theta$ specifies the dynamics in discrete causal theory, just as Feynman's phase map specifies the dynamics in continuum theory, via the classical action. I will not attempt to conceal that I do not presently know precisely what the corresponding ``action" in the discrete causal context should be. As mentioned in section \hyperref[subsectionapproach]{\ref{subsectionapproach}} near the beginning of this paper, in the discussion of causal set phenomenology, this problem has been studied in the case of causal sets by Fay Dowker and a few others \cite{DowkerScalarCurvature10}. Under the strong form of the causal metric hypothesis (\hyperref[cmh]{CMH}), the phase map $\Theta$ ought to be somehow determined by ``underlying directed structure," perhaps via natural ``mediating structures," such as the valence fields introduced at the end of section \hyperref[subsectiontopology]{\ref{subsectiontopology}} above. In particular, in the background-independent context, the elements of $R$ represent co-relative histories, whose entire internal structures as equivalence classes of transitions are relevant to the choice of phase map under the {\it quantum causal metric hypothesis} (\hyperref[qcmh]{QCMH}), introduced in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} below. It is reasonable to surmise that {\it entropy} plays a significant role in the action in the background independent setting. In the present abstract context, I make simplifying assumptions regarding $T$ and $\Theta$, while retaining enough generality to yield reasonably interesting and relevant {\it causal Schr\"{o}dinger-type equations,} derived in section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} below. In particular, I assume that the target object $T$ is a commutative ring with unit, and that the phase map $\Theta$ is generated under the concatenation product on $\tn{Ch}(R)$ by a relation function $\theta:R\rightarrow T$. The latter assumption means that the phase $\Theta(\gamma)$ of any finite chain $\gamma=r_1\sqcup...\sqcup r_N$ in $R$ is the product \begin{equation}\label{chainfactorization} \Theta(\gamma)=\prod_{i=1}^N\theta(r_i). \end{equation} Since $R$ is a finite directed set, any chain $\gamma$ in $R$ has a unique factorization of this form.\footnotemark\footnotetext{Again, this simplification does not apply to general paths in $R$.} At a more technical level, $\Theta$ may be regarded as a ``semicategory morphism," from the chain concatenation semicategory $\big(\tn{Ch}(R),\sqcup\big)$ over $R$, to the target object $T$. Factorization of $\gamma$ is analogous to the partitioning of a continuum ``path" $\gamma$ into ``path increments" $\gamma_i$, via a partition $\Delta$, in Feynman's construction. The product formula for $\Theta(\gamma)$ in equation \hyperref[chainfactorization]{\ref{chainfactorization}} is analogous to Feynman's formula \[\Theta(\gamma)=\prod_{i=1}^{N}\Theta(\gamma_i)=\prod_{i=1}^{N}e^{\frac{i}{\hbar}\ms{S}(\gamma_i)}=\tn{exp}\Bigg(\sum_{i=1}^N\frac{i}{\hbar}\ms{S}(\gamma_i)\Bigg).\] As mentioned above, much more general treatments are possible. For example, the target object $T$ might be replaced by some higher-level algebraic structure, such as a {\it monoidal category.} \refstepcounter{textlabels}\label{abstractamp} Under these assumptions, the analogue of Feynman's quantum amplitude $\psi(X;\ms{L})$ is an element $\psi(R;\theta)$ of $T$, given by summing phases over the set $\Gamma$ of maximal chains in $R$. In the present simplified setting, the existence of $\psi(R;\theta)$ is automatic. The choice of notation for this {\it generalized quantum amplitude} reflects the fact that the relation function $\theta$ is {\it information-theoretically analogous to a Lagrangian.} It is useful to define an element $\Psi(R;\theta)$ of the chain concatenation algebra $T^\sqcup[R]$ to serve as an {\it algebraic precursor} to $\psi(R;\theta)$.\footnotemark\footnotetext{There is nothing to prevent consideration of such precursor functionals in continuum theory, but they are {\it a priori} cumbersome as formal sums due to their large numbers of terms.} As a causal path functional, $\Psi(R;\theta)$ is merely the {\it restriction to} $\Gamma$ of the phase map $\Theta$ generated by the relation function $\theta$. The corresponding amplitude $\psi(R;\theta)$ is then given by applying the global evaluation map $e_R$ to $\Psi(R;\theta)$. The following definition makes these notions precise:\\ \begin{defi}\label{maximalfunctional} Let $M'=(M',R',i',t')$ be a locally finite acyclic multidirected set with relation space $\ms{R}(M')$, and let $R$ be a finite subset of $\ms{R}(M')$, viewed as a full subobject. Let $\Gamma=\tn{Ch}_{\tn{\fsz max}}(R)$ be the set of maximal chains in $R$, viewed as a subset of the chain concatenation semicategory $\big(\tn{Ch}(R),\sqcup\big)$. Let $T$ be a commutative ring with unit, and let $\Theta:\tn{Ch}(R)\rightarrow T$ be a multiplicative phase map generated by a relation function $\theta:R\rightarrow T$. \begin{enumerate} \item The {\bf maximal chain functional} $\Psi(R;\theta)$ of $R$ with respect to $\theta$ is the causal path functional $\Gamma\rightarrow T$ given by restricting the phase map $\Theta$ to $\Gamma$: \begin{equation}\label{precursoramplitude} \Psi(R;\theta)=\Theta|_\Gamma=\sum_{\gamma\in \Gamma}\Theta(\gamma)\gamma. \end{equation} \item The {\bf (generalized quantum) amplitude} $\psi(R;\theta)$ of $R$ with respect to $\theta$ is the element of $T$ given by applying the global evaluation map $e_R$ to $\Psi(R;\theta)$: \begin{equation}\label{amplitude} \psi(R;\theta)=e_R\big(\Psi(R;\theta)\big)=\sum_{\gamma\in\Gamma}\Theta(\gamma)=\sum_{\gamma \in \Gamma}\Bigg(\prod_{i=1}^{N_\gamma}\theta(r_{\gamma,i})\Bigg), \end{equation} where $\gamma=r_{\gamma,1}\sqcup...\sqcup r_{\gamma,N_\gamma}$ is the unique factorization of $\gamma$ in $R\subset \big(\tn{Ch}(R),\sqcup\big)$. \end{enumerate} \end{defi} The sum $\sum_{\gamma \in \Gamma}()$ appearing in equation \hyperref[amplitude]{\ref{amplitude}} is analogous to the integral $\int_\mbf{x}()d\mbf{x}$ in Feynman's continuum construction, appearing in equation \hyperref[continuumpathintegral]{\ref{continuumpathintegral}} above. However, in the discrete causal context, this sum is exact; no limiting process is necessary. \subsection{Schr\"{o}dinger-Type Equations in Quantum Causal Theory}\label{subsectionschrodinger} \refstepcounter{textlabels}\label{schrodhistory} {\bf Schr\"{o}dinger-type Equations via the Histories Approach.} Schr\"{o}dinger's equation describes the dynamics of nonrelativistic quantum systems. It may be derived in a conceptually satisfying manner via path integral methods, as demonstrated by Feynman. Since these methods represent a special case of the general histories approach to quantum theory, Schr\"{o}dinger's equation is much more than merely a ``low-energy approximation of quantum field theory." So far as the histories approach itself is valid, any shortcomings in Schr\"{o}dinger's equation arise from {\it shortcomings in the choice of history configuration space} used in this approach. In nonrelativistic quantum theory, histories are represented in a manifestly unrealistic fashion, by paths traced out on a fixed copy of Euclidean spacetime. Supplying better underlying models in a similar conceptual context may be expected to produce superior dynamical laws of the same general type. This expectation is already borne out to some extent in more sophisticated modern implementations of path integration in quantum field theory and various approaches to quantum gravity. In this section, I examine this problem in the context of discrete causal theory. I begin by reviewing Feynman's derivation of Schr\"{o}dinger's equation, then derive discrete causal analogues. Equation \hyperref[causalschrodinger]{\ref{causalschrodinger}} below is an example of a Schr\"{o}dinger-type equation derived via this method. \refstepcounter{textlabels}\label{conschrodequ} {\bf Continuum Version of Schr\"{o}dinger's Equation via Path Integrals.} One of the major achievements of Feynman's paper \cite{FeynmanSOH48} is the recovery, via path integral methods, of the {\bf nonrelativistic Schr\"{o}dinger equation}: \begin{equation}\label{hamiltonschrod} i\hbar\frac{\partial\psi^-}{\partial t}=\mathbf{H}\psi^-, \end{equation} which describes the mathematical behavior of the {\it past wave function} $\psi^-$, associated with the motion of a particle through a distinguished region $X$ of Euclidean spacetime, as described in section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} above. This is the ``usual wave function" encountered in elementary quantum theory. The adjective ``past" indicates that it depends only on events {\it preceding} a specified choice of ``present," given by an appropriate spatial section of $X$. Here $\mathbf{H}$ is the {\bf Hamiltonian operator}, defined in the single-particle case by the formula \begin{equation}\label{classham} \mathbf{H}\psi^-=\frac{-\hbar^2}{2m}\frac{\partial^2\psi^-}{\partial x^2}+V\psi^-, \end{equation} where $m$ is the mass of the particle and $V$ is its potential energy. Schr\"{o}dinger's equation is a {\it dynamical law} for nonrelativistic quantum theory, permitting quantitative predictions of observable physical phenomena. To derive Schr\"{o}dinger's equation, Feynman first expressed his quantum amplitude $\psi(X;\ms{L})$ as the {\it complex inner product} of the past wave function $\psi^-$ and a corresponding {\it future wave function} $\psi^+$: \begin{equation}\label{innerprod} \psi(X;\ms{L})=\big\langle \psi^+(x',t'),\psi^-(x',t')\big\rangle:=\int_{\sigma'} \psi^+(x',t')^*\psi^-(x',t')dx'. \end{equation} The domain of integration of the variable $x'$ in equation \hyperref[innerprod]{\ref{innerprod}} is a {\it single} spatial section $\sigma'$ of $X$ at time $t=t'$. This spatial section partitions the spacetime region $X$ into a disjoint union $X=X^-\coprod\sigma'\coprod X^+$, where $X^{\pm}$ are the {\bf past} and {\bf future regions} with respect to the choice of ``present" represented by $\sigma'$. The past and future wave functions $\psi^{\pm}$ depend on {\it both position and time}; hence, different choices of $\sigma'$ produce different values of $\psi^{\pm}$. However, the quantum amplitude $\psi(X;\ms{L})$ is a complex number depending {\it only on the spacetime region $X$ and the Lagrangian $\ms{L}$,} so different choices of $\sigma'$ produce the same inner product in equation \hyperref[innerprod]{\ref{innerprod}}. \refstepcounter{textlabels}\label{contwavefunc} Without belaboring the details, the value $\psi^-(x',t')$ of the past wave function at a point $(x',t')$ in $X$ is determined by ``integrating over all maximal paths" in the past region $X^-$ of $X$ {\it terminating} at $(x',t')$. Similarly, the value of the future wave function at $(x',t')$ is determined by ``integrating over all maximal paths" in the future region $X^+$ of $X$ {\it beginning} at $(x',t')$. Like Feynman's description of how to determine his quantum amplitude $\psi(X;\ms{L})$, given in his first ``postulate" (\hyperref[f1]{F1}) above, these are {\it heuristic descriptions of a limiting process.} Every path contributing to the quantum amplitude $\psi(X;\ms{L})$ is given by ``joining paths in $X^-$ and $X^+$ at a common point $(x',t')$ in the spatial section $\sigma'$," since $\sigma'$ is a Cauchy surface. Hence, combining the wave functions $\psi^{\pm}$, via the inner product formula in equation \hyperref[innerprod]{\ref{innerprod}}, suffices to recover $\psi(X;\ms{L})$, with the $x'$-independence ``integrated out." The {\bf past wave function} $\psi^-$ may be {\it defined} as a limit: \begin{equation}\label{psiminuslim} \psi^-(x',t'):=\lim_{|\Delta^-|\to0}\int_{\mbf{x}^-}\psi^-(\Delta^-;\mathbf{x}^-;\ms{L})d\mbf{x}^-, \end{equation} where $\{\Delta^-\}$ is a sequence of partitions of the time interval corresponding to the past region $X^-$ of $X$, bounded above by the spatial section $\sigma'$. The integration is performed over all sequences $\mbf{x}^-$ of values of spatial coordinates corresponding to such partitions, subject to the additional condition that these sequences must terminate at the specific spatial value $x'$. The integrand $\psi^-(\Delta^-;\mathbf{x}^-;\ms{L})$ is given, up to a constant, by products of phases of ``path increments" determined by $\Delta^-$ and $\mbf{x}^-$. \refstepcounter{textlabels}\label{contschrodequ} Schr\"{o}dinger's equation \hyperref[hamiltonschrod]{\ref{hamiltonschrod}} is derived by comparing the values of the past wave function $\psi^-$ for ``nearby" time values $t'$ and $t''$, where $t'<t''$, then taking the limit as the difference $t''-t'$ approaches zero. For simplicity, I replace the limit notation with the ``approximation symbol" $\approx$ below, since no limiting process is necessary in the discrete causal case. Let $t'$ and $t''$ be ``nearby" time values, and let $\sigma'$ and $\sigma''$ be the corresponding spatial sections. Let $\Delta^-$ be a partition of the time interval up to and including $t'$. This setup is illustrated in figure \hyperref[feynman]{\ref{feynman}}b of section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} above. The approximation \[\psi^-(x',t')\approx\int_{\mbf{x}^-}\psi^-(\Delta^-;\mathbf{x}^-;\ms{L})d\mbf{x}^-,\] computed at time $t=t'$, given by dropping the limit in equation \hyperref[psiminuslim]{\ref{psiminuslim}} above, leads to the ``subsequent" approximation \[\psi^-(x'',t'')\approx\int_{\sigma'}\int_{\mbf{x}^-}\psi^-(\Delta^-;\mathbf{x}^-;\ms{L})d\mbf{x}^-\tn{exp}\Big(\frac{i}{\hbar}\ms{S}(\delta\gamma)\Big)dx',\] computed at time $t=t''$, where $\delta\gamma$ ranges over all ``path increments" beginning somewhere on the spatial section $\sigma'$, and terminating at the specific point $(x'',t'')$ of the spatial section $\sigma''$. This expression, in turn, leads to the ``approximate recursion" \begin{equation}\label{schrodinger} \psi^-(x'',t'')\approx\int_{\sigma'}\psi^-(x',t')\tn{exp}\Big(\frac{i}{\hbar}\ms{S}(\delta\gamma)\Big)dx', \end{equation} which yields Schr\"{o}dinger's equation \hyperref[hamiltonschrod]{\ref{hamiltonschrod}} in the limit. {\bf Discrete Causal Analogues of Schr\"{o}dinger's Equation.} Returning to the discrete causal context, let $M'$, $R$, $\sigma$, $R^\pm$, $T$, and $\Theta$ be as designated in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} above, and let $r$ be an element of $\sigma$. Let $\gamma$ be a maximal chain in $R^-$ admitting {\it extension by} $r$; i.e., a maximal chain in $R^-$ such that $\gamma'=\gamma\sqcup r$ is a maximal chain in $R^-\coprod\sigma$. In what follows, $r$, $\gamma$, and $\gamma'$ are allowed to vary as needed. For example, $r$ is allowed to vary over $\sigma$, and later, over all of $R$. This setup is illustrated in figure \hyperref[figcausalschrodinger]{\ref{figcausalschrodinger}}a below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{16cm}{5.6cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,-.1)(0,5.5) \pgfxyline(10.5,-.1)(10.5,5.5) \pgfxyline(16.5,-.1)(16.5,5.5) \pgfxyline(0,5.5)(16.5,5.5) \pgfputat{\pgfxy(.15,5.2)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(10.65,5.2)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{-.25cm}{-.2cm}} \begin{pgfmagnify}{1}{1} \comment{ \begin{colormixin}{50!white} \pgfxyline(0,0)(12,0) \pgfxyline(0,.5)(12,.5) \pgfxyline(0,1)(12,1) \pgfxyline(0,1.5)(12,1.5) \pgfxyline(0,2)(12,2) \pgfxyline(0,2.5)(12,2.5) \pgfxyline(0,3)(12,3) \pgfxyline(0,3.5)(12,3.5) \pgfxyline(0,4)(12,4) \pgfxyline(0,4.5)(12,4.5) \pgfxyline(0,5)(12,5) \pgfxyline(0,5.5)(12,5.5) \pgfxyline(0,6)(12,6) \pgfxyline(0,0)(0,6) \pgfxyline(1,0)(1,6) \pgfxyline(2,0)(2,6) \pgfxyline(3,0)(3,6) \pgfxyline(4,0)(4,6) \pgfxyline(5,0)(5,6) \pgfxyline(6,0)(6,6) \pgfxyline(7,0)(7,6) \pgfxyline(8,0)(8,6) \pgfxyline(9,0)(9,6) \pgfxyline(10,0)(10,6) \pgfxyline(11,0)(11,6) \pgfxyline(12,0)(12,6) \end{colormixin} \pgfputat{\pgfxy(0,.25)}{\pgfbox[center,center]{0}} \pgfputat{\pgfxy(1,.25)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(2,.25)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(3,.25)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(4,.25)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(5,.25)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(6,.25)}{\pgfbox[center,center]{6}} \pgfputat{\pgfxy(7,.25)}{\pgfbox[center,center]{7}} \pgfputat{\pgfxy(8,.25)}{\pgfbox[center,center]{8}} \pgfputat{\pgfxy(9,.25)}{\pgfbox[center,center]{9}} \pgfputat{\pgfxy(10,.25)}{\pgfbox[center,center]{10}} \pgfputat{\pgfxy(11,.25)}{\pgfbox[center,center]{11}} \pgfputat{\pgfxy(12,.25)}{\pgfbox[center,center]{12}} \pgfputat{\pgfxy(0,1)}{\pgfbox[center,center]{1}} \pgfputat{\pgfxy(0,2)}{\pgfbox[center,center]{2}} \pgfputat{\pgfxy(0,3)}{\pgfbox[center,center]{3}} \pgfputat{\pgfxy(0,4)}{\pgfbox[center,center]{4}} \pgfputat{\pgfxy(0,5)}{\pgfbox[center,center]{5}} \pgfputat{\pgfxy(0,6)}{\pgfbox[center,center]{6}} } \begin{colormixin}{25!white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0104}[fill]{\pgfxy(2.5,.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0105}[fill]{\pgfxy(1.6,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0110}[fill]{\pgfxy(3.7,.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0203}[fill]{\pgfxy(.8,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0206}[fill]{\pgfxy(1.3,3.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0207}[fill]{\pgfxy(2.1,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0405}[fill]{\pgfxy(2.3,1.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0603}[fill]{\pgfxy(1.7,4.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0703}[fill]{\pgfxy(2.4,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0803}[fill]{\pgfxy(2.9,5.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0904}[fill]{\pgfxy(3.1,.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2026}[fill]{\pgfxy(7.6,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2122}[fill]{\pgfxy(7.4,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2127}[fill]{\pgfxy(7.8,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2420}[fill]{\pgfxy(7.4,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2533}[fill]{\pgfxy(9.1,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2730}[fill]{\pgfxy(8.5,4.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2833}[fill]{\pgfxy(9.5,1.9)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2927}[fill]{\pgfxy(8.5,3.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2930}[fill]{\pgfxy(9,4.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2935}[fill]{\pgfxy(9.4,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3125}[fill]{\pgfxy(8.4,1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3128}[fill]{\pgfxy(9.4,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3132}[fill]{\pgfxy(10.2,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3226}[fill]{\pgfxy(8.3,2.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3233}[fill]{\pgfxy(10.2,2.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3435}[fill]{\pgfxy(10.2,4)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3530}[fill]{\pgfxy(9.6,4.8)}{\pgfxy(0.17,0.17)} \pgfnodeconnline{Node0102}{Node0203} \pgfnodeconnline{Node0102}{Node0206} \pgfnodeconnline{Node0102}{Node0207} \pgfnodeconnline{Node0104}{Node0405} \pgfnodeconnline{Node0104}{Node0414} \pgfnodeconnline{Node0105}{Node0502} \pgfnodeconnline{Node0105}{Node0506} \pgfnodeconnline{Node0105}{Node0507} \pgfnodeconnline{Node0105}{Node0508} \pgfnodeconnline{Node0105}{Node0514} \pgfnodeconnline{Node0110}{Node1007} \pgfnodeconnline{Node0110}{Node1013} \pgfnodeconnline{Node0206}{Node0603} \pgfnodeconnline{Node0207}{Node0703} \pgfnodeconnline{Node0207}{Node0708} \pgfnodeconnline{Node0405}{Node0502} \pgfnodeconnline{Node0405}{Node0506} \pgfnodeconnline{Node0405}{Node0507} \pgfnodeconnline{Node0405}{Node0508} \pgfnodeconnline{Node0405}{Node0514} \pgfnodeconnline{Node0414}{Node1412} \pgfnodeconnline{Node0414}{Node1415} \pgfnodeconnline{Node0502}{Node0203} \pgfnodeconnline{Node0502}{Node0206} \pgfnodeconnline{Node0502}{Node0207} \pgfnodeconnline{Node0506}{Node0603} \pgfnodeconnline{Node0507}{Node0703} \pgfnodeconnline{Node0507}{Node0708} \pgfnodeconnline{Node0508}{Node0803} \pgfnodeconnline{Node0514}{Node1412} \pgfnodeconnline{Node0514}{Node1415} \pgfnodeconnline{Node0708}{Node0803} \pgfnodeconnline{Node0904}{Node0405} \pgfnodeconnline{Node0904}{Node0414} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0703} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1208}{Node0803} \pgfnodeconnline{Node1307}{Node0703} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1620}{Node2026} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2026}{Node2621} \pgfnodeconnline{Node2026}{Node2629} \pgfnodeconnline{Node2122}{Node2219} \pgfnodeconnline{Node2122}{Node2223} \pgfnodeconnline{Node2127}{Node2730} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2420}{Node2013} \pgfnodeconnline{Node2420}{Node2017} \pgfnodeconnline{Node2420}{Node2018} \pgfnodeconnline{Node2420}{Node2022} \pgfnodeconnline{Node2420}{Node2026} \pgfnodeconnline{Node2517}{Node1714} \pgfnodeconnline{Node2521}{Node2122} \pgfnodeconnline{Node2521}{Node2127} \pgfnodeconnline{Node2529}{Node2927} \pgfnodeconnline{Node2529}{Node2930} \pgfnodeconnline{Node2529}{Node2935} \pgfnodeconnline{Node2533}{Node3329} \pgfnodeconnline{Node2533}{Node3334} \pgfnodeconnline{Node2621}{Node2122} \pgfnodeconnline{Node2621}{Node2127} \pgfnodeconnline{Node2629}{Node2927} \pgfnodeconnline{Node2629}{Node2930} \pgfnodeconnline{Node2629}{Node2935} \pgfnodeconnline{Node2833}{Node3329} \pgfnodeconnline{Node2833}{Node3334} \pgfnodeconnline{Node2927}{Node2730} \pgfnodeconnline{Node2935}{Node3530} \pgfnodeconnline{Node3125}{Node2517} \pgfnodeconnline{Node3125}{Node2521} \pgfnodeconnline{Node3125}{Node2529} \pgfnodeconnline{Node3125}{Node2533} \pgfnodeconnline{Node3128}{Node2833} \pgfnodeconnline{Node3132}{Node3226} \pgfnodeconnline{Node3132}{Node3229} \pgfnodeconnline{Node3132}{Node3233} \pgfnodeconnline{Node3226}{Node2621} \pgfnodeconnline{Node3226}{Node2629} \pgfnodeconnline{Node3229}{Node2927} \pgfnodeconnline{Node3229}{Node2930} \pgfnodeconnline{Node3229}{Node2935} \pgfnodeconnline{Node3233}{Node3329} \pgfnodeconnline{Node3233}{Node3334} \pgfnodeconnline{Node3329}{Node2927} \pgfnodeconnline{Node3329}{Node2930} \pgfnodeconnline{Node3329}{Node2935} \pgfnodeconnline{Node3334}{Node3435} \pgfnodeconnline{Node3435}{Node3530} \comment{ \color{white} \pgfnoderect{Node0102}[fill]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[fill]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[fill]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[fill]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[fill]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[fill]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[fill]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[fill]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[fill]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[fill]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[fill]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[fill]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[fill]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[fill]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} \color{black} \pgfnoderect{Node0102}[stroke]{\pgfxy(.9,1.2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0414}[stroke]{\pgfxy(3.2,2)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0502}[stroke]{\pgfxy(1.4,2.6)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0506}[stroke]{\pgfxy(1.8,3.3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0507}[stroke]{\pgfxy(2.3,2.7)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0508}[stroke]{\pgfxy(2.8,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node0514}[stroke]{\pgfxy(3,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2529}[stroke]{\pgfxy(8.7,2.5)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2621}[stroke]{\pgfxy(7.9,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2629}[stroke]{\pgfxy(8.4,3)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3229}[stroke]{\pgfxy(9.2,2.8)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3329}[stroke]{\pgfxy(9.7,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node3334}[stroke]{\pgfxy(10.4,3.1)}{\pgfxy(0.17,0.17)} \pgfnoderect{Node2521}[stroke]{\pgfxy(7.5,2.7)}{\pgfxy(0.17,0.17)} } \end{colormixin} \begin{colormixin}{40!white} \pgfnoderect{Node0708}[fill]{\pgfxy(3.2,4.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0917}[fill]{\pgfxy(5.4,1.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1208}[fill]{\pgfxy(3.3,4.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1215}[fill]{\pgfxy(4.3,5.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1412}[fill]{\pgfxy(3.9,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1415}[fill]{\pgfxy(4.6,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1617}[fill]{\pgfxy(6,1.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1930}[fill]{\pgfxy(6,5)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2017}[fill]{\pgfxy(6.2,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2219}[fill]{\pgfxy(6.5,4.5)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2223}[fill]{\pgfxy(7.1,4.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2517}[fill]{\pgfxy(7,2.1)}{\pgfxy(0.22,0.22)} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node0910}{Node1007} \pgfnodeconnline{Node0913}{Node1307} \pgfnodeconnline{Node0913}{Node1312} \pgfnodeconnline{Node0913}{Node1314} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node0913}{Node1319} \pgfnodeconnline{Node0917}{Node1714} \pgfnodeconnline{Node1007}{Node0708} \pgfnodeconnline{Node1013}{Node1307} \pgfnodeconnline{Node1013}{Node1312} \pgfnodeconnline{Node1013}{Node1314} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node1013}{Node1319} \pgfnodeconnline{Node1113}{Node1307} \pgfnodeconnline{Node1113}{Node1312} \pgfnodeconnline{Node1113}{Node1314} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1113}{Node1319} \pgfnodeconnline{Node1307}{Node0708} \pgfnodeconnline{Node1312}{Node1208} \pgfnodeconnline{Node1312}{Node1215} \pgfnodeconnline{Node1314}{Node1412} \pgfnodeconnline{Node1314}{Node1415} \pgfnodeconnline{Node1318}{Node1814} \pgfnodeconnline{Node1318}{Node1819} \pgfnodeconnline{Node1318}{Node1822} \pgfnodeconnline{Node1318}{Node1823} \pgfnodeconnline{Node1319}{Node1930} \pgfnodeconnline{Node1412}{Node1208} \pgfnodeconnline{Node1412}{Node1215} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node1610}{Node1007} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1617}{Node1714} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node1620}{Node2017} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1620}{Node2022} \pgfnodeconnline{Node1714}{Node1412} \pgfnodeconnline{Node1714}{Node1415} \pgfnodeconnline{Node1814}{Node1412} \pgfnodeconnline{Node1814}{Node1415} \pgfnodeconnline{Node1819}{Node1930} \pgfnodeconnline{Node1822}{Node2219} \pgfnodeconnline{Node1822}{Node2223} \pgfnodeconnline{Node2013}{Node1307} \pgfnodeconnline{Node2013}{Node1312} \pgfnodeconnline{Node2013}{Node1314} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node2013}{Node1319} \pgfnodeconnline{Node2017}{Node1714} \pgfnodeconnline{Node2018}{Node1814} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node2018}{Node1822} \pgfnodeconnline{Node2018}{Node1823} \pgfnodeconnline{Node2022}{Node2219} \pgfnodeconnline{Node2022}{Node2223} \pgfnodeconnline{Node2219}{Node1930} \pgfnodeconnline{Node2517}{Node1714} \end{colormixin} \color{white} \pgfnoderect{Node1007}[fill]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[fill]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[fill]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[fill]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[fill]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[fill]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[fill]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[fill]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[fill]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[fill]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \color{black} \pgfnoderect{Node1007}[stroke]{\pgfxy(3.6,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1307}[stroke]{\pgfxy(3.9,2.8)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1312}[stroke]{\pgfxy(4.1,3.3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1314}[stroke]{\pgfxy(4.5,3.4)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1319}[stroke]{\pgfxy(5.4,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1714}[stroke]{\pgfxy(6,2.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1814}[stroke]{\pgfxy(4.8,3.6)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1819}[stroke]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1822}[stroke]{\pgfxy(6.5,3.7)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1823}[stroke]{\pgfxy(6.2,3.9)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2022}[stroke]{\pgfxy(7,3)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.2,0.2)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.2,0.2)} \pgfsetlinewidth{1pt} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1113}{Node1318} \pgfsetdash{{1.5pt}{1.5pt}}{0pt} \pgfnodeconnline{Node1318}{Node1819} \end{pgfmagnify} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{8.25cm}{-.2cm}} \begin{pgfmagnify}{1}{1} \pgfnoderect{Node0910}[fill]{\pgfxy(4.2,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1610}[fill]{\pgfxy(4.6,1.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node0913}[fill]{\pgfxy(5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1013}[fill]{\pgfxy(4.1,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1113}[fill]{\pgfxy(5,2.1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1318}[fill]{\pgfxy(5.45,2.75)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1611}[fill]{\pgfxy(5.5,1)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node1620}[fill]{\pgfxy(6.6,1.2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2013}[fill]{\pgfxy(5.8,2)}{\pgfxy(0.22,0.22)} \pgfnoderect{Node2018}[fill]{\pgfxy(6.4,2.3)}{\pgfxy(0.22,0.22)} \color{white} \pgfnoderect{Node1819}[fill]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \color{black} \pgfnoderect{Node1819}[stroke]{\pgfxy(5.8,4.2)}{\pgfxy(0.22,0.22)} \pgfnodeconnline{Node0910}{Node1013} \pgfnodeconnline{Node1610}{Node1013} \pgfnodeconnline{Node0913}{Node1318} \pgfnodeconnline{Node1013}{Node1318} \pgfnodeconnline{Node2013}{Node1318} \pgfnodeconnline{Node1620}{Node2013} \pgfnodeconnline{Node2018}{Node1819} \pgfnodeconnline{Node1620}{Node2018} \pgfnodeconnline{Node1611}{Node1113} \pgfnodeconnline{Node1113}{Node1318} \pgfnodeconnline{Node1318}{Node1819} \end{pgfmagnify} \end{pgftranslate} \pgfputat{\pgfxy(14.2,4.15)}{\pgfbox[left,center]{\large{$r$}}} \pgfputat{\pgfxy(5.7,4.15)}{\pgfbox[left,center]{\large{$r$}}} \pgfputat{\pgfxy(8.25,3)}{\pgfbox[left,center]{\large{$\ms{R}(M')$}}} \pgfputat{\pgfxy(5.75,1.6)}{\pgfbox[left,center]{\large{$R^-$}}} \pgfputat{\pgfxy(3.75,4)}{\pgfbox[left,center]{\large{$R^+$}}} \pgfputat{\pgfxy(4.55,1.5)}{\pgfbox[left,center]{\large{$\gamma$}}} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Setup for deriving Schr\"{o}dinger-type equation in the discrete causal context: $R$ represented by large nodes; $\sigma$ represented by large open nodes; b) multiplication on the right by $r$ selects chains admitting extension by $r$.} \label{figcausalschrodinger} \end{figure} \vspace*{-.5cm} Now consider the restriction $\Theta|_\sigma$ of the phase map $\Theta:\tn{Ch}(R)\rightarrow T$ to the maximal antichain $\sigma$, viewed as an element of the chain concatenation algebra $T^\sqcup[R]$ over $R$: \[\Theta|_\sigma=\sum_{r\in\sigma}\theta(r)r,\] where I have replaced $\Theta$ in the sum by its generating relation function $\theta$, since the two coincide on individual relations; i.e., individual elements of $R$. Consider also the restriction $\Theta|_{\Gamma^-}$ of $\Theta$ to the set $\Gamma^-:=\tn{Ch}_{\tn{\fsz max}}(R^-)$ of maximal chains in the past region $R^-$: \[\Theta|_{\Gamma^-}=\sum_{\gamma\in \Gamma^-}\Theta(\gamma)\gamma,\] where I have chosen not to factor $\Theta(\gamma)$ in terms of $\theta$ in the summand. I claim that the product in $T^\sqcup[R]$ of these two elements, suitably ordered, is the element representing the restriction of $\Theta$ to the set $\Gamma_\sigma^-:=\tn{Ch}_{\tn{\fsz max}}(R^-\coprod\sigma)$ of maximal chains in $R^-\coprod\sigma$: \begin{equation}\label{restrictions} \Theta|_{\Gamma^-}\sqcup\Theta|_\sigma=\sum_{\gamma'\in \Gamma_\sigma^-}\Theta(\gamma')\gamma'. \end{equation} To see that this is true, observe that for any chain $\gamma$ in $\Gamma^-$, and for any relation $r$ belonging to $\sigma$, the concatenation product $\gamma'=\gamma\sqcup r$ is nonzero in $T^\sqcup[R]$ if and only if $\gamma$ admits extension by $r$. An equivalent condition is that the terminal relation $r^-$ of $\gamma$ directly precedes $r$ in $R$. Hence, the summand $\theta(r)r$ in $\Theta|_\sigma$, multiplying the element $\Theta|_{\Gamma^-}$ on the right, annihilates all chains $\gamma$ in $\Gamma^-$ {\it except those admitting extension by $r$,} as illustrated in figure \hyperref[figcausalschrodinger]{\ref{figcausalschrodinger}}b above. The surviving terms in the product $\Theta|_{\Gamma^-}\sqcup\theta(r)r$ are therefore chains $\gamma'$ in $\Gamma_\sigma^-$ terminating at $r$, weighted by their appropriate phases, since $\Theta$ is generated by $\theta$.\footnotemark\footnotetext{The qualifier ``weighted by their appropriate phases" makes this statement true even when $\theta(r)=0_T$.} Letting $r$ vary over $\sigma$, it remains to show that {\it every} such chain $\gamma'$ is included in the product $\Theta|_{\Gamma^-}\sqcup\Theta|_\sigma$, appearing on the left-hand side of equation \hyperref[restrictions]{\ref{restrictions}}. This follows from the impermeability of the maximal antichain $\sigma$, guaranteed by theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} of section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above, which implies that every maximal chain $\gamma$ in $R^-$ terminates in a relation $r^-$ directly preceding some relation $r$ belonging to $\sigma$. \refstepcounter{textlabels}\label{abstractpathfunct} \refstepcounter{textlabels}\label{abstractwavefunct} Now let $r$ vary over all of $R$. Denote by $\Gamma_r^-:=\tn{Ch}_{\tn{\fsz max}}^-(R;r)$ the set of all maximal chains in $R$ terminating at $r$, and by $\Gamma_r^+:=\tn{Ch}_{\tn{\fsz max}}^+(R;r)$ the set of all maximal chains in $R$ beginning at $r$. Proceeding as in the construction of the maximal chain functional $\Psi(R;\theta)$ and the abstract quantum amplitude $\psi(R;\theta)$ in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} above, I define {\it past and future chain functionals} $\Psi_{R;\theta}^\pm(r)$ on the chain sets $\Gamma_r^\pm$, viewed as elements of the chain concatenation algebra $T^\sqcup[R]$, which then serve as algebraic precursors to {\it past and future wave functions} $\psi_{R;\theta}^\pm(r)$, obtained from the functionals $\Psi_{R;\theta}^\pm(r)$ by application of the global evaluation map $e_R$.\\ \begin{defi}\label{pastfuturepathfunctionals} Let $M'=(M',R',i',t')$ be a locally finite acyclic multidirected set, with relation space $\ms{R}(M')$, and let $R$ be a finite subset of $\ms{R}(M')$, viewed as a full subobject. Let $r$ be an element of $R$, and let $\Gamma_r^-=\tn{Ch}_{\tn{\fsz max}}^-(R;r)$ and $\Gamma_r^+=\tn{Ch}_{\tn{\fsz max}}^+(R;r)$ be the sets of maximal chains in $R$ terminating at $r$ and beginning at $r$, respectively, viewed as subsets of the chain concatenation semicategory $\big(\tn{Ch}(R),\sqcup\big)$. Let $T$ be a commutative ring with unit, and let $\Theta:\tn{Ch}(R)\rightarrow T$ be a multiplicative phase map generated by a relation function $\theta:R\rightarrow T$. \begin{enumerate} \item The elements \[\Psi_{R;\theta}^-(r):=\sum_{\gamma\in \Gamma_r^-}\Theta(\gamma)\gamma\hspace*{.5cm}\tn{and}\hspace*{.5cm}\Psi_{R;\theta}^+(r):=\sum_{\gamma\in \Gamma_r^+}\Theta(\gamma)\gamma,\] of the chain concatenation algebra $T^\sqcup[R]$, are called the {\bf past} and {\bf future chain functionals} of $r$ in $R$ with respect to $\theta$, respectively. \item The relation functions \[\psi_{R;\theta}^-(r):=\sum_{\gamma\in \Gamma_r^-}\Theta(\gamma)\hspace*{.5cm}\tn{and}\hspace*{.5cm}\psi_{R;\theta}^+(r):=\sum_{\gamma\in \Gamma_r^+}\Theta(\gamma),\] given by applying the global evaluation map $e_R$ to the past and future chain functionals $\Psi_{R;\theta}^-(r)$ and $\Psi_{R;\theta}^+(r)$, respectively, are called the {\bf past} and {\bf future wave functions} over $R$ with respect to $\theta$. \end{enumerate} \end{defi} Given two different elements $r$ and $\overline{r}$ of $R$, $\Psi_{R;\theta}^-(r)$ and $\Psi_{R;\theta}^-(\overline{r})$ are {\it two different functionals,} mapping two different chain sets $\Gamma_r^-$ and $\Gamma_{\overline{r}}^-$ to the common target object $T$, respectively, while $\psi_{R;\theta}^-(r)$ and $\psi_{R;\theta}^-(\overline{r})$ are merely {\it two different values of the same past wave function} $\psi_{R;\theta}^-$, mapping $R$ to $T$. Similar statements apply to the future chain functionals $\Psi_{R;\theta}^+(r)$ and $\Psi_{R;\theta}^+(\overline{r})$, and the corresponding values $\psi_{R;\theta}^+(r)$ and $\psi_{R;\theta}^+(\overline{r})$ of the future wave function $\psi_{R;\theta}^+$. Heuristically, passage from past and future chain functionals to values of past and future wave functions is a ``diagonal process," similar to taking a {\it category-theoretic limit,} reducing a ``function-like entity" at one level of algebraic hierarchy to a ``value-like entity" at the next lower level of hierarchy. \refstepcounter{textlabels}\label{causalschrodequ} Now consider in more detail the past chain functional $\Psi_{R;\theta}^-(r)$ of $r$ in $R$ with respect to $\theta$. If $r$ is a minimal element of $R$, then $\Psi_{R;\theta}^-(r)$ consists of the single summand $\theta(r)r$. Otherwise, $\Psi_{R;\theta}^-(r)$ may be {\it decomposed} in terms of the set $D_0^-(R;r)$ of direct predecessors of $r$ in $R$; i.e., those elements $r^-$ in $R$ such that $r^-\prec r$.\footnotemark\footnotetext{The reason for using the notation $D_0^-(R;r)$, rather than merely $D_0^-(r)$, is to make it clear that only direct predecessors of $r$ belonging to $R$ are considered here. Other direct predecessors of $r$ in the ``ambient relation space" $\ms{R}(M')$ are excluded.} Every maximal chain $\gamma'$ in $R$ terminating at $r$ may be expressed as a concatenation product $\gamma\sqcup r$ for an appropriate maximal chain $\gamma$ terminating at some $r^-\in D_0^-(R;r)$. Fixing $r^-$, and summing in $T^\sqcup[R]$ over all phase-weighted chains $\gamma$ of this form, yields the past chain functional $\Psi_{R;\theta}^-(r^-)$ of $r^-$ in $R$ with respect to $\theta$. Now letting $r^-$ vary over {\it all} direct predecessors of $r$ in $R$, the past chain functional $\Psi_{R;\theta}^-(r)$ of $r$ in $R$ with respect to $\theta$ may be written as a product: \begin{equation}\label{causschrodprecursor} \Psi_{R;\theta}^-(r)=\Big(\sum_{r^-\prec r}\Psi_{R;\theta}^-(r^-)\Big)\theta(r)r. \end{equation} Equation \hyperref[causschrodprecursor]{\ref{causschrodprecursor}} serves as an algebraic precursor to the causal Schr\"{o}dinger-type equation introduced in definition \hyperref[causschrod]{\ref{causschrod}} below, just as the functionals $\Psi(R;\theta)$ and $\Psi_{R;\theta}^{\pm}$ are algebraic precursors to the generalized quantum amplitude $\psi(R;\theta)$ and the past and future wave functions $\psi_{R;\theta}^{\pm}$, respectively.\\ \begin{defi}\label{causschrod} Under the hypotheses of definition \hyperref[pastfuturepathfunctionals]{\ref{pastfuturepathfunctionals}}, the equation \begin{equation}\label{causalschrodinger} \psi_{R;\theta}^-(r)=\theta(r)\sum_{r^-\prec r}\psi_{R;\theta}^-(r^-), \end{equation} given by applying the evaluation map $e_R$ to equation \hyperref[causschrodprecursor]{\ref{causschrodprecursor}}, is called the {\bf causal Schr\"{o}dinger equation} over $R$ with respect to $\theta$. \end{defi} The method used to derive equation \hyperref[causschrodprecursor]{\ref{causschrodprecursor}}, in which the past chain functional of a relation is decomposed in terms of its direct predecessors, may be iterated, in a process I refer to as {\bf past fractal decomposition.} The adjective {\it fractal} refers to the similarity of the nested chain sets $\Gamma_r^-$ across the various levels of the decomposition. Past fractal decomposition leads to {\it higher-order analogues} of Schr\"{o}dinger's equation. \refstepcounter{textlabels}\label{causalfeynamp} To conclude this section, I briefly discuss two analogues of Feynman's complex inner product formula, given in equation \hyperref[innerprod]{\ref{innerprod}} above, which expresses his quantum amplitude $\psi(X;\ms{L})$ in terms of the past and future wave functions $\psi^\pm$. To construct the first analogue, begin by expressing the maximal chain functional $\Psi(R;\theta)$ as a path algebra element in the usual way; i.e., as the sum of all chains in $\Gamma=\tn{Ch}_{\tn{\fsz max}}(R)$, weighted by their appropriate phases. Let $\sigma$ be a maximal antichain in $R$. By the impermeability of $\sigma$, every term in the sum for $\Psi(R;\theta)$ is a concatenation product of the form $\Theta(\gamma^-)\gamma^-\sqcup\theta(r)r\sqcup\Theta(\gamma^+)\gamma^+$ for appropriate chains $\gamma^\pm$ in $\Gamma^\pm=\tn{Ch}_{\tn{\fsz max}}^\pm(R)$, and an appropriate relation $r$ in $\sigma$. The sum of all such terms is therefore \begin{equation} \Psi(R;\theta)=\Psi(R^-;\theta)\sqcup\Theta|_\sigma\sqcup\Psi(R^+;\theta). \end{equation} To construct the second analogue, observe that for any relation $r$ in $R$, the directed product $\Psi_{R;\theta}^-(r)\vee\Psi_{R;\theta}^+(r)$ {\it almost} coincides with the sum of phase-weighted maximal chains in the total domain of influence $D(r)=D^-(r)\cup\{r\}\cup D^+(r)$ of $r$ in $R$. The only difference between the product and the sum is that the phase of $r$ is counted twice in the product. Given a maximal antichain $\sigma$ in $R$, every maximal path in $R$ passes through a unique relation $r$ in $\sigma$. Therefore, $\Psi(R;\theta)$ is given by the sum: \begin{equation}\label{directedinnerprod} \Psi(R;\theta)=\sum_{r\in\sigma}\frac{1}{\theta(r)}\Psi_{R;\theta}^-(r)\vee\Psi_{R;\theta}^+(r), \end{equation} where the factors of $\theta(r)$ in the denominators correct for the corresponding extra factors in the products. Equation \hyperref[directedinnerprod]{\ref{directedinnerprod}} easily reduces to an expression for the corresponding amplitude $\psi(R;\theta)$. To obtain this expression, note that since every path contributing to $\Psi_{R;\theta}^-(r)$ precedes every path contributing to $\Psi_{R;\theta}^+(r)$ in the directed product relation, the global evaluation map ``commutes with the directed product" in this case. Hence, \begin{equation} \psi(R;\theta)=\sum_{r\in\sigma}\frac{1}{\theta(r)}\psi_{R;\theta}^-(r)\psi_{R;\theta}^+(r), \end{equation} which is directly analogous to equation \hyperref[innerprod]{\ref{innerprod}}. \subsection{Kinematic Schemes}\label{subsectionkinematicschemes} \refstepcounter{textlabels}\label{pathsumbackground} {\bf Path Summation in the Background-Independent Context.} The principle of path summation (\hyperref[ps]{PS}), appearing at the beginning of section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} above, states that the abstract quantum causal theory outlined in sections \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} and \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} may be used to construct a {\it background independent} version of the histories approach to quantum theory in the discrete causal context. In this section, I describe the rudiments of how this construction may be carried out. The proper objects over which to perform path summation in this context are {\it relation spaces over ``higher-level multidirected sets," whose ``elements" are directed sets.} This approach combines the theory of relation space, developed in section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, with the principle of iteration of structure (\hyperref[is]{IS}), introduced in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above. The ``higher-level multidirected sets" involved in this approach are called {\it kinematic schemes}. Their ``relation spaces" are {\it spaces of co-relative histories,} viewed as relationships between pairs of classical universes. The theory of kinematic schemes formalizes the {\it evolutionary viewpoint} in discrete causal theory, generalizing Sorkin and Rideout's theory of sequential growth dynamics for causal sets \cite{SorkinSequentialGrowthDynamics99}. Kinematic schemes play a structural role similar to that of categories, and their ``relation spaces" play a role similar to that of morphism categories. The abstract structure of a kinematic scheme is generally more complicated than the structures of its individual members. For example, kinematic schemes of finite acyclic directed sets generally have multidirected, rather than merely directed, structures, as illustrated by McKay's example appearing in figure \hyperref[corelativeexamples]{\ref{corelativeexamples}} of section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above. Kinematic schemes of countable acyclic directed sets generally have ``higher-level cycles." \refstepcounter{textlabels}\label{kinversusdyn} \refstepcounter{textlabels}\label{kinpreschemes} {\bf Kinematic Preschemes; Classes of Co-Relative Histories.} In classical mechanics, {\bf kinematics} is the study of the general types of behavior {\it permissible} within a theory, while {\bf dynamics} is the study of which behaviors are {\it determined} or {\it favored} under specified conditions. Under the strong form of the causal metric hypothesis (\hyperref[cmh]{CMH}), all physical behavior is represented by the structure of directed sets. {\it Permissible behavior} is therefore defined by specifying a distinguished class $\ms{K}$ of directed sets.\footnotemark\footnotetext{As usual, only the isomorphism classes of directed sets are significant. For simplicity, I work with representatives, with the implicit understanding that at most one representative of each isomorphism class appears in $\ms{K}$.} This class may be endowed with a ``higher-level multidirected structure," by specifying a class $\mc{H}$ of distinguished co-relative histories between pairs of members of $\ms{K}$, which represent possible instances of evolution from one directed set into another. An arbitrary choice of $\mc{H}$ and $\ms{K}$ yields a structure called a {\it kinematic prescheme.} Imposing additional conditions, described below, yields a {\it kinematic scheme.} Dynamics is then supplied by specifying which co-relative histories are ``favored," in a suitable sense. This may be accomplished by identifying a distinguished phase map, or a suitable higher-level analogue.\\ \begin{defi}\label{defikinprescheme} A {\bf kinematic prescheme} is a pair $(\ms{K},\mc{H})$, where $\ms{K}$ is a distinguished class of directed sets, and $\mc{H}$ is a distinguished class of co-relative histories between pairs of members of $\ms{K}$. \end{defi} The class $\ms{K}$ is called the {\bf object class} of the kinematic prescheme $(\ms{K},\mc{H})$, and the class $\mc{H}$ is called the {\bf class of co-relative histories} of $(\ms{K},\mc{H})$. This definition may be immediately generalized to define kinematic preschemes of {\it multidirected} sets, but the physical relevance of such kinematic preschemes is unclear, at least in the context of causal theory. Following the precedent of section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, I focus attention on {\it proper, full, originary} co-relative histories, which are represented by equivalence classes of transitions, as expressed in definition \hyperref[deficorelative]{\ref{deficorelative}}. A kinematic prescheme whose class of co-relative histories $\ms{H}$ includes only proper, full, originary co-relative histories is called a {\bf proper, full, originary kinematic prescheme.} All kinematic preschemes in this section are assumed to be proper, full, and originary, unless stated otherwise. The class $\mc{H}$ of co-relative histories of a kinematic prescheme $(\ms{K},\mc{H})$ ``bundles equivalent transitions together into single morphism-like entities," for the purpose of distinguishing essential physical information from ``coordinate-like" or ``gauge-like" information. Despite this, it is often useful to {\it remember the internal structures} of co-relative histories in a kinematic prescheme; i.e., to preserve explicit knowledge of their constituent transitions, rather than simply treating them as ``abstract arrows." One motivation for maintaining this viewpoint is that co-relative histories exhibit ``unfamiliar behavior" from a category-theoretic perspective, as discussed in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above. For example, the ``composition" of two co-relative histories $h:D\Rightarrow D'$ and $h':D'\Rightarrow D''$ is generally {\it not} a co-relative history $D\Rightarrow D''$, but rather a {\it family of such co-relative histories.} The reason for this is that different representatives of $h$ and $h'$ may compose to yield inequivalent transitions $D\rightarrow D''$; i.e., equivalence of transitions is not a congruence relation under composition in the category $\ms{D}$ of directed sets. An example of this phenomenon appears in figure \hyperref[transitionnotcongruence]{\ref{transitionnotcongruence}} of section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above. Transitions, on the other hand, are morphisms in $\ms{D}$, so they compose to yield unique transitions. For such reasons, it is best to regard a kinematic prescheme $(\ms{K},\mc{H})$ as a ``category-like entity" whose ``objects" and ``morphisms" both encode nontrivial ``internal" structure. Whenever any aspect of this structure is not needed, it may be systematically discarded via a quotient operation, or a {\it decategorification-like procedure,} as explained below. \refstepcounter{textlabels}\label{underlyingdirclass} {\bf Underlying Directed Sets and Multidirected Sets of Kinematic Preschemes.} I have repeatedly referred to kinematic schemes, and more generally, kinematic preschemes, as ``higher-level multidirected sets." In fact, a variety of different ``higher-level directed or multidirected structures" may be ascribed to a kinematic prescheme $(\ms{K},\mc{H}$), each defined by taking ``relations" between members of its object class $\ms{K}$ to correspond to one of several types of ``morphism-like entities" associated with its class $\mc{H}$ of co-relative histories. The ``preferred" notion of a ``morphism-like entity" in this context is just a member of $\mc{H}$, but variations of this notion are sometimes useful. Here, I describe three such structures, in increasing order of complexity. \vspace*{.2cm} \begin{defi}\label{defiunderlying} Let $(\ms{K},\mc{H})$ be a kinematic prescheme. \begin{enumerate} \item The {\bf underlying directed set} $\ms{U}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$ is the directed set possessing one element $x(D)$ for each directed set $D$ in $\ms{K}$, and one relation $x(D_i)\prec x(D_t)$ for each pair of directed sets $D_i$ and $D_t$ in $\ms{K}$ with a co-relative history between them in $\mc{H}$. \item The {\bf underlying multidirected set} $\ms{V}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$ is the multidirected set possessing one element $x(D)$ for each directed set $D$ in $\ms{K}$, one relation $r(h)$ for each co-relative history $h:D_i\Rightarrow D_t$ in $\mc{H}$, and initial and terminal element maps $i$ and $t$ sending $r(h)$ to $x(D_i)$ and $x(D_t)$, respectively. \item The {\bf underlying transition structure} $\ms{W}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$ is the multidirected set possessing one element $x(D)$ for each directed set $D$ in $\ms{K}$, one relation $r(\tau)$ for each transition $\tau$ representing a co-relative history $h:D_i\Rightarrow D_t$ in $\mc{H}$, and initial and terminal element maps $i$ and $t$ sending $\tau$ to $x(D_i)$ and $x(D_t)$, respectively. \end{enumerate} \end{defi} Additional ``higher-level" directed or multidirected sets may be associated with a kinematic prescheme $(\ms{K},\mc{H})$, each defined by preserving or ignoring different amounts and/or types of structure. For example, relations may be chosen to represent {\it image-fixed co-relative histories,} mentioned near the end of section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above. However, the three alternatives appearing in definition \hyperref[defiunderlying]{\ref{defiunderlying}} suffice for the purposes of this paper. The underlying multidirected set $\ms{V}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$ is a quotient of its underlying transition structure $\ms{W}(\ms{K},\mc{H})$, given by identifying relations corresponding to transitions representing the same co-relative history in $\mc{H}$. The underlying directed set $\ms{U}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$, in turn, is a quotient of $\ms{V}(\ms{K},\mc{H})$, given by identifying all relations corresponding to co-relative histories in $(\ms{K},\mc{H})$ sharing the same cobase and target. All three sets possess the same abstract element set $\{x(D)|D\in\ms{K}\}$. $\ms{U}, \ms{V}$ and $\ms{W}$ may be viewed more formally in a functorial sense, but I do not elaborate on this viewpoint in this paper. The directed set $\ms{U}(\ms{K},\mc{H})$, and the multidirected sets $\ms{V}(\ms{K},\mc{H})$ and $\ms{W}(\ms{K},\mc{H})$, each {\it forget internal structure} associated with members of $\ms{K}$ and $\mc{H}$. Passage to $\ms{U}, \ms{V}$ or $\ms{W}$ may be viewed as a {\it decategorification-like procedure,} systematically reducing the ``category-like entity" $(\ms{K},\mc{H})$ to an ``object-like entity." It is important to emphasize that $\ms{U}, \ms{V}$ and $\ms{W}$ all derive their relations from the class of co-relative histories $\mc{H}$ of $(\ms{K},\mc{H})$, {\it not} from the ambient category $\ms{D}$ of directed sets. In particular, there may be pairs of directed sets $D$ and $D'$ in $\ms{K}$ related by morphisms in $\ms{D}$, or even by transitions in $\ms{D}$, whose corresponding co-relative histories are {\it not chosen} as members of $\mc{H}$, and whose underlying elements $x(D)$ and $x(D')$ are therefore not related in $\ms{U}, \ms{V}$ or $\ms{W}$. The abstract structural similarities between underlying directed or multidirected sets of kinematic preschemes, and the individual members of their object classes, is the essence of the principle of iteration of structure (\hyperref[is]{IS}) in discrete quantum causal theory. Application of the relation space functor $\ms{R}$ to such underlying directed or multidirected sets renders this similarity ``perfect," in the sense that all structures under consideration are then directed sets ``at different levels of algebraic hierarchy," since $\ms{R}$ reduces multidirected structure to directed structure. However, a more practical reason for applying $\ms{R}$ in this context is that it circumvents the problem of permeability of maximal antichains in multidirected sets, as proven in theorem \hyperref[theoremrelimpermeable]{\ref{theoremrelimpermeable}} of section \hyperref[subsectionrelation]{\ref{subsectionrelation}}. This allows for effective implementation of the abstract quantum causal theory outlined in sections \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} and \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} above. \refstepcounter{textlabels}\label{kinschemes} {\bf Kinematic Schemes.} A kinematic prescheme $(\ms{K},\mc{H})$ is called a {\bf kinematic scheme} if it satisfies the following two additional properties: \refstepcounter{textlabels}\label{hereditary} \refstepcounter{textlabels}\label{h} \hspace*{.35cm}H. \hspace*{.3cm}{\bf Hereditary Property}: {\it $\ms{K}$ is closed under the formation of proper, full, originary \\ \hspace*{1.1cm} subobjects.} \refstepcounter{textlabels}\label{weakaccessibility} \refstepcounter{textlabels}\label{wa} \hspace*{.3cm}WA. {\bf Weak Accessibility}: {\it Suppose that $D$ and $D'$ are members of $\ms{K}$. If there exists a \\ \hspace*{1.1cm} transition $\tau:D\rightarrow D'$ in $\ms{D}$ with finite complement, then there exists a chain from $x(D)$ \\ \hspace*{1.1cm} to $x(D')$ in $\ms{W}(\ms{K},\mc{H})$.} The intuition underlying the hereditary property is that if a given directed set is a permissible model of causal structure, then so are its ``ancestors;" i.e., its ``earlier stages of development." The intuition underlying weak accessibility is that any permissible directed set should be {\it realizable,} in the sense that it should have at least one ``evolutionary pathway" leading to it. The qualifier {\it weak} is included because of the finite complement condition in the definition; without this caveat, certain pairs of directed sets differing by an infinite number of elements would perforce be ``connected by a finite number of evolutionary steps" in $(\ms{K},\mc{H})$. While there are interesting kinematic schemes that {\it do} exhibit such behavior, it is nonetheless too restrictive to include as part of the definition. Weak accessibility is closely related to the notion of {\it inaccessible cardinals} in order theory. In particular, it is analogous to the fact that one may access any positive integer in a finite number of steps by taking minimal successors, but may {\it not} access Cantor's first transfinite ordinal in this manner. Note that weak accessibility does not depend on the choice to use the underlying transition structure $\ms{W}(\ms{K},\mc{H})$ of $(\ms{K},\mc{H})$ in the definition, rather than one of its quotients $\ms{U}(\ms{K},\mc{H})$ or $\ms{V}(\ms{K},\mc{H})$. The two classes $\ms{K}$ and $\mc{H}$ making up a kinematic scheme $(\ms{K},\mc{H})$ stand on very different footings. The object class $\ms{K}$ of $(\ms{K},\mc{H})$ embodies {\it a hypothesis about the structure of the physical universe,} while the class $\mc{H}$ of co-relative histories of $(\ms{K},\mc{H})$ is chosen for convenience among many equally valid alternatives. This difference is analogous to the difference in general relativity between choosing a class of pseudo-Riemannian manifolds as models of classical spacetime, and choosing particular families of coordinate systems on these manifolds. The arbitrary choice of the class of co-relative histories in a kinematic scheme requires an appropriate notion of {\it covariance,} to ensure that the physical predictions of the theory do not depend on this choice. This topic is more subtle than one might expect, essentially because the ``frames of reference" encoded by kinematic schemes tend to be much more general and varied than relativistic coordinate systems.\footnotemark\footnotetext{Sorkin and Rideout discuss ``general discrete covariance" in the context of sequential growth dynamics. This is closely related, but not identical, to the notion required here.} \refstepcounter{textlabels}\label{quantumcausalmetric} {\bf Quantum Causal Metric Hypothesis.} Just as the classical causal metric hypothesis (\hyperref[ccmh]{CCMH}) takes the properties of classical spacetime to arise from directed structure, so the {\it quantum causal metric hypothesis} takes the properties of quantum spacetime to arise from higher-level multidirected structure, via the principle of iteration of structure (\hyperref[is]{IS}). The theory of kinematic schemes enables a precise statement of this idea. \hspace*{.3cm} QCMH.\refstepcounter{textlabels}\label{qcmh} {\bf Quantum Causal Metric Hypothesis.} {\it The properties of quantum spacetime\\ \hspace*{1.8cm} arise from the structure of a kinematic scheme of directed sets.} As noted above, the class $\mc{H}$ of co-relative histories of a kinematic scheme $(\ms{K},\mc{H})$ involves a {\it choice,} similar to a choice of coordinate system. Hence, the quantum causal metric hypothesis may be realized via {\it any suitable kinematic scheme.} Uniqueness may be achieved by using a {\it universal kinematic scheme,} as described below, but this is not always convenient. \refstepcounter{textlabels}\label{pathsumkinscheme} \refstepcounter{textlabels}\label{corelkin} {\bf Path Summation over a Kinematic Scheme.} Path summation over a kinematic scheme $(\ms{K},\mc{H})$, or more precisely, over the relation space of an underlying directed or multidirected set of $(\ms{K},\mc{H})$, provides a precise realization of the quantum causal metric hypothesis (\hyperref[qcmh]{QCMH}). Although the formal aspects of this approach are automatic, due to the principle of path summation (\hyperref[ps]{PS}) stated in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} above, faithful organization and interpretation of the corresponding physical information is a subtle matter, and I can only briefly sketch the rudiments here. Let $(\ms{K},\mc{H})$ be a kinematic scheme, $D$ and $D'$ objects of $\ms{K}$, and $\tau:D\rightarrow D'$ a transition in $\ms{D}$. Suppose, for simplicity, that $\tau$ has finite complement. By weak accessibility (\hyperref[wa]{WA}), there exists at least one chain $\gamma$ in $\ms{W}(\ms{K},\mc{H})$ from $x(D)$ to $x(D')$. Such a chain represents a {\it generalized order extension,} in which ``relativity of simultaneity" is broken by arbitrary choices of succession among causally unrelated subsets of ``histories lying between $D$ and $D'$ in $(\ms{K},\mc{H})$." Hence, in addition to specifying the co-relative history $h$ represented by $\tau$, such a chain also provides a ``kinematic account," or {\it generalized frame of reference,} for $h$. I refer to such a chain as a {\bf maximal co-relative kinematics for $h$ in} $(\ms{K},\mc{H})$.\footnotemark\footnotetext{There is an interesting analogy here arising from algebraic geometry: a co-relative kinematics may be compared to a point in a {\it flag variety,} since it involves a ``distinguished sequence of subobjects in a space."} Analogous chains in the ``coarser" sets $\ms{U}(\ms{K},\mc{H})$ and $\ms{V}(\ms{K},\mc{H})$ correspond to equivalence classes of co-relative kinematics, representing less-detailed generalized order extensions. Generalizing this, any directed path in an underlying directed or multidirected set associated with the kinematic scheme $(\ms{K},\mc{H})$, with or without initial and terminal elements, may be called a {\bf kinematics} in $(\ms{K},\mc{H})$. Path summation over a kinematic scheme $(\ms{K},\mc{H})$ is therefore better characterized as summation over {\it kinematics} than as summation over {\it histories.} The underlying reason for this is simply that any particular path in $(\ms{K},\mc{H})$ involves arbitrary information, analogous to the arbitrary information treated by covariance and/or gauge theory in ordinary physics. Sorkin's quote about {\it upward directed paths in a tree of causal sets,} in the context of quantum measure theory, cited in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, foreshadows the general idea of co-relative kinematics. Path summation over a kinematic scheme $(\ms{K},\mc{H})$ requires specification of a suitable phase map $\Theta$. In order to respect the quantum causal metric hypothesis, $\Theta$ must depend only on the class $\mc{H}$ of co-relative histories of $(\ms{K},\mc{H})$. In this way, the information contained in $\Theta$ derives ultimately from the causal preorders of the individual members of $\ms{K}$. The notion of a phase map must be accorded a broad interpretation in this context. In particular, since the ``elements" in the ``domain" of $\Theta$ represent paths in a space of co-relative histories, whose individual relations represent relationships between pairs of classical universes, it may be asking too much to expect individual ``numbers;" i.e., elements of a ring, to preserve enough information as images in this context. Hence, $\Theta$ may be a ``higher-level map," such as a generalized functor, and path sums may be ``higher-level evaluations," such as generalized category-theoretic limits. \refstepcounter{textlabels}\label{possequkinscheme} \refstepcounter{textlabels}\label{relsorkinrideout} {\bf Positive Sequential Kinematic Scheme.} The prototypical example of a kinematic scheme $(\ms{K},\mc{H})$ appears in Sorkin and Rideout's theory of sequential growth dynamics of causal sets \cite{SorkinSequentialGrowthDynamics99}. In this case, $\ms{K}$ is the class $\ms{C}_{\tn{\fsz{fin}}}$ of finite causal sets, and $\mc{H}$ is the class $\mc{H}_1$ of co-relative histories $h:C_i\Rightarrow C_t$ such that $C_t$ differs from $C_i$ by the addition of a single element. In section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, I briefly discussed this kinematic scheme in a simplified manner, denoting it by the single symbol $\ms{K}$, and describing its abstract structure in terms of transitions, rather than co-relative histories. Due to the structural shortcomings of causal set theory, I use a slightly different kinematic scheme for illustrative purposes in this section, replacing $\ms{C}_{\tn{\fsz{fin}}}$ with the class $\ms{A}_{\tn{\fsz{fin}}}$ of finite acyclic directed sets, and enlarging the class of co-relative histories $\mc{H}_1$ to accommodate the larger class of objects, without changing its abstract definition or its notation. I refer to $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ as the {\bf positive sequential kinematic scheme}, since each originary chain\footnotemark\footnotetext{That is, each directed path beginning at $x(\oslash)$, the element of $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ corresponding to the empty set.} in its underlying transition structure $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ defines a total sequential labeling of the elements of its terminal acyclic directed set, or of its countably infinite {\it limit set,} in the case of an infinite chain. A portion of the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ is illustrated in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} below. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{14cm}{11.8cm} \begin{pgfmagnify}{1.03}{1.03} \begin{pgfscope} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfmoveto{\pgfxy(6.5,-.1)} \pgfcurveto{\pgfxy(6.75,1.15)}{\pgfxy(7.5,1.45)}{\pgfxy(8.25,1.45)} \pgfcurveto{\pgfxy(9,1.45)}{\pgfxy(9.75,1.15)}{\pgfxy(10,-.1)} \pgfstroke \pgfmoveto{\pgfxy(5.25,-.1)} \pgfcurveto{\pgfxy(6,2.55)}{\pgfxy(7.5,2.95)}{\pgfxy(8.25,2.95)} \pgfcurveto{\pgfxy(9,2.95)}{\pgfxy(10.5,2.55)}{\pgfxy(11.25,-.1)} \pgfstroke \pgfmoveto{\pgfxy(3.5,-.1)} \pgfcurveto{\pgfxy(4.5,4.35)}{\pgfxy(7.5,4.35)}{\pgfxy(8.25,4.35)} \pgfcurveto{\pgfxy(9,4.35)}{\pgfxy(12,4.35)}{\pgfxy(13,-.1)} \pgfstroke \pgfmoveto{\pgfxy(1.5,-.1)} \pgfcurveto{\pgfxy(2.5,6.5)}{\pgfxy(7.5,6.5)}{\pgfxy(8.25,6.5)} \pgfcurveto{\pgfxy(9,6.5)}{\pgfxy(14,6.5)}{\pgfxy(15,-.1)} \pgfstroke \end{pgfscope} \pgfputat{\pgfxy(7.25,.65)}{\pgfbox[center,center]{\huge{$0$}}} \pgfputat{\pgfxy(6.2,.9)}{\pgfbox[center,center]{\huge{$1$}}} \pgfputat{\pgfxy(4.75,1.3)}{\pgfbox[center,center]{\huge{$2$}}} \pgfputat{\pgfxy(3,1.8)}{\pgfbox[center,center]{\huge{$3$}}} \pgfputat{\pgfxy(.9,2.4)}{\pgfbox[center,center]{\huge{$4$}}} \color{white} \pgfmoveto{\pgfxy(11,-.1)} \pgflineto{\pgfxy(11,3.5)} \pgflineto{\pgfxy(16.5,3.5)} \pgflineto{\pgfxy(16.5,-.1)} \pgflineto{\pgfxy(11,-.1)} \pgffill \color{black} \pgfxyline(0,-.1)(0,11.6) \pgfxyline(16.5,-.1)(16.5,11.6) \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,11.6)(16.5,11.6) \pgfxyline(11,-.1)(11,3.5) \pgfxyline(11,3.5)(16.5,3.5) \begin{pgftranslate}{\pgfpoint{0cm}{.7cm}} \begin{colormixin}{20!white} \pgfnodecircle{Node6}[fill]{\pgfxy(6.25,4.1)}{.55cm}% \pgfnodecircle{Node12}[fill]{\pgfxy(.7,4.9)}{.55cm}% \pgfnodecircle{Node13}[fill]{\pgfxy(1.8,5.2)}{.55cm}% \pgfnodecircle{Node14}[fill]{\pgfxy(1.2,6.4)}{.55cm}% \pgfnodecircle{Node15}[fill]{\pgfxy(2.4,6.2)}{.55cm}% \pgfnodecircle{Node16}[fill]{\pgfxy(.8,8.1)}{.55cm}% \pgfnodecircle{Node17}[fill]{\pgfxy(2.2,7.8)}{.55cm}% \pgfnodecircle{Node18}[fill]{\pgfxy(1.8,9.2)}{.55cm}% \pgfnodecircle{Node19}[fill]{\pgfxy(5.25,6.2)}{.55cm}% \pgfnodecircle{Node23}[fill]{\pgfxy(8.4,10)}{.55cm}% \pgfnodecircle{Node24}[fill]{\pgfxy(8,7)}{.55cm}% \pgfnodecircle{Node25}[fill]{\pgfxy(6.1,9.5)}{.55cm}% \pgfnodecircle{Node27}[fill]{\pgfxy(4.7,8.8)}{.55cm}% \pgfnodecircle{Node31}[fill]{\pgfxy(13.4,7.9)}{.55cm}% \pgfnodecircle{Node36}[fill]{\pgfxy(12.75,6.6)}{.55cm}% \pgfnodecircle{Node38}[fill]{\pgfxy(14.1,6)}{.55cm}% \end{colormixin} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,0)}{.55cm}% \pgfnodecircle{Node2}[stroke]{\pgfxy(8.25,1.5)}{.55cm}% \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.55cm}% \pgfnodecircle{Node4}[stroke]{\pgfxy(9.25,2.75)}{.55cm}% \pgfnodecircle{Node5}[stroke]{\pgfxy(5,3.5)}{.55cm}% \pgfnodecircle{Node6}[stroke]{\pgfxy(6.25,4.1)}{.55cm}% \pgfnodecircle{Node7}[stroke]{\pgfxy(7.5,4.5)}{.55cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.55cm}% \pgfnodecircle{Node9}[stroke]{\pgfxy(10.25,4.1)}{.55cm}% \pgfnodecircle{Node10}[stroke]{\pgfxy(11.5,3.5)}{.55cm}% \pgfnodecircle{Node11}[stroke]{\pgfxy(1.5,4)}{.55cm}% \pgfnodecircle{Node12}[stroke]{\pgfxy(.7,4.9)}{.55cm}% \pgfnodecircle{Node13}[stroke]{\pgfxy(1.8,5.2)}{.55cm}% \pgfnodecircle{Node14}[stroke]{\pgfxy(1.2,6.4)}{.55cm}% \pgfnodecircle{Node15}[stroke]{\pgfxy(2.4,6.2)}{.55cm}% \pgfnodecircle{Node16}[stroke]{\pgfxy(.8,8.1)}{.55cm}% \pgfnodecircle{Node17}[stroke]{\pgfxy(2.2,7.8)}{.55cm}% \pgfnodecircle{Node18}[stroke]{\pgfxy(1.8,9.2)}{.55cm}% \pgfnodecircle{Node19}[stroke]{\pgfxy(5.25,6.2)}{.55cm}% \pgfnodecircle{Node20}[stroke]{\pgfxy(6.8,7.5)}{.55cm}% \pgfnodecircle{Node21}[stroke]{\pgfxy(4.15,7.7)}{.55cm}% \pgfnodecircle{Node22}[stroke]{\pgfxy(3.1,8.7)}{.55cm}% \pgfnodecircle{Node23}[stroke]{\pgfxy(8.4,10)}{.55cm}% \pgfnodecircle{Node24}[stroke]{\pgfxy(8,7)}{.55cm}% \pgfnodecircle{Node25}[stroke]{\pgfxy(6.1,9.5)}{.55cm}% \pgfnodecircle{Node26}[stroke]{\pgfxy(4.9,10)}{.55cm}% \pgfnodecircle{Node27}[stroke]{\pgfxy(4.7,8.8)}{.55cm}% \pgfnodecircle{Node28}[stroke]{\pgfxy(7.3,9.5)}{.55cm}% \pgfnodecircle{Node29}[stroke]{\pgfxy(9.2,8.2)}{.55cm}% \pgfnodecircle{Node30}[stroke]{\pgfxy(13.8,9.2)}{.55cm}% \pgfnodecircle{Node31}[stroke]{\pgfxy(13.4,7.9)}{.55cm}% \pgfnodecircle{Node32}[stroke]{\pgfxy(12.5,9)}{.55cm}% \pgfnodecircle{Node33}[stroke]{\pgfxy(11.6,9.8)}{.55cm}% \pgfnodecircle{Node34}[stroke]{\pgfxy(10.4,9)}{.55cm}% \pgfnodecircle{Node35}[stroke]{\pgfxy(9.7,10)}{.55cm}% \pgfnodecircle{Node36}[stroke]{\pgfxy(12.75,6.6)}{.55cm}% \pgfnodecircle{Node37}[stroke]{\pgfxy(14.6,7.1)}{.55cm}% \pgfnodecircle{Node38}[stroke]{\pgfxy(14.1,6)}{.55cm}% \pgfnodecircle{Node39}[stroke]{\pgfxy(15.4,5.8)}{.55cm}% \pgfnodecircle{Node40}[stroke]{\pgfxy(14.7,4.7)}{.55cm}% \pgfnodecircle{Node41}[stroke]{\pgfxy(15.8,4.3)}{.55cm}% \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node4}{Node9} \pgfnodeconnline{Node4}{Node10} \pgfnodeconnline{Node5}{Node11} \pgfnodeconnline{Node5}{Node12} \pgfnodeconnline{Node5}{Node13} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node5}{Node19} \pgfnodeconnline{Node5}{Node20} \pgfnodeconnline{Node5}{Node21} \pgfnodeconnline{Node5}{Node22} \pgfnodeconnline{Node6}{Node15} \pgfnodeconnline{Node6}{Node16} \pgfnodeconnline{Node6}{Node17} \pgfnodeconnline{Node6}{Node18} \pgfnodeconnline{Node6}{Node19} \pgfnodeconnline{Node6}{Node23} \pgfnodeconnline{Node6}{Node24} \pgfnodeconnline{Node6}{Node25} \pgfnodeconnline{Node7}{Node20} \pgfnodeconnline{Node7}{Node24} \pgfnodeconnline{Node7}{Node26} \pgfnodeconnline{Node7}{Node27} \pgfnodeconnline{Node7}{Node28} \pgfnodeconnline{Node7}{Node29} \pgfnodeconnline{Node8}{Node21} \pgfnodeconnline{Node8}{Node23} \pgfnodeconnline{Node8}{Node29} \pgfnodeconnline{Node8}{Node30} \pgfnodeconnline{Node8}{Node31} \pgfnodeconnline{Node8}{Node32} \pgfnodeconnline{Node8}{Node33} \pgfnodeconnline{Node8}{Node34} \pgfnodeconnline{Node9}{Node33} \pgfnodeconnline{Node9}{Node35} \pgfnodeconnline{Node9}{Node36} \pgfnodeconnline{Node9}{Node37} \pgfnodeconnline{Node9}{Node38} \pgfnodeconnline{Node9}{Node39} \pgfnodeconnline{Node10}{Node34} \pgfnodeconnline{Node10}{Node39} \pgfnodeconnline{Node10}{Node40} \pgfnodeconnline{Node10}{Node41} \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.48cm}% \pgfnodecircle{Node9}[fiill]{\pgfxy(10.25,4.1)}{.48cm}% \pgfnodecircle{Node10}[stroke]{\pgfxy(11.5,3.5)}{.48cm}% \begin{pgfscope} \pgfsetlinewidth{2pt} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,0)}{.53cm}% \pgfnodecircle{Node2}[stroke]{\pgfxy(8.25,1.5)}{.53cm}% \pgfnodecircle{Node4}[stroke]{\pgfxy(9.25,2.75)}{.53cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.53cm}% \pgfnodecircle{Node33}[stroke]{\pgfxy(11.6,9.8)}{.55cm}% \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node8}{Node33} \end{pgfscope} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{-.3cm}} \begin{pgftranslate}{\pgfpoint{0cm}{0cm} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,1)}{0.13cm} \pgfxyline(8.12,.87)(8.38,1.13) \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{0cm}{1.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-1cm}{2.75cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1cm}{2.75cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.25cm}{3.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.15,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.15,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-2cm}{4.1cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.75cm}{4.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.2)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.2)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.75cm}{4.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2cm}{4.1cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,.8)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.8)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.25cm}{3.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(7.9,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.6,1)}{0.07cm} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-6.75cm}{4cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-7.55cm}{4.9cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-6.45cm}{5.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-7.05cm}{6.4cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-5.85cm}{6.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-7.45cm}{8.1cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.6)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.875)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1.125)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.4)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-6.05cm}{7.8cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-6.45cm}{9.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.85)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,1.15)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.1,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3cm}{6.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.95)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(7.95,1.2)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.55,1.2)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-1.45cm}{7.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-4.1cm}{7.7cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-5.15cm}{8.7cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.95)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(7.95,1.2)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.55,1.2)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.15cm}{10cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.1,1.3)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.25cm}{7cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.1,1.3)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-2.15cm}{9.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.95)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.35cm}{10cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-3.55cm}{8.8cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{-.95cm}{9.5cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1.1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1.1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{.95cm}{8.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.2,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.2)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.4,1.2)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.6,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node1}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.55cm}{9.2cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,.65)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(7.95,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.55,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.15cm}{7.9cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8.1,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.1,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.1,1.3)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node4}{Node3} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.25cm}{9cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.75)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{3.35cm}{9.8cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.75)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{2.15cm}{9cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{1.45cm}{10cm} \pgfnodecircle{Node1}[fill]{\pgfxy(7.95,.8)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.55,.8)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1.05)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.35)}{0.07cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{4.5cm}{6.6cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.7)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.35cm}{7.1cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.75)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.5,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node3}{Node2} \pgfnodeconnline{Node3}{Node4} \pgfnodeconnline{Node1}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{5.85cm}{6cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.7)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.5,.7)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.25,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.3)}{0.07cm} \pgfnodeconnline{Node1}{Node3} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7.15cm}{5.8cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.2,1.25)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.4,.75)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.6,1)}{0.07cm} \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node3}{Node2} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{6.45cm}{4.7cm} \pgfnodecircle{Node1}[fill]{\pgfxy(8,.75)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,.75)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.5,.75)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.25,1.25)}{0.07cm} \pgfnodeconnline{Node1}{Node4} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node4} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{7.55cm}{4.3cm} \pgfnodecircle{Node1}[fill]{\pgfxy(7.85,1)}{0.07cm} \pgfnodecircle{Node2}[fill]{\pgfxy(8.125,1)}{0.07cm} \pgfnodecircle{Node3}[fill]{\pgfxy(8.375,1)}{0.07cm} \pgfnodecircle{Node4}[fill]{\pgfxy(8.65,1)}{0.07cm} \end{pgftranslate} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{11.3cm}{.1cm}} \begin{pgfmagnify}{.3}{.3} \begin{colormixin}{50!white} \pgfnodecircle{Node1}[virtual]{\pgfxy(8.25,0)}{.35cm}% \pgfnodecircle{Node2}[virtual]{\pgfxy(8.25,1.5)}{.35cm}% \pgfnodecircle{Node3}[virtual]{\pgfxy(7.25,2.75)}{.35cm}% \pgfnodecircle{Node4}[virtual]{\pgfxy(9.25,2.75)}{.35cm}% \pgfnodecircle{Node5}[virtual]{\pgfxy(5,3.5)}{.35cm}% \pgfnodecircle{Node6}[virtual]{\pgfxy(6.25,4.1)}{.35cm}% \pgfnodecircle{Node7}[virtual]{\pgfxy(7.5,4.5)}{.35cm}% \pgfnodecircle{Node8}[virtual]{\pgfxy(9,4.5)}{.35cm}% \pgfnodecircle{Node9}[virtual]{\pgfxy(10.25,4.1)}{.35cm}% \pgfnodecircle{Node10}[virtual]{\pgfxy(11.5,3.5)}{.35cm}% \pgfnodecircle{Node11}[virtual]{\pgfxy(1.5,4)}{.35cm}% \pgfnodecircle{Node12}[virtual]{\pgfxy(.7,4.9)}{.35cm}% \pgfnodecircle{Node13}[virtual]{\pgfxy(1.8,5.2)}{.35cm}% \pgfnodecircle{Node14}[virtual]{\pgfxy(1.2,6.4)}{.35cm}% \pgfnodecircle{Node15}[virtual]{\pgfxy(2.4,6.2)}{.35cm}% \pgfnodecircle{Node16}[virtual]{\pgfxy(.8,8.1)}{.35cm}% \pgfnodecircle{Node17}[virtual]{\pgfxy(2.2,7.8)}{.35cm}% \pgfnodecircle{Node18}[virtual]{\pgfxy(1.8,9.2)}{.35cm}% \pgfnodecircle{Node19}[virtual]{\pgfxy(5.25,6.2)}{.35cm}% \pgfnodecircle{Node20}[virtual]{\pgfxy(6.8,7.5)}{.35cm}% \pgfnodecircle{Node21}[virtual]{\pgfxy(4.15,7.7)}{.35cm}% \pgfnodecircle{Node22}[virtual]{\pgfxy(3.1,8.7)}{.35cm}% \pgfnodecircle{Node23}[virtual]{\pgfxy(8.4,10)}{.35cm}% \pgfnodecircle{Node24}[virtual]{\pgfxy(8,7)}{.35cm}% \pgfnodecircle{Node25}[virtual]{\pgfxy(6.1,9.5)}{.35cm}% \pgfnodecircle{Node26}[virtual]{\pgfxy(4.9,10)}{.35cm}% \pgfnodecircle{Node27}[virtual]{\pgfxy(4.7,8.8)}{.35cm}% \pgfnodecircle{Node28}[virtual]{\pgfxy(7.3,9.5)}{.35cm}% \pgfnodecircle{Node29}[virtual]{\pgfxy(9.2,8.2)}{.35cm}% \pgfnodecircle{Node30}[virtual]{\pgfxy(13.8,9.2)}{.35cm}% \pgfnodecircle{Node31}[virtual]{\pgfxy(13.4,7.9)}{.35cm}% \pgfnodecircle{Node32}[virtual]{\pgfxy(12.5,9)}{.35cm}% \pgfnodecircle{Node33}[virtual]{\pgfxy(11.6,9.8)}{.35cm}% \pgfnodecircle{Node34}[virtual]{\pgfxy(10.4,9)}{.35cm}% \pgfnodecircle{Node35}[virtual]{\pgfxy(9.7,10)}{.35cm}% \pgfnodecircle{Node36}[virtual]{\pgfxy(12.75,6.6)}{.35cm}% \pgfnodecircle{Node37}[virtual]{\pgfxy(14.6,7.1)}{.35cm}% \pgfnodecircle{Node38}[virtual]{\pgfxy(14.1,6)}{.35cm}% \pgfnodecircle{Node39}[virtual]{\pgfxy(15.4,5.8)}{.35cm}% \pgfnodecircle{Node40}[virtual]{\pgfxy(14.7,4.7)}{.35cm}% \pgfnodecircle{Node41}[virtual]{\pgfxy(15.8,4.3)}{.35cm}% \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node3} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node3}{Node5} \pgfnodeconnline{Node3}{Node6} \pgfnodeconnline{Node3}{Node7} \pgfnodeconnline{Node3}{Node8} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node4}{Node9} \pgfnodeconnline{Node4}{Node10} \pgfnodeconnline{Node5}{Node11} \pgfnodeconnline{Node5}{Node12} \pgfnodeconnline{Node5}{Node13} \pgfnodeconnline{Node5}{Node14} \pgfnodeconnline{Node5}{Node19} \pgfnodeconnline{Node5}{Node20} \pgfnodeconnline{Node5}{Node21} \pgfnodeconnline{Node5}{Node22} \pgfnodeconnline{Node6}{Node15} \pgfnodeconnline{Node6}{Node16} \pgfnodeconnline{Node6}{Node17} \pgfnodeconnline{Node6}{Node18} \pgfnodeconnline{Node6}{Node19} \pgfnodeconnline{Node6}{Node23} \pgfnodeconnline{Node6}{Node24} \pgfnodeconnline{Node6}{Node25} \pgfnodeconnline{Node7}{Node20} \pgfnodeconnline{Node7}{Node24} \pgfnodeconnline{Node7}{Node26} \pgfnodeconnline{Node7}{Node27} \pgfnodeconnline{Node7}{Node28} \pgfnodeconnline{Node7}{Node29} \pgfnodeconnline{Node8}{Node21} \pgfnodeconnline{Node8}{Node23} \pgfnodeconnline{Node8}{Node29} \pgfnodeconnline{Node8}{Node30} \pgfnodeconnline{Node8}{Node31} \pgfnodeconnline{Node8}{Node32} \pgfnodeconnline{Node8}{Node33} \pgfnodeconnline{Node8}{Node34} \pgfnodeconnline{Node9}{Node33} \pgfnodeconnline{Node9}{Node35} \pgfnodeconnline{Node9}{Node36} \pgfnodeconnline{Node9}{Node37} \pgfnodeconnline{Node9}{Node38} \pgfnodeconnline{Node9}{Node39} \pgfnodeconnline{Node10}{Node34} \pgfnodeconnline{Node10}{Node39} \pgfnodeconnline{Node10}{Node40} \pgfnodeconnline{Node10}{Node41} \end{colormixin} \color{white} \pgfnodecircle{Node1}[fill]{\pgfxy(8.25,0)}{.4cm}% \pgfnodecircle{Node2}[fill]{\pgfxy(8.25,1.5)}{.4cm}% \pgfnodecircle{Node3}[fill]{\pgfxy(7.25,2.75)}{.4cm}% \pgfnodecircle{Node4}[fill]{\pgfxy(9.25,2.75)}{.4cm}% \pgfnodecircle{Node5}[fill]{\pgfxy(5,3.5)}{.4cm}% \pgfnodecircle{Node6}[fill]{\pgfxy(6.25,4.1)}{.4cm}% \pgfnodecircle{Node7}[fill]{\pgfxy(7.5,4.5)}{.4cm}% \pgfnodecircle{Node8}[fill]{\pgfxy(9,4.5)}{.4cm}% \pgfnodecircle{Node9}[fill]{\pgfxy(10.25,4.1)}{.4cm}% \pgfnodecircle{Node10}[fill]{\pgfxy(11.5,3.5)}{.4cm}% \pgfnodecircle{Node11}[fill]{\pgfxy(1.5,4)}{.4cm}% \pgfnodecircle{Node12}[fill]{\pgfxy(.7,4.9)}{.4cm}% \pgfnodecircle{Node13}[fill]{\pgfxy(1.8,5.2)}{.4cm}% \pgfnodecircle{Node14}[fill]{\pgfxy(1.2,6.4)}{.4cm}% \pgfnodecircle{Node15}[fill]{\pgfxy(2.4,6.2)}{.4cm}% \pgfnodecircle{Node16}[fill]{\pgfxy(.8,8.1)}{.4cm}% \pgfnodecircle{Node17}[fill]{\pgfxy(2.2,7.8)}{.4cm}% \pgfnodecircle{Node18}[fill]{\pgfxy(1.8,9.2)}{.4cm}% \pgfnodecircle{Node19}[fill]{\pgfxy(5.25,6.2)}{.4cm}% \pgfnodecircle{Node20}[fill]{\pgfxy(6.8,7.5)}{.4cm}% \pgfnodecircle{Node21}[fill]{\pgfxy(4.15,7.7)}{.4cm}% \pgfnodecircle{Node22}[fill]{\pgfxy(3.1,8.7)}{.4cm}% \pgfnodecircle{Node23}[fill]{\pgfxy(8.4,10)}{.4cm}% \pgfnodecircle{Node24}[fill]{\pgfxy(8,7)}{.4cm}% \pgfnodecircle{Node25}[fill]{\pgfxy(6.1,9.5)}{.4cm}% \pgfnodecircle{Node26}[fill]{\pgfxy(4.9,10)}{.4cm}% \pgfnodecircle{Node27}[fill]{\pgfxy(4.7,8.8)}{.4cm}% \pgfnodecircle{Node28}[fill]{\pgfxy(7.3,9.5)}{.4cm}% \pgfnodecircle{Node29}[fill]{\pgfxy(9.2,8.2)}{.4cm}% \pgfnodecircle{Node30}[fill]{\pgfxy(13.8,9.2)}{.4cm}% \pgfnodecircle{Node31}[fill]{\pgfxy(13.4,7.9)}{.4cm}% \pgfnodecircle{Node32}[fill]{\pgfxy(12.5,9)}{.4cm}% \pgfnodecircle{Node33}[fill]{\pgfxy(11.6,9.8)}{.4cm}% \pgfnodecircle{Node34}[fill]{\pgfxy(10.4,9)}{.4cm}% \pgfnodecircle{Node35}[fill]{\pgfxy(9.7,10)}{.4cm}% \pgfnodecircle{Node36}[fill]{\pgfxy(12.75,6.6)}{.4cm}% \pgfnodecircle{Node37}[fill]{\pgfxy(14.6,7.1)}{.4cm}% \pgfnodecircle{Node38}[fill]{\pgfxy(14.1,6)}{.4cm}% \pgfnodecircle{Node39}[fill]{\pgfxy(15.4,5.8)}{.4cm}% \pgfnodecircle{Node40}[fill]{\pgfxy(14.7,4.7)}{.4cm}% \pgfnodecircle{Node41}[fill]{\pgfxy(15.8,4.3)}{.4cm}% \begin{colormixin}{20!white} \color{black} \pgfnodecircle{Node6}[fill]{\pgfxy(6.25,4.1)}{.4cm}% \pgfnodecircle{Node12}[fill]{\pgfxy(.7,4.9)}{.4cm}% \pgfnodecircle{Node13}[fill]{\pgfxy(1.8,5.2)}{.4cm}% \pgfnodecircle{Node14}[fill]{\pgfxy(1.2,6.4)}{.4cm}% \pgfnodecircle{Node15}[fill]{\pgfxy(2.4,6.2)}{.4cm}% \pgfnodecircle{Node16}[fill]{\pgfxy(.8,8.1)}{.4cm}% \pgfnodecircle{Node17}[fill]{\pgfxy(2.2,7.8)}{.4cm}% \pgfnodecircle{Node18}[fill]{\pgfxy(1.8,9.2)}{.4cm}% \pgfnodecircle{Node19}[fill]{\pgfxy(5.25,6.2)}{.4cm}% \pgfnodecircle{Node23}[fill]{\pgfxy(8.4,10)}{.4cm}% \pgfnodecircle{Node24}[fill]{\pgfxy(8,7)}{.4cm}% \pgfnodecircle{Node25}[fill]{\pgfxy(6.1,9.5)}{.4cm}% \pgfnodecircle{Node27}[fill]{\pgfxy(4.7,8.8)}{.4cm}% \pgfnodecircle{Node31}[fill]{\pgfxy(13.4,7.9)}{.4cm}% \pgfnodecircle{Node36}[fill]{\pgfxy(12.75,6.6)}{.4cm}% \pgfnodecircle{Node38}[fill]{\pgfxy(14.1,6)}{.4cm}% \end{colormixin} \color{black} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,0)}{.4cm}% \pgfnodecircle{Node2}[stroke]{\pgfxy(8.25,1.5)}{.4cm}% \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.4cm}% \pgfnodecircle{Node4}[stroke]{\pgfxy(9.25,2.75)}{.4cm}% \pgfnodecircle{Node5}[stroke]{\pgfxy(5,3.5)}{.4cm}% \pgfnodecircle{Node6}[stroke]{\pgfxy(6.25,4.1)}{.4cm}% \pgfnodecircle{Node7}[stroke]{\pgfxy(7.5,4.5)}{.4cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.4cm}% \pgfnodecircle{Node9}[stroke]{\pgfxy(10.25,4.1)}{.4cm}% \pgfnodecircle{Node10}[stroke]{\pgfxy(11.5,3.5)}{.4cm}% \pgfnodecircle{Node11}[stroke]{\pgfxy(1.5,4)}{.4cm}% \pgfnodecircle{Node12}[stroke]{\pgfxy(.7,4.9)}{.4cm}% \pgfnodecircle{Node13}[stroke]{\pgfxy(1.8,5.2)}{.4cm}% \pgfnodecircle{Node14}[stroke]{\pgfxy(1.2,6.4)}{.4cm}% \pgfnodecircle{Node15}[stroke]{\pgfxy(2.4,6.2)}{.4cm}% \pgfnodecircle{Node16}[stroke]{\pgfxy(.8,8.1)}{.4cm}% \pgfnodecircle{Node17}[stroke]{\pgfxy(2.2,7.8)}{.4cm}% \pgfnodecircle{Node18}[stroke]{\pgfxy(1.8,9.2)}{.4cm}% \pgfnodecircle{Node19}[stroke]{\pgfxy(5.25,6.2)}{.4cm}% \pgfnodecircle{Node20}[stroke]{\pgfxy(6.8,7.5)}{.4cm}% \pgfnodecircle{Node21}[stroke]{\pgfxy(4.15,7.7)}{.4cm}% \pgfnodecircle{Node22}[stroke]{\pgfxy(3.1,8.7)}{.4cm}% \pgfnodecircle{Node23}[stroke]{\pgfxy(8.4,10)}{.4cm}% \pgfnodecircle{Node24}[stroke]{\pgfxy(8,7)}{.4cm}% \pgfnodecircle{Node25}[stroke]{\pgfxy(6.1,9.5)}{.4cm}% \pgfnodecircle{Node26}[stroke]{\pgfxy(4.9,10)}{.4cm}% \pgfnodecircle{Node27}[stroke]{\pgfxy(4.7,8.8)}{.4cm}% \pgfnodecircle{Node28}[stroke]{\pgfxy(7.3,9.5)}{.4cm}% \pgfnodecircle{Node29}[stroke]{\pgfxy(9.2,8.2)}{.4cm}% \pgfnodecircle{Node30}[stroke]{\pgfxy(13.8,9.2)}{.4cm}% \pgfnodecircle{Node31}[stroke]{\pgfxy(13.4,7.9)}{.4cm}% \pgfnodecircle{Node32}[stroke]{\pgfxy(12.5,9)}{.4cm}% \pgfnodecircle{Node33}[stroke]{\pgfxy(11.6,9.8)}{.4cm}% \pgfnodecircle{Node34}[stroke]{\pgfxy(10.4,9)}{.4cm}% \pgfnodecircle{Node35}[stroke]{\pgfxy(9.7,10)}{.4cm}% \pgfnodecircle{Node36}[stroke]{\pgfxy(12.75,6.6)}{.4cm}% \pgfnodecircle{Node37}[stroke]{\pgfxy(14.6,7.1)}{.4cm}% \pgfnodecircle{Node38}[stroke]{\pgfxy(14.1,6)}{.4cm}% \pgfnodecircle{Node39}[stroke]{\pgfxy(15.4,5.8)}{.4cm}% \pgfnodecircle{Node40}[stroke]{\pgfxy(14.7,4.7)}{.4cm}% \pgfnodecircle{Node41}[stroke]{\pgfxy(15.8,4.3)}{.4cm}% \color{black} \pgfnodecircle{Node3}[stroke]{\pgfxy(7.25,2.75)}{.25cm}% \pgfnodecircle{Node9}[stroke]{\pgfxy(10.25,4.1)}{.25cm}% \pgfnodecircle{Node10}[stroke]{\pgfxy(11.5,3.5)}{.25cm}% \pgfsetlinewidth{4pt} \pgfnodecircle{Node1}[stroke]{\pgfxy(8.25,0)}{.4cm}% \pgfnodecircle{Node2}[stroke]{\pgfxy(8.25,1.5)}{.4cm}% \pgfnodecircle{Node33}[stroke]{\pgfxy(11.6,9.8)}{.4cm}% \pgfnodecircle{Node4}[stroke]{\pgfxy(9.25,2.75)}{.4cm}% \pgfnodecircle{Node8}[stroke]{\pgfxy(9,4.5)}{.4cm}% \pgfnodeconnline{Node1}{Node2} \pgfnodeconnline{Node2}{Node4} \pgfnodeconnline{Node4}{Node8} \pgfnodeconnline{Node8}{Node33} \end{pgfmagnify} \end{pgftranslate} \comment{ \pgfputat{\pgfxy(9.3,4.3)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(6.9,7.25)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(7.95,6.8)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(11.4,9)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(12.85,6.25)}{\pgfbox[center,center]{$2$}} \pgfputat{\pgfxy(13.1,5.35)}{\pgfbox[center,center]{$3$}} \pgfputat{\pgfxy(11.2,6.3)}{\pgfbox[center,center]{$3$}} } \pgfputat{\pgfxy(12.3,11.1)}{\pgfbox[center,center]{$A$}} \pgfputat{\pgfxy(14.3,.15)}{\pgfbox[center,center]{\fsz{$x(\oslash)$}}} \pgfputat{\pgfxy(15.25,3.28)}{\pgfbox[center,center]{\fsz{$x(A)$}}} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{A portion of the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$; inset shows the underlying multidirected set $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$; large-font numbers indicate generations.} \label{figsequentialscheme} \end{figure} \vspace*{-.5cm} The reader should compare figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} to figure \hyperref[sequential]{\ref{sequential}} of section \hyperref[subsectionapproach]{\ref{subsectionapproach}} above, and also to figure $1$ of Sorkin and Rideout's paper {\it Classical sequential growth dynamics for causal sets} \cite{SorkinSequentialGrowthDynamics99}, which illustrate portions of Sorkin and Rideout's kinematic scheme $(\ms{C}_{\tn{\fsz{fin}}},\mc{H}_1)$. The large numbers $0$ through $4$ in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} indicate how each member set $D$ of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ may be assigned to a positive integer-valued {\it generation} by its cardinality, or equivalently, by the common length of all originary chains terminating at $x(D)$ in the underlying multidirected set $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$. The nodes shaded grey in the figure represent member sets which are {\it nontransitive}, and which therefore belong to $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ but {\it not} to Sorkin and Rideout's kinematic scheme $(\ms{C}_{\tn{\fsz{fin}}},\mc{H}_1)$. In particular, generations $0$ through $3$ of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ differ from the corresponding portion of $(\ms{C}_{\tn{\fsz{fin}}},\mc{H}_1)$ by a single member, the nontransitive chain of length three. However, about half of the fourth generation of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ consists of nontransitive acyclic directed sets, and subsequent generations are dominated by them. The inset in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} illustrates the portion of the underlying multidirected set $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ corresponding to the portion of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ appearing in the main figure. $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ is a {\it a countably infinite, locally finite, acyclic multidirected set.} The portion shown here does not contain multiple edges between individual pairs of elements, but multiple edges begin to appear by the seventh and eighth generations. The fact that $\ms{A}_{\tn{\fsz{fin}}}$, and hence $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, is countable, was proven in section \hyperref[settheoretic]{\ref{settheoretic}} above, while the local finiteness and acyclicity of $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ follow from the finite cardinality of the member sets of $\ms{A}_{\tn{\fsz{fin}}}$, and the definition of a co-relative history. The remaining labels in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} are to aid in the discussion below. \refstepcounter{textlabels}\label{transindices} \refstepcounter{textlabels}\label{causalgalois} Underlying directed sets and multidirected sets of kinematic schemes generally suffer from the problem of permeability of maximal antichains in multidirected sets, discussed in section \hyperref[subsectionrelation]{\ref{subsectionrelation}} above. The underlying multidirected set $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ of the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ is no exception. For example, the three nodes distinguished by double circles in figure \hyperref[figsequentialscheme]{\ref{figsequentialscheme}} represent a maximal antichain in $\ms{V}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, permeated by the chain from $x(\oslash)$ to $x(A)$ indicated by the thickened nodes and edges. This indicates that passage to relation space remains a crucial tool in the context of kinematic schemes. A variety of natural relation functions may be associated with underlying directed or multidirected sets of kinematic schemes. Perhaps the most obvious choices for such relation functions involve the automorphism groups of the directed sets corresponding to the initial and terminal elements of each relation. For example, the subgroup indices of the {\it causal Galois groups,} mentioned in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, provide examples of such relation functions. Analogous ``higher-level" relation functions are given by simply assigning the entire causal Galois groups to their corresponding relations. As mentioned in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}, such relation functions may be viewed as partial precursors for the phase maps appearing in the path summation approach to abstract quantum causal theory. \refstepcounter{textlabels}\label{genkinematics} {\bf Generational Kinematics for $\ms{A}_{\tn{\fsz{fin}}}$.} The positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ is not the only interesting kinematic scheme whose object class is the set $\ms{A}_{\tn{\fsz{fin}}}$ of isomorphism classes of finite acyclic directed sets. An alternative kinematic scheme over $\ms{A}_{\tn{\fsz{fin}}}$, which provides closer analogues of relativistic frames of reference than $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, is the {\bf generational kinematic scheme} $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_{\tn{\fsz{gen}}})$. The class $\mc{H}_{\tn{\fsz{gen}}}$ of co-relative histories of the generational kinematic scheme includes every co-relative history adding a single {\it generation;} i.e., antichain, to its cobase, rather than merely a single element. The generational kinematic scheme provides formal analogues of {\it spacelike foliations} of members of $\ms{A}_{\tn{\fsz{fin}}}$, and of countably infinite limit sets of $\ms{A}_{\tn{\fsz{fin}}}$, discussed below. The underlying multidirected set of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_{\tn{\fsz{gen}}})$ is acyclic, but is not locally finite, since each new generation may have any positive integer cardinality. \refstepcounter{textlabels}\label{completions} {\bf Completions of Kinematic Schemes.} In addition to organizing the set of isomorphism classes of finite acyclic directed sets into a higher-level multidirected structure, the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ also encodes the structure of certain isomorphism classes of {\it countably infinite} acyclic directed sets, corresponding to {\it equivalence classes of maximal chains in the underlying transition structure} $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$. These countably infinite sets may be viewed as {\it limits,} in a generalized category-theoretic sense. At a more elementary level, they are analogous to limits of Cauchy sequences of rational numbers. More precisely, let \[x(\oslash)\prec x(A_1)\prec x(A_2)\prec...\prec x(A_n)\prec...\] be a maximal chain in the underlying transition structure $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, where each precursor symbol $\prec$ represents a {\it specific} transition. Each finite acyclic directed set $A_n$ in this chain may be regarded as a subset of the succeeding set $A_{n+1},$, and each relation in $A_n$ may be regarded as a relation in $A_{n+1}$. Hence, the union $A_\infty:=\cup_{n=1}^\infty A_n$ is a well-defined, countably infinite, acyclic directed set. The maximal chain in $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ corresponding to $A_\infty$ determines a total labeling of $A_\infty$, which corresponds to a bijective morphism $A_\infty\rightarrow \mathbb{N}$. Conversely, every bijective morphism from a directed set\footnotemark\footnotetext{The existence of such a morphism guarantees that the source is in fact a countably infinite acyclic directed set.} into $\mathbb{N}$ determines a unique maximal chain in $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$. Hence, the class of maximal chains in $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ corresponds bijectively to the class of countably infinite relative directed sets over $\mathbb{N}$, viewed as morphisms.\footnotemark\footnotetext{As usual, only isomorphism classes are significant.} Defining an appropriate equivalence relation on maximal chains in $\ms{W}(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$, corresponding to the equivalence relation identifying relative directed sets over $\mathbb{N}$ sharing common sources, identifies a class of countably infinite acyclic directed sets called {\bf limits} of the kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$. Augmenting $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ by its class of limits yields a new kinematic scheme $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$ called the {\bf chain completion} of $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$. The object class of $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$ is the familiar set of isomorphism classes of countable acyclic directed sets representable as relative directed sets over $\mathbb{N}$, which has cardinality $2^{\aleph_0}$. As suggested by its name, its definition in terms of maximal chains, and its cardinality, $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$ is related to $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ in a manner analogous to the relationship between the real numbers $\mathbb{R}$ and the rational numbers $\mathbb{Q}$. More generally, let $(\ms{K},\mc{H})$ be a kinematic scheme, and let $\widetilde{\tn{Ch}_{\tn{\fsz{max}}}}=\widetilde{\tn{Ch}_{\tn{\fsz{max}}}}(\ms{K},\mc{H})$ be the class of limits of $(\ms{K},\mc{H})$; i.e., the class of directed sets represented by equivalence classes of maximal chains in $\ms{W}(\ms{K},\mc{H})$.\footnotemark\footnotetext{Care is required when $\ms{W}(\ms{K},\mc{H})$ includes cycles.} Then the {\bf chain completion} $\overline{(\ms{K},\mc{H})}$ of $(\ms{K},\mc{H})$ is defined to be the kinematic prescheme $\big(\ms{K}\cup\widetilde{\tn{Ch}_{\tn{\fsz{max}}}},\mc{H}\big)$ given by enlarging the object class $\ms{K}$ to include equivalence classes of maximal chains, {\it without adding new co-relative histories.} It is easy to see that the chain completion of a kinematic scheme is a kinematic scheme: the addition of directed sets corresponding to maximal chains preserves the hereditary property (\hyperref[h]{H}) by construction, while the finite complement condition in the definition of weak accessibility (\hyperref[wa]{WA}) exempts the added sets from the necessity of accessibility by a finite chain in $\ms{W}(\ms{K},\mc{H})$. This, in fact, is one of the principal reasons for imposing the finite complement condition. \refstepcounter{textlabels}\label{kinschemecountablyinf} {\bf Kinematic Schemes of Countably Infinite Acyclic Directed Sets.} The chain completion $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$ of the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ provides a simple example of a kinematic scheme including infinite directed sets among its objects. However, these infinite sets are of a very restricted type, even in the locally finite context, being representable as relative directed sets over $\mathbb{N}$. Further, they are ``mere limits," in the sense that no co-relative histories involving these sets, either as cobases or targets, are admitted into the higher-level multidirected structure of $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$. There are good physical reasons to consider more complex kinematic schemes including countably infinite member sets; for example, kinematic schemes whose object classes are the classes $\ms{C}$ and $\ms{A}_{\aleph_0}$ of {\it all} countably infinite causal sets and locally finite acyclic directed sets, respectively. Such kinematic schemes have the same cardinality $2^{\aleph_0}$ as $\overline{(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)}$, but are generally much more complicated. For example, they usually contain cycles, such as the ``self-transition" of causal sets $-\mathbb{N}\rightarrow-\mathbb{N}$ sending $-n$ to $-n-1$, whose ``new" element is the maximal element $0$ in the second copy of $-\mathbb{N}$. \refstepcounter{textlabels}\label{morphkinscheme} {\bf ``Functors" of Kinematic Schemes.} The principle of iteration of structure (\hyperref[is]{IS}) may be extended to apply in more general contexts than the relationship between directed sets and kinematic schemes. At least one higher level of hierarchy is also physically relevant. This level of hierarchy involves {\it relationships between pairs of kinematic schemes,} following Grothendieck's relative viewpoint (\hyperref[rv]{RV}). These relationships may be viewed as ``higher-level morphisms," or ``higher-level co-relative histories." Drawing on the relativistic analogy, they represent a vast generalization of coordinate transformations. From an algebraic perspective, in which kinematic schemes are viewed as generalized categories, relationships between kinematic schemes may be viewed as generalized functors. In this paper, I do not attempt to undertake any systematic analysis of physically essential structure for such relationships, as I did for co-relative histories in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} above, nor do I even give general definitions. Instead, I briefly discuss a particular class of examples, for purely illustrative purposes. Let $(\ms{K},\mc{H})$ and $(\ms{K}',\mc{H}')$ be kinematic schemes such that $\ms{K}\subset\ms{K}'$ and $\mc{H}\subset\mc{H}'$. In this case, there is a natural inclusion ``functor" $\mc{I}:(\ms{K},\mc{H})\rightarrow (\ms{K}',\mc{H}')$. I focus on the case where the inclusion of objects is proper, and where the inclusion of co-relative histories is full, meaning that any ``new" co-relative histories involve ``new" objects. Further, I assume that every object of $\ms{K}$ is finite, and every object of $\ms{K}'$ is countable. In this case, the cardinalities of $\ms{K}$ and $\ms{K}'$ are ``typically" $\aleph_0$ and $2^{\aleph_0}$, respectively. The ``inclusion functor" $\mc{I}$ is a higher-level analogue of a transition, since it embeds $(\ms{K},\mc{H})$ into $(\ms{K}',\mc{H}')$ as a ``proper, full, originary subscheme." The prototypical example is the inclusion of the positive sequential kinematic scheme into its chain completion. \refstepcounter{textlabels}\label{numberanalogies} Interesting analogies exist between such ``functors" and inclusions of number systems of cardinality $\aleph_0$, such as the integers or rational numbers, into number systems of cardinality $2^{\aleph_0}$, such as the real numbers, $p$-adic numbers, complex numbers, etc. For example, consider the set of rational numbers in the unit interval $[0,1]$, represented by irreducible fractions, with binary relation $\prec$ induced by the usual linear order on $\mathbb{Q}$ via the {\it Farey sequences,} in which each fraction is related to ``the next two irreducible fractions bracketing it in the usual order." The resulting acyclic directed set $F=(F,\prec)$ is a {\it tree} from its ``second generation" onward, sometimes called the {\it Farey tree}. A small portion of the Farey tree is illustrated in figure \hyperref[farey]{\ref{farey}}a below. A maximal chain in $F$ represents a Cauchy sequence of rational numbers in $[0,1]$, and hence an irrational number in the interval $[0,1]$. Moreover, every irrational number in $[0,1]$ is represented by a maximal chain in $(F,\prec)$ in this way. In this particular case, the representation is unique, precisely because $F$ is a tree from its second generation onward. In a similar way, a maximal chain in the underlying transition structure $\ms{W}(\ms{K},\mc{H})$ of a kinematic scheme $(\ms{K},\mc{H})$ represents an object in its chain completion. In this case, the representation is generally {\it not} unique, since $\ms{W}(\ms{K},\mc{H})$ is generally not a tree. \begin{figure}[H] \begin{pgfpicture}{0cm}{0cm}{17cm}{5.1cm} \begin{pgfmagnify}{1.03}{1.03} \pgfxyline(0,-.1)(16.5,-.1) \pgfxyline(0,5.1)(16.5,5.1) \pgfxyline(0,-.1)(0,5.1) \pgfxyline(16.5,-.1)(16.5,5.1) \pgfxyline(9.5,-.1)(9.5,5.1) \pgfxyline(10.4,1.5)(15.3,1.5) \pgfxyline(10.4,2.7)(15.3,2.7) \begin{pgfscope} \pgfsetlinewidth{1.5pt} \pgfxyline(13.6,.5)(13.6,4.5) \pgfxyline(10.4,3.9)(15.3,3.9) \end{pgfscope} \pgfputat{\pgfxy(.15,4.8)}{\pgfbox[left,center]{a)}} \pgfputat{\pgfxy(9.65,4.8)}{\pgfbox[left,center]{b)}} \begin{pgftranslate}{\pgfpoint{-7.3cm}{-.4cm}} \pgfputat{\pgfxy(9,1)}{\pgfbox[center,center]{\fsz{$0$}}} \pgfputat{\pgfxy(15,1)}{\pgfbox[center,center]{\fsz{$1$}}} \pgfputat{\pgfxy(12,2.1)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{1}{2}$}}} \pgfputat{\pgfxy(10.5,3.2)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{1}{3}$}}} \pgfputat{\pgfxy(13.5,3.2)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{2}{3}$}}} \pgfputat{\pgfxy(9.8,4.3)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{1}{4}$}}} \pgfputat{\pgfxy(14.2,4.3)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{3}{4}$}}} \pgfputat{\pgfxy(11.2,4.3)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{2}{5}$}}} \pgfputat{\pgfxy(12.8,4.3)}{\pgfbox[center,center]{\fsz{$\displaystyle\frac{3}{5}$}}} \pgfnodecircle{Node11}[stroke]{\pgfxy(9,1)}{.4cm} \pgfnodecircle{Node12}[stroke]{\pgfxy(15,1)}{.4cm} \pgfnodecircle{Node21}[stroke]{\pgfxy(12,2.1)}{.4cm} \pgfnodecircle{Node31}[stroke]{\pgfxy(10.5,3.2)}{.4cm} \pgfnodecircle{Node32}[stroke]{\pgfxy(13.5,3.2)}{.4cm} \pgfnodecircle{Node41}[stroke]{\pgfxy(9.8,4.3)}{.4cm} \pgfnodecircle{Node42}[stroke]{\pgfxy(11.2,4.3)}{.4cm} \pgfnodecircle{Node43}[stroke]{\pgfxy(12.8,4.3)}{.4cm} \pgfnodecircle{Node44}[stroke]{\pgfxy(14.2,4.3)}{.4cm} \pgfnodecircle{Node51}[virtual]{\pgfxy(9.4,5.3)}{.1cm} \pgfnodecircle{Node52}[virtual]{\pgfxy(10.2,5.3)}{.1cm} \pgfnodecircle{Node53}[virtual]{\pgfxy(10.8,5.3)}{.1cm} \pgfnodecircle{Node54}[virtual]{\pgfxy(11.6,5.3)}{.1cm} \pgfnodecircle{Node55}[virtual]{\pgfxy(12.4,5.3)}{.1cm} \pgfnodecircle{Node56}[virtual]{\pgfxy(13.2,5.3)}{.1cm} \pgfnodecircle{Node57}[virtual]{\pgfxy(13.8,5.3)}{.1cm} \pgfnodecircle{Node58}[virtual]{\pgfxy(14.6,5.3)}{.1cm} \pgfnodeconnline{Node11}{Node21} \pgfnodeconnline{Node12}{Node21} \pgfnodeconnline{Node21}{Node31} \pgfnodeconnline{Node21}{Node32} \pgfnodeconnline{Node31}{Node41} \pgfnodeconnline{Node31}{Node42} \pgfnodeconnline{Node32}{Node43} \pgfnodeconnline{Node32}{Node44} \pgfsetdash{{2pt}{2pt}}{0pt} \pgfnodeconnline{Node41}{Node51} \pgfnodeconnline{Node41}{Node52} \pgfnodeconnline{Node42}{Node53} \pgfnodeconnline{Node42}{Node54} \pgfnodeconnline{Node43}{Node55} \pgfnodeconnline{Node43}{Node56} \pgfnodeconnline{Node44}{Node57} \pgfnodeconnline{Node44}{Node58} \end{pgftranslate} \begin{pgftranslate}{\pgfpoint{10.2cm}{-.6cm}} \pgfputat{\pgfxy(1.7,4.9)}{\pgfbox[center,center]{$(\ms{K},\mc{H})$}} \pgfputat{\pgfxy(4.2,4.9)}{\pgfbox[center,center]{$\mathbb{Q}$}} \pgfputat{\pgfxy(1.7,4.1)}{\pgfbox[center,center]{\small{minimal nonempty}}} \pgfputat{\pgfxy(1.7,3.7)}{\pgfbox[center,center]{{\small objects }}} \pgfputat{\pgfxy(4.2,3.9)}{\pgfbox[center,center]{$\mathbb{Z}$}} \pgfputat{\pgfxy(1.7,2.7)}{\pgfbox[center,center]{{\small objects}}} \pgfputat{\pgfxy(4.2,2.7)}{\pgfbox[center,center]{$\mathbb{Q}$}} \pgfputat{\pgfxy(1.7,1.6)}{\pgfbox[center,center]{\small{maximal chains}}} \pgfputat{\pgfxy(4.2,1.6)}{\pgfbox[center,center]{$\mathbb{R}-\mathbb{Q}$}} \end{pgftranslate} \begin{colormixin}{15!white} \begin{pgfmagnify}{.6}{.6} \pgfputat{\pgfxy(27,.2)}{\pgfbox[center,center]{\tiny{BDCT}}} \pgfputat{\pgfxy(27,0)}{\pgfbox[center,center]{\tiny{TGGA}}} \end{pgfmagnify} \end{colormixin} \end{pgfmagnify}\end{pgfpicture} \caption{a) Farey ``tree" $(F,\prec)$; chain completion is $[0,1]\subset\mathbb{R}$; b) analogies between a countable kinematic scheme and familiar number systems.} \label{farey} \end{figure} \vspace*{-.5cm} \refstepcounter{textlabels}\label{universalkinschemes} {\bf Universal Kinematic Schemes; Kinematic Spaces.} For a given class $\ms{K}$ of directed sets, considered up to isomorphism, there are generally many different classes $\mc{H}$ of co-relative between pairs of members of $\ms{K}$, and in particular, many different kinematic schemes $(\ms{K},\mc{H})$. The class $\mc{H}(\ms{K})$ of {\it all} proper, full, originary co-relative histories in $\ms{D}$ between pairs of members of $\ms{K}$ defines a ``transitive" kinematic scheme over $\ms{K}$, which is {\it maximal} in the class of all proper, full, originary kinematic preschemes over $\ms{K}$. Due to this property, I refer to $(\ms{K},\mc{H}(\ms{K}))$ as the {\bf universal kinematic scheme} over $\ms{K}$. Despite its uniqueness, $(\ms{K},\mc{H}(\ms{K}))$ can be cumbersome to work with, since it encodes ``many redundant accounts of each evolutionary process." At the opposite extreme are ``irreducible" kinematic schemes, in which descriptions of evolutionary processes are ``as unique as possible." In the locally finite case, a kinematic scheme may be reduced to an irreducible kinematic scheme by applying the skeleton operation to its underlying multidirected set, since this operation does not effect the hereditary property (\hyperref[h]{H}) or weak accessibility (\hyperref[wa]{WA}) in this context. For example, the skeleton of the universal kinematic scheme $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}(\ms{A}_{\tn{\fsz{fin}}}))$ over the set $\ms{A}_{\tn{\fsz fin}}$ of isomorphism classes of finite acyclic directed sets is the positive sequential kinematic scheme $(\ms{A}_{\tn{\fsz fin}},\mc{H}_1)$. Since ``transitive" kinematic schemes over classes of {\it finite} directed sets are recoverable from their skeleta, $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}_1)$ inherits a universal property of sorts from $(\ms{A}_{\tn{\fsz{fin}}},\mc{H}(\ms{A}_{\tn{\fsz{fin}}}))$. This provides a purely structural reason for regarding the positive sequential kinematic scheme as important. Every proper, full, originary kinematic prescheme with object class $\ms{K}$ is a subobject of the universal kinematic scheme $(\ms{K},\mc{H}(\ms{K}))$ over $\ms{K}$. In terms of underlying multidirected sets, these kinematic preschemes are given by deleting relations of $\ms{V}(\ms{K},\mc{H}(\ms{K}))$, without altering its class of elements. The class of proper, full, originary kinematic preschemes over $\ms{K}$ is then naturally ordered by proper inclusion of classes of co-relative histories, with $(\ms{K},\mc{H}(\ms{K}))$ being the unique maximal element. These inclusions correspond to ``functors" of kinematic schemes, bijective on objects, but neither full nor originary. They induce a natural partial order $\prec_{\ms{K}}$ on the class of proper, full, originary kinematic preschemes over $\ms{K}$, in which $(\ms{K},\mc{H})\prec_{\ms{K}}(\ms{K},\mc{H}')$ if and only if there is a proper inclusion $\mc{H}\subset\mc{H}'$. Restricting to kinematic schemes over $\ms{K}$ yields a structure $(\tn{\bf Kin}(\ms{K}),\prec_{\ms{K}})$, called the {\bf kinematic space over} $\ms{K}$, whose ``elements" correspond to proper, full, originary kinematic schemes over $\ms{K}$. \newpage \section{Conclusions}\label{sectionconclusions} This final section of the paper is devoted to two main tasks. The first, undertaken in section \hyperref[subsectionalternative]{\ref{subsectionalternative}}, is to summarize the principal conclusions of the preceding sections. Included are a short recapitulation of the axiomatic analysis of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}}; a specific proposal for a new set of axioms for discrete causal theory; a brief discussion of possible alternatives to this proposal, both conservative and radical; and a concise review of the new perspectives and technical methods introduced here. The second task, pursued in section \hyperref[subsectionomitted]{\ref{subsectionomitted}}, is more difficult and open-ended. It consists of briefly outlining important omitted topics. Some of these topics were included in earlier versions of the paper, but were excised due to length considerations, or due to my dissatisfaction with my own understanding of them. Other omitted topics involve elaboration of the material presented here. Included in this section are further remarks about the axioms of countability and irreflexivity/acyclicity, not examined in detail in the main body of the paper; an additional word or two about covariance; a glimpse of algebraic and structural topics only encountered obliquely in the paper; a speculative elaboration regarding phase maps; a short description of the mathematical field of {\it random graph dynamics,} likely to play a crucial role in mechanisms of emergence in discrete causal theory; and an outline of alternatives to the theory of power spaces for generalizing order theory in physics, particularly in the context of classical holism. The section concludes with a list of other physical and mathematical theories with potentially interesting connections to the ideas developed in this paper. Finally, section \hyperref[subsectionacknowledgements]{\ref{subsectionacknowledgements}} consists of acknowledgements and a few miscellaneous notes. \subsection{New Axioms, Perspectives, and Technical Methods}\label{subsectionalternative} \refstepcounter{textlabels}\label{axiomaticanalysis} {\bf Summary of Axiomatic Analysis.} Sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}} of this paper present the case that the causal set axioms of {\it transitivity} (\hyperref[tr]{TR}) and {\it interval finiteness} (\hyperref[if]{IF}) are unsuitable for discrete causal theory under Sorkin's version of the {\it causal metric hypothesis} (\hyperref[cmh]{CMH}). Section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} addresses the shortcomings of transitivity, focusing on the case of directed sets, and particularly acyclic directed sets. The principal conclusion is that transitive causal orders, or generalizations of such orders containing cycles, should be viewed as derivative constructs of generally nontransitive binary relations called {\it causal preorders,} physically interpreted via the {\it independence convention} (\hyperref[ic]{IC}), introduced in section \hyperref[subsectionchains]{\ref{subsectionchains}}. Section \hyperref[subsectiontransitivitydeficient]{\ref{subsectiontransitivitydeficient}} presents six arguments that transitive relations are information-theoretically deficient. Section \hyperref[sectioninterval]{\ref{sectioninterval}} details the many shortcomings of interval finiteness as a local finiteness condition for directed sets and multidirected sets. The proposed solution is to replace interval finiteness with the alternative axiom of {\it local finiteness} (\hyperref[lf]{LF}). Sections \hyperref[subsectionintervalfinitenessdeficient]{\ref{subsectionintervalfinitenessdeficient}} and \hyperref[subsectionarglocfin]{\ref{subsectionarglocfin}} summarize a few of the principal arguments against interval finiteness and in favor of local finiteness. The remaining axioms of causal set theory are the {\it binary axiom} (\hyperref[b]{B}), the {\it measure axiom} ({\hyperref[m]{M}), {\it countability} (\hyperref[c]{C}), and {\it irreflexivity} (\hyperref[ir]{IR}). None of these axioms seem to present problems necessitating their outright abandonment, but more subtle changes are worth considering. Most obviously, irreflexivity no longer rules out causal cycles in the absence of transitivity, leading to a choice to either admit cycles, or elevate acyclicity (\hyperref[ac]{AC}) to an axiom. Most of the results of this paper do {\it not} depend on acyclicity, but some of the most important results of section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} are developed and presented only in the acyclic case. Turning to the measure axiom, sections \hyperref[subsectioncausalmetric]{\ref{subsectioncausalmetric}} and \hyperref[subsectionintervalfiniteness]{\ref{subsectionintervalfiniteness}} include brief remarks regarding the possibility of weakening this axiom to admit consideration of {\it local causal structure} in the emergent definition of spacetime volume. In the locally finite case, this change is not so radical as to annul the basic insight that counting can serve as a proxy for the missing conformal factor in Malament's metric recovery theorem. Regarding the binary axiom, section \hyperref[sectionbinary]{\ref{sectionbinary}} provides ample illustrations of how this axiom {\it could} be altered; for example, by invoking holistic power spaces. However, as elaborated below, the primary attitude toward the binary axiom adopted in this paper is to {\it expand} the catalogue of structures that may be viewed in essentially the same way, culminating in the principle of iteration of structure (\hyperref[is]{IS}) in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}. Finally, there are excellent reasons to retain the axiom of countability (\hyperref[c]{C}), at least in the locally finite case. \refstepcounter{textlabels}\label{suggestedalt} {\bf Suggested Alternative Axioms.} The following set of axioms for classical discrete causal theory represents, in my judgment, a reasonable balance of relevance and generality, in light of the analysis presented in sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} of this paper. The inclusion of acyclicity as an axiom is a conservative, provisional choice, based on the lack of clear physical evidence for the existence of causal cycles. Due to this choice, the underlying set referenced in the following list of axioms is denoted by the letter $A$, for ``acyclic directed set." For comparison, the letter $C$, for ``causal set," is used in some of the same axioms in section \hyperref[subsectionaxioms]{\ref{subsectionaxioms}}. \hspace*{.4cm}B. \hspace*{.2cm}{\bf Binary Axiom}: {\it Classical spacetime may be modeled as a set $A$, whose elements \\ \hspace*{1cm} represent spacetime events, together with a binary relation $\prec$ on $A$, whose elements \\ \hspace*{1cm} represent causal relations between pairs of spacetime events.} \hspace*{.35cm}M$^*$. \hspace*{0cm}{\bf Measure Axiom}: {\it Elements of $A$ are assigned volume via a discrete measure \\ \hspace*{1.1cm} $\mu:\ms{P}(A)\rightarrow\mathbb{R}^+$, whose value depends only on local causal structure.} \hspace*{.35cm}C. \hspace*{.25cm}{\bf Countability}: {\it $A$ is countable.} \hspace*{.3cm}LF. \hspace*{.1cm}{\bf Local Finiteness}: {\it Every element of $A$ has a finite number of direct predecessors\\ \hspace*{1.1cm}and successors.} \hspace*{.3cm}AC. \hspace*{.05cm}{\bf Acyclicity}: {\it $(A,\prec)$ has no cycles.} Note the change to the measure axiom to admit consideration of local causal structure, as compared to the original measure axiom (\hyperref[m]{M}). Also observe that local finiteness is stated here in terms of {\it direct predecessors and successors,} rather than relation sets, since no multidirected structure is involved. Since the binary axiom is included in this list of axioms, the resulting alternative formulation of classical discrete causal theory is {\it not irreducibly holistic.} However, relation spaces, path spaces, and other important directed sets and multidirected sets whose elements do not correspond to spacetime events, remain crucially important in this context. \refstepcounter{textlabels}\label{conservativealt} {\bf Conservative Alternative.} The alternative list of axioms presented above is conservative, excluding both causal cycles and classical holism. However, it is worthwhile to consider an even more conservative theory, incorporating only those conclusions from sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}} of this paper {\it absolutely necessary} to repair the most serious structural shortcomings of causal set theory. This hyper-conservative approach permits only {\it irreducible} relations at the level of the causal preorder, thus obviating the need for the independence convention (\hyperref[ic]{IC}). Interval finiteness is also included, so that only relative acyclic directed sets over $\mathbb{Z}$ are considered. Local finiteness is {\it added,} to rule out pathologies such as the infinite bouquet. Many existing methods and results from causal set theory may be carried over, or adjusted, with a minimal amount of effort, under these axioms. The technical methods of relation space, path spaces, and kinematic schemes may then be brought to bear, without drastically altering the information content of causal set theory. \hspace*{.4cm}B. \hspace*{.25cm}{\bf Binary Axiom}: {\it Classical spacetime may be modeled as a set $A$, whose elements \\ \hspace*{1.1cm} represent spacetime events, together with a binary relation $\prec$ on $A$, whose elements \\ \hspace*{1.1cm} represent causal relations between pairs of spacetime events.} \hspace*{.4cm}M. \hspace*{.2cm}{\bf Measure Axiom}: {\it The volume of a spacetime region corresponding to a subset \\ \hspace*{1.1cm} $S$ of $A$ is equal to the cardinality of $S$ in fundamental units, up to Poisson-type \\ \hspace*{1.1cm} fluctuations.} \hspace*{.4cm}C. \hspace*{.25cm}{\bf Countability}: {\it $A$ is countable.} \hspace*{.25cm}IRR. \hspace*{0cm}{\bf Irreducibility}: {\it All relations between pairs of elements of $A$ are irreducible.} \hspace*{.35cm}IF. \hspace*{.18cm}{{\bf Interval Finiteness}: {\it For every pair of elements $x$ and $z$ in $A$, the open interval \\ \hspace*{1.1cm}$\llangle x,z\rrangle:=\{y\in A\hspace*{.1cm}|\hspace*{.1cm} x\prec y\prec z\}$ has finite cardinality.} \hspace*{.35cm}LF. \hspace*{.1cm}{\bf Local Finiteness}: {\it Every element of $A$ has a finite number of direct predecessors\\ \hspace*{1.1cm}and successors.} \hspace*{.35cm}AC. {\bf Acyclicity}: {\it $(A,\prec)$ has no cycles.} The three axioms of countability, acyclicity, and interval finiteness may be replaced by the single axiom that $(A,\prec)$ admits a morphism of finite index into the integers. \refstepcounter{textlabels}\label{radicalalt} {\bf Radical Alternatives.} Many, but not all, of the methods and results of this paper, also apply to more speculative approaches to classical discrete causal theory, involving either causal cycles, modification of the binary axiom to admit irreducible classical holism, or a combination of the two. Consideration of causal cycles has strong supporting precedent from general relativity. Perhaps the most reasonable alternative along these lines is to simply drop the axiom of acyclicity from the list of suggested alternative axioms above. Regarding classical holism, probably the simplest versions, at least from a conceptual standpoint, involve ``families of events influencing other families of events" in an irreducible fashion, a notion admitting straightforward formalization in terms of power spaces, as described in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}}. Modified versions of the measure axiom, countability, local finiteness, and acyclicity still make sense in this context. Alternatively, one might prefer to regard ``spacetime events" as merely a way of referencing some aspect of structure defined in other terms. Penrose's twistor theory provides classical precedent for this perspective. Grothendieck's relative viewpoint (\hyperref[rv]{RV}), Isham's topos theory, and other structural notions from modern mathematics provide many alternatives, besides those arising from causal theory itself. \refstepcounter{textlabels}\label{summaryperspectives} {\bf Summary of New Perspectives and Technical Methods.} The axiomatic changes suggested above serve to correct what I view as actual {\it deficiencies} in the current formulation of causal set theory. However, the potential exists to further improve discrete causal theory by adopting a broader structural viewpoint and introducing new technical methods. This paper includes a variety of suggestions along these lines. The principles of natural philosophy listed in section \hyperref[naturalphilosophy]{\ref{naturalphilosophy}} provide a general preamble to this effort. The guiding physical idea, throughout the paper, is the {\it causal metric hypothesis} (\hyperref[cmh]{CMH}), presented in section \hyperref[subsectioncausalmetric]{\ref{subsectioncausalmetric}}, which states that the properties of the physical universe are manifestations of causal structure. This idea itself is not new; in particular, Sorkin's phrase, {\it``order plus number equals geometry,"} represents a version of it. However, the choice to explicitly isolate and study the causal metric hypothesis in its own right, independently of any specific choice of mathematical model, enables a much clearer, more general, and more flexible approach, incorporating many new features. In particular, it facilitates the recognition of distinct classical (\hyperref[ccmh]{CCMH}) and quantum (\hyperref[qcmh]{QCMH}) versions of this hypothesis, representing different hierarchical aspects of discrete causal theory. A number of other basic structural concepts, transcending any specific choice of model, appear in the latter part of section \hyperref[sectionaxioms]{\ref{sectionaxioms}}. On an individual basis, these concepts are standard, or elementary, or both, but they may be combined and utilized in novel ways in the context of discrete causal theory. These include the {\it independence convention} (\hyperref[ic]{IC}), the {\it order extension principle} (\hyperref[oep]{OEP}), the principle of {\it hidden structure} (\hyperref[hs]{HS}), and Grothendieck's {\it relative viewpoint} (\hyperref[rv]{RV}). ``New physics" is introduced beginning in section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} of this paper. The first step is to reassign the fundamental role of the transitive causal order to a generally nontransitive binary relation called the {\it causal preorder,} which generates the causal order under the operation of transitive closure. The terms {\it order} and {\it preorder} are accorded nonstandard generalized meanings in this context, to allow for the possibility of causal cycles. The causal preorder provides a more nuanced view of fundamental causal structure than the causal order in the discrete setting. In particular, it accommodates the possibilities of {\it irreducibility} and {\it independence} of relations between pairs of events, features which are absent in the interpolative continuum setting where the transitive paradigm originated. Previous work of Finkelstein and Raptis involves structures formally equivalent to acyclic causal preorders, but with different physical interpretations. Among other improvements, recognition of the causal preorder rectifies the {\it Kleitman-Rothschild pathology,} in which configuration spaces of causal sets are dominated by members of negligible ``temporal size." The second major change of perspective introduced in this paper is a new viewpoint regarding {\it local structure,} presented in section \hyperref[sectioninterval]{\ref{sectioninterval}}. This viewpoint is much different than the interval-theoretic viewpoint prevailing in the current formulation of causal set theory, expressed, for example, in the causal set axiom of interval finiteness (\hyperref[if]{IF}). It is based on the concept of {\it causal locality,} under which a condition is taken to be local if and only if it may be checked by examining the {\it independent causes and effects associated with each event.} Again, there are formal precursors in the work of Finkelstein and Raptis. The foremost application of causal locality in section \hyperref[sectioninterval]{\ref{sectioninterval}} is the replacement of interval finiteness with the alternative axiom of {\it local finiteness} (\hyperref[lf]{LF}), as discussed in the summary of axiomatic analysis above. The {\it star topology,} introduced in section \hyperref[subsectiontopology]{\ref{subsectiontopology}}, provides a technical means to implement this new local viewpoint. The theory of {\it relative multidirected sets over a fixed base,} appearing in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}}, is the first major application of Grothendieck's relative viewpoint (\hyperref[rv]{RV}) appearing in this paper. The interplay of relative multidirected sets and the concept of causal locality provides an exceptionally clear picture of the structural limitations of causal sets, revealing how drastically the supposedly ``local" condition of interval finiteness restricts {\it global} structure. Section \hyperref[sectionbinary]{\ref{sectionbinary}} greatly amplifies the new perspectives developed in sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}} and \hyperref[sectioninterval]{\ref{sectioninterval}}. The structural theme of this section is to expand the viewpoint embodied in the binary axiom (\hyperref[b]{B}) of causal set theory, to encompass ``causal relationships" between pairs of structures more complex than individual events, following Grothendieck's relative viewpoint (\hyperref[rv]{RV}). The simplest example of this theme is the theory of {\it relation space}, explored in section \hyperref[subsectionrelation]{\ref{subsectionrelation}}, which studies ``causal relationships" between pairs of {\it relations} in a multidirected set. The global structure arising from all such relationships elevates this set of relations to the status of a directed set in its own right, called the {\it relation space} of the original multidirected set. This viewpoint leads to the solution of important technical problems in discrete causal theory, such as the {\it permeability of maximal antichains} in multidirected sets. Relation spaces have appeared previously in purely mathematical contexts as {\it line digraphs,} but do not seem to have been invoked in the study of fundamental spacetime structure. Comparison of the {\it relation space functor} and the {\it abstract element space functor} reveals the striking fact that ``typical" multidirected sets are {\it nearly interchangeable} with their relation spaces. In particular, the relation space functor {\it reduces multidirected structure to directed structure,} while {\it preserving information, except at the boundary.} Section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} introduces {\it power spaces,} which provide many more examples and applications of natural relationships between pairs of nontrivial structures in discrete causal theory. An example is the theory of {\it causal atoms,} already used in section \hyperref[relativeacyclicdirected]{\ref{relativeacyclicdirected}} to give a conceptually satisfying proof that causal sets are relative directed sets over the integers. Power spaces also provide a natural framework for the study of {\it classical holism} in the discrete causal context. {\it Causal path spaces,} introduced in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}}, are important generalizations of relation spaces, formalizing the flow of information or causal influence in a wide variety of contexts. Modern algebraic methods provide powerful techniques for studying causal path spaces. In particular, algebraic objects called {\it causal path semicategories} and {\it causal path algebras} play central roles in the general theory {\it path summation over a multidirected set,} introduced in section \hyperref[subsectionpathsummation]{\ref{subsectionpathsummation}}. Section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} brings these new perspectives and methods to bear on the quantum theory of spacetime. The crucial link between the classical and quantum realms in discrete causal theory is the principle of {\it iteration of structure} (\hyperref[is]{IS}), introduced in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}}, which permits effective implementation of the {\it histories approach to quantum theory} in the {\it background-independent context,} via path summation over multidirected sets. Iteration of structure formalizes the observation that configuration spaces of directed sets possess natural {\it higher-level multidirected structures,} collectively induced by their member sets, and by relationships between pairs of these sets. I refer to such relationships as {\it co-relative histories.} Of particular interest in discrete causal theory are {\it proper, full, originary co-relative histories,} which may be viewed as equivalence classes of special morphisms called {\it transitions,} representing the ``evolution" of one directed set into another. In section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}}, I carry out a discrete causal adaptation of the histories approach to quantum theory, motivated by Feynman's continuum path integral. This adaptation is {\it abstract,} in the sense that the multidirected sets involved are not {\it a priori} assigned any specific physical interpretation. The {\it path summation principle} (\hyperref[ps]{PS}), motivated by iteration of structure, states that this abstract approach applies to {\it both} the background-dependent theory of particles and fields on directed sets, and the background-independent theory of co-relative histories. In section \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}}, I apply this theory to derive {\it causal Schr\"{o}dinger-type equations,} which serve as dynamical laws for discrete quantum causal theory. The impermeability of maximal antichains in relation space implies that these equations adequately account for the information encoded in the underlying causal structure. Section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} introduces the theory of {\it kinematic schemes,} which are special configuration spaces of directed sets encoding ``evolutionary pathways" for all ``permissible histories." Kinematic schemes are endowed with higher-level multidirected structures supplied by classes of co-relative histories between pairs of their member sets, providing a prototypical example of iteration of structure (\hyperref[is]{IS}). They provide ``global formalizations" of the {\it evolutionary viewpoint} in background independent discrete causal theory, generalizing Sorkin and Rideout's theory of sequential growth dynamics for causal sets. The abstract quantum causal theory developed in sections \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} and \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} may be applied to kinematic schemes by isolating this higher-level multidirected structure and moving to relation space. The resulting Schr\"{o}dinger-type equations provide dynamical laws for discrete quantum spacetime. \subsection{Omitted Topics and Future Research Directions}\label{subsectionomitted} A number of topics of obvious interest in discrete causal theory are omitted from this paper, due both to length considerations and to my own inability to locate or develop certain results. In this section, I make remarks about a few of these topics, and give some hints regarding interesting directions for future research. \refstepcounter{textlabels}\label{remainingax} {\bf Further Remarks on Countability, Irreflexivity, and Acyclicity.} As stated in section \hyperref[subsectionalternative]{\ref{subsectionalternative}} above, I know of no compelling reasons to abandon the causal set axioms of countability (\hyperref[c]{C}) and irreflexivity (\hyperref[ir]{IR}), though of course the latter axiom loses much of its strength in the nontransitive context, no longer implying acyclicity (\hyperref[ac]{AC}), for example. Since this paper contains no significant analysis of these axioms, I give here a brief outline of what such analysis might entail. Deeper examination of the countability axiom might include the following: 1) explanation of how Grothendieck's relative viewpoint (\hyperref[rv]{RV}) applies to the topic of cardinality; this explains why cardinality conditions are less objectionable as global constraints than order-theoretic conditions; 2) elaboration of the distinguished role of {\it finite} directed sets in the theory of experimentation and measurement; 3) further investigation of {\it entropy issues,} such as the Kleitman-Rothschild pathology; 4) discussion of the {\it continuum hypothesis} and related set-theoretic and order-theoretic topics; 5) fuller treatment of {\it qualitative} structural differences between classes of directed sets and multidirected sets satisfying different cardinality conditions. \refstepcounter{textlabels}\label{cycles} Turning to irreflexivity and acyclicity, the choice to abstain from assumptions about causal cycles, even at the classical level, throughout much of sections \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}, \hyperref[sectioninterval]{\ref{sectioninterval}}, and \hyperref[sectionbinary]{\ref{sectionbinary}} of this paper, reflects my conviction that their status is far from a settled issue. General relativity admits many solutions involving continuum analogues of such cycles; i.e., closed timelike curves, and it seems premature to entirely dismiss these solutions as pathologies that may be expected to disappear in a suitable quantum theory of gravity. After all, general relativity is the only well-supported theory available which {\it could} shed any light on the status of classical causal cycles, since it is the only such theory that leaves their existence to be determined dynamically. The choice to rule out such cycles, by including acyclicity in the suggested list of alternative axioms in section \hyperref[subsectionalternative]{\ref{subsectionalternative}} above, is therefore provisional, and somewhat reluctant. The argument tipping the balance in favor of this choice is simply that it is best to explore simple alternatives first. However, it does go against my descriptive instincts. General {\it types} of objections to causal cycles include: 1) lack of observational evidence for their existence; 2) logical ``problems" such as the grandfather paradox; 3) the fact that certain existing theories ``conspire in unforeseen ways to avoid cycles," an example being the {\it no-cloning theorem} in quantum information theory; and 4) technical difficulties arising in theories permitting cycles, such as those discussed in the context of path summation in section \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} of this paper. The strong form of the causal metric hypothesis (\hyperref[cmh]{CMH}) completely resolves the second type of objection, since it leads to theories with perfect background independence. The other types of objections, however, remain relevant in causal theory. Arguments in favor of considering cycles in discrete causal theory include the following: 1$^*$) general relativity permits closed timelike curves; 2$^*$) regardless of their classical status, cycles may be unavoidable at higher levels of hierarchy, appearing, for example, in the abstract structure of kinematic schemes of countable acyclic directed sets; 3$^*$) abstract element spaces over acyclic multidirected sets generally include cycles; 4$^*$) cycles are generally nonlocal and therefore do not generally present any local pathology; 5$^*$) cycles are reasonable candidates for {\it subclassical structure;} i.e., ``internal structure of classical elements." \refstepcounter{textlabels}\label{covariance} {\bf Covariance.} Whenever a physical theory involves arbitrary choices, it must provide mechanisms to ensure that its substance remains unchanged if these choices are made otherwise. This is the essence of the principle of {\it covariance} in physical law. In the classical continuum context, covariance is closely related to the arbitrary choice of a coordinate system; for example, dynamical laws expressible as tensor equations are often called ``covariant," because their abstract forms are not coordinate-dependent. In causal theory, this type of covariance generalizes in a straightforward manner in terms of generalized frames of reference for directed sets, represented by generalized order extensions. The prototypical example of this idea appears in Sorkin and Rideout's theory of sequential growth dynamics for causal sets \cite{SorkinSequentialGrowthDynamics99}, as the principle of {\it discrete general covariance.} The purpose of this principle is to capture the notion that ``all evolutionary accounts of a causal set are equally valid." At a formal level, the arbitrary choice associated with discrete general covariance may be viewed as a {\it particular choice of kinematics in a particular kinematic scheme;} in this case, Sorkin and Rideout's kinematic scheme $(\ms{C}_{\tn{\fsz{fin}}},\mc{H}_1)$. However, a {\it prior} arbitrary choice, in the context of background-independent quantum causal theory, is the choice of an appropriate class of co-relative histories, necessary to {\it define} a kinematic scheme. Indeed, as mentioned in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} of this paper, the object class $\ms{K}$ of a kinematic scheme $(\ms{K},\mc{H})$ represents a hypothesis about the structure of the physical universe, and should therefore be uniquely determined, but the class of co-relative histories $\mc{H}$ is selected for convenience; i.e., arbitrarily, from among many equally valid choices. The meaning of covariance in this context is that the essential physical properties of the theory should not depend on the choice of $\mc{H}$. In practical terms, this means that {\it if one changes kinematic schemes, one must change phase maps to compensate.} \refstepcounter{textlabels}\label{algstructure} {\bf Algebraic Structure and Hierarchy.} The analogy between kinematic schemes and categories prompts a closer examination of their similarities and differences. This, in turn, leads to a broader consideration of algebraic structure in causal theory, and in theoretical physics more generally, with a particular emphasis on questions of {\it hierarchy}. Isham's topos-theoretic framework for physics provides valuable and wide-ranging contributions in this direction. The principle of iteration of structure (\hyperref[is]{IS}), introduced in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} of this paper, raises the issue of hierarchy explicitly, but in a limited context, arising more or less directly from the ordinary superposition principle of quantum theory. However, much more general notions of hierarchy are hinted at throughout sections \hyperref[sectionbinary]{\ref{sectionbinary}} and \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}, particularly in the contexts of relation spaces, causal path spaces, and kinematic spaces. One may discern at least four distinct ``vertical" levels of hierarchy, represented, for example, by elements, directed sets, kinematic schemes, and kinematic spaces. An {\it infinite} number of distinct ``horizontal" levels of hierarchy are readily apparent, represented, for example, by spaces of paths of various lengths or isomorphism types in directed or multidirected sets. A deeper examination of algebraic structure and hierarchy in discrete causal theory might include the following: 1) more detailed comparison of causal theory and category theory, particularly topos theory; 2) more adequate identification and explanation of the relevance and utility of various category-theoretic properties in physical contexts; 3) general discussion of {\it local, global, vertical,} and {\it horizontal} hierarchy; 4) the use of {\it binary trees} to encode algebraic hierarchy; 5) detailed comparison of categories with operations, such as {\it monoidal categories,} and analogous structures in discrete causal theory; 6) systematic examination of causal analogues of the theory of {\it categorification} and {\it decategorification,} generalizing the relationship between a kinematic scheme and its underlying multidirected set; 7) discussion of possible physical interpretations of additional levels of vertical hierarchy, such as {\it subclassical} and {\it hyperquantum} theories; 8) analysis of the fate of Lie representation theory in theoretical physics under the causal metric hypothesis (\hyperref[cmh]{CMH}), including bifurcation into {\it group-theoretic} and {\it non-group-theoretic} aspects in the theory of ``external" symmetries. \refstepcounter{textlabels}\label{phasetheory} {\bf Phase Theory.} When confronted with a causal Schr\"{o}dinger-type equation such as equation \hyperref[causalschrodinger]{\ref{causalschrodinger}}: \[\psi_{R;\theta}^-(r)=\theta(r)\sum_{r^-\prec r}\psi_{R;\theta}^-(r^-),\] the most basic question is the following: \hspace*{.3cm}{\bf Question:} {\it How is the relation function $\theta$ generating the phase map determined, and what is \\ \hspace*{.3cm}its target object?} More general versions of this question arise for arbitrary phase maps, not necessarily generated by relation functions. To investigate these questions, it is natural to begin by considering Feynman's phase map $\Theta$ in the continuum context, appearing in equation \hyperref[feynmanphase]{\ref{feynmanphase}}: \[\Theta(\gamma)=e^{\frac{i}{\hbar}\ms{S}(\gamma)}.\] Feynman's phase map depends on the classical action $\ms{S}$, which in turn depends on the Lagrangian. Hence, conversion of equation \hyperref[causalschrodinger]{\ref{causalschrodinger}}, or a more general analogue of this equation, into a form that can actually be used to make predictions, involves identification of a suitable ``action" or ``Lagrangian" for a suitable class of directed sets or kinematic schemes of directed sets. Interesting work has already been done in this direction in the special case of causal sets, of which I have mentioned just one example: the recent paper {\it Scalar Curvature of a Causal Set} \cite{DowkerScalarCurvature10}, by Dionigi Benincasa and Fay Dowker. I have not attempted to adapt or generalize these efforts here. However, it is possible to make a few general remarks on elementary mathematical and information-theoretic grounds. Feynman's phase map takes values in the unit circle $S^1$, viewed as a subset of the complex numbers $\mathbb{C}$. These values are summed in $\mathbb{C}$ to yield quantum amplitudes, modulo appropriate constants of proportionality, which are then translated into real-valued probabilities. $S^1$ and $\mathbb{C}$ are natural objects in the continuum setting: both admit continuum manifold structures, and $\mathbb{C}$ is the algebraic closure of $\mathbb{R}$. In discrete causal theory, however, one would expect objects such as $S^1$ and $\mathbb{C}$ to somehow {\it emerge as continuum limits of more primitive target objects.} For example, working with a kinematic scheme over the class $\ms{A}_{\tn{\fsz{fin}}}$ of finite acyclic directed sets, it would be reasonable to expect phases to belong to some {\it finite algebraic object.} Speculating further, different phases might have different natural targets, perhaps organized into a {\it filtered system,} whose category-theoretic limit might resemble a countable subset of $\mathbb{C}$. {\it Counting} or {\it entropy} considerations may be expected to contribute in this context. Covariance also plays a role; {\it a priori,} different kinematic schemes require families of targets. Finally, as suggested in section \hyperref[subsectionkinematicschemes]{\ref{subsectionkinematicschemes}} of this paper, it may be preferable in the background-independent context to view phase maps as {\it generalized functors,} and path sums as {\it generalized limits.} A basic constraint on all these considerations is that one must eventually {\it interpret} the resulting generalized quantum amplitudes physically, in terms of probabilities or some generalization thereof. \refstepcounter{textlabels}\label{randomgraph} {\bf Random Graph Dynamics.} In the late 1950's, mathematicians Paul Erd\"{o}s and Alfr\'{e}d R\'{e}nyi introduced the theory of {\it random graphs} in a long series of papers, beginning with \cite{ErdosRenyi59}. A relatively modern reference on the subject is {\it Random Graphs} \cite{Bollobas01} by B\'{e}la Bollob\'{a}s, one of the leaders in the field. The original context of this work was pure graph theory, with information-theoretic and physical applications appearing only afterwards. A simple example of an {\it undirected} random graph may be constructed via the following process: beginning with a finite vertex set $V$ of cardinality $N$, consider each pair of elements $(x,y)$ of $V$, in some order, adding an edge between $x$ and $y$ with some {\it fixed} probability $p$. When every pair of vertices has been considered, the process stops, and the result is a {\bf random graph} of type $(N,p)$. The evolutionary viewpoint in discrete causal theory, epitomized by Sorkin and Rideout's theory of sequential growth dynamics for causal sets \cite{SorkinSequentialGrowthDynamics99}, is closely related to random graph dynamics. Sorkin and Rideout explicitly mention this connection in their 2000 paper {\it Evidence of a Continuum Limit in Causal Set Dynamics} \cite{SorkinEvidenceofContLimVer208}. Transitions, co-relative histories, and kinematic schemes, discussed in section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}} of this paper, represent vastly generalized versions of random graph processes. A number of complicating factors distinguish these physically-oriented constructions from the simple models studied by Erd\"{o}s and R\'{e}nyi. For example, the graphs involved are directed, their vertex sets have no fixed cardinality, more complex structures than single edges are added at each step of the process, and the ``probabilities" at each step depend on the previous steps. \refstepcounter{textlabels}\label{phasetrans} One reason why the analogy between random graph dynamics and the evolutionary viewpoint in causal theory is interesting is because {\it peculiar and physically suggestive phenomena} occur in random graph dynamics, phenomena that no one seems to have even guessed at prior to Erd\"{o}s and R\'{e}nyi's papers. One class of phenomena of particular interest is graph-dynamical {\it phase transitions.}\footnotemark\footnotetext{Use of the term {\it phase} here is entirely distinct from its use in the context of {\it phase maps.}} In very general terms, a {\bf phase transition} is an abrupt change in some qualitative property of a class of random graphs that occurs at some critical value of some associated parameter. For example, consider the class of random graphs of type $(N,p)$, mentioned above, and let $N$ and $p$ vary in such a way that $p$ is a function of $N$. One may then consider the limit of the product $Np$ as $N$ approaches infinity; in particular, one may consider varying $p$ in such a way that the limit varies from less than $1$ to greater than $\log N$. A sequence of phase transitions is then observed: for sufficiently small values of $p$, the largest connected component of a typical random graph of type $(N,p)$ has approximately $\log N$ vertices or fewer. As the limit of $Np$ reaches the critical value of $1$, a single {\it giant component} with roughly $N^{\frac{2}{3}}$ vertices suddenly emerges, and rapidly increases in size until it contains an appreciable fraction of all the vertices, while still leaving many small components and isolated vertices. This qualitative picture remains unchanged until the limit passes another critical value at $\log N$, at which point the typical graph suddenly becomes completely connected. Such striking behavior raises the possibility that certain qualitative properties of spacetime might be attributed to graph-dynamical processes in the context of discrete causal theory. For example, one of the principal arguments in favor of the hypothesis of {\it cosmic inflation} in the early universe is the uniformity of the {\it cosmic microwave background,} which, following a conventional interpretation, suggests the likelihood that regions of spacetime with large separation were causally connected in the past. A possible alternative to the inflationary hypothesis is the idea that {\it causal structure abruptly became sparser} in some sense, possibly due to a phase transition. \refstepcounter{textlabels}\label{holalt} {\bf Alternatives to Power Spaces.} Power spaces are only one of many ways to generalize binary relations. For example, the causal path spaces studied in section \hyperref[subsectionpathspaces]{\ref{subsectionpathspaces}} of this paper are not power spaces, since the morphism defining a path is generally not information-theoretically reducible to its image. However, even if one restricts consideration to {\it subobjects of individual structured sets,} rather than incorporating information about relationships between structured sets, a variety of alternatives remain. In the theory of power spaces, the basic ``binary" device of relations between {\it pairs} is retained; it is the {\it members} of these pairs that are allowed greater complexity. An alternative viewpoint is to exchange pairs for {\it $N$-tuples.} For example, a {\it ternary relation} on a set $S$ is a subset of the third Cartesian power $S\times S\times S$; its elements are {\it ordered triples}. Similarly, an {\it $N$-ary relation} is a subset of the $N$th Cartesian power $S^N$. A ``multidirected version" of a set equipped with an $N$-ary relation is an $(N+2)$-tuple of the form $S=(S,R,e_1,...,e_N)$, where $S$ and $R$ are sets, and where the maps $e_i:R\rightarrow S$ are ``generalized initial and terminal element maps." In this context, an element of $R$ may be viewed as an $(N-1)$-dimensional {\it simplex,} and the total structure of $S$ may be viewed as an $(N-1)$-dimensional {\it simplicial complex}. One may also allow $N$ to vary, thereby obtaining simplicial complexes of nonconstant dimension. In either case, random graph theory is subsumed by {\it random simplicial theory,} and mathematical problems such as {\it Whitehead's conjecture} come into play. The associated physical theory may be viewed as a generalization of simplicial power spaces, discussed in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} in the context of higher induced relations. Another way to generalize binary relations is to consider {\it multiple independent} binary or $N$-ary relations on a single set. This may be accomplished in the context of multidirected sets, or $N$-ary analogues thereof, by adding {\it colors} to elements of relation sets or generalized relation sets. One obvious motivation for such a construction in the discrete context would be the belief that the causal metric hypothesis (\hyperref[cmh]{CMH}), or at least the strong form of this hypothesis, is {\it false.} For example, one might regard {\it spatial} as well as causal structure as fundamental. One might then invoke {\it two} relations: a {\it symmetric} relation encoding spatial structure, and an {\it antisymmetric} relation encoding causal structure. Two relatively familiar theories involving combinations of spatial and temporal structure on discrete manifolds at the fundamental scale are {\it causal dynamical triangulations} and {\it loop quantum gravity.} At a broad conceptual level, there is at least a partial {\it duality} involving spatial/temporal structure and symmetric/antisymmetric binary relations in physics, observed already in relativity. The extent and significance of this duality remains unsettled. In particular, one finds ``time-first" theories, such as causal set theory, ``space-first" theories, such as shape dynamics, and hybrid theories, such as causal dynamical triangulations, all competing for the same ground. It is conceivable that the basic approaches represented by these theories are to some degree interchangeable. Another possible application of multiple binary or $N$-ary relations might be to encode {\it different types of spatial or temporal structure.} For example, a pair of independent antisymmetric relations on a set might be interpreted as a {\it pair of timelike dimensions.} More generally, power spaces, $N$-ary relations, and colors may be combined in various ways as a ``theme and variations." If nothing else, this creates fertile ground for speculation. For example, one might invoke a whole family of independent relations, with ``time" emerging as a ``residual drift." If history is any judge, however, the simplest nontrivial versions of ideas are usually the most important. \refstepcounter{textlabels}\label{othertheories} {\bf Connections with Other Physical Theories.} Interesting analogies and potential connections exist between the version of discrete causal theory I have developed in this paper and a variety of other approaches to physics at the fundamental scale. Although some of these connections have been partially explored for existing versions of discrete causal theory, the new perspectives presented here might justify reconsideration. These theories include: \begin{enumerate} \item {\bf Loop Quantum Gravity}. Analogies involving spin networks and spin foams might be examined in detail. \item {\bf Causal Dynamical Triangulations}. This theory {\it can} be expressed in terms of special types of directed sets, and methods described in this paper {\it can} be applied, setting aside questions of actual suitability. Significant work has been done on the histories approach to quantum theory for causal dynamical triangulations, and it would be interesting to compare this work to the approach presented here. \newpage \item {\bf Domain Theory.} An interesting general project with potential applications in domain theory would be to examine which of the ideas presented in this paper admit ``limiting versions" in the interpolative context. \item {\bf Physical Applications of Category Theory.} Connections are cited throughout the paper. \item {\bf Quantum Information Theory}. Quantum circuits may be modeled as directed sets equipped with matrix-valued phase maps. This leads to an interesting approach to quantum information theory involving ideas of a {\it relativistic} flavor, such as frames of reference, relativity of simultaneity, and so on. Na\"{i}vely, one might hope to turn this relationship around and use quantum computers as ``virtual Planck-scale laboratories." \item {\bf Quantum Automaton Theory.} G. M. D'Ariano and Paolo Perinotti \cite{DArianoDirac13} have recently recovered the {\it Dirac equation,} the prototypical ``relativistic Schr\"{o}dinger-type equation," via this approach. It would be interesting to compare their approach to the method of deriving causal Schr\"{o}dinger-type equations presented in this paper. \item {\bf Physical Applications of Noncommutative Geometry}. Noncommutative geometry gives a completely different geometric perspective regarding causal path algebras than that provided by the causal metric hypothesis (\hyperref[cmh]{CMH}). Ioannis Raptis and Roman Zapatrin discuss related ideas in their paper {\it Algebraic description of spacetime foam} \cite{RaptisSpacetimeFoam01}. There are also some relevant physical notions underlying the theory {\it deformed special relativity} that involve noncommutative geometry, in spite of the general lack of success of this theory to date. \item {\bf Twistor Theory}. Analogies mentioned in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} of this paper might be further explored. \item {\bf Entropic Gravity}. Causal theory is {\it at least} a theory of gravity, and entropic considerations are of interest in the context of phase theory. \item {\bf Shape Dynamics}. As mentioned in section \hyperref[subsectionpowerset]{\ref{subsectionpowerset}} of this paper, the viewpoint of this ``space-first" theory is nearly opposite to the ``time-first" approach of causal theory. On occasion, such ``opposite approaches" are related by duality theories, and it would be interesting to determine if such a relationship exists in this case. \end{enumerate} More tenuous ``connections" with additional physical theories may also be exhibited. For example, it is mathematically tempting to ``promote" directed sets to {\it tubular diagrams,} to obtain a ``background-independent causal theory of interacting closed spatial loops." Similar diagrams appear in string theory and topological quantum field theory. In this context, the direct future $D_0^+(x)$ of an element $x$ might be replaced by a {\it cobordism of $S^1$ to a disjoint union of copies of $S^1$}. Beyond this, however, the analogy seems to break down. For example, the status of individual relations is unclear, and a natural analogue of iteration of structure (\hyperref[is]{IS}) is not evident. \refstepcounter{textlabels}\label{othermath} {\bf Connections with Mathematical Topics.} Although this paper touches on many different mathematical fields, there is much more to be said about many of these relationships. In addition, there are other potentially fruitful connections deserving at least brief mention. \begin{enumerate} \item {\bf Graph Theory.} Numerous connections are cited throughout the paper, but many more exist. \item {\bf Category Theory.} Many additional connections exist besides those involving existing physical applications. \item {\bf Noncommutative Algebra; Noncommutative Geometry; ``Quantum Groups."} Again, there are additional relevant mathematical ideas not specifically associated with any existing physical theory. \newpage \item {\bf Matroid Theory.} As mentioned in section \hyperref[subsectionchains]{\ref{subsectionchains}} of this paper, matroid theory involves a notion of {\it independence} similar to linear independence of vectors, and distinct from the notion of independence specified by the independence convention (\hyperref[ic]{IC}). Its potential interest in the context of discrete causal theory stems from the existence of many interesting and physically suggestive matroids arising in graph theory. For example, {\it gammoids} are obviously relevant. \item {\bf Numerical Methods.} Usually appearing in the context of {\it approximation} in the fields of mathematical analysis and partial differential equations, such methods acquire new algebraic and arithmetic connotations when they are considered {\it exact,} as in the derivation of causal Schr\"{o}dinger-type equations. Particularly relevant are {\it finite element methods} involving {\it variational techniques.} \item {\bf Cohomology Theories.} Algebra cohomology theories such as {\it Hochschild cohomology} and {\it cyclic cohomology} are relevant to the study of causal path algebras and phase theory. There are results to proven concerning relationships between such cohomology theories and corresponding theories involving more primitive objects such as semicategories.\footnotemark\footnotetext{James Madden \cite{MaddenPrivate13} is ultimately responsible for making me aware of these connections.} Relative cohomology plays a role via congruences. Lie cohomology and algebraic $K$-theory also enter the picture. \item {\bf Partial Differential Equations, Integral Equations, Boundary Value Problems.} There is a huge and sophisticated existing literature on Schr\"{o}dinger-type partial differential equations. The concepts involved in sections \hyperref[subsectionquantumpathsummation]{\ref{subsectionquantumpathsummation}} and \hyperref[subsectionschrodinger]{\ref{subsectionschrodinger}} of this paper, involving ``flux through a boundary," the relationship between local (i.e., ``differential") and global (i.e., ``integral") equations, generalized Lagrangians, the superposition principle, etc., are all central to this field. \item {\bf Random Topology.} This may be viewed as a generalization of random graph dynamics. \item {\bf Representation Theory.} Particularly important in modern physical continuum theories, especially quantum field theories, is the representation theory of Lie groups. If continuum geometry is shown to break down, and if the causal metric hypothesis (\hyperref[cmh]{CMH}) withstands scrutiny, generalized order-theoretic analogues of Lie representation theory may eventually take over part of this workload. Other connections to representation theory arise via the theory of semicategory algebras, by abstracting the role of groups algebras in group representation theory. \item {\bf Galois Theory.} The brief remarks about causal Galois groups and Galois connections in section \hyperref[subsectionquantumprelim]{\ref{subsectionquantumprelim}} of this paper may be greatly elaborated. \end{enumerate} \subsection{Acknowledgements; Personal Notes}\label{subsectionacknowledgements} \refstepcounter{textlabels}\label{ack} {\bf Acknowledgements.} Rafael Sorkin was kind enough to provide detailed answers to several queries about foundational issues, and to allow me to quote personal correspondence. David Finkelstein answered questions about his papers on causal nets, and sent me some of his unpublished work. David Rideout supplied informative answers to multiple questions about causal sets. Jorge Pullin provided background on general relativity and cosmology, and gave me some valuable opinions about various approaches to quantum gravity. Jimmie Lawson provided useful advice, both general and specific, and contributed significantly to my understanding of related topics in quantum information theory. George Ellis made encouraging remarks about my suggestion to combine causal theory with classical holism via power spaces. Manfred Droste wrote me a very nice response to a question about universal causal sets. Giacomo Mauro D'Ariano provided useful information from the related viewpoint of quantum automata, as well as general encouragement. James Madden supplied important underlying ideas about algebraic notions involving multiplicative structures more primitive than groups, as well as order-theoretic background. Brendan McKay and James Oxley provided helpful advice on graph-theoretic issues. Marcel Ern\'{e} was kind enough to scan and send me copies of his unpublished notes on order theory. Joel David Hamkins provided some helpful answers via the MathOverflow forum. My junior colleague Thomas Naugle, and my sister Stephanie Dribus, endured much tedious speculation on my part and made helpful suggestions. Steph also helped me check some of the proofs,\footnotemark\footnotetext{Remaining errors should be attributed to her.} and Tommy pointed out some very useful references. Al Vitter, as my first algebraic geometry teacher, acquainted me with the work and philosophy of Grothendieck. Jerome W. Hoffman, as my thesis advisor, provided much of my general mathematical background, via years of advice and hundreds of references on algebraic geometry, algebraic $K$-theory, and category theory; it may surprise him to know how many I actually read! Finally, my sister Marian Dribus, and my mother Virginia Dribus, provided invaluable editing assistance. \refstepcounter{textlabels}\label{personal} {\bf Personal Notes.} I began to develop the ideas behind this paper in 2009, with little knowledge of existing causal theory beyond a superficial awareness of domains, acquired during my presence at Tulane University when Keye Martin was on the faculty. While no serious researcher would deliberately abstain from reading existing results relevant to a project, it is perhaps not such a disadvantage to devote considerable thought to a subject in a state of accidental ignorance, provided that one is willing afterwards to read thousands of pages of literature with an open mind, despite the disappointment of repeatedly discovering that cherished ideas are either already known, or are wrong. My first interaction with the ``outside world" on the subject of discrete causal theory came in June 2010, when I emailed some questions about the current state of causal set theory to a leading exponent of the field. In January of 2011, I suggested to him the basic premises of sections 3 and 4 of this paper, but never learned his view of the matter. In August 2012, I outlined many of the principal ideas presented in this paper in an essay \cite{DribusFoundations12} for the Foundational Questions Institute (FQXi), which was one of the winners in their essay contest ``Questioning the Foundations." \newpage \begin{appendix} \section{Index of Notation}\label{appendixindex} The following index gives a section-by-section listing of the notation used in this paper. Some duplication of symbols was unavoidable; for example, the mathscript symbol $\ms{C}$ means an arbitrary category in one context, but the specific category of causal sets in another. The Greek letter $\Gamma$ stands for a number of different sets or spaces of paths or chains. To make this index easier to use on a section-by-section basis, some definitions are repeated. My hope is that the notation in the paper is at least {\it locally} unambiguous. \vspace*{.3cm} {\bf Section \hyperref[sectionaxioms]{\ref{sectionintroduction}}:} \begin{multicols}{2} \begin{itemize} \item $C$ \hspace*{2.18cm} causal set \item $\prec$ \hspace*{2.25cm} binary relation on $C$ \item $x,y$ \hspace*{1.9cm} spacetime events or\\ \hspace*{2.6cm} elements of a causal set \item $\mu$ \hspace*{2.2cm} discrete measure on $C$ \item $S$ \hspace*{2.2cm} subset of $C$ \item $X,X'$ \hspace*{1.5cm} pseudo-Riemannian \\ \hspace*{2.6cm} manifolds \item $g,g'$ \hspace*{1.8cm} pseudo-Riemannian \\ \hspace*{2.6cm} metrics \item $\ms{K}$ \hspace*{2.1cm} Sorkin and Rideout's \\ \hspace*{2.55cm} kinematic scheme \\ \hspace*{2.55cm} for sequential growth \item $\tau:C_i\rightarrow C_t$ \hspace*{.65cm} transition \item $\theta(\tau)$ \hspace*{1.7cm} transition weight or phase \item $\{x_0,x_1,x_2,...\}$ \hspace*{.1cm} total labeling \\ \hspace*{2.6cm} of a causal set \item $C,A,D,M$ \hspace*{.7cm} structured sets \\ \hspace*{2.6cm} (causal, acyclic directed, \\ \hspace*{2.6cm} directed, multidirected) \item $\ms{C},\ms{A},\ms{D},\ms{M}$ \hspace*{.75cm} categories of \\ \hspace*{2.6cm} structured sets \item $\mathbb{N},\mathbb{Z},\mathbb{Q},\mathbb{R},\mathbb{C}$ \hspace*{.45cm} nonnegative integers, \\ \hspace*{2.6cm} integers, rationals, \\ \hspace*{2.6cm} reals, complex numbers \item $\mathbb{Z}_n$ \hspace*{1.98cm} integers modulo $n$ \end{itemize} \end{multicols} {\bf Section \hyperref[sectionaxioms]{\ref{sectionaxioms}}:} \begin{multicols}{2} \begin{itemize} \item CMH \hspace*{1.55cm} causal metric hypothesis \item CCMH \hspace*{1.25cm} classical causal metric \\ \hspace*{2.6cm} hypothesis \item $C$ \hspace*{2.15cm} causal set \item B \hspace*{2.2cm} binary axiom \item M \hspace*{2.15cm} measure axiom \item C \hspace*{2.2cm} (axiom of) countability \item TR \hspace*{1.9cm} (axiom of) transitivity \item $\llangle w,y\rrangle$ \hspace*{1.4cm} open interval in \\ \hspace*{2.6cm} a directed or \\ \hspace*{2.6cm} multidirected set \item IF \hspace*{2.05cm} (axiom of) interval \\ \hspace*{2.6cm} finiteness \item IR \hspace*{2.05cm} (axiom of) irreflexivity \item $\mu$ \hspace*{2.2cm} discrete measure on $C$ \item $\ms{P}(C)$ \hspace*{1.6cm} power set of $C$ \item $\mathbb{R}^+$ \hspace*{2cm} set of positive \\ \hspace*{2.6cm} real numbers \item $(C,\prec)$ \hspace*{1.5cm} causal set with \\ \hspace*{2.6cm} explicit binary relation \item $S$ \hspace*{2.2cm} subset of a causal set \item $\ms{C}$ \hspace*{2.2cm} category of causal sets \item $\phi$ \hspace*{2.2cm} morphism of causal sets, \\ \hspace*{2.6cm} acyclic directed sets, \\ \hspace*{2.6cm} or directed sets \item $\phi^{-1}(x')$ \hspace*{1.2cm} fiber of $\phi$ over $x'$ \item $\preceq$ \hspace*{2.25cm} partial order \newpage \item $(P,\preceq)$ \hspace*{1.5cm} partially ordered \\ \hspace*{2.6cm} set with explicit \\ \hspace*{2.6cm} binary relation \item $A$ \hspace*{2.2cm} acyclic directed set \item AC \hspace*{1.9cm} (axiom of) acyclicity \item $(A,\prec)$ \hspace*{1.45cm} acyclic directed set with \\ \hspace*{2.6cm} explicit binary relation \item $\ms{A}$ \hspace*{2.15cm} category of acyclic\\ \hspace*{2.6cm} directed sets \item $D$ \hspace*{2.2cm} directed set \item $(D,\prec)$ \hspace*{1.45cm} directed set with \\ \hspace*{2.6cm} explicit binary relation \item $\ms{D}$ \hspace*{2.15cm} category of \\ \hspace*{2.6cm} directed sets \item $M$ \hspace*{2.05cm} multidirected set \item $R$ \hspace*{2.15cm} relation set of a \\ \hspace*{2.6cm} multidirected set \item $i,t$ \hspace*{2cm} initial and terminal \\ \hspace*{2.6cm} element maps of a\\\hspace*{2.6cm} multidirected set \item $x\prec y$ \hspace*{1.6cm} means there exists \\ \hspace*{2.6cm} $r\in R$ such that \\ \hspace*{2.6cm} $i(r)=x$ and $t(r)=y$ \item $(M,R,i,t)$ \hspace*{.65cm} multidirected set \\ \hspace*{2.6cm} expressed as an \\ \hspace*{2.6cm} ordered quadruple \item $\ms{M}$ \hspace*{2.1cm} category of \\ \hspace*{2.6cm} multidirected sets \item $\phi=(\phi_{\tn{\fsz elt}},\phi_{\tn{\fsz rel}})$ \hspace*{.1cm} morphism of \\ \hspace*{2.6cm} multidirected sets \item $\phi_{\tn{\fsz elt}}$ \hspace*{1.85cm} map of elements \item $\phi_{\tn{\fsz rel}}$ \hspace*{1.85cm} map of relations \item $\gamma$ \hspace*{2.2cm} a chain in $M$ \item $\tn{Ch}(M)$ \hspace*{1.2cm} chain set of $M$ \item $\tn{Ch}_n(M)$ \hspace*{1.05cm} set of chains of \\ \hspace*{2.6cm} length $n$ in $M$ \item $\sigma$ \hspace*{2.2cm} an antichain in $M$ \item $V$ \hspace*{2.15cm} a vector space \item $S$ \hspace*{2.2cm} linearly independent \\ \hspace*{2.6cm} subset of $V$ \item $\Gamma, \Gamma'$ \hspace*{1.75cm} sets of chains in\\ \hspace*{2.6cm} definition \hyperref[definitionindependence]{\ref{definitionindependence}} \item IC \hspace*{2.05cm} independence convention \item $J(x)$ \hspace*{1.55cm} total domain of influence \\\hspace*{2.6cm} of a spacetime event $x$ \item $I^\pm(x)$ \hspace*{1.5cm} chronological past and\\ \hspace*{2.6cm} future of a spacetime \\\hspace*{2.6cm} event $x$ \item $J^\pm(x)$ \hspace*{1.5cm} causal past and future\\ \hspace*{2.6cm} of a spacetime event $x$ \item $D(x)$ \hspace*{1.58cm} total domain of influence \\\hspace*{2.6cm} of an element $x$ in\\\hspace*{2.6cm} a multidirected set \item $D^\pm(x)$ \hspace*{1.35cm} past and future of $x$ \item $D_0^\pm(x)$ \hspace*{1.37cm} direct past and future of $x$ \item $D_{\tn{\fsz{max}}}^-(x),$ \hspace*{.88cm} maximal past and minimal \\ $D_{\tn{\fsz{min}}}^+(x)$\hspace*{1.2cm} future of $x$ \item $\partial M$ \hspace*{1.8cm} boundary of $M$ \item $\tn{Int}(M)$ \hspace*{1.2cm} interior of $M$ \item $\ms{A}_{\tn{\fsz fin}}$ \hspace*{1.75cm} class of isomorphism\\ \hspace*{2.6cm} classes of finite \\ \hspace*{2.6cm} acyclic directed sets \item $\aleph_0$ \hspace*{2.05cm} countably infinite \\ \hspace*{2.6cm} cardinal \item $\ms{A}_{\aleph_0}$ \hspace*{1.8cm} class of isomorphism\\ \hspace*{2.6cm} classes of countable\\ \hspace*{2.6cm} acyclic directed sets \item $\ms{C}_{\tn{LF}}$ \hspace*{1.83cm} class of isomorphism\\ \hspace*{2.6cm} classes of locally\\ \hspace*{2.6cm} finite causal sets \item $\ms{A}_{\aleph_0,\tn{LF}}$ \hspace*{1.35cm} class of isomorphism\\ \hspace*{2.6cm} classes of countable locally \\ \hspace*{2.6cm} finite acyclic directed sets \item $S_{\aleph_0}$ \hspace*{1.9cm} fixed countably infinite set \item $2^{\aleph_0}$ \hspace*{1.95cm} cardinality of $\ms{P}(S_{\aleph_0})$ \item OEP \hspace*{1.65cm} order extension principle \item $\ms{C}$ \hspace*{2.3cm} a category; e.g., causal sets \item $\tn{Obj}(\ms{C})$ \hspace*{1.25cm} objects of $\ms{C}$ \item $\tn{Mor}(\ms{C})$ \hspace*{1.2cm} morphisms of $\ms{C}$ \item $\mbf{i},\mbf{t}$ \hspace*{2cm} initial and terminal \\ \hspace*{2.6cm} object maps of a \\ \hspace*{2.6cm} category \item $\gamma_C$ \hspace*{2.05cm} identity morphism of $C$ \item RV \hspace*{1.95cm} Grothendieck's relative \\ \hspace*{2.6cm} viewpoint \item HS \hspace*{1.9cm} (principle of) hidden \\ \hspace*{2.6cm} structure \end{itemize} \end{multicols} {\bf Section \hyperref[sectiontransitivity]{\ref{sectiontransitivity}}:} \begin{multicols}{2} \begin{itemize} \item $\mbf{C}$ \hspace*{2.15cm} Finkelstein's transitive \\ \hspace*{2.6cm} causal relation \item $\mbf{c}$ \hspace*{2.25cm} Finkelstein's nontransitive \\ \hspace*{2.6cm} local causal relation \item $\prec$ \hspace*{2.25cm} binary relation, viewed as \\ \hspace*{2.6cm} a causal preorder \item $\tn{tr}(D)$ \hspace*{1.5cm} transitive closure of $D$ \item $\prec_{\tn{\fsz tr}}$ \hspace*{2cm} binary relation on $\tn{tr}(D)$ \item $\tn{sk}(D)$ \hspace*{1.5cm} skeleton of $D$ \item $\prec_{\tn{\fsz sk}}$ \hspace*{1.9cm} binary relation on $\tn{sk}(D)$ \item $\ms{D}_{\tn{\fsz tr}}$ \hspace*{1.9cm} category of transitive\\ \hspace*{2.6cm} directed sets \item $\tn{Mor}_{\ms{D}}\big(D,D'\big)$ \hspace*{.25cm} set of morphisms \\ \hspace*{2.6cm} from $D$ to $D'$ in $\ms{D}$ \end{itemize} \end{multicols} {\bf Section \hyperref[sectioninterval]{\ref{sectioninterval}}:} \begin{multicols}{2} \begin{itemize} \item $\ms{T}, \ms{T}'$ \hspace*{1.7cm} topologies \item $\overline{S}$ \hspace*{2.2cm} enlarged set containing $S$ \item $\overline{\ms{T}}$ \hspace*{2.2cm} topology on $\overline{S}$ \item $\ms{T}_{\tn{\fsz dis}}$ \hspace*{1.8cm} discrete topology \item $\ms{T}_{\tn{\fsz int}}$ \hspace*{1.8cm} interval topology \item $R(x)$ \hspace*{1.6cm} relation set at $x$ \item $E(x)$ \hspace*{1.6cm} unit intervals \\ \hspace*{2.6cm} glued at $x$ \item $M_\mathbb{R}$ \hspace*{1.9cm} continuum model of $M$ \item $\ms{T}_\mathbb{R}$ \hspace*{2cm} continuum topology \item $\tn{St}_\rho(x)$ \hspace*{1.4cm} star of radius $\rho$ at $x$ \item $U$ \hspace*{2.15cm} neighborhood of $x$\\ \hspace*{2.6cm} in continuum topology \item $M_\star$ \hspace*{1,9cm} star model of $M$ \item $\tn{St}(x)$ \hspace*{1.5cm} star at $x$ in $M_\star$ \item $\ms{T}_\star$ \hspace*{2.05cm} star topology \item $\ms{T}_{\star,\tn{\fsz irr}}$ \hspace*{1.6cm} irreducible star topology \item $\tn{St}_{\tn{\fsz irr}}(x)$ \hspace*{1.15cm} irreducible star \\ \hspace*{2.6cm} at $x$ in $M_\star$ \item $v(x)$ \hspace*{1.65cm} valence of $x$ \item $v$ \hspace*{2.25cm} cardinal-valued \\ \hspace*{2.6cm} scalar field on $M$ \item $v_M$ \hspace*{2cm} valence field on $M$ \item $v_M^{\pm}$ \hspace*{1.95cm} past and future \\ \hspace*{2.6cm} valence fields on $M$ \item LF \hspace*{2cm} local finiteness \item LFT \hspace*{1.7cm} local finiteness in the \\ \hspace*{2.6cm} transitive closure \item LFS \hspace*{1.75cm} local finiteness in the \\ \hspace*{2.6cm} skeleton \item $\alpha$ \hspace*{2.2cm} causal atom \\ \hspace*{2.6cm} in theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} \item $\llangle \alpha, y_{k+1}\rrangle$ \hspace*{.93cm} generalized open interval \\ \hspace*{2.6cm} in theorem \hyperref[theoremcausalsetrelint]{\ref{theoremcausalsetrelint}} \item $\mathbb{N}\coprod-\mathbb{N}$ \hspace*{1.25cm} linearly ordered base \\ \hspace*{2.6cm} for Jacob's ladder \item $[n]$ \hspace*{2cm} finite simplex $\{0,...,n\}$. \item EF \hspace*{1.95cm} element finiteness \item LEF \hspace*{1.75cm} local element finiteness \item RF \hspace*{1.95cm} relation finiteness \item PRF \hspace*{1.7cm} pairwise relation finiteness \item CF \hspace*{1.95cm} chain finiteness \item AF \hspace*{1.95cm} antichain finiteness \end{itemize} \end{multicols} \newpage {\bf Section \hyperref[sectionbinary]{\ref{sectionbinary}}:} \begin{multicols}{2} \begin{itemize} \item $\ms{R}(M)$ \hspace*{1.45cm} relation space over \\ \hspace*{2.6cm} $M=(M,R,i,t)$ \item $\prec$ \hspace*{2.25cm} induced relation on $\ms{R}(M)$ \item $r,s,\overline{r},\overline{s}$ \hspace*{1.25cm} elements of relation \\ \hspace*{2.6cm} spaces \item $\ms{R}:\ms{M}\rightarrow\ms{D}$ \hspace*{.7cm} relation space functor \item $\phi=(\phi_{\tn {\fsz elt}},\phi_{\tn {\fsz rel}})$ \hspace*{.1cm} morphism of \\ \hspace*{2.6cm} multidirected sets \item $\ms{R}(\phi)$ \hspace*{1.6cm} induced morphism of \\ \hspace*{2.6cm} relation spaces \item $R=(R,\prec)$ \hspace*{.8cm} arbitrary directed set, \\ \hspace*{2.6cm} viewed as a relation set \item $\ms{E}(R)$ \hspace*{1.6cm} abstract element space \\ \hspace*{2.6cm} over $R$ \item $M^{\pm}$ \hspace*{1.9cm} abstract sets yielding \\ \hspace*{2.6cm} $\ms{E}(R)$ as a quotient \item $\widetilde{x_r},\widetilde{y_r}$ \hspace*{1.6cm} abstract elements \\ \hspace*{2.6cm} in $\ms{E}(R)$ of $r\in R$ \item $\ms{M}_{\overline{\partial}},\ms{D}_{\overline{\partial}}$ \hspace*{1.25cm} subcategories of $\ms{M}$ \\ \hspace*{2.6cm} and $\ms{D}$ whose objects\\ \hspace*{2.6cm} have empty boundary \item $\tn{Id}_{\ms{M}_{\overline{\partial}}},\tn{Id}_{\ms{D}_{\overline{\partial}}}$ \hspace*{.7cm} identity functors on \\ \hspace*{2.6cm} $\ms{M}_{\overline{\partial}}$ and $\ms{D}_{\overline{\partial}}$ \item $\sigma$ \hspace*{2.2cm} Cauchy surface in \\ \hspace*{2.6cm} spacetime, or maximal \\ \hspace*{2.6cm} antichain in directed or \\ \hspace*{2.6cm} multidirected set \item $W,X,Y,Z$ \hspace*{.75cm} objects of a category \item $a,b,f,g$ \hspace*{1.2cm} morphisms of a category \item $(M_\star,R_\star,i_\star, t_\star)$ \hspace*{0cm} star model of $M$ as \\ \hspace*{2.65cm} a multidirected set \item $M_2$ \hspace*{1.9cm} set of $2$-element \\ \hspace*{2.6cm} subsets of $M$ \item $\ms{R}^2(A)$ \hspace*{1.35cm} second relation space \\ \hspace*{2.6cm} over $A$ \item $\prec_n$ \hspace*{2.05cm} $n$th induced relation \item $\overline{\gamma},\overline{\delta}$ \hspace*{1.9cm} simplices in $M$ \item $\alpha,\alpha'$ \hspace*{1.7cm} causal atoms \item $M^n$ \hspace*{1.85cm} $n$th degree causal atomic\\ \hspace*{2.6cm} decomposition of $M$ \item $\mbf{M}$ \hspace*{2.05cm} causal atomic \\ \hspace*{2.6cm} resolution of $M$ \item $\lambda_n$ \hspace*{2.05cm} $n$th growth factor for $\mbf{M}$ \item $\Lambda(\mbf{M})$ \hspace*{1.45cm} growth sequence for $\mbf{M}$ \item $T$ \hspace*{2.15cm} twistor space over \\ \hspace*{2.6cm} Minkowski spacetime $X$ \item $\mathbb{C}\mathbb{P}^3$ \hspace*{1.75cm} complex projective space \item $L,L',L''$ \hspace*{1.05cm} linear directed sets \item $\alpha, \beta, \gamma, \delta$ \hspace*{1.1cm} paths in a directed \\ \hspace*{2.6cm} or multidirected set \item $\Gamma(M)$ \hspace*{1.45cm} path set of $M$ \item $\Gamma_{\tn{\fsz irr}}(M)$ \hspace*{1.1cm} irreducible path\\ \hspace*{2.6cm} set of $M$ \item $\ell_+,\ell_-'$ \hspace*{1.6cm} terminal element of $L$ and\\ \hspace*{2.6cm} initial element of $L'$ in\\ \hspace*{2.6cm} definitions \hyperref[deficoncatprod]{\ref{deficoncatprod}} and \hyperref[defidirprod]{\ref{defidirprod}}. \item $L\sqcup L'$ \hspace*{1.4cm} linear directed set given\\ \hspace*{2.6cm} by joining $L,L'$ via \\ \hspace*{2.6cm} connecting relation $\ell_+\prec\ell_-'$ \item $\alpha\sqcup\beta$ \hspace*{1.6cm} concatenation product \\ \hspace*{2.6cm} of $\alpha$ and $\beta$ \item $\prec_\sqcup$ \hspace*{2.05cm} concatenation relation \\ \hspace*{2.6cm} on $\Gamma(D)$ \item $\big(\Gamma(D),\prec_\sqcup\big)$ \hspace*{.55cm} causal concatenation space \\ \hspace*{2.6cm} over $D=(D,\prec)$ \item $L\vee L'$ \hspace*{1.45cm} linear directed set given\\ \hspace*{2.6cm} by identifying $\ell_+$ with $\ell_-'$ \item $\alpha\vee\beta$ \hspace*{1.55cm} directed product \\ \hspace*{2.6cm} of $\alpha$ and $\beta$ \item $\prec_\vee$ \hspace*{2.05cm} directed product \\ \hspace*{2.6cm} relation on $\Gamma(M)$ \item $\big(\Gamma(M),\prec_\vee\big)$ \hspace*{.45cm} causal directed product \\ \hspace*{2.6cm} space over $M=(M,R,i,t)$ \item $\ms{S}$ \hspace*{2.2cm} object class of a \\ \hspace*{2.6cm} semicategory \item $\Gamma(\ms{S})$ \hspace*{1.65cm} morphism class of a \\ \hspace*{2.6cm} semicategory \item $\mbf{i},\mbf{t}$ \hspace*{2cm} initial and terminal \\ \hspace*{2.6cm} object maps of a \\ \hspace*{2.6cm} semicategory \item $\big(\Gamma(D),\sqcup\big)$ \hspace*{.8cm} causal concatenation \\ \hspace*{2.6cm} semicategory over $D$ \item $\big(\Gamma(M),\vee\big)$ \hspace*{.7cm} causal directed product \\ \hspace*{2.6cm} semicategory over $M$ \item $\big(\tn{Ch}(D),\sqcup\big)$ \hspace*{.5cm} chain concatenation \\ \hspace*{2.6cm} semicategory over $D$ \item $\big(\tn{Ch}(M),\vee\big)$ \hspace*{.45cm} chain directed product \\ \hspace*{2.6cm} semicategory over $M$ \item $\sim$ \hspace*{2.25cm} congruence relation \\ \hspace*{2.6cm} on a semicategory \item $\mathbb{R}[x]$ \hspace*{1.7cm} algebra of polynomials \\ \hspace*{2.6cm} with real coefficients \item $T$ \hspace*{2.2cm} a ring, usually \\ \hspace*{2.6cm} commutative with unit \item $t_\gamma$ \hspace*{2.2cm} element of $T$, appearing \\ \hspace*{2.6cm} as coefficient of path $\gamma$ \\ \hspace*{2.6cm} in path algebra \item $T^{\sqcup}(D)$ \hspace*{1.35cm} concatenation algebra over\\ \hspace*{2.6cm} $D$ with coefficients in $T$ \item $T^{\vee}(M)$ \hspace*{1.25cm} directed product \\ \hspace*{2.6cm} algebra over $M$ with \\ \hspace*{2.6cm} coefficients in $T$ \item $T^{\sqcup}[D]$ \hspace*{1.35cm} chain concatenation \\ \hspace*{2.6cm} algebra over $D$ with \\ \hspace*{2.6cm} coefficients in $T$ \item $T^{\vee}[M]$ \hspace*{1.3cm} chain directed product \\ \hspace*{2.6cm} algebra over $M$ with \\ \hspace*{2.6cm} coefficients in $T$ \item $P(x)$\hspace*{1.75cm} punctual module of $x$ \item $|x\rangle\langle y|$ \hspace*{1.6cm} ket-bra notation for \\ \hspace*{2.6cm} Raptis' incidence algebra \item $\delta_{uv}$ \hspace*{1.95cm} Kronecker delta function \item $\ms{L}$\hspace*{2.3cm} Lagrangian \item $\ms{S}$\hspace*{2.35cm} classical action \item $\Theta$\hspace*{2.3cm} phase map \item $\hbar$ \hspace*{2.25cm} Planck's reduced \\ \hspace*{2.6cm} constant $h/2\pi$ \item $\theta$\hspace*{2.4cm} relation function \item $T$\hspace*{2.3cm} target object of relation \\ \hspace*{2.6cm} function or phase map \item $\sum_{\gamma\in\Gamma(M)}\Theta(\gamma)\gamma$\hspace*{.05cm} phase map $\Theta$ as a path \\ \hspace*{2.6cm} algebra element \item $T^{\Gamma(D)}, T^{\Gamma(M)}$\hspace*{.35cm} mapping spaces of maps \\ \hspace*{2.6cm} $\Gamma(D)\rightarrow T$ and $\Gamma(M)\rightarrow T$ \item $\Theta,\Psi$\hspace*{1.8cm} pair of elements of \\ \hspace*{2.6cm} $T^{\Gamma(D)}$ or $T^{\Gamma(M)}$ \item $\Phi\vee\Psi, \Phi\sqcup\Psi$\hspace*{.6cm} convolutions; i.e., products \\ \hspace*{2.6cm} in path algebras \\ \hspace*{2.6cm} $T^{\sqcup}(D)$ and $T^{\sqcup}(M)$ \item $f\star g$\hspace*{1.85cm} convolution of \\ \hspace*{2.6cm} real-valued functions \item $\alpha-\beta$\hspace*{1.75cm} path such that\\ \hspace*{2.6cm} $\alpha\sqcup(\alpha-\beta)=\beta$ \item $T^{\tn{Ch}(D)},T^{\tn{Ch}(M)}$\hspace*{.05cm} chain mapping spaces \item $0_T$\hspace*{2.2cm} additive identity of $T$ \item $e_M$\hspace*{2.1cm} global evaluation \\ \hspace*{2.6cm} map $T^{\Gamma(M)}\dashedrightarrow T$ \item $e_\gamma$\hspace*{2.2cm} indicator function \item $\Theta(M)$\hspace*{1.53cm} shorthand for path\\ \hspace*{2.6cm} sum $\sum_{\gamma\in\Gamma(M)}\Theta(\gamma)$ \end{itemize} \end{multicols} {\bf Section \hyperref[subsectionquantumcausal]{\ref{subsectionquantumcausal}}:} \begin{multicols}{2} \begin{itemize} \item $\ms{Q}$ \hspace*{2.2cm} category in Isham's \\ \hspace*{2.6cm} quantization \item $Q,Q'$ \hspace*{1.6cm} objects of $\ms{Q}$ \item IS \hspace*{2.05cm} (principle of) iteration \\ \hspace*{2.6cm} of structure \item $\tau$ \hspace*{2.25cm} transition in $\ms{D}$ \item $D_i$ \hspace*{2cm} initial object, or source, \\ \hspace*{2.6cm} of a transition \item $D_t$ \hspace*{2cm} terminal object, or target, \\ \hspace*{2.6cm} of a transition \item $\tn{Aut}(D_i),$ \hspace*{.95cm} automorphism groups \\ $\tn{Aut}(D_t)$ \item $\alpha,\beta$ \hspace*{1.8cm} automorphisms of \\ \hspace*{2.6cm} directed sets \item $F_\tau$ \hspace*{2.05cm} group of extensions \\ \hspace*{2.6cm} of symmetries of $D_i$ \item $\rho_\tau$ \hspace*{2.05cm} restriction homomorphism \\ \hspace*{2.6cm} $F_\tau\rightarrow \tn{Aut}(D_i)$ \item $K_\tau,G_\tau$ \hspace*{1.3cm} kernel and cokernel \\ \hspace*{2.6cm} of $\rho_\tau$, also called \\ \hspace*{2.6cm} causal Galois groups \item $\iota$ \hspace*{2.3cm} inclusion $K_\tau\rightarrow F_\tau$ \item $\overline{\rho_\tau}$ \hspace*{2.1cm} quotient homomorphism \\ \hspace*{2.6cm} $G_\tau\rightarrow \tn{Aut}(D_i)$ \item $[F_\tau:G_\tau]$ \hspace*{.95cm} subgroup index of \\ \hspace*{2.6cm} $G_\tau$ in $F_\tau$ \item $h:D_i\Rightarrow D_t$ \hspace*{.55cm} co-relative history \item $D_i,D_t$ \hspace*{1.4cm} cobase and target \\ \hspace*{2.6cm} of $h$ \item $\mbf{D}_i$ \hspace*{2.05cm} cobase family \item PS \hspace*{1.95cm} path summation principle \item $X$ \hspace*{2.15cm} finite subset of\\ \hspace*{2.6cm} Euclidean spacetime \item $t=t^{\pm}$ \hspace*{1.6cm} spatial sections\\ \hspace*{2.6cm} bounding $X$ \item $\sigma^{\pm}$ \hspace*{2.05cm} upper and lower\\ \hspace*{2.6cm} boundaries of $X$ \item $\Gamma=\Gamma_{\tn{\fsz max}}(X)$ \hspace*{.3cm} space of maximal\\ \hspace*{2.6cm} directed paths in $X$ \item $\gamma$ \hspace*{2.25cm} element of $\Gamma_{\tn{\fsz max}}(X)$ \item $\ms{L}$ \hspace*{2.15cm} Lagrangian for single\\ \hspace*{2.6cm} particle moving in $X$ \item $\ms{S}$ \hspace*{2.25cm} classical action \\ \hspace*{2.6cm} corresponding to \\ \hspace*{2.6cm} Lagrangian $\ms{L}$ \item F1, F2 \hspace*{1.35cm} Feynman's ``postulates" \item $\psi(X;\ms{L})$ \hspace*{1.05cm} Feynman's quantum \\ \hspace*{2.6cm} amplitude \item $\hbar$ \hspace*{2.25cm} Planck's reduced \\ \hspace*{2.6cm} constant $h/2\pi$ \item $\Theta(\gamma)=e^{\frac{i}{\hbar}\ms{S}(\gamma)}$ \hspace*{.2cm} Feynman's phase map \item $\Delta$ \hspace*{2.2cm} partition of $[t^-,t^+]$ \item $\mbf{x}$ \hspace*{2.3cm} sequence of \\ \hspace*{2.6cm} position values for $\Delta$ \item $\gamma_i$ \hspace*{2.1cm} ``path increment" \\ \hspace*{2.6cm} from $(x_i,t_i)$ to $(x_{i+1},t_{i+1})$ \item $\sigma_i=\sigma_i(\Delta)$ \hspace*{.75cm} spatial section $t=t_i$ \\ \hspace*{2.6cm} for $t_i\in\Delta$ \item $|\Delta|$ \hspace*{2cm} norm of partition $\Delta$ \item $\psi(\Delta;\mbf{x};\ms{L})$ \hspace*{.72cm} amplitude for a\\ \hspace*{2.6cm} particular sequence $\mbf{x}$ \\ \hspace*{2.6cm} of position values \item $C$ \hspace*{2.15cm} proportionality factor \\ \hspace*{2.6cm} in Feynman's integral \item $R$ \hspace*{2.15cm} finite subset of\\ \hspace*{2.6cm} relation space $\ms{R}(M')$ \item $\big(\tn{Ch}(R),\sqcup\big)$ \hspace*{.55cm} chain concatenation \\ \hspace*{2.6cm} semicategory over $R$ \item $\Gamma=\tn{Ch}_{\tn{\fsz max}}(R)$ \hspace*{.1cm} subset of maximal \\ \hspace*{2.6cm} chains in $R$ \item $\gamma$ \hspace*{2.3cm} element of $\tn{Ch}_{\tn{\fsz max}}(R)$ \item $\sigma$ \hspace*{2.25cm} maximal \\ \hspace*{2.6cm} antichain in $R$ \item $R^-\coprod\sigma\coprod R^+$ \hspace*{.37cm} separation of $R$ \\ \hspace*{2.6cm} into past and future \\ \hspace*{2.6cm} regions $R^\pm$ by $\sigma$ \item $T$ \hspace*{2.2cm} target object \\ \hspace*{2.6cm} for path sums \item $\Theta$ \hspace*{2.15cm} phase map \\ \hspace*{2.6cm} $\tn{Ch}(R)\rightarrow T$ \item $\theta$ \hspace*{2.25cm} relation function \\ \hspace*{2.6cm} generating $\Theta$ \item $\psi(R;\theta)$ \hspace*{1.25cm} generalized quantum \\ \hspace*{2.6cm} amplitude \item $T^\sqcup[R]$ \hspace*{1.4cm} chain concatenation \\ \hspace*{2.6cm} algebra over $R$ \item $\Psi(R;\theta)$ \hspace*{1.2cm} maximal chain \\ \hspace*{2.6cm} functional of $R$ \\ \hspace*{2.6cm} with respect to $\theta$ \item $\Theta|_\Gamma$ \hspace*{1.9cm} restriction of $\Theta$ \\ \hspace*{2.6cm} to $\Gamma=\tn{Ch}_{\tn{\fsz max}}(R)$ \item $\psi^-$ \hspace*{2cm} past wave function in \\ \hspace*{2.6cm} Schr\"{o}dinger's equation \item $\mbf{H}$ \hspace*{2.1cm} Hamiltonian operator \item $V$ \hspace*{2.15cm} potential energy \item $m$ \hspace*{2.15cm} particle mass \item $\big\langle,\big\rangle$ \hspace*{2.05cm} complex inner product \item $\sigma'$ \hspace*{2.2cm} spatial section $t=t'$ \item $X^-\coprod\sigma'\coprod X^+$ \hspace*{.15cm} separation of $X$ \\ \hspace*{2.6cm} into past and future \\ \hspace*{2.6cm} regions $X^\pm$ by $\sigma'$ \item $\Delta^-$ \hspace*{1.95cm} partition of time \\ \hspace*{2.6cm} interval for $X^-$ \item $\approx$ \hspace*{2.25cm} signifies approximation \item $t''$ \hspace*{2.1cm} a time shortly after $t'$ \item $\delta\gamma$ \hspace*{2.05cm} path increment \item $r$ \hspace*{2.3cm} element of $\sigma$ \item $\Gamma^\pm$ \hspace*{2.05cm} sets of maximal chains \\ \hspace*{2.6cm} in $R^\pm$, respectively \item $\gamma$ \hspace*{2.25cm} element of $\Gamma^-$ \item $\Theta|_\sigma$ \hspace*{1.9cm} restriction of phase \\ \hspace*{2.6cm} map $\Theta$ to $\sigma$ \item $\Gamma_\sigma^{\pm}$ \hspace*{2cm} sets of maximal chains in \\ \hspace*{2.6cm} $R$ terminating on $\sigma$ $(-)$; \\ \hspace*{2.6cm} beginning on $\sigma$ $(+)$ \item $\Gamma_r^{\pm}$ \hspace*{2cm} sets of maximal chains in\\ \hspace*{2.6cm} $R$ terminating at $r$ $(-)$; \\ \hspace*{2.6cm} beginning at $r$ $(+)$ \item $\Psi^{\pm}_{R,\theta}$ \hspace*{1.7cm} past and future \\ \hspace*{2.6cm} chain functionals \item $\psi^{\pm}_{R,\theta}$ \hspace*{1.7cm} past and future \\ \hspace*{2.6cm} wave functions \item $r^-$ \hspace*{2.1cm} maximal predecessor of $r$ \item $(\ms{K},\mc{H})$ \hspace*{1.3cm} kinematic prescheme \\ \hspace*{2.6cm} or scheme \item $\ms{K}$ \hspace*{2.1cm} object class of $(\ms{K},\mc{H})$ \item $\mc{H}$ \hspace*{2.15cm} class of co-relative \\ \hspace*{2.6cm} histories of $(\ms{K},\mc{H})$ \item $\ms{U}(\ms{K},\mc{H})$ \hspace*{1cm} underlying directed set \\ \hspace*{2.6cm} of $(\ms{K},\mc{H})$ \item $\ms{V}(\ms{K},\mc{H})$ \hspace*{1.05cm} underlying multidirected \\ \hspace*{2.6cm} set of $(\ms{K},\mc{H})$ \item $\ms{W}(\ms{K},\mc{H})$ \hspace*{.95cm} underlying transition \\ \hspace*{2.6cm} structure of $(\ms{K},\mc{H})$ \item $x(D_i),x(D_t)$ \hspace*{.3cm} elements in an \\ \hspace*{2.6cm} underlying directed or \\ \hspace*{2.6cm} multidirected set of $(\ms{K},\mc{H})$ \item $r(\tau),r(h)$ \hspace*{.8cm} relations in an \\ \hspace*{2.6cm} underlying directed or \\ \hspace*{2.6cm} multidirected set of $(\ms{K},\mc{H})$ \item H \hspace*{2.2cm} hereditary property \item WA \hspace*{1.85cm} weak accessibility \item QCMH \hspace*{1.25cm} quantum causal metric \\ \hspace*{2.6cm} hypothesis \item $(\ms{A}_{\tn{\fsz fin}},\mc{H}_1)$ \hspace*{.8cm} positive sequential \\ \hspace*{2.6cm} kinematic scheme \item $\mc{H}_1$ \hspace*{2cm} class of transitions\\ \hspace*{2.6cm} adding a single element \\ \hspace*{2.6cm} to a directed set \item $(\ms{C}_{\tn{\fsz fin}},\mc{H}_1)$ \hspace*{.85cm} Sorkin and Rideout's \\ \hspace*{2.6cm} kinematic scheme \item $(\ms{A}_{\tn{\fsz fin}},\mc{H}_{\tn{\fsz gen}})$ \hspace*{.45cm} generational \\ \hspace*{2.6cm} kinematic scheme \item $\mc{H}_{\tn{\fsz gen}}$ \hspace*{1.65cm} class of transitions\\ \hspace*{2.6cm} adding a single generation \\ \hspace*{2.6cm} to a directed set \item $\widetilde{\tn{Ch}_{\tn{\fsz{max}}}}(\ms{K},\mc{H})$ \hspace*{.15cm} class of limits of $(\ms{K},\mc{H})$ \item $\overline{(\ms{K},\mc{H})}$ \hspace*{1.3cm} chain completion \\ \hspace*{2.6cm} of $(\ms{K},\mc{H})$ \item $\mc{I}$ \hspace*{2.2cm} inclusion ``functor" of \\ \hspace*{2.6cm} kinematic schemes \item $(F,\prec)$ \hspace*{1.45cm} Farey ``tree" \item $\mc{H}(\ms{K})$ \hspace*{1.45cm} class of all transitions \\ \hspace*{2.6cm} between pairs of \\ \hspace*{2.6cm} objects of $\ms{K}$ \item $\big(\ms{K},\mc{H}(\ms{K})\big)$ \hspace*{.6cm} universal kinematic \\ \hspace*{2.6cm} scheme over $\ms{K}$ \item $\big(\mbf{Kin}(\ms{K}),\prec_{\ms{K}}\big)$ \hspace*{0cm} kinematic space over $\ms{K}$ \end{itemize} \end{multicols} {\bf Section \hyperref[sectionconclusions]{\ref{sectionconclusions}}:} \begin{multicols}{2} \begin{itemize} \item M$^*$ \hspace*{1.95cm} modified measure axiom \item $e_i$ \hspace*{2.17cm} generalized initial or \\ \hspace*{2.6cm} terminal element map \item $V$ \hspace*{2.15cm} vertex set of a \\ \hspace*{2.6cm} random graph \item $(N,p)$ \hspace*{1.45cm} type of random graph \item $N$ \hspace*{2.1cm} number of vertices \\ \hspace*{2.6cm} in a random graph \item $p$ \hspace*{2.25cm} probability of an edge \\ \hspace*{2.6cm} between two vertices \\ \hspace*{2.6cm} in a random graph \end{itemize} \end{multicols} \end{appendix} \newpage \setcounter{secnumdepth}{0} \section{References} In the information age it is easy, and therefore not very helpful, to exhibit hundreds of papers on any given scientific subject not wholly obscure. I have chosen instead to restrict attention to a minimal number of reliable and well-written references, subject to the requirement of conveying the proper priority of authorship. Many of the authors cited here have produced numerous other equally relevant papers, but I have made an effort to distill the necessary concepts and quotations from a more manageable cross-section. A few of these references fall short in quality and/or clarity, but demand mention due to their priority in treating certain crucial topics. I have cited peer-reviewed and/or edited versions of references whenever possible, but have also included blue-colored links to arXiv preprints when available; the reader should always be aware of the possibility of differences between the two, and of the transient nature of links in general. Only the names of the actual articles cited are in italics; journal names, books containing articles, etc., are in normal font. Bold digits in typical journal citations indicate the {\it volume} of the journal cited; succeeding digits indicate the {\it number} or {\it issue,} the {\it page numbers} (preceded by ``pp.") or {\it article number,} and finally the {\it year published.} References are listed in the order in which they are {\it first cited} in the paper. The section headings below indicate in which sections these initial citations appear. I beg indulgence for any unintended slights due to my own ignorance. \vspace*{-1.5cm} \renewcommand{\refname}{}
\section{Introduction} Non-standard gravity models provide an alternative possibility towards understanding the accelerated expansion of the Universe (see \cite{Teg04} and references therein). The physical mechanism which is responsible for the present accelerating stage of the universe can be driven by a modification of the Einstein-Hilbert action, while the matter content of the universe remains the same (relativistic and cold dark matter). In the literature there are plenty of modified gravity models proposed by different authors, such as the braneworld Dvali, Gabadadze and Porrati \cite{Dvali2000} model, $f(R)$ gravity \cite{Sot10}, scalar-tensor theories \cite{scal}, Gauss-Bonnet gravity \cite{gauss}, Ho\v{r}ava-Lifshitz gravity \cite{hor}, nonlinear massive gravity \cite{Hinterbichler:2011tt} etc. Another gravitational scenario which has recently gained a lot of attention is the so called $f(T)$ gravity. The intrinsic properties of this scenario are based on the rather old formulation of the teleparallel equivalent of General Relativity (TEGR) \cite{ein28,Hayashi79,Maluf:1994ji}. Specifically, instead of using the torsion-less Levi-Civita connection of the classical General Relativity (GR) one utilizes the curvature-less Weitzenb{\"{o}}ck connection in which the corresponding dynamical fields are the four linearly independent {\it vierbeins.} Therefore, all the information concerning the gravitational field are included in the torsion tensor. Within this framework, considering invariance under general coordinate transformations, global Lorentz-parity transformations, and requiring up to second order terms of the torsion tensor, one can write down the corresponding Lagrangian density $T$ \cite{Hayashi79} by using some suitable contractions. A natural generalization of TEGR gravity is $f(T)$ gravity which is based on the fact that we allow the Lagrangian to be a function of $T$ \cite{Ferraro,Ferraro:2006jd,Linder:2010py}, inspired, of course, by the well-known extension of $f(R)$ Einstein-Hilbert action. However, $f(T)$ gravity does not coincide with $f(R)$ extension, but it rather consists a different class of modified gravity. It is interesting to mention that the torsion tensor includes only products of first derivatives of the vierbeins, giving rise to second-order field differential equations in contrast with the $f(R)$ gravity that provides fourth-order equations which potentially may lead to some problems, for example in the well-position and well-formulation of Cauchy problem \cite{vignolo}. Despite the fact that TEGR coincides completely with GR, both at the background and perturbation levels, it has been shown that $f(T)$ gravity provides different structural properties with respect to GR as well as different black-hole solutions and cosmological features \cite{Ferraro,Ferraro:2006jd,Linder:2010py,Myrzakulov:2010vz, Wu:2010mn,Bengochea001,Iorio:2012cm,Wang:2011xf}. An important question here is what classes of $f(T)$ extensions are allowed. From the phenomenological viewpoint, the aforementioned cosmological and spherical analysis lead to a variety of such expressions. Using cosmological \cite{Wu:2010mn,Bengochea001,Zhang:2012jsa} and Solar System \cite{Iorio:2012cm} observations, one can show that the deviations from TEGR must be small. In this work, we use a model-independent selection rule based on first integrals, due to Noether symmetries of the equations of motion, in order to identify the viability of $f(T)$ gravity in the context of flat Friedmann-Lema\^\i tre-Robertson-Walker (FLRW) cosmologies. Actually, the idea to use Noether symmetries in cosmology is not new and indeed there is a lot of work in the literature (see \cite{Cap96,Cap97,RubanoSFQ,Sanyal05,Szy06,Cap07,Capa07,Bona07,Cap08,Cap09,Vakili08,Yi09,WeiHao}) along this line. In this context, recently we have shown (see Basilakos et al. \cite{Basilakos11}; Paliathanasis et al. \cite{Tsafr}) that the existence of Noether symmetries can be used as a selection criterion in order to distinguish the scalar dark energy models \cite{Basilakos11} as well as the $f(R)$ gravity models \cite{Tsafr}. Inspired by the above works in the current article, we would like to estimate the Noether symmetries of the $f(T)$ gravity. The aim here is $(a)$ to identify the $f(T)$ functional forms which accommodate extra Noether symmetries, and $(b)$ for these models, to solve the system of the resulting field equations and derive analytically the main cosmological functions (the scale factor, the Hubble expansion rate, deceleration parameter and growth factor) and finally to compare with other cosmological patterns which are outside and inside GR. The structure of the article is as follows: In Sec. II, we discuss the issue of torsion in GR and its connection with unholonomic frames. This discussion is useful in order to clarify some misunderstandings on the role of torsion that are present in literature. In particular, we shall discuss its dependence on the frame where observations are made. In Sec. III, we give the basic FLRW cosmological equations in the framework of $f(T)$ gravity. The main properties and theorems of the Noether Symmetry Approach are summarized in Sec. IV. Noether symmetries for $f(T)$ cosmology are discussed in Sec. V. In Sec VI we provide analytical solutions for $f(T)$ models that admit non trivial Noether symmetries. A comparison with analogous $f(R)$ cosmology is pursued putting in evidence similarities and differences. We draw conclusions in Sec.VI. \section{The role of torsion in General Relativity} Before starting our considerations on $f(T)$ gravity and its cosmological realization, it is useful to discuss in detail the role of torsion in GR considering, in particular, how it behaves with respect to holonomic and unholonomic frames. Let us start with some definitions. In an $n-$dimensional manifold ${\cal M}$ consider a coordinate neighborhood ${\cal U}$ with a coordinate system $\{x^{\mu }\}.$ At each point $P \in {\cal U}$, we have the resulting holonomic frame $\{\partial _{\mu }\}.$ We define in ${\cal U}$ a new frame $\{e_{a}(x^{\mu })\}$ which is related to the holonomic frame \{\partial _{a}\}$ as follows:% \begin{equation} e_{a}(x^{\mu })=h_{a}^{\mu }\partial _{\mu }~\ \ a,\mu =1,2,...,n \label{H.1} \end{equation}% where the quantities $h_{a}^{\mu }(x)$ are in general functions of the coordinates (i.e. depend on the point $P$). Notice that Latin indexes count vectors, while Greek indexes are tensor indices. We assume that $\det h_{a}^{\mu }\neq 0$ which guaranties that the vectors $\{e_{a}(x^{\mu })\}$ form a set of linearly independent vectors. We define the ''inverse'' quantities $h_{a }^{\mu}$ by means of the following ''orthogonality'' relations: \begin{equation} h_{a}^{\mu }h_{\nu }^{a}=\delta _{\nu }^{\mu }\;\;\;,h_{b}^{\mu }h_{\mu }^{c}=\delta _{b}^{c}. \label{H.1a} \end{equation} The commutators of the vectors $\{e_{a}\}$ are not in general all zero. If they are zero, then there exists a new coordinate system in ${\cal U}$, $\{y^{b}\}$ so that ${\displaystyle e_{b}=\frac{\partial }{\partial y^{b}}}$, i.e. the new frame is holonomic. If there are commutators $[e_{a},e_{b}]\neq 0$ then the new frame $\{e_{b}\}$ is called unholonomic and at least a number of vectors $e_{b}$ cannot be written in the form $e_{b}=\partial _{b}.$ The quantities which characterize an unholonomic frame are the objects of unholonomicity or Ricci rotation coefficients $\Omega _{\text{ }bc}^{a}$ defined by the relation% \begin{equation} \lbrack e_{a},e_{b}]=\Omega _{\text{ }ab}^{c}e_{c}. \label{H.3} \end{equation} Let us compute:% \begin{equation*} \lbrack e_{a},e_{b}]=[h_{a}^{\mu }\partial _{\mu },h_{b}^{\nu }\partial _{\nu }]=\left[ h_{a}^{\mu }h_{b,\mu }^{\nu }h_{\nu }^{c}-h_{b}^{\nu }h_{a,\nu }^{\mu }h_{\mu }^{c}\right] e_{c} \end{equation*}% from which follows that the Ricci rotation coefficients of the frame $% \{e_{a}\}$ are: \begin{equation} \Omega _{\text{ }bc}^{a}=2h_{[b}^{\mu }h_{c],\mu }^{\nu }h_{\nu }^{a}. \label{H.2} \end{equation}% The condition for $\{e_{a}\}$ to be a holonomic basis is $\Omega _{\text{ }% bc}^{a}=0$ at all points $P\in {\cal U}.$ This is a set of linear partial differential equations whose solution defines all holonomic frames and all coordinate systems in ${\cal U}.$ One obvious solution is $h_{b}^{c}=\delta _{b}^{c} $. The set of all coordinate systems in ${\cal U}$, equipped with the operation of composition of transformations, has the structure of an infinite dimensional Lie group which is called the {\it Manifold Mapping Group} \cite{StephaniB}. Let us consider now the special unholonomic frames which satisfy the Jacobi identity:% \begin{equation} \lbrack \lbrack e_{a},e_{b}],e_{c}]+[[e_{b},e_{c}],e_{a}]+[[e_{c},e_{a}],e_{b}]=0. \label{H.4} \end{equation}% These frames are the generators of a Lie algebra, therefore they have an extra role to play. Replacing the commutator in terms of the unholonomicity objects, we find the following identity:% \begin{equation} \Omega _{\text{ }ab,c}^{d}+\Omega _{\text{ }ba,a}^{d}+\Omega _{\text{ }% ca,b}^{d}-\Omega _{\text{ }ab}^{l}\Omega _{cl}^{d}-\Omega _{\text{ }% bc}^{l}\Omega _{al}^{d}-\Omega _{\text{ }ca}^{l}\Omega _{bl}^{d}=0. \end{equation} Using the definition of the covariant derivative we write:% \begin{equation} \nabla _{e_{i}}e_{j}=\Gamma _{ij}^{k}e_{k} \label{H.20} \end{equation}% where $\Gamma _{ij}^{k}$ are the connection coefficients in the frame $% \{e_{i}\}.$ If we compute the $\Gamma _{ij}^{k}$ assuming \begin{equation*} \lbrack e_{i},e_{j}]=C_{.ij}^{k}e_{k} \end{equation*}% it follows that% \begin{equation*} C_{.ij}^{k}=\Omega _{.jk}^{k}. \end{equation*} Let us consider now three vector fields $X,Y,Z$ and the covariant derivative of the metric vector $X.$ Then we have: \begin{equation} \nabla _{X}g(Y,Z)=X(g(Y,Z))-g(\nabla _{X}Y,Z)-g(Y,\nabla _{Y}Z) \label{H.21} \end{equation}% and by interchanging the role of $X,Y,Z:$% \begin{equation} \nabla _{Y}g(Z,X)=Y(g(Z,X))-g(\nabla _{Y}Z,X)-g(Z,\nabla _{Z}X) \label{H.22} \end{equation} \begin{equation} \nabla _{Z}g(X,Y)=Z(g(X,Y))-g(\nabla _{Z}X,Y)-g(X,\nabla _{X}Y). \label{H.23} \end{equation} Adding Eqs. (\ref{H.21}), (\ref{H.22}) and subtracting (\ref{H.23}), one obtains: \begin{eqnarray*} \nabla _{X}g(Y,Z)+\nabla _{Y}g(Z,X)-\nabla _{Z}g(X,Y) &=&X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))+ \\ &&-\left[ g(\nabla _{X}Y,Z)+g(\nabla _{Y}Z,X)-g(\nabla _{Z}X,Y)\right] + \\ &&-\left[ g(Y,\nabla _{X}Z)+g(Z,\nabla _{Y}X)-g(X,\nabla _{Z}Y)\right] \end{eqnarray*}% then% \begin{eqnarray*} \nabla _{X}g(Y,Z)+\nabla _{Y}g(Z,X)-\nabla _{Z}g(X,Y) &=&X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))+ \\ &&-\left[ g(\nabla _{X}Y,Z)+g(Z,\nabla _{Y}X)\right] + \\ &&-\left[ g(\nabla _{Y}Z,X)-g(X,\nabla _{Z}Y)\right] + \\ &&-\left[ g(Y,\nabla _{X}Z)-g(\nabla _{Z}X,Y)\right] \end{eqnarray*}% that is% \begin{eqnarray*} \nabla _{X}g(Y,Z)+\nabla _{Y}g(Z,X)-\nabla _{Z}g(X,Y) &=&X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))+ \\ &&-\left[ g(Z,\nabla _{X}Y+\nabla _{Y}X)+g(X,\nabla _{Y}Z-\nabla _{Z}Y)+g(Y,\nabla _{X}Z-\nabla _{Z}X)\right] \,, \end{eqnarray*}% where \begin{equation*} g(Z,\nabla _{X}Y+\nabla _{Y}X)=2g\left( Z,\nabla _{X}Y\right) +g\left( Z,\nabla _{Y}X-\nabla _{X}Y\right) . \end{equation*}% Replacing in the last relation and solving for $2g\left( Z,\nabla _{X}Y\right) $, we find% \begin{eqnarray*} 2g\left( Z,\nabla _{X}Y\right) &=&\left[ X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))% \right] + \\ &&-\left[ \nabla _{X}g(Y,Z)+\nabla _{Y}g(Z,X)-\nabla _{Z}g(X,Y)\right] + \\ &&-\left[ g\left( Z,\nabla _{Y}X-\nabla _{X}Y\right) +g(X,\nabla _{Y}Z-\nabla _{Z}Y)+g(Y,\nabla _{X}Z-\nabla _{Z}X)\right] \end{eqnarray*}% or% \begin{eqnarray*} 2g\left( Z,\nabla _{X}Y\right) &=&\left[ X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))% \right] + \\ &&-\left[ \nabla _{X}g(Y,Z)+\nabla _{Y}g(Z,X)-\nabla _{Z}g(X,Y)\right] + \\ &&-\left[ g\left( Z,\nabla _{Y}X-\nabla _{X}Y-\left[ Y,X\right] \right) +g(X,\nabla _{Y}Z-\nabla _{Z}Y-\left[ Y,Z\right] )+g(Y,\nabla _{X}Z-\nabla _{Z}X-\left[ X,Z\right] )\right] + \\ &&-\left[ g\left( Z,\left[ Y,X\right] \right) +g\left( X,\left[ Y,Z\right] \right) +g\left( Y,\left[ X,Z\right] \right) \right] . \end{eqnarray*}% At this point, we can define the quantities% \begin{eqnarray*} T_{\nabla }(X,Y) = \nabla _{X}Y-\nabla _{Y}X-[X,Y]\,,\qquad A_{\nabla }(X,Y,Z) = \nabla _{X}g(Y,Z)\,. \end{eqnarray*}% The tensors $T_{\nabla }$ and $A_{\nabla }$ are called the {\it torsion} ($T_{\nabla }\equiv T$) and the {\it metricity} of the connection $\nabla $ respectively. Last relation in terms of the fields $T_{\nabla }$ and $A_{\nabla }$ is written as follows$:$% \begin{eqnarray} 2g\left( Z,\nabla _{X}Y\right) &=&\left[ X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))% \right] + \notag \\ &&-\left[ A_{\nabla }\left( X,Y,Z\right) +A_{\nabla }\left( Y,Z,X\right) -A_{\nabla }\left( Z,X,Y\right) \right] + \notag \\ &&-\left[ g\left( Z,T_{\nabla }\left( Y,X\right) \right) +g\left( X,T_{\nabla }\left( Y,Z\right) \right) +g\left( Y,T_{\nabla }\left( X,Z\right) \right) \right] \notag \\ &&-\left[ g\left( Z,\left[ Y,X\right] \right) +g\left( X,\left[ Y,Z\right] \right) +g\left( Y,\left[ X,Z\right] \right) \right] . \label{H.24} \end{eqnarray} Let $X=e_{l}~,~Y=e_{j}$ and $Z=e_{k}.$ Contracting with $\frac{1}{2}g^{il}$, we have% \begin{equation*} 2g\left( Z,\nabla _{X}Y\right) \rightarrow \Gamma _{jk}^{i} \end{equation*}% \begin{equation*} \left[ X(g(Y,Z))+Y(g(Z,X))-Z(g(X,Y))\right] \rightarrow \left\{ _{jk}^{i}\right\} \end{equation*}% \begin{equation*} g\left( X,T_{\nabla }\left( Y,Z\right) \right) \rightarrow Q_{.kj}^{i} \end{equation*}% \begin{equation*} g\left( Z,T_{\nabla }\left( Y,X\right) \right) +g\left( Y,T_{\nabla }\left( X,Z\right) \right) \rightarrow g^{il}(g_{tj}Q_{kl}^{t}+g_{tk}Q_{jl}^{t})=-% \bar{S}_{.kj}^{i} \end{equation*}% \begin{equation*} g\left( X,\left[ Y,Z\right] \right) \rightarrow \frac{1}{2}C_{.jk}^{i} \end{equation*}% \begin{equation*} g\left( Z,\left[ Y,X\right] \right) +g\left( Y,\left[ X,Z\right] \right) =% \frac{1}{2}g^{il}(g_{tj}C_{lk}^{t}+g_{tk}C_{jl}^{t})=-S_{.kj}^{i} \end{equation*}% and% \begin{equation*} A_{\nabla }\left( X,Y,Z\right) +A_{\nabla }\left( Y,Z,X\right) -A_{\nabla }\left( Z,X,Y\right) \rightarrow \frac{1}{2}g^{il}\Delta _{jkl} \end{equation*}% Replacing in Eq.(\ref{H.24}), we find the connection coefficients in the frame $% \{e_{i}\}$, that is \begin{equation} \Gamma _{jk}^{i}=\left\{ _{jk}^{i}\right\} +\bar{S}_{.kj}^{i}+S_{.kj}^{i}-% \frac{1}{2}g^{il}\Delta _{jkl}+Q_{jk}^{i}-\frac{1}{2}C_{.jk}^{i} \label{H25} \end{equation}% where $\left\{ _{jk}^{i}\right\} $ are the standard Levi-Civita connection coefficients (i.e. the Christofell symbols). This is the most general expression for the connection coefficients in terms of the fields $\left\{ _{jk}^{i}\right\} ,$ $T_{\nabla },$ $A_{\nabla }$ and $C_{jk}^{i}$. Concerning the symmetric and antisymmetric part, we have:% \begin{eqnarray} \Gamma _{.(jk)}^{i} &=&\left\{ _{jk}^{i}\right\} +\bar{S}% _{.jk}^{i}+S_{.jk}^{i}-\frac{1}{2}g^{il}\Delta _{jkl} \label{H.26} \\ \Gamma _{.[jk]}^{i} &=&Q_{.jk}^{i}-\frac{1}{2}C_{.jk}^{i}\,, \label{H.27} \end{eqnarray} and then we can draw the following conclusions: \begin{enumerate} \item The connection coefficients in a frame $\{e_{i}\}$ are determined from the metric, the torsion, the metricity and the unholonomicity objects (equivalently the commutators) of the frame. \item The symmetric part $\Gamma _{.(jk)}^{i}$ of $\Gamma _{jk}^{i}$ depends on all fields. This means that the geodesics and the autoparallels in a given frame depend on the geometric properties of the underlying manifold (fields $g_{ij},$ $% Q_{.kj}^{i},g_{ij|k})$ and the unholonomicity of the frame (field $% C_{.jk}^{i})$. \item The antisymmetric part $\Gamma _{.[jk]}^{i}$ of $\Gamma _{jk}^{i}$ depends only on all fields $Q_{.kj}^{i}$ and $C_{.jk}^{i}.$ \item The objects of unholonomicity $C_{.jk}^{i}$ behave in the same way as the components of torsion. This means that even in a Riemannian space where $% Q_{.kj}^{i}=0,g_{ij|k}=0$ in an unholonomic basis the antisymmetric part $% \Gamma _{.[jk]}^{i}=-\frac{1}{2}C_{.jk}^{i}\neq 0.$ \end{enumerate} This result has lead to the misunderstanding that when one works in an unholonomic frame then the torsion is introduced. This statement is clearly not correct. This misunderstanding has important consequences because the effects one will observe in an unholonomic frame will be frame dependent and not covariant effects. Therefore all conclusions made in a specific unholonomic frame must be restricted to that frame only. \section{$f(T)$ gravity and cosmology} With the above considerations in mind, let us consider TEGR and its straightforward extension $f(T)$. Teleparallelism uses as dynamical objects the {\it vierbiens} as unholonomic frames in spacetime. Following the definitions in the previous section, they are defined by the requirement $% g(e_{i},e_{j})=e_{i}.e_{j}=\eta _{ij}$, where $\eta _{ij}={\rm diag}(-1,+1,+1,+1)$ is the Lorentz metric in canonical form. Obviously $g_{\mu \nu }(x)=\eta _{ij}h_{\mu }^{i}(x)h_{\nu }^{j}(x)$ where $e^{i}(x)=h_{\mu }^{i}(x)dx^{i}$ is the dual basis. Differing from GR, which uses the torsionless Levi-Civita connection, Teleparallelism utilizes the curvatureless {\it Weitzenb\"{o}ck connection}, whose non-null torsion tensor is defined as \begin{equation}\label{Wein} T_{\mu\nu}^{\beta} =\hat{\Gamma}_{\nu\mu}^{\beta} -\hat{\Gamma}_{\mu\nu}^{\beta} =h_{i}^{\beta} (\partial_\mu h_{\nu}^{i} - \partial_\nu h^{i}_{\mu}) \;. \end{equation} Notice the Ricci rotation coefficients are $\Omega _{jk}^{i}=T_{jk}^{i}$ and encompass all the information concerning the gravitational field. The TEGR Lagrangian for the gravitational field equations (Einstein equations) is assumed to be:% \begin{equation} T={S_{\beta}}^{\mu \nu } {T^{\beta}}_{\mu \nu }, \label{lagrangian} \end{equation}% where \begin{equation} {S_{\beta}}^{\mu \nu }=\frac{1}{2}({K^{\mu \nu }}_{\beta}+\delta^{\mu} _{\beta }{T^{\theta \nu }}_{\theta}-\delta^{\nu}_{\beta}{T^{\theta \mu }}_{\theta}) \label{s} \end{equation}% and ${K^{\mu \nu }}_{\beta}$ is the {\it contorsion} tensor \begin{equation} {K^{\mu \nu }}_{\beta}=-\frac{1}{2}({T^{\mu \nu }}_{\beta}-{T^{\nu \mu} }_\beta -{T_{\beta}}^{\mu \nu }), \label{contorsion} \end{equation}% which equals the difference of the Levi Civita connection in the holonomic and the unholonomic frame (see Sec. II for details). Here, the gravitational field will be driven by a Lagrangian density which is a function of the trace $T$. Therefore, the corresponding action of $f(T)$ gravity reads as \begin{equation} \label{action} \mathcal{A}_T=\frac{1}{16\pi G}\int {d^{4}xef(T)} \end{equation}% where $e=det(e_{\mu }^{i}\cdot e_{\nu }^{i})=\sqrt{-g}$. Obviously, TEGR and thus GR, are restored for $f(T)=T$. In order to construct a realistic theory of gravity, we have to incorporate the matter and radiation fields too. Therefore, the total action is written as \begin{eqnarray} \label{action11} A_{\rm tot}= {\cal A}_T+\frac{1}{16\pi G }\int d^4x e \left(L_m+L_r\right), \end{eqnarray} where the matter and radiation Lagrangians are assumed to correspond to perfect fluids with energy densities $\rho_m$, $\rho_r$ and pressures $p_m$, $p_r$ respectively. If matter couples to the metric in the standard form then the variation of the action with respect to the vierbein leads to the equations \cite{Ferraro} \begin{eqnarray} &&e^{-1}\partial _{\mu }(e{S}_{i}^{\mu \nu })f^{\prime }(T)-h_{i}^{\lambda } T_{\mu \lambda }^{\beta}S_{\beta}^{\nu \mu }f^{\prime }(T) \notag \\ &&+S_{i}^{\mu \nu }\partial _{\mu }(T)f^{\prime \prime }(T)+\frac{1}{4}% h_{i}^{\nu }f(T)=4\pi Gh_{i}^{\beta}{T^{(m)}}_{\beta}^{\nu } \label{equations} \end{eqnarray}% where a prime denotes differentiation with respect to $T$, ${S_{i}}^{\mu \nu }={h_{i}}^{\beta}S_{\beta }^{\mu \nu }$ and $T^{(m)}_{\mu \nu }$ is the matter energy-momentum tensor. It is easy to show that, for $f(T)=T$, Eqs.(\ref{equations}) reduce to the standard Einstein equations \cite{miao}. In order to consider the related $f(T)$ cosmology, let us assume a spatially flat FLRW metric which, in the holonomic (comoving) frame $\{\partial t,\partial x,\partial y,\partial z\}$, assumes the form \begin{equation*} ds^{2}=-dt^{2}+a^{2}(t)(dx^{2}+dy^{2}+dz^{2}) \end{equation*}% where $a(t)$ is the cosmological scale factor. In this space we define the vierbein (unholonomic frame) $\{e_{i}\}$ which becomes: \begin{equation} h_{\mu }^{i}(t)={\rm diag}(1,a(t),a(t),a(t)), \label{metric} \end{equation} In order to derive the cosmological equations in a FLRW metric, we need to deduce a point-like Lagrangian from the action (\ref{action}). As a consequence, the infinite degrees of freedom of the original field theory will be reduced to a finite number as in mechanical systems. This fact allows to deal with minisuperspaces of finite dimensions (see \cite{hamilton} for details). In this framework, considering $\{a,T\}$ as the canonical variables of the configuration space the $f(T)$ action becomes formally: \begin{equation*} \mathcal{A}_T=\int \mathcal{L}(a,{\dot{a}},T,\dot{T})dt \;. \end{equation*}% Due to the fact that $T$, in GR, reduces to \begin{equation} \label{H.50} T=-6\left( \frac{\dot{a}}{a}\right)^2 =-6H^{2} \end{equation}% where $H$ is the Hubble parameter \cite{Myrzakulov:2010vz}, one can rewrite the $f(T)$ action using a Lagrange multiplier $\lambda_{\cal L}$ as follows: \begin{equation} \mathcal{A}_T=2\pi ^{2}\int dt\left\{ f(T)a^{3}-\lambda_{\cal L} \left[ T+6\left( \frac{% \dot{a}^{2}}{a^{2}}\right) \right] \right\} \,. \label{H.51} \end{equation} In order to determine $\lambda_{\cal L}$, we need to vary the $f(T)$ action with respect to $T$, that is \begin{equation*} a^{3}\frac{df(T)}{dT}\delta T-\lambda_{\cal L} \delta T=0\, \end{equation*}% from which follows \begin{equation*} \lambda_{\cal L} =a^{3}f^{\prime }(T)\,. \end{equation*}% Replacing in the Lagrangian we find:% \begin{equation} \mathcal{L}=a^{3}\left[ f(T)-Tf^{\prime }(T)\right] -6\dot{a}^{2}af^{\prime }(T)\,, \label{H.52} \end{equation} which is canonical in the variables $\{a,T\}$. Also, the substitution of the vierbein (\ref{metric}) in Eq.(\ref{equations}) for $i=\nu=0 $ (as well as the energy condition) yields \begin{equation} \label{friedmann} 12H^{2}f^{\prime }(T)+f(T)=16\pi G\rho \;. \end{equation}% Besides, for $i=\nu=1$ Eq.(\ref{equations}) gives \begin{equation} 48H^{2}\dot{H}f^{\prime \prime }(T)-4(\dot{H}+3H^2)f^{\prime}(T)-f(T)=16\pi Gp \label{acceleration} \end{equation}% where $\rho=\rho_{m}+\rho_{r}$ and $p=p_{m}+p_{r}$ are the total energy density and pressure respectively which they have been measured in the unholonomic frame. It is important to stress that Eqs.(\ref{friedmann}), (\ref{acceleration}) can be derived by the Euler-Lagrange equations \begin{equation} E_{\mathcal{L}}=\frac{\partial \mathcal{L}}{\partial \dot{a}}\dot{a}+\frac{\partial \mathcal{L}}{\partial \dot{T}}\dot{T}-\mathcal{L}\,, \end{equation} and \begin{equation} \frac{d}{dt}\frac{\partial \mathcal{L}}{\partial \dot{a}}=\frac{\partial \mathcal{L}}{\partial a}\,, \end{equation} respectively. The Euler-Lagrange equation \begin{equation} \frac{d}{dt}\frac{\partial \mathcal{L}}{\partial \dot{T}}=\frac{\partial \mathcal{L}}{\partial T}\,, \end{equation} gives the constraint (\ref{H.50}). In this sense, the point-like Lagragian (\ref{H.52}) completely defines the related dynamical system in the minisuperspace $\{a,T \}$. It is interesting to mention that using the conservation equation $\dot{\rho}+3H(\rho +p)=0$ one can rewrite Eqs. (\ref{friedmann}) and (\ref{acceleration}) in the Friedmann-Einstein form \begin{equation} H^{2}=\frac{8\pi G}{3}(\rho +\rho _{T}), \label{modfri} \end{equation}% \begin{equation} 2\dot{H}+3H^{2}=-8\pi G(p+p_{T}) \label{modacce} \end{equation}% where \begin{equation} \rho _{T}=\frac{1}{16\pi G}[2Tf^{\prime }(T)-f(T)-T], \label{rhoT} \end{equation}% \begin{equation} p_{T}=\frac{1}{16\pi G}\left\{4\dot{H}[2Tf^{\prime\prime}(T)+f^{\prime}(T)-1]\right\}-\rho_T \label{pT} \end{equation}% are the unholonomicity contributions to the energy density and pressure that disappear as son as $f(T)=T$. Finally, $f(T)$ gravity can mimic, under specific circumstances, the scalar field for dark energy \cite{Myrzakulov:2010vz}. In order to address this crucial question, we need to derive an effective equation-of-state parameter $w(a)$ for the $f(T)$ cosmology. Indeed, utilizing Eqs. (\ref{rhoT}) and (\ref{pT}), we can easily obtain the effective unholonomicity equation of state as \begin{equation} \label{omegaeff} \omega _{T}\equiv \frac{p_{T}}{\rho _{T}}=-1+\frac{4\dot{H}[2Tf^{\prime \prime }(T)+f^{\prime }(T)-1]}{2Tf^{\prime }(T)-f(T)-T} \;. \end{equation} It is easy to see that possible deviations from $\Lambda$CDM model can be addressed by the second term in such an equation. \section{Noether symmetries} Generally, Noether symmetries play an important role in physics because they can be used to simplify a given system of differential equations as well as to determine the integrability of the system. In general, the existence of a Noether symmetry can be related to a conserved quantity bringing a physical meaning. The so called {\it Noether Symmetry Approach} results extremely useful in cosmology in order to find out exact solutions (see \cite{Cap96} for a comprehensive review of the method). We would like to remind the reader that a fundamental approach to derive the Noether symmetries for a given dynamical problem (in a Riemannian space) has been published recently by Tsamparlis \& Paliathanasis \cite{Tsam10} (a similar analysis can be found in \cite{Kalotas,Olver,StephaniB,MoyoLeach,Tsamparlis2010,Tsama10}). Let us consider the Hamiltonian ${\cal H}$ which depends on one independent$~$variable~$\left\{ t\right\} $ and $n$ dependent variables~$\left\{ x^{i}(t):i=1...n\right\} $, i.e. ${\cal H}={\cal H}\left( t,x^{k},\dot{x}^{k},...,x^{\left[ n\right] k}\right)$ where a dot over a symbol means differentiation with respect to $t$. We perform the one parameter point transformation \begin{equation} \label{Ls.01} \bar{t}=\Xi \left( t,x^{k},\varepsilon \right)\;,~~\bar{x}^{A}=\Phi \left( t,x^{k},\varepsilon \right) \;. \end{equation}% In that case, the generating vector of the one parameter point transformation is \begin{equation} X=\xi \left( t,x^{k},\varepsilon \right) \partial _{t}+\eta ^{i}\left( t,x^{k},\varepsilon \right) \partial _{i} \label{Ls.02} \end{equation}% where% \begin{equation*} \xi \left( t,x^{k}\right) =\frac{\partial \Xi ^{i}\left( t,x^{k},\varepsilon \right) }{\partial \varepsilon }|_{\varepsilon \rightarrow 0}~~,~~\eta ^{i}\left( t,x^{k}\right) =\frac{\partial \Phi \left( t,x^{k},\varepsilon \right) }{\partial \varepsilon }|_{\varepsilon \rightarrow 0}. \end{equation*} The extension of the generator vector in the jet space $B_{M}=\left\{ t,x^{k},\dot{x}^{k},\ddot{x}^{k}...,x^{\left[ n\right] k}\right\} $ is \cite% {StephaniB} \begin{equation*} X^{\left[ n\right] }=X+\eta _{i}^{A}\partial _{u_{i}}+...+\eta _{ij..i_{n}}^{A}\partial _{u_{ij..i_{n}}} \end{equation*}% where% \begin{equation} \eta ^{\left[ 1\right] i}=\frac{d}{dt}\eta ^{i}-\dot{x}^{i}\frac{d}{dt}\xi \end{equation}% \begin{equation} \eta ^{\left[ n\right] i}=\frac{d}{dt}\eta ^{\left[ n-1\right] i}-x^{\left[ n% \right] ^{i}}\frac{d}{dt}\xi \end{equation}% $X^{\left[ n\right] }$ is called the nth prolongation of the generator (\ref{Ls.02}). We say that the function ${\cal H}\left( t,x^{k},\dot{x}^{k},\ddot{x}^{k}...,x^{\left[ n\right] k}\right) =0$ is invariant under the transformation of Eq.(\ref{Ls.01}) if and only if there is a function $\lambda_{\cal L}$ such as the following condition holds \begin{equation} X^{\left[ n\right] }\left( {\cal H}\right) =\lambda_{\cal L} {\cal H}\;,\;\;mod{\cal H}=0 \label{Ls.03} \end{equation}% where $\lambda_{\cal L}$ is a function to be determined \cite{IbraB}. Moreover, the generating vector (\ref{Ls.02}) is a Lie symmetry of the function ${\cal H}\left( t,x^{k},\dot{x}^{k},\ddot{x}^{k}...,x^{\left[ n\right] k}\right)$. In the following sections we are interested on systems of second order which implies that the Hamiltonian becomes ${\cal H}={\cal H}\left( t,x^{k},\dot{x}^{k},\ddot{x}^{k}\right)$. \subsection{Noether Theorems} Let $\mathcal{L}\left( t,x^{k},\dot{x}^{k}\right)$ be a function which describes the dynamics of a system. The equations of motion of the dynamical system follow from the action of the Euler Lagrange vector $E_{i}$ on the function $\mathcal{L}$, i.e.% \begin{equation} E_{i}\left( L\right) =0. \label{Ls.04} \end{equation}% where the Euler Lagrange vector is \begin{equation} E_{i}=\frac{d}{dt}\frac{\partial }{\partial \dot{x}^{i}}-\frac{\partial }{% \partial x^{i}} \end{equation} If the Lagrangian is invariant under the action of the transformation (\ref% {Ls.01}), namely $X^{\left[ 1\right] }\mathcal{L}=0$ then, it is easy to see that the Euler Lagrange equations (\ref{Ls.04}) are also invariant under the transformation (\ref{Ls.01}). In general we have the following theorem. \cite{StephaniB} \textbf{\emph{Theorem 1:}} \textit{Let \begin{equation} X=\xi \left( t,x^{k}\right) \partial _{t}+\eta ^{i}\left( t,x^{k}\right) \partial _{i} \label{Ls.05} \end{equation}% be the infinitesimal generator of the transformation (\ref{Ls.01}) and \begin{equation} \mathcal{L}=\mathcal{L}\left( t,x^{k},\dot{x}^{k}\right) \label{Ls.06} \end{equation}% be a Lagrangian describing the dynamical system (\ref{Ls.04}). The action of the transformation (\ref{Ls.01}) on (\ref{Ls.06}) leaves the Euler Lagrange equations (\ref{Ls.04}) invariant, if and only if there exist a function $g=g\left( t,x^{k}\right) $ such that the following condition holds \begin{equation} X^{\left[ 1\right] }L+L\frac{d\xi }{dt}=\frac{dg}{dt} \label{Ns.03} \end{equation}% where $X^{\left[ 1\right] }$ is the first prolongation of (\ref{Ls.05}). } If the generator of Eq.(\ref{Ls.05}) satisfies Eq.(\ref{Ns.03}) then the generator (\ref{Ls.05}) is a Noether symmetry of the dynamical system described by the Lagrangian (\ref{Ls.06}). Noether symmetries form a Lie algebra called the Noether algebra. We also have the result \textbf{\emph{Theorem 2:}} \textit{For any Noether symmetry (\ref{Ls.05}) of the Lagrangian (% \ref{Ls.06}) there corresponds a function $I\left( t,x^{k},\dot{x}% ^{k}\right) $ \begin{equation} I=\xi \left( \dot{x}^{i}\frac{\partial L}{\partial \dot{x}^{i}}-L\right) -\eta ^{i}\frac{\partial L}{\partial x^{i}}+g \label{Ls.08} \end{equation}% which is a first integral i.e. $\frac{dI}{dt}=0$. The function (\ref{Ls.08}) is called a Noether integral (first integral) of the dynamical system (\ref{Ls.04}).} \section{Noether Symmetries for $f\left( T\right) $ cosmology} In this section we apply the Noether symmetries approach to $f(T)$ cosmology in which the corresponding Lagrangian of the field equations is given by Eq.(\ref{H.52}). Here we consider a one parameter point transformation in the space $\left\{ t,a,T\right\}$ and the generator is written as \begin{equation*} X=\xi \left( t,a,T\right) \partial _{t}+\eta _{1}\left( t,a,T\right) \partial _{a}+\eta _{2}\left( t,a,T\right) \partial _{T} \;. \end{equation*}% Notice that the Lagrangian (\ref{H.52}) is a singular Lagrangian (the Hessian vanishes), hence the jet space is $\bar{B}_{M}=\left\{ t,a,T,\dot{a}\right\}$ and thus the first prolongation of $X$ in the jet space $\bar{B}_{M}~$is \begin{equation} X^{\left[ 1\right] }=\xi \partial _{t}+\eta _{1}\partial _{a}+\eta _{2}\partial _{T}+\eta _{1}^{\left[ 1\right] }\partial _{\dot{a}} \end{equation}% where $\eta _{1}^{\left[ 1\right] }=\dot{\eta}_{1}-\dot{a}\dot{\xi}$ \cite{IbraB,Havelkova, Ziping,Chris2}. Now we compute each term in the symmetry condition (\ref{Ns.03}). The term $X^{\left[ 1\right] }L$ gives% \begin{eqnarray*} X^{\left[ 1\right] }L &=&\left[ 3a^{2}\eta _{1}\left( f_{T}T-f\right) +a^{3}f_{TT}T\eta _{2}\right] + \\ &&+\left[ 12f_{T}a\eta _{1,t}\right] \dot{a}+\left[ 12f_{T}a\xi _{,a}\right] \dot{a}^{3}+ \\ &&+6\left[ f_{T}\eta _{1}+f_{TT}a\eta _{2}+2f_{T}a\eta _{1,a}-2f_{T}a\xi _{,t}\right] \dot{a}^{2}+ \\ &&+\left[ 12f_{T}a\eta _{1,T}\right] \dot{a}\dot{T}+\left[ 12f_{T}a\xi _{,T}% \right] \dot{a}^{2}\dot{T}. \end{eqnarray*} The second term $L\dot{\xi}$ gives% \begin{eqnarray*} L\dot{\xi} &=&\left[ a^{3}\left( f_{T}T-f\right) \xi _{,t}\right] +\left[ a^{3}\left( f_{T}T-f\right) \xi _{,a}\right] \dot{a}+ \\ &&+\left[ a^{3}\left( f_{T}T-f\right) \xi _{,T}\right] \dot{T}+\left[ 6f_{T}a\xi _{,t}\right] \dot{a}^{2}+ \\ &&+\left[ 6f_{T}a\xi _{,a}\right] \dot{a}^{3}+\left[ 6f_{T}a\xi _{,T}\right] \dot{a}^{2}\dot{T}. \end{eqnarray*}% Finally the rhs of Eq.(\ref{Ns.03}) is% \begin{equation*} \dot{g}=g_{,t}+g_{,a}\dot{a}+g_{,T}\dot{T}. \end{equation*} Replacing the results in Eq.(\ref{Ns.03}) and setting the terms with the powers of $ \dot{a}$ and $\dot{T}$ equal to zero in order to select the Lie vector (see \cite{Cap96} for details), we find the following set of Noether symmetry conditions% \begin{equation} \xi _{,a}=0~,~\xi _{,T}=0~,~\eta _{1,T}=0 \label{NC.01} \end{equation}% \begin{equation} a^{3}\left( f_{T}T-f\right) \xi _{,T}=g_{,T} \label{NC.02} \end{equation}% \begin{equation} 3a^{2}\eta _{1}\left( f_{T}T-f\right) +a^{3}f_{TT}T\eta _{2}+a^{3}\left( f_{T}T-f\right) \xi _{,t}=g_{,t} \label{NC.03} \end{equation}% \begin{equation} 12f_{T}a\eta _{1,t}+a^{3}\left( f_{T}T-f\right) \xi _{,a}=g_{,a} \label{NC.04} \end{equation}% \begin{equation} f_{T}\eta _{1}+f_{TT}Ta\eta _{2}+2f_{T}a\eta _{1,a}-f_{T}a\xi _{,t}=0 \label{NC.05} \end{equation} From equations (\ref{NC.01}), (\ref{NC.02}) it follows% \begin{equation*} \xi =\xi \left( t\right) ,~\eta _{1}=\eta _{1}\left( t,a\right) ~,~g=g\left( t,a\right) . \end{equation*}% Then Eq.(\ref{NC.04}) becomes $12f_{T}a\eta _{1,t}=g_{,a}$ Because $\eta _{1},g$ are independent of $T$ which follows that \begin{equation*} \eta _{1}=\eta _{1}\left( a\right) ~,~g=g\left( t\right) . \end{equation*} Dividing Eq.(\ref{NC.05}) with $af_{T}$ we find \begin{equation} 2\eta _{1,a}+\frac{\eta _{1}}{a}+\frac{f_{TT}}{f_{T}}\eta _{2}-\xi _{,t}=0 \label{NC.05a} \end{equation}% from which follows that \begin{equation*} \eta _{2}=\frac{f_{T}}{f_{TT}}S\left( a,t\right) \end{equation*}% where $S$ is an arbitrary function of its arguments. Taking this result into consideration the conditions (\ref{NC.03}) and (\ref{NC.05a}) become respectively \begin{equation} 2\eta _{1,a}+\frac{\eta _{1}}{a}+S\left( a,t\right) -\xi _{,t}=0 \label{NC.06} \end{equation}% \begin{equation} 3a^{2}\eta _{1}\left( f_{T}T-f\right) +a^{3}f_{TT}T~S+a^{3}\left( f_{T}T-f\right) \xi _{,t}=g_{,t}. \label{NC.07} \end{equation}% From Eq.(\ref{NC.06}) follows that $S\left( a,t\right) =M\left( a\right) +N\left( t\right) $ hence we have the final symmetry conditions (where $% f\neq e^{kT}$ $k=$constant):% \begin{equation} 2\eta _{1,a}+\frac{\eta _{1}}{a}+M+N-\xi _{,t}=0 \label{NC.06a} \end{equation} \begin{equation} 3\frac{\eta _{1}}{a}+\frac{f_{T}T}{f_{T}T-f}~M+\frac{f_{T}T}{f_{T}T-f}% ~~N+\xi _{,t}=\frac{1}{a^{3}\left( f_{T}T-T\right) }g_{,t}. \label{NC.07a} \end{equation} It is obvious that equations (\ref{NC.06}), (\ref{NC.07}) hold for arbitrary $f\left( T\right)$ as long as $\xi =c_{0}$ and $\eta _{1}=\eta _{2}=0$ (i.e. $S=0$). In this case the corresponding Noether integral is the Hamiltonian ${\cal H}$, implying that the dynamical system is autonomous. Moreover, the conditions (\ref{NC.06a}),(\ref{NC.07a}) give the following system of equations% \begin{equation} \frac{f_{T}T}{f_{T}T-f}=\frac{n}{n-1} \label{NC.08} \end{equation}% and% \begin{equation*} g_{,t}=0~~,~~N=c+\xi _{,t} \end{equation*}% \begin{equation*} 2\eta _{1,a}+\frac{\eta _{1}}{a}+M=c \end{equation*}% \begin{equation*} 3\frac{\eta _{1}}{a}+\frac{n}{n-1}M=m \end{equation*}% \begin{equation*} \frac{n}{1-n}N-\xi _{,t}=m \end{equation*}% Solving the first equation of the system (\ref{NC.08}) we find that \begin{equation} f\left( T\right)=f_{0}T^{n} \end{equation} where $f_{0}$ is the integration constant. In this context we can obtain the Nother symmetries. Specifically, in the case of $n\neq \frac{1}{2},\frac{3}{2}$, the Noether symmetry vector is \begin{eqnarray*} X_{1} &=&\left( \frac{3C}{2n-1}t\right) \partial _{t}+\left( Ca+c_{3}a^{1-% \frac{3}{2n}}\right) \partial _{a}+ \\ &&+\left[ \frac{1}{n}\left( \left( c-m\right) n+3c_{3}a^{-\frac{3}{2n}% }\right) +\frac{3C}{2n-1}+c\right] T\partial _{T} \end{eqnarray*}% as well as the corresponding Noether integral is \begin{equation*} I_{1}=\left( \frac{3C}{2n-1}t\right) {\cal H}-12f_{0}n\left( Ca^{2}+c_{3}a^{2-\frac{% 3}{2n}}\right) T^{n-1}\dot{a} \end{equation*}% where ${\displaystyle C=\frac{m\left( 1-n\right) +nc}{3}}$. For ${\displaystyle n=\frac{3}{2},}$ the Noether symmetry is given by \begin{eqnarray} X_{2} &=&\frac{1}{5}\left( 3c-2m\right) t\partial _{t}+\left[ \left( \frac{c% }{2}-\frac{m}{6}\right) a+c_{4}\right] \partial _{a}+ \notag \\ &&+\left[ \left( m+11c\right) -\frac{c_{4}}{a}+\frac{2}{5}\left( 8c-2m\right) \right] T\partial _{T} \end{eqnarray}% with corresponding Noether integral \begin{equation*} I_{2}=\frac{1}{5}\left( 3c-2m\right) t{\cal H}-18f_{0}\left[ \left( \frac{c}{2}-% \frac{m}{6}\right) a^{2}+c_{4}a\right] T^{\frac{1}{2}}\dot{a} \;. \end{equation*}% Finally for ${\displaystyle n=\frac{1}{2},}$ the Noether symmetry becomes \begin{equation*} X_{3}=c_{1}t\partial _{t}+\left( -2c_{1}+c_{3}a^{\frac{1}{4}}\right) \partial _{a}+\left( 4c_{1}+c_{2}+\frac{3c_{3}}{2}a^{-\frac{3}{4}}\right) T\partial _{T} \end{equation*}% and the Noether integral is \begin{equation*} I_{3}=c_{1}t{\cal H}-6f_{0}\left( -2c_{1}a+c_{3}a^{\frac{3}{4}}\right) T^{-\frac{1}{% 2}}\dot{a}. \end{equation*} We would like to stress that our results are in agreement with those of \cite{WeiHao} but they are richer because we have considered the term $\xi \partial _{t}$ in the generator which is not done in \cite{WeiHao}. To this end it becomes evident that $f\left( T\right)=f_{0}T^{n}$ is the only form that admits extra Noether symmetries implying the existence of exact analytical solutions (see next section). \section{Exact cosmological solutions} In this section we proceed in an attempt to analytically solve the basic cosmological equations of the $f\left( T\right)=f_{0}T^{n}$ gravity model. In particular from the Lagrangian (\ref{H.52}), we obtain the main field equation \begin{equation} \label{FF} \ddot{a}+\frac{1}{2a}\dot{a}^{2}+\frac{f^{\prime \prime }}{f^{\prime }}\dot{a% }\dot{T}-\frac{1}{4}a\frac{f^{\prime }T-f}{f^{\prime }}=0 \;. \end{equation} Also differentiating Eq.(\ref{H.50}) we find \begin{equation} \label{FFa} \dot{T}=12\left[\left( \frac{\dot{a}}{a}\right) ^{3}-\frac{\dot{a}\ddot{a}}{a^{2}} \right] \;. \end{equation} Finally, inserting $f\left( T\right)=f_{0}T^{n}$, Eq.(\ref{H.50}) and Eq.(\ref{FFa}) into Eq.(\ref{FF}) we derive, after some algebra, that \begin{equation} \left( 2n-1\right) \left[ \ddot{a}-\frac{\dot{a}^{2}}{2a}\frac{\left( 2n-3\right) }{n}\right] =0 \end{equation} a solution of which is \begin{equation} \label{aHE} a(t)=a_{0}t^{2n/3} \;\;\;\;H(t)=\frac{\dot{a}}{a}=\frac{2n}{3t} \end{equation} or \begin{equation} \label{HE} H=H_{0}a^{-3/2n}=H_{0}(1+z)^{3/2n} \end{equation} where $n\in {\cal R}^{\star}_{+}-\{\frac{1}{2}\}$, $a(z)=(1+z)^{-1}$ and $H_{0}$ is the Hubble constant in agreement with \cite{WeiHao}. Also using Eq.\eqref{HE} the deceleration parameter is given by \begin{equation}\label{eq:qnu} q=-1-\frac{{\rm dln}H}{{\rm dln}a}=-1+\frac{3}{2n} \;. \end{equation} From Eq.(\ref{aHE}) it is evident that this cosmological model has no inflection point. Therefore, the main drawback of the $f(T)=f_{0}T^{n}$ gravity model is that the deceleration parameter preserves sign, and therefore the universe always accelerates or always decelerates depending on the value of $n$. Indeed, if we consider $n=1$ (TEGR) then the above solution boils down to the Einstein de Sitter model as it should. On the other hand, the accelerated expansion of the universe ($q<0$) is recovered for $n>\frac{3}{2}$. The latter points that even if we would admit $n>\frac{3}{2}$ as a mere phenomenological possibility, we would be also admitting that the universe has been accelerating forever, which is of course difficult to accept. Now, we proceed to provide the growth factor of the $f(T)=f_{0}T^{n}$. In general, the basic equation which governs the evolution of the matter fluctuations in the linear regime is given by \begin{equation} \label{odedelta} \ddot{\delta}_{m}+ 2H\dot{\delta}_{m}-4 \pi G_{\rm eff} \rho_{m} \delta_{m}=0 \end{equation} where $\rho_{m}$ is the matter density and $G_{\rm eff}$ is the effective Newton's parameter which is written as \cite{Zheng:2010am} \begin{eqnarray} \label{Geff} G_{\rm eff}=\frac{G}{f^{\prime}(T)} \;. \end{eqnarray} Note, that $G$ denotes Newton's gravitational constant. On the other hand, using Eqs. (\ref{modfri}) and (\ref{rhoT}) one can easily write \begin{eqnarray} \label{reff} 4\pi G \rho_{m}=\frac{3H^{2}}{2}-4\pi G \rho_{T}= \frac{3H^{2}}{2}-\frac{2Tf^{\prime}(T)-f(T)-T}{4} \;. \end{eqnarray} Therefore, inserting Eqs.(\ref{H.50}), (\ref{Geff}) and (\ref{reff}) into Eq.(\ref{odedelta}) we have the following general equation \begin{equation} \label{odedelta1} \ddot{\delta}_{m}+ 2H\dot{\delta}_{m}+\frac{2Tf^{\prime}(T)-f(T)}{4f^{\prime}(T)}\delta_{m}=0 \;. \end{equation} We focus now on the $f(T)=f_{0}T^{n}$ gravity model. First of all for GR ($n=1$) we have $G_{\rm eff}=G$ and thus, without losing the generality, we can set $f_{0}=1$\footnote{If $f(T)=f_{0}T$ then the Newton's constant is just rescaled to be $G_{\rm eff}=G/f_{0}$ which is also constant in time. This result comes directly from the action (\ref{action}) (see also \cite{Zheng:2010am}).}. Therefore, Eq.(\ref{odedelta1}) becomes \begin{equation} \label{odedelta2} \ddot{\delta}_{m}+ \frac{4n}{3t}\dot{\delta}_{m}- \frac{2n(2n-1)}{3t^{2}} \delta_{m}=0 \;. \end{equation} Notice, that in order to derive Eq.(\ref{odedelta2}) we have utilized Eqs.(\ref{H.50}) and (\ref{aHE}). Interestingly, the above differential equation modifies that of the Einstein de-Sitter model in which $n=1$ (GR). From the mathematical point of view, Eq.(\ref{odedelta2}) is of Euler type whose general solution is \begin{equation} \delta_{m}(t)=C_{1}t^{2n/3}+C_{2}t^{1-2n} \end{equation} or \begin{equation} \delta_{m}(a)={\tilde C}_{1}a+{\tilde C}_{2}a^{3(1-2n)/2n} \end{equation} where ${\tilde C}_{1}=C_{1}/a_{0}^{3/2n}$ and ${\tilde C}_{2}=C_{2}/a_{0}^{3(1-2n)/2n}$. In the case of $0<n <\frac{1}{2}$ we have two growth factors while for $n>\frac{1}{2}$ the only growth factor is $D_{+}=a \propto t^{2n/3}$. It is interesting to mention that if we write the growth factor as a function of the scale factor then mathematically it coincides with that of the Einstein de-Sitter model \cite{Peeb93}. This result means that the growth rate of clustering $f_{+}(a)=d{\rm ln}D_{+}/d{\rm ln}a$ remains constant and equal to unity for every scale factor, implying that the present growth data disfavor the $f(T)=f_{0}T^{n}$ gravity. Indeed, in Fig.1 we plot the growth data as collected by Basilakos et al. (see \cite{BasNes13} and references therein) with the estimated growth rate function, $f_{+}(z)\sigma_{8}(z)$ [see $f(T)$ - solid line and $\Lambda$CDM - dashed line]. Notice, that the theoretical $\sigma_{8}(z)$ is given by $\sigma_{8}(z)=\sigma_{8}D_{+}(z)$, where $\sigma_{8}$ is the rms mass fluctuation on $R_{8}=8 h^{-1}$ Mpc scales at redshift $z=0$. \begin{figure} \mbox{\epsfxsize=14.2cm \epsffile{Fig1.ps}} \caption{Comparison of the observed (solid points) and theoretical evolution of the growth rate $f_{+}(z)\sigma_{8}(z)$. The solid and dashed lines correspond to $f(T)=f_{0}T^{n}$ and $\Lambda$CDM. As in Basilakos et al. \cite{BasNes13} we use $\sigma_{8}=0.8$ while for the $\Lambda$CDM case we set $\Omega_{m0}=0.272$.} \end{figure} \subsection{Cosmological analogue to other models} In this section (assuming flatness) we present the cosmological equivalence at the background level between the current $f(T)$ gravity with $f(R)$ modified gravity and dark energy, through a specific reconstruction of the $f(R)$ and vacuum energy density namely, $f(R)=R^{n}$ and $\Lambda(H)=3\gamma H^{2}$. In the case of $f(R)=R^{n}$ it has been found by Paliathanasis (see Appendix in \cite{Palia12}) that the corresponding scale factor obeys Eq.(\ref{aHE}), where $n\in {\cal R}^{\star}_{+}-\{2,\frac{3}{2},\frac{7}{8}\}$ \footnote{The Lagrangian here is ${\cal L}_{R}=6naR^{n-1}\dot{a}^{2}+6n(n-1)a^{2}R^{n-2}\dot{a}\dot{R}+(n-1)a^{3}R^{n}$, where $R$ is the Ricci scalar. For $n=1$ the solution of the Euler-Lagrange equations is the Einstein de-Sitter model [$a(t)\propto t^{2/3}$] as it has to be. Note, that for $n=2$ one can find a de-Sitter solution ($a(t) \propto e^{H_{0}t}$, see \cite{Palia12}).}. In \cite{prado1,prado2}, it has been shown that the particular model $f(R)\propto R^{3/2}$ has the cosmological solution ${\displaystyle a(t)=\sqrt{a_4 t^4 + a_3 t^3 + a_2 t^2 + a_1 t}}$ capable of addressing both dark-energy and dark-matter dominated phases. However, despite of the analogies, we have to point out that $f(R)$ gravity is a fourth-order theory while $f(T)$ gravity remains of second order. On the other hand, considering a spatially flat FLRW metric in the context of GR, the combination of the Friedmann equations with the total (matter+vacuum) energy conservation in the matter dominated era provides (for more details see \cite{BPS09}) \begin{equation}\label{LLAA} {\dot H}+\frac{3}{2}H^{2}=\frac{\Lambda}{2} \;. \end{equation} Solving Eq.(\ref{LLAA}) for $\Lambda(H)=3\gamma H^{2}$ (see Refs. \cite{FreeseET87,CarvalhoET92,ArcuriWaga94}) we end up with \begin{equation} \label{HE11} H=H_{0}a^{-3(1-\gamma)/2}=H_{0}(1+z)^{3(1-\gamma)/2} \;. \end{equation} Now, comparing Eqs.(\ref{HE}) and (\ref{HE11}) and connecting the above coefficients such as $n^{-1}=1-\gamma$, we find that the $f(T)=f_{0}T^{n}$ and the flat $\Lambda(H)=3\gamma H^{2}$ models can be viewed as equivalent cosmologies as far as the Hubble expansion is concerned, despite the fact that the current time varying vacuum model adheres to GR. However, if the $\Lambda(H)=3\gamma H^{2}$ cosmological model is confronted with the current observations provides a poor fit\,\cite{BPS09}. Since the current time varying vacuum model shares exactly the same Hubble parameter with the $f(T)=f_{0}T^{n}$ gravity model, this fact implies that the latter is also under observational pressure when we compare against the background cosmological data (SnIa, BAOs and CMB data). The same observational situation holds also for $f(R)=R^{n}$ modified gravity. \section{Conclusions} In this paper, we present a general study of Noether symmetries for $f(T)$ gravity and discuss the role of torsion and unholonomic frames in the context of teleparallel gravity and its straightforward extension. In particular, we point out the misunderstanding that when one works in an unholonomic frame, the torsion is introduced showing that this statement is not correct. The misunderstanding consists in the fact that the effects one observes in an unholonomic frame are frame dependent and not covariant effects. Therefore all conclusions made in a specific unholonomic frame must be restricted to that frame only. Coming to the specific {\it Noether Symmetry Approach}, this article extends the works by Basilakos et al. \cite{Basilakos11}, Paliathanasis et al. \cite{Tsafr} and Wei et al. \cite{WeiHao}. We confirm the result of \cite{WeiHao} that amongst the variety of $f(T)$ modified gravity theories, $f(T)=f_{0}T^{n}$ gravity admits Noether symmetries (integrals of motion). However, we provide here a more general family of Noether integrals with respect to that of \cite{WeiHao}. From the mathematical viewpoint the existence of extra integrals of motion points out the existence of further analytical solutions. Based on the $f(T)=f_{0}T^{n}$ models, we derive analytical solutions and thus we find the evolution of the main cosmological functions, namely the scale factor of the universe, the Hubble parameter, the deceleration parameter and for the first time to our knowledge the growth of matter fluctuations in the linear regime. Furthermore, we discuss the linear matter fluctuations from these background solutions. The analysis of the deceleration parameter points out that the $f(T)=f_{0}T^{n}$ gravity models include an intrinsic problem namely, the fact that the expansion of the universe always accelerates or always decelerates without spanning the different trends of cosmic evolution. Another basic problem is related to the fact that the growth rate of clustering is constant and always equal to unity which means that the present growth data cannot accommodate the $f(T)=f_{0}T^{n}$ gravity. As shown in \cite{aviles}, a robust cosmographic reconstruction of $f(T)$ cosmology needs more complicated models to address data. Finally, we find that flat $f(T)=f_{0}T^{n}$ cosmologically models are perfectly equivalent to the cosmic expansion history of the flat $f(R)=R^{n}$ modified gravity and the flat time varying vacuum model $\Lambda(H)=3\gamma H^{2}$ (where $n^{-1}=1-\gamma$), despite the fact that the three models live in a completely different geometrical background. This fact is a further indication of the high degeneracy problem affecting cosmological models capable of addressing the dark energy issue. \begin{acknowledgments} SB acknowledges support by the Research Center for Astronomy of the Academy of Athens in the context of the program ``{\it Tracing the Cosmic Acceleration}''. SC and MDL are supported by INFN (iniziative specifiche NA12 and OG51). \end{acknowledgments} \bigskip
\section{Introduction} Since the seminal work of Rytov and co-workers \cite{Rytov:1989ur}, it is known that thermal radiation has a different behaviour when the involved characteristic lengths are large or small compared to the thermal wavelength \cite{Joulain:2005ih,Volokitin:2007el,Dorofeyev:2011bg}. For example, the heat flux transferred between bodies separated by a subwavelength distance can exceed by far the one between black bodies \cite{Polder:1971uu,BenAbdallah:2010hp}. Energy density \cite{Shchegrov:2000td} and coherence properties \cite{Henkel:2000tr} are also strongly affected in the near field, especially close to materials exhibiting resonances such as polaritons. Knowing precisely how the electromagnetic field behaves close to a surface is therefore an important issue in order to address potential applications involving near-field heat transfer. From an experimental point of view, the coherence properties of near-field radiation have been utilized to produce directional and monochromatic thermal sources \cite{Greffet:2002ur,Lee:2006cj,Biener:2008cj}. The enhancement of radiative heat transfer at short distances has been demonstrated recently between two macroscopic surfaces \cite{Ottens:2011kh,Kralik:2012}, but probe microscopy techniques are still playing a prominent role \cite{Kittel:2005fr,Narayanaswamy:2008gj,Rousseau:2009es}. Near-field thermal flux imaging has been operated with a scanning thermal microscope \cite{Kittel:2008bc,Wischnath:2008hp}. A scanning near-field optical microscope (SNOM) without external illumination, termed thermal radiation scanning tunneling microscope (TRSTM), has also been used to image surfaces \cite{DeWilde:2006kt,Kajihara:2010fo,Kajihara:2011uu}. Very recently, local spectra have also been measured \cite{Babuty:dYmzb9en,Jones:2012fx}. Most of these experimental techniques use a small probe brought in the vicinity of the sample surface. Its response to the sample's near field is given by a polarizability. The induced multipoles are sources that radiate into the far field, thus providing the TRSTM signal. At short (sub-wavelength) distances however, the mutual interaction between the probe and the surface modifies the local electromagnetic field~\cite{GarciadeAbajo:2007eb,Intravaia:2010gp,BenAbdallah:2011be,Castanie:2011ue}, and this actually changes the probe's optical properties such as the polarizability. These interactions also complicate the data analysis for the near-field techniques mentioned above. In particular, one is often interested in the sample's optical properties, as encoded in the electromagnetic local density of states (EM-LDOS)~% \cite{Joulain:2003hc,Kittel:2008bc}. Due to the tip-sample interaction, it is no longer obvious how the TRSTM signal scattered by the tip into the far field is related to the EM-LDOS. In particular, can a SNOM detecting thermal radiation be the electromagnetic equivalent of the scanning tunneling microscope detecting the electronic LDOS \cite{Tersoff:1985wm}? Moreover, one can ask what information can be extracted from the exchanged heat flux between the probe and the sample. If some of these questions have already been addressed in the past~\cite{Joulain:2003hc,Mulet:2001kp}, our goal is here to clarify remaining interrogations. Following previous similar works~\cite{Knoll:2000wm,Sun:2007cl}, we will first see how the particle polarizability can be replaced by an effective or {\it dressed} polarizability taking into account multiple reflections between the probe and the surface. % This is more general than the image-dipole model~\cite{Knoll:2000wm} whose range of validity is very restricted in the infrared. We will then use the theory to calculate the signal detected in the far field when the near field is scattered by a probe dipole. An expression for the SNOM signal is calculated and illustrated by scanning a surface excited either by a plasmon or by broadband thermal radiation (TRSTM mode). In this paper, the probe tip is modeled by both electric and magnetic dipoles. This approach fails to capture field inhogeneities across the tip that occur at short distances (comparable to the tip size) and excite higher multipoles. It has the advantage, however, of providing relatively simple expressions that can be physically interpreted. The model has also been shown to reproduce the main physical ingredients in the case of a TRSTM tip \cite{Babuty:dYmzb9en}. It is powerful since the analysis of the results based on analytical expressions is straightforward. It is sure that a more accurate modeling of the tip, e.g. with a cone, would be better. However, the analysis of the various detected components (polarization, electric vs magnetic, etc) may be much more complicated in this case. The dipole approach could also be included as a building block into more flexible numerical schemes like the coupled dipole method~\cite{Keller:1993,Lax:1951tb,GarciadeAbajo:2007eb}, the multiple multipole method~\cite{Hafner:1999} or the discrete dipole approximation \cite{Yurkin:2007,Loke:2011} that has been very recently adapted to thermal near field radiation \cite{Edalatpour:2013ve}. It is also an alternative to more complicated but exact numerical methods such as surface-integral methods~\cite{Rodriguez:2011ki}. We conclude the paper by analyzing the signal detected in far field due to a heated probe when accounting for probe-surface interactions. Finally, the radiative cooling of a particle in the near field and the spectrum of the heat flux are analyzed. \section{Dressed polarizabilities} \label{s:dressed-polarizabilities} We propose here to calculate the {\it dressed} polarizability of a dipolar particle when it is placed in an environment which is different from free space. Indeed, when a particle is added to a system, the electromagnetic field present in the system illuminates the particle and induces a dipole moment (Fig.~\ref{systeme}). This dipole radiates a field everywhere that scatters back to the particle position. This interaction between the particle and the system modifies the total electromagnetic field, which is no longer given by the field in absence of the perturbing probe. In other words, the probe is no longer a passive test dipole. Our aim is to show that we can work with the unperturbed electromagnetic field if we ascribe to the particle a dressed polarizability. \begin{figure}[h] \centering \includegraphics[width=10cm]{Fig1.eps} \caption{Sketch of the system.} \label{systeme} \end{figure} Let us call ${\bf E}^0$ the electromagnetic field in the system without the particle (i.e., the ``non-perturbed field''). When a particle (tip) is placed in the system at position ${\bf r}_t$, an electric dipole ${\bf p}$ and a magnetic dipole ${\bf m}$ will be induced in the particle (tip). These dipoles radiate a field: the total field ${\bf E}^{tot}$ is the sum of ${\bf E}^0$ and of the field radiated by the dipoles. In the following, we use Green tensors to express the field radiated by a dipole: \begin{equation} \label{ } {\bf E}^{tot}({\bf r}) = {\bf E}^{0}({\bf r}) + {\stackrel{\leftrightarrow}{\bf G}}^{EE}({\bf r},{\bf r}_t)\cdot{\bf p} + {\stackrel{\leftrightarrow}{\bf G}}^{EH}({\bf r},{\bf r}_t)\cdot{\bf m} \end{equation} where ${\stackrel{\leftrightarrow}{\bf G}}^{EE}$ and ${\stackrel{\leftrightarrow}{\bf G}}^{EH}$ take into account the reflection (scattering) by the sample. This can also be written \begin{equation} \label{eloc} {\bf E}^{tot}({\bf r})={\bf E}^{0}({\bf r})+{\stackrel{\leftrightarrow}{\bf G}}^{EE}({\bf r},{\bf r}_t)\cdot\alpha{\bf E}^{tot}({\bf r}_t)+{\stackrel{\leftrightarrow}{\bf G}}^{EH}({\bf r},{\bf r}_t)\cdot\beta{\bf H}^{tot}({\bf r}_t) \end{equation} where $\alpha$ and $\beta$ are the ``bare'' electric and magnetic polarizabilities of the particle: they do not know about the surrounding sample and describe its reaction to the local field ${\bf E}^{tot}({\bf r}_t)$, ${\bf H}^{tot}({\bf r}_t)$. For a spherical particle, they are scalars. The generalization to anisotropic particles where $\alpha$ and $\beta$ become tensors is straightforward, see, e.g., Ref.\cite{Huth:2010fg}. Analogous expressions exist for the magnetic field \begin{equation} \label{} {\bf H}^{tot}({\bf r})={\bf H}^{0}({\bf r})+{\stackrel{\leftrightarrow}{\bf G}}^{HE}({\bf r},{\bf r}_t)\cdot{\bf p}+{\stackrel{\leftrightarrow}{\bf G}}^{HH}({\bf r},{\bf r}_t)\cdot{\bf m} \end{equation} that can also be written \begin{equation} \label{hloc} {\bf H}^{tot}({\bf r})={\bf H}^{0}({\bf r})+{\stackrel{\leftrightarrow}{\bf G}}^{HE}({\bf r},{\bf r}_t)\cdot\alpha{\bf E}_{tot}({\bf r}_t)+{\stackrel{\leftrightarrow}{\bf G}}^{HH}({\bf r},{\bf r}_t)\cdot\beta{\bf H}^{tot}({\bf r}_t) \end{equation} The Green tensors used here come in four types: ${\stackrel{\leftrightarrow}{\bf G}}^{EE}({\bf r},{\bf r}_t)$ gives the electric field at position ${\bf r}$ when an electric dipole source is placed at ${\bf r}_t$. In the same way, ${\stackrel{\leftrightarrow}{\bf G}}^{EH}({\bf r},{\bf r}_t)$ gives the electric field at position ${\bf r}$ when a magnetic dipole is placed at ${\bf r}_t$. ${\stackrel{\leftrightarrow}{\bf G}}^{HE}$ and ${\stackrel{\leftrightarrow}{\bf G}}^{HH}$ respectively give the magnetic field of an electric and a magnetic dipole. These Green tensors, which can all be calculated from ${\stackrel{\leftrightarrow}{\bf G}}^{EE}({\bf r},{\bf r}_t)$ \cite{Joulain:2010bq}, are the sum of a direct contribution (i.e., the Green tensor \emph{in vacuo}) and of a contribution due to scattering from the sample. The latter is labelled in the rest of the paper by a subscript $R$. When one considers the electromagnetic field at the particle position ${\bf r}_t$, the direct contribution leads to renormalized parameters (radiative line width, Lamb shift) that we suppose already included in $\alpha$, $\beta$, so that we may focus on the scattered Green tensors only. By solving the system (\ref{eloc}, \ref{hloc}), we find the local field in the form \begin{eqnarray} {\bf E}^{tot}({\bf r}_t) & = &{\stackrel{\leftrightarrow}{\bf A}}\ {\bf E}^0({\bf r}_t)+{\stackrel{\leftrightarrow}{\bf B}} \ {\bf H}^0({\bf r}_t) \label{elocfull} \\ {\bf H}^{tot}({\bf r}_t) & = & {\stackrel{\leftrightarrow}{\bf C}}\ {\bf E}^0({\bf r}_t)+{\stackrel{\leftrightarrow}{\bf D}}\ {\bf H}^0({\bf r}_t)\label{hlocfull} \end{eqnarray} where \begin{eqnarray} \label{eq:A-tensor} {\stackrel{\leftrightarrow}{\bf A}} & = & \left[[{\stackrel{\leftrightarrow}{\bf I}}-\alpha{\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)]-\alpha\beta{\stackrel{\leftrightarrow}{\bf G}}^{EH}_R({\bf r}_t,{\bf r}_t)[{\stackrel{\leftrightarrow}{\bf I}}-\beta{\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)]^{-1}{\stackrel{\leftrightarrow}{\bf G}}^{HE}_R({\bf r}_t,{\bf r}_t)\right]^{-1} \\ {\stackrel{\leftrightarrow}{\bf B}} & = &\left[-\alpha{\stackrel{\leftrightarrow}{\bf G}}^{HE}_R({\bf r}_t,{\bf r}_t)+\beta^{-1}[{\stackrel{\leftrightarrow}{\bf I}}-\beta{\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)][{\stackrel{\leftrightarrow}{\bf G}}^{EH}_R({\bf r}_t,{\bf r}_t)]^{-1}[{\stackrel{\leftrightarrow}{\bf I}}-\alpha{\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)]\right]^{-1}\\ {\stackrel{\leftrightarrow}{\bf C}} & = & \left[-\beta{\stackrel{\leftrightarrow}{\bf G}}^{EH}_R({\bf r}_t,{\bf r}_t)+\alpha^{-1}[{\stackrel{\leftrightarrow}{\bf I}}-\alpha{\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)][{\stackrel{\leftrightarrow}{\bf G}}^{HE}_R({\bf r}_t,{\bf r}_t)]^{-1}[{\stackrel{\leftrightarrow}{\bf I}}-\beta{\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)]\right]^{-1}\\ {\stackrel{\leftrightarrow}{\bf D}}& = & \left[[{\stackrel{\leftrightarrow}{\bf I}}-\beta{\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)]-\alpha\beta{\stackrel{\leftrightarrow}{\bf G}}^{HE}_R({\bf r}_t,{\bf r}_t)[{\stackrel{\leftrightarrow}{\bf I}}-{\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)]^{-1}{\stackrel{\leftrightarrow}{\bf G}}^{EH}_R({\bf r}_t,{\bf r}_t)\right]^{-1} \label{eq:D-tensor} \end{eqnarray} The preceding equations are fully general for the total field at the dipole position. Apart from the fact that the magnetic dipole is taken into account, the reasoning used here to obtain the total field (also known as ``self-consistent field'') is very similar to previous works using the so-called coupled dipole theory, see, e.g.~\cite{Keller:1993,Lax:1951tb,GarciadeAbajo:2007eb,Intravaia:2010gp,BenAbdallah:2011be}. Using the bare polarizabilities, the induced dipoles can be related to the non-perturbed fields at the tip position \begin{equation} \label{eq:def-dressed-alpha} \left(\begin{array}{c} {\bf p} \\ {\bf m} \end{array}\right)=\left(\begin{array}{cc} \alpha{\stackrel{\leftrightarrow}{\bf A}} & \alpha{\stackrel{\leftrightarrow}{\bf B}} \\ \beta{\stackrel{\leftrightarrow}{\bf C}} & \beta{\stackrel{\leftrightarrow}{\bf D}} \end{array}\right)\left(\begin{array}{c} {\bf E}^0( {\bf r}_t ) \\ {\bf H}^0( {\bf r}_t ) \end{array}\right) \end{equation} The four sub-matrices can be seen as dressed polarizabilities that depend on the particle position ${\bf r}_t$; they will be anisotropic in general. We now highlight the case of a simple system made of a planar interface separating a material and vacuum. The corresponding Green tensors are well known \cite{Sipe:1987td,Joulain:2010bq} and depend on the material's optical properties. In this plane-parallel geometry with $\hat{\bf z} = (0, 0, 1)^T$ normal to the interface, it is convenient to bring the tensors into block-diagonal form by identifying suitable sub-spaces, in particular to perform the inversions in Eqs.(\ref{eq:A-tensor}--\ref{eq:D-tensor}). The resulting forms are \begin{equation} \label{ } {\stackrel{\leftrightarrow}{\bf G}}^{HE}_R({\bf r}_t,{\bf r}_t)=\mu_0{\stackrel{\leftrightarrow}{\bf G}}^{EH}_R({\bf r}_t,{\bf r}_t)=\left(\begin{array}{ccc} 0 & a & 0 \\ -a & 0 & 0 \\0 & 0 & 0\end{array}\right) \end{equation} \begin{equation} \label{eq:ee-and-hh-Green} {\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)=\left(\begin{array}{ccc} b & 0 & 0 \\ 0 & b & 0 \\0 & 0 & C\end{array}\right), \qquad {\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)=\left(\begin{array}{ccc} d & 0 & 0 \\ 0 & d & 0 \\0 & 0 & f\end{array}\right) \end{equation} with matrix elements \begin{eqnarray} a & = & \frac{\omega}{8\pi} \int_0^\infty KdK(r^s-r^p)e^{2i\gamma z_t} \label{eq:HExy-Green} \\ b & = & \frac{i\mu_0\omega^2}{8\pi}\int_0^\infty\frac{KdK}{\gamma}(r^s-r^p\frac{\gamma^2}{k_0^2})e^{2i\gamma z_t} \\ C & = & \frac{i\mu_0\omega^2}{4\pi}\int_0^\infty\frac{K^3dK}{\gamma k_0^2}r^pe^{2i\gamma z_t}\\ d & = & \frac{i\omega^2}{8\pi c^2}\int_0^\infty\frac{KdK}{\gamma}(r^p-r^s\frac{\gamma^2}{k_0^2})e^{2i\gamma z_t} \label{eq:HHxx-Green}\\ f & = & \frac{i\omega^2}{4\pi c^2}\int_0^\infty\frac{K^3dK}{\gamma k_0^2}r^se^{2i\gamma z_t} \label{eq:HHzz-Green} \end{eqnarray} In these integrals over plane waves, rotational symmetry in the $xy$-plane has been exploited: the wave vector projected onto the surface has length $K$, its perpendicular component is $\gamma=\sqrt{k_0^2-K^2}$ and $k_0 = \omega / c$ is the wavenumber \emph{in vacuo}. $r^s$ and $r^p$ denote the Fresnel reflection amplitudes at the vacuum-material interface in the principal polarizations s (or TE) and p (TM). The expression~(\ref{eq:def-dressed-alpha}) for the electric and magnetic dipoles becomes in terms of the dressed polarizabilities \begin{eqnarray} {\bf p}({\bf r}_t) & = & {\stackrel{\leftrightarrow}{\bf \alpha}}^{EE}{\bf E}^0({\bf r}_t)+{\stackrel{\leftrightarrow}{\bf \alpha}}^{EH}{\bf H}^0({\bf r}_t) \\ {\bf m}({\bf r}_t)& = & {\stackrel{\leftrightarrow}{\bf \beta}}^{HE}{\bf E}^0({\bf r}_t) + {\stackrel{\leftrightarrow}{\bf \beta}}^{HH}{\bf H}^0({\bf r}_t) \end{eqnarray} and the latter read as follows in the our planar setting: \begin{eqnarray} \label{alpheeff} {\stackrel{\leftrightarrow}{\bf \alpha}}^{EE} &=& \left(\begin{array}{ccc} \frac{\alpha}{1-\alpha b+\frac{\mu_0\alpha\beta a^2}{(1-\beta d)}} & 0 & 0 \\ 0 & \frac{\alpha}{1-\alpha b+\frac{\mu_0\alpha\beta a^2}{(1-\beta d)}} & 0 \\0 & 0 & \frac{\alpha}{1-\alpha C}\end{array}\right) \\ \label{} {\stackrel{\leftrightarrow}{\bf \alpha}}^{EH} &=& \left(\begin{array}{ccc} 0 & \frac{\mu_0\alpha\beta a}{(1-\alpha b)(1-\beta d)-\mu_0\alpha\beta a^2} & 0 \\ -\frac{\mu_0\alpha\beta a}{(1-\alpha b)(1-\beta d)-\mu_0\alpha\beta a^2} & 0& 0 \\0 & 0 & 0\end{array}\right) = \mu_0 {\stackrel{\leftrightarrow}{\bf \beta}}^{HE}, \\ \label{betaeff} {\stackrel{\leftrightarrow}{\bf \beta}}^{HH} &=&\left(\begin{array}{ccc} \frac{\beta}{1-\beta d+\frac{\mu_0\alpha\beta a^2}{(1-\alpha b)}} & 0 & 0 \\ 0 & \frac{\beta}{1-\beta d+\frac{\mu_0\alpha\beta a^2}{(1-\alpha b)}} & 0 \\0 & 0 & \frac{\beta}{1-\beta f}\end{array}\right). \end{eqnarray} Due to multiple reflections between particle and interface, these polarizabilities are anisotropic tensors. Note the dimensionless parameters $\alpha b$ and $\alpha C$ for the electric case, and $\beta d$ and $\beta f$ for the magnetic one. We shall see below that an upper limit to these parameters scales, in order of magnitude, as $(R_t/z_t)^3$ where $R_t$ is the radius of the probe particle. The dipole approximation is valid when the distance $z_t$ is sufficiently large compared to the radius. Otherwise the image field would be significantly inhomogeneous across the particle volume, inducing quadrupole and higher multipoles. We therefore restrict to $(R_t/z_t)^3 \le 1/8 \ll 1$, so that the anisotropy of ${\stackrel{\leftrightarrow}{\bf \alpha}}^{EE}$ is weak. The magneto-electric cross-polarizabilities ${\stackrel{\leftrightarrow}{\bf \alpha}}^{EH}$, ${\stackrel{\leftrightarrow}{\bf \beta}}^{HE}$ scale with the parameter $\mu_0 \alpha \beta a^2$ which is typically small for the same reason. We see below that we can safely neglect them. In the plots of the following Section, we compare the different polarizabilities after normalizing them by the volume $V = 4\pi R_t^3/3$ of the particle, namely ${\stackrel{\leftrightarrow}{\bf \alpha}}^{EE} / (\epsilon_0 V)$, $c {\stackrel{\leftrightarrow}{\bf \alpha}}^{EH} / V$, $\mu_0 c {\stackrel{\leftrightarrow}{\bf \beta}}^{HE} / V$, and ${\stackrel{\leftrightarrow}{\bf \beta}}^{HH} / V$. \section{Parametric study of the dressed polarizabilities} In this section, we analyze how the dressed polarizabilities depend on the material's optical properties, particle sizes and particle-surface distance. For a spherical particle, the bare polarizabilities are well established in Mie theory. They involve the ratio of the particle radius $R_t$ to the wavelength $\lambda$ (Mie parameter $x = 2\pi R_t / \lambda = k_0 R_t$) and the corresponding ratio $y = x \sqrt{\epsilon( \omega )}$ inside the particle, whose dielectric function is $\epsilon( \omega )$. For a non-magnetic material ($\mu=1$), the following expressions given by Chapuis \& al.~\cite{Chapuis:2008kcb} apply if the wavelength is much larger than $R_t$: \begin{equation} \label{eq:Mie-alpha} \alpha(\omega)=\epsilon_02\pi R_t^3\frac{2\left[\sin(y)-y\cos(y)\right]-x^2\left[\frac{-\sin(y)}{y^2}+\frac{\cos(y)}{y}+\sin(y)\right]}{\left[\sin(y)-y\cos(y)\right]+x^2\left[\frac{-\sin(y)}{y^2}+\frac{\cos(y)}{y}+\sin(y)\right]} \end{equation} and \begin{equation} \label{eq:Mie-beta} \beta(\omega)=-2\pi R_t^3\left[\left(1-\frac{x^2}{10}\right)+\left(-\frac{3}{y^2}+\frac{3}{y}\cot (y)\right)\left(1-\frac{x^2}{6}\right)\right] \end{equation} For metallic particles or near a resonance of $\epsilon( \omega )$, the parameter $y$ can be of order unity, but we always assume $x \ll 1$. The next order in the multipole series is smaller by a factor $x^{2}$~\cite{VanDeHulst:1981}. A discussion of the validity of the dipole approximation can be found in Sec.\ref{s:check-dipole-approx} below. We simply note here that our aim is to compare the theory to apertureless Scanning Near-field Optical Microscopy (SNOM) measurements where scattering probes with a size of one micron or smaller were used, clearly smaller than the wavelength in the near infrared range (whence $x \ll 1$). The partial screening of the field inside the particle (skin effect) is in fact taken into account in the polarizabilities~(\ref{eq:Mie-alpha}, \ref{eq:Mie-beta}) via the parameter $y$, essentially the ratio between particle size and skin depth in the material. Note that our work here is limited to particles with an isotropic polarizability, but can be generalized straightforwardly. For the more complicated case of spheroids, see Refs.\cite{Biehs:2010kp,Huth:2010fg}. Since many experimental devices use tungsten tips, we consider tungsten as material. We study the dressed polarisabilities for two spherical tip sizes (100 nm and 500 nm radii) and above three materials: SiC and SiO2, being both dielectrics, and gold. The dielectric functions are taken from tabulated data~\cite{Palik:1985}. \subsection{Contributions to the dressed polarizabilities} Let us consider a tungsten sphere of 100 nm radius located 200 nm above the sample surface. This configuration allows to observe multiple interactions with the surface without being in a distance regime where multipolar interactions are expected. In the case of a dipolar particle close to a plane interface, five different contributions to the polarizabilities are identified: the parallel electric ($\alpha^{EE}_{xx}$), perpendicular electric ($\alpha^{EE}_{zz}$), parallel magnetic ($\beta^{HH}_{xx}$), perpendicular magnetic ($\beta^{HH}_{zz}$) and crossed polarizability ($\alpha^{EH}_{xy}$ or $\beta^{HE}_{xy}$). The first four contributions are plotted in Fig.~\ref{aleff100200}. The crossed polarizability is not represented since it is two orders of magnitude smaller than the other contributions. \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig2.eps} \caption{Dressed polarizability tensor of a 100nm-radius tungsten sphere above a plane surface of SiC, SiO$_{2}$ or gold. The sphere center is at a distance $z = 200\,{\rm nm}$ from the surface. Panels (a,b): electric polarizability ${\stackrel{\leftrightarrow}{\bf \alpha}}^{EE}$, parallel and perpendicular components. Panels (c,d): magnetic polarizability ${\stackrel{\leftrightarrow}{\bf \beta}}^{HH}$. All curves show absolute values, normalized to the volume of the particle. The SiC surface shows a surface phonon polariton resonance at $948\,{\rm cm}^{-1}$. } \label{aleff100200} \end{center} \end{figure} We observe that both perpendicular and parallel dressed electrical polarizabilities are quite different from the bare polarizability of the single particle. In particular, resonant features appear close to the frequencies of the SiC and SiO$_2$ surface phonon polariton mode. This is not surprising since the single-interface Green tensors appear in the dressed polarizabilities. The latter involve the reflection coefficients which diverge at the plasmon resonance. This is particularly pronounced at small distances from the interface. Thus the dressed polarizabilities are greatly affected in the near field and close to surface resonances. In contrast, no peak is seen in the spectra of gold dressed polarizabilities since gold does not exhibit resonances in the studied frequency range (mid-infrared). Very differently from the electric case, the magnetic polarizability is only weakly modified by the dressing. The main reason is that the bare magnetic polarizability shows a different scaling at small radius $R_t$. While the electrical polarizability behaves as $R_t^3$ (Clausius-Mossotti limit), the magnetic is proportional to $(R_{p}^{5} / \lambda^{2}) \left(\epsilon -1 \right)$ \cite{Chapuis:2008kcb}. In the present case, as the particle radius is much smaller than the wavelength, the magnetic polarizability is smaller; this implies that the dressing correction is also smaller, in particular for the dielectrics SiC and SiO$_2$. The correction is much more significant for gold because of its large dielectric constant in the infrared. This is a consequence of the fact that magnetic near fields are stronger at metals than above dielectrics. Let us now study the case of a larger particle (500 nm) at $1\,\mu{\rm m}$ above the sample. The dressed polarizabilities are represented in Fig. \ref{aleff5001000}. Here again crossed polarizabilities are not represented since they are smaller by two orders of magnitude. We note significant corrections for both dressed electric and magnetic polarizabilities. These corrections are once again quite prominent around surface resonance frequencies. The impact on the magnetic polarizabilities is now much more striking than for small particles due to the fact that here $R_t / \lambda$ is larger. We will see in the next section that the different behavior at $1\,\mu{\rm m}$ is due to the onset of retardation. This can be noticed by analyzing carefully the dressed polarizabilities. \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig3.eps} \caption{ Dressed polarizabilities. Same as Fig.~\ref{aleff100200}, but for a larger particle: radius $R_t = 500\,{\rm nm}$, distance $z = 1\,\mu{\rm m}$. } \label{aleff5001000} \end{center} \end{figure} \subsection{Asymptotic expressions} We now simplify the dressed polarizabilities in order to get simpler expressions valid in the near field regime. We first start with Green tensors ${\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)$ and ${\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)$ that appear as integrals over the parallel wavevector [Eqs.(\ref{eq:ee-and-hh-Green}--\ref{eq:HHzz-Green})]. We suppose that the distance $z_t$ is much smaller than the wavelength, so that the exponential $e^{ 2 i \gamma z_t }$ allows for large values of $K$ to contribute to the integrals~(\ref{eq:HExy-Green}--\ref{eq:HHzz-Green}). This is valid in a regime sometimes called ``extreme near field"~\cite{Henkel:2000tr}. In this regime, $r^p \approx (\epsilon - 1) / (\epsilon + 1)$ and $r^s \approx (\epsilon - 1) k_0^2 / 4 K^2$. Integration over the parallel wavevector $K$ is then easy since $\gamma^2 = k_0^2 - K^2 \approx -K^2$ so that $e^{ 2 i \gamma z_t } \approx e^{ - 2 K z_t }$. For corrections to this regime that may even appear in the near field, see Ref.\cite{Chapuis:2008kca}. Adopting the extreme near field regime, the Green tensors become \begin{equation} \label{ } {\stackrel{\leftrightarrow}{\bf G}}^{EE}_R({\bf r}_t,{\bf r}_t)\approx\frac{1}{32\pi \epsilon_0 \, z_t^3} \frac{\epsilon-1}{\epsilon+1} \left(\begin{array}{ccc} 1 & 0 & 0 \\0 & 1 & 0 \\0 & 0 & 2 \end{array}\right) \end{equation} and \begin{equation} \label{ } {\stackrel{\leftrightarrow}{\bf G}}^{HH}_R({\bf r}_t,{\bf r}_t)=\frac{k_0^2}{8\pi z_t}\left(\begin{array}{ccc}\frac{1}{2}\left(\frac{\epsilon-1}{4} + \frac{\epsilon-1}{\epsilon+1}\right)& 0 & 0 \\0 & \frac{1}{2}\left(\frac{\epsilon-1}{4} + \frac{\epsilon-1}{\epsilon+1}\right) & 0 \\0 & 0 &\frac{\epsilon-1}{4}\end{array}\right) \end{equation} These expressions are the asymptotic approximations of the Green tensors. For the electric term, it is also known as the electrostatic limit, which means that retardation is not taken into account. The dressed polarizabilities hence take the form: \begin{eqnarray} \label{alphaeffapprox} {\stackrel{\leftrightarrow}{\bf \alpha}}^{EE} &\approx& \left(\begin{array}{ccc}\frac{\alpha}{1-\frac{\alpha(\epsilon-1)}{32\epsilon_0\pi z_t^3(\epsilon+1)}} & 0 & 0 \\0 & \frac{\alpha}{1-\frac{\alpha(\epsilon-1)}{32\epsilon_0\pi z_t^3(\epsilon+1)}} & 0 \\0 & 0 &\frac{\alpha}{1-\frac{\alpha(\epsilon-1)}{16\epsilon_0\pi z_t^3(\epsilon+1)}}\end{array}\right) \\ \label{eq:beta-approx} {\stackrel{\leftrightarrow}{\bf \beta}}^{HH} &\approx& \left(\begin{array}{ccc}\frac{\beta}{1-\frac{\beta k_0^2}{16\pi z_t}\left(\frac{\epsilon-1}{4} + \frac{\epsilon-1}{\epsilon+1}\right)} & 0 & 0 \\0 & \frac{\beta}{1-\frac{\beta k_0^2}{16\pi z_t}\left(\frac{\epsilon-1}{4} + \frac{\epsilon-1}{\epsilon+1}\right)} & 0 \\0 &0 &\frac{\beta}{1 - \frac{\beta k_0^2}{32\pi z_t}(\epsilon-1)}\end{array}\right) \end{eqnarray} This approximation for the electric case is identical to previous work of Knoll and Keilmann\cite{Knoll:2000wm}. In Fig.~\ref{compkeil}, the example of a tungsten spherical particle above SiO$_2$ shows that the electrostatic approximation is very good for small particles at short distances. For a larger particle ($1\,\mu{\rm m}$ radius), deviations are visible between the full calculation and the extreme near field (electrostatic) approximation. This is due to retardation that is clearly not negligible even though we are at subwavelength distances here. From Eq.(\ref{eq:beta-approx}), the dressing correction to the magnetic polarizability vanishes if retardation is discarded due to the factor $k_0^2$ in the denominators. This is illustrated in Fig.~\ref{compkeil} for a small particle $R_t = 100\,{\rm nm}$. The extreme near field approximation breaks down as we take a radius $R_t = 500\,{\rm nm}$ because of the scaling $\sim R_t^5$ of the polarizability. The calculations of the following sections are therefore based on the full Green tensors, as given by the integrals in Eqs.(\ref{eq:HHxx-Green},\ref{eq:HHzz-Green}). \begin{figure} \begin{center} \includegraphics[width=7cm]{Fig4a.eps} \includegraphics[width=7cm]{Fig4b.eps} \end{center} \caption{Accuracy of the electrostatic approximation (extreme near field limit, Eqs.(\ref{alphaeffapprox}, \ref{eq:beta-approx})) for the dressed polarizabilities. Left panel: small sphere with 100nm radius at $z=200\,{\rm nm}$. Right panel: 500 nm radius at $z = 1\,\mu{\rm m}$. The panels (a, b) give the parallel and perpendicular components of the electric polarizability, panel (c) the magnetic polarizability. In all cases, the values are normalized to the particle volume and plotted in absolute value. The sphere is made from tungsten, the surface is SiO$_2$. } \label{compkeil} \end{figure} \section{Radiation scattered into the far field} The spherical particles studied so far provide a simple model for a SNOM tip. This kind of setup aims at detecting the near field of a sample, at a distance much smaller than the wavelength. The tip scatters this field and converts it into radiation that propagates to a detector placed in the far field. See Fig.~\ref{detscheme} for a sketch of the geometry where the emission is collected around a direction parametrized by the spherical coordinates $\theta, \varphi$. \begin{figure} \begin{center} \includegraphics[width=6.5cm]{Fig6a.eps} \includegraphics[width=6.5cm]{Fig6b.eps} \caption{Detection system scheme (left). Tip and detector position (right).} \label{detscheme} \end{center} \end{figure} \subsection{Signal at the detector} The local electromagnetic field ${\bf E}^0( {\bf r}_t ), {\bf H}^0( {\bf r}_t )$ at the tip induces in the latter electric and magnetic dipole moments, as described by the polarizabilites discussed before. The far field at the detector (distance $R$ from the tip) is just the electromagnetic radiation of these dipoles, taking into account the reflection at the interface. The signal is calculated in the far-field (Fraunhofer) approximation when $R$ is large enough compared to the size $R_d$ of the scatterer: $R \gg R_d^2 / \lambda$. In this limit, the field at the detector can be considered a plane wave so that its power (averaged over one period) becomes \begin{equation} \label{signal1} \left< S^d(\omega) \right> =\frac{\epsilon_0c}{2}|{\bf E}^d( \omega )|^2 r^2 d\Omega \end{equation} where $d\Omega$ is the solid angle subtended by the detector. Let us introduce the unit vector ${\bf u}_d$ pointing from the tip to the detector and decompose it as ${\bf u}_d = \hat{\bf u}_{d\parallel} \sin\theta + \hat{\bf z} \cos\theta$ where $\hat{\bf z}$ is the outward unit normal to the surface and the angle $\theta$ is shown in Fig.~\ref{detscheme}. We define ${\bf u}_{d}^{-} = \hat{\bf u}_{d\parallel} \sin\theta - \hat{\bf z} \cos\theta$ (the direction of a ray from tip to surface before reflection) and the (non-normalized) polarization vectors $\hat{s}_d = \hat {\bf u}_{d\parallel}\times{\bf e}_z$, $\hat{p}^+_d = {\bf u}_d\times\hat{s}_d$, and $\hat{p}^-_d = -{\bf u}_d^-\times\hat{s}_d$. These vectors and the Fresnel coefficients $r^{s,p}( \theta )$, evaluated for waves at the angle $\theta$, are the building blocks for two tensors ${\stackrel{\leftrightarrow}{\bf \Gamma}}^{E,H}( {\bf u}_d )$ that give the classical expression for the detector field ${\bf E}^d$: \begin{eqnarray} {\stackrel{\leftrightarrow}{\bf \Gamma}}^E({\bf u}_d) &=& \hat{s}_d\hat{s}_d+\hat{p}^+_d\hat{p}^+_d + (\hat{s}_d r^s( \theta ) \hat{s}_d+\hat{p}^+_d r^p( \theta ) \hat{p}^-_d)e^{i \phi} \\ {\stackrel{\leftrightarrow}{\bf \Gamma}}^H({\bf u}_d) &=& -\hat{s}_d\hat{p}^+_d+\hat{p}^+_d\hat{s}_d + (-\hat{s}_d r^s( \theta ) \hat{p}^-_d+\hat{p}^+_d r^p( \theta ) \hat{s}_d) e^{i \phi} \\ {\bf E}^d & = & \frac{\mu_0\omega^2}{4\pi}\frac{e^{ikR}}{R} \left[ {\stackrel{\leftrightarrow}{\bf \Gamma}}^E( {\bf u}_d ) {\stackrel{\leftrightarrow}{\bf \alpha}}^{EE}{\bf E}^0({\bf r}_t)\nonumber + \frac{ 1 }{ c } {\stackrel{\leftrightarrow}{\bf \Gamma}}^H( {\bf u}_d ) {\stackrel{\leftrightarrow}{\bf \beta}}^{HH}{\bf H}^0({\bf r}_t) \right] \end{eqnarray} where $\phi = 2 k_0 z_t \cos\theta$ is the phase difference between the `direct ray' from tip to detector and the ray once reflected at the surface. We thus find the following expression for the detector signal~(\ref{signal1}): \begin{equation} \label{eq:signal-and-Gamma} \left< S^d( \omega ) \right> = \frac{\mu_0\omega^4d\Omega}{32\pi^2c} \sum_{i,j,k} \left( \Gamma^E_{ij} \alpha^{EE}_{jj} E^0_j({\bf r}_t) + \frac{ 1 }{ c } \Gamma^H_{ij} \beta^{HH}_{jj} H^0_j({\bf r}_t) \right) \left( \Gamma^{E*}_{ik} \alpha^{EE*}_{kk} E^{0*}_k({\bf r}_t) + \frac{ 1 }{ c } \Gamma^{H*}_{ik} \beta^{HH*}_{kk} H^{0*}_k({\bf r}_t) \right) \end{equation} where we have used that the polarizabilities are diagonal, even when dressed. This is a bilinear combination of electromagnetic field components, and obviously not simply proportional to the electromagnetic energy density at the tip. \subsection{Expressions of the tensors ${\stackrel{\leftrightarrow}{\bf \Gamma}}^E$ and ${\stackrel{\leftrightarrow}{\bf \Gamma}}^H$} For definiteness, we consider a situation where the detector is placed in the $yz$ plane, at an angle $\theta$ to the surface normal, see Fig.~\ref{detscheme}. The tensors in Eq.(\ref{eq:signal-and-Gamma}) then become \begin{equation} \label{ } {\stackrel{\leftrightarrow}{\bf \Gamma}}^E=\left(\begin{array}{ccc} 1+r^s(\theta)e^{i \phi} & 0 & 0\\ 0 & \cos^2\theta(1-r^p(\theta)e^{i \phi}) & -\sin\theta\cos\theta(1+r^p(\theta)e^{i \phi}) \\ 0 & -\sin\theta\cos\theta(1-r^p(\theta)e^{i \phi}) & \sin^2\theta(1+r^p(\theta)e^{i \phi}) \end{array}\right) \end{equation} and \begin{equation} \label{ } {\stackrel{\leftrightarrow}{\bf \Gamma}}^H=\left(\begin{array}{ccc} 0 & \cos\theta(1-r^s(\theta)e^{i \phi}) & -\sin\theta(1+r^s(\theta)e^{i \phi})\\ -\cos\theta(1+r^p(\theta)e^{i \phi}) & 0 & 0 \\ \sin\theta(1+r^p(\theta)e^{i \phi}) &0 & 0\end{array}\right) \end{equation} where the interference between the direct and reflected rays is manifest. We therefore expect to see a signal that oscillates when the distance $z_t$ is comparable to the wavelength. \section{Application to apertureless SNOM experiments} \subsection{Coherently excited surface polariton} We now calculate the signal detected by an apertureless SNOM above a material supporting surface modes. Typical materials are metals \cite{Raether:1988ty} and more generally all materials with a dielectric constant smaller than $-1$. Surface polaritons are electromagnetic modes bound to a planar surface and only exist for $p$ (or TM) polarization. They can be found by looking for a pole of the reflection amplitude $r^p$. In the coordinates of Fig.~\ref{detscheme}, the magnetic and electric fields of the surface mode propagating in the $y$-direction are given by \begin{equation} \label{ } {\bf H}^0({\bf r}) = H_0 \left(\begin{array}{c}1 \\0 \\0\end{array}\right) e^{i(K y + \gamma z)} ,\qquad {\bf E}^0({\bf r}) = \frac{ H_0 }{ \omega\epsilon_0 } \left(\begin{array}{c}0 \\ -\gamma \\ K \end{array}\right) e^{i(Ky + \gamma z)} \end{equation} where the pair $(K, \omega)$ satisfies the surface mode dispersion relation~\cite{Raether:1988ty}. This leads to $\Im \gamma > 0$ (evanescent mode) and a complex $K$ if we force $\omega$ to be real. The detector signal becomes \begin{eqnarray} \left<S^d( \omega )\right> & = & \frac{\mu_0 \omega^4 d\Omega }{ 32\pi^2 c^3 } |H_0|^2e^{-2 \Im(K) y_t}e^{-2 \Im( \gamma ) z_t} \label{sigdet} \\ && {} \times \left[ \cos^2\theta|1-r^p(\theta)e^{i \phi}|^2 |\alpha^E_{yy}|^2 \frac{| \gamma |^2}{ k_0^2 } + \sin^2\theta|1+r^p(\theta) e^{i \phi}|^2 |\alpha^E_{zz}|^2 \frac{|K|^2}{ k_0^2 }\right. \nonumber\\ && {} + |1+r^p(\theta)e^{i \phi}|^2|\beta^H_{xx}|^2 \nonumber\\ && {} + 2\sin\theta\cos\theta \, \Re\left( (1-r^p(\theta)e^{i \phi})(1+r^{p*}(\theta)e^{-i \phi}) \alpha^E_{yy}\alpha^{E*}_{zz} \frac{ \gamma K^* }{ k_0^2 } \right) \nonumber\\ && {} + 2\cos\theta \, \Re\left( (1-r^p(\theta)e^{i \phi})(1+r^{p*}(\theta)e^{-i \phi})\alpha^E_{yy}\beta^{H*}_{xx} \frac{ \gamma }{ k_0 } \right) \nonumber\\ && {} + \left. 2\sin\theta \, \Re\left( |1+r^p(\theta)e^{i \phi}|^2\alpha^E_{zz}\beta^{H*}_{xx}\frac{ K }{ k_0 } \right) \right] \nonumber \end{eqnarray} where ${\stackrel{\leftrightarrow}{\bf \alpha}}^{E} = {\stackrel{\leftrightarrow}{\bf \alpha}}^{EE} / \epsilon_0$, ${\stackrel{\leftrightarrow}{\bf \beta}}^{H} = {\stackrel{\leftrightarrow}{\bf \beta}}^{HH}$, both with dimension volume. The detector signal from Eq.(\ref{sigdet}) is plotted in Fig.~\ref{Theoplasm} as a function of distance $z_t$ (dashed curve). A $1\,\mu{\rm m}$-radius tungsten sphere is taken as a model for a SNOM tip. The oscillations in the lower panel illustrate the interference between the direct and reflected rays. The period is indeed around half the excitation wavelength $\lambda / 2 = 3.75\,\mu{\rm m}$. In such experiments, the signal is rather weak and is extracted from the noise with a lock-in amplifier, see Ref.~\cite{DeWilde:2006kt}. One modulates the tip-surface distance and detects the signal at the modulation frequency or its second harmonic (typical frequencies are in the kHz range). The signal calculated for this modulation technique is also shown in Fig.~\ref{Theoplasm} (black solid line). As expected, the lock-in signal qualitatively arises from the derivative of the signal at fixed distance. The modulation amplitude is rather large ($150\,{\rm nm}$) and at the shortest distances shown in the plot, the dipole approximation is probably no longer reliable. We have also plotted in Fig.~\ref{Expplasm} the experimental TRSTM data obtained as described above. The experimental setup is similar to the one used in Refs.\cite{Tetienne:2010gy, Babuty:dYmzb9en}. Here the plasmon is excited in an integrated plasmonic device by end-fire coupling between one end facet of a quantum cascade laser cavity at 7.5 $\mu$m and the surface of a planar gold strip. The wavelength is very much below the plasma resonance of gold, therefore the surface plasmon mode penetrates significantly outside the surface. From the Drude parameters for gold, we estimate a extension into the vacuum of $1 / \Im \gamma \sim 60\,\mu{\rm m}$. The overall behaviour of the experimental signal is quite close to the theoretical prediction. The experimental curve decays somewhat faster with distance. We attribute this to the different signal collection: indeed, the detector involves a Cassegrain objective that collects a finite range of angles $\theta$. This averages over interference fringes with different periods (recall the $\cos\theta$ in the exponential in Eq.(\ref{sigdet})). The modelling of the tip by a sphere may also reach its limits here. By comparing the experimental signal to the theory calculated for a spherical particle, we find an effective tip radius of around $1\,\mu{\rm m}$. This is about the same value as recently found \cite{Babuty:dYmzb9en}. The volume of this spherical tip matches roughly the effective scattering volume of the tip whose estimation we discuss in Sec.\ref{s:check-dipole-approx}, given the tip's conical shape. \begin{figure} \begin{center} \includegraphics[width=10cm]{Theo_Exp_plasmon_mod.eps} \caption{TRSTM signal (theory and experiment) for a gold sample with a surface plasmon mode excited at $\lambda = 7.5\,\mu{\rm m}$. The excitation is performed with a quantum cascade laser cavity integrated in the sample surface, as in the setup of Refs.\cite{Tetienne:2010gy, Babuty:dYmzb9en}. We plot the intensity scattered into the far field at an angle $\theta = 30^\circ$, as a function of tip-sample distance $z_t$. Experiment (data points): detection through a Cassegrain objective (numerical aperture 0.5). The tungsten tip (conical shape with half opening angle $\approx 20^\circ$) oscillates with an amplitude of 150\,nm, and the amplitude of the signal oscillating in phase with the tip is plotted. Lines: theory, the tip is modeled by a tungsten sphere of radius $1\,\mu{\rm m}$. Solid line: simulation of the lock-in detection with the same modulation amplitude. Dashed line: distance $z_t$ fixed. } \label{Theoplasm} \label{Expplasm} \end{center} \end{figure} \subsection{Thermal emission from a heated surface} \subsubsection{Thermal fields above the sample} When the electromagnetic field above the surface is only the one due to thermal emission from the sample surface, several simplifications occur. The TRSTM signal of Eq.(\ref{eq:signal-and-Gamma}) is calculated using correlation functions in thermal equilibrium for the non-perturbed fields ${\bf E}^0( {\bf r}_t )$, ${\bf H}^0( {\bf r}_t )$ \cite{Joulain:2005ih,Volokitin:2007el,Joulain:2010bq}. Most of the cross correlations functions vanish due to the fact that thermal currents decorrelate for different directions \cite{Joulain:2005ih}. Due to rotational symmetry around the $z$-axis, we are left with $\langle|E^0_x({\bf r}_t)|^2\rangle = \langle|E^0_y({\bf r}_t)|^2\rangle$, $\langle|E^0_z({\bf r}_t)|^2\rangle$, $\langle|H^0_x({\bf r}_t)|^2\rangle = \langle|H^0_y({\bf r}_t)|^2\rangle$, $\langle|H^0_z({\bf r}_t)|^2\rangle$, and $\langle E^0_x( {\bf r}_t ) H^{0*}_y( {\bf r}_t ) \rangle = - \langle E^0_y( {\bf r}_t ) H^{0*}_x( {\bf r}_t ) \rangle$. In these conditions, the signal at the detector in a direction making an angle $\theta$ with the $z$ axis reads \begin{eqnarray} \left< S^d( \omega ) \right> & = & \frac{ \omega^4 }{32\pi c^3 }d\Omega \left\{ \left( \cos^2\theta|1-r^p(\theta)e^{i \phi}|^2 + |1+r^s(\theta)e^{i \phi}|^2 \right) |\alpha^E_{xx}|^2 \langle \epsilon_0 |E^0_x({\bf r}_t)|^2 \rangle \right.\nonumber\\ && {} + \sin^2\theta|1+r^p( \theta ) e^{i \phi}|^2 |\alpha^E_{zz}|^2 \langle \epsilon_0 |E^0_z({\bf r}_t)|^2 \rangle \nonumber \\ && {} + \left( \cos^2\theta|1-r^s(\theta)e^{i \phi}|^2 + |1+r^p(\theta)e^{i \phi}|^2 \right) |\beta^H_{xx}|^2 \langle \mu_0 |H^0_x({\bf r}_t)|^2 \rangle \nonumber\\ && {} + \sin^2\theta|1+ r^s( \theta ) e^{i \phi}|^2 |\beta^H_{zz}|^2 \langle \mu_0 |H^0_z({\bf r}_t)|^2 \rangle \\ && {} + 2\cos\theta\,\Re\left[ \alpha^E_{xx}\beta^{H*}_{xx} \langle E^0_x({\bf r}_t) H^{0*}_y({\bf r}_t) / c \rangle \right.\nonumber\\ &&{} \times \left.\left.\left( (1+ r^s( \theta ) e^{i \phi})(1- r^{s*}( \theta ) e^{-i \phi}) + (1- r^p( \theta ) e^{i \phi})(1+ r^{p*}( \theta ) e^{-i \phi}) \right)\right]\right\} \nonumber \end{eqnarray} The thermal radiation above a plane interface is well known in the literature and yields the field correlations \cite{Joulain:2005ih,Volokitin:2007el,Dorofeyev:2011bg}: \begin{eqnarray} \left<|E^0_x({\bf r}_t)|^2\right> & = & \frac{\mu_0\omega \Theta(\omega,T) }{2\pi^2} \Im\left( i\int_0^\infty\frac{K dK}{ \gamma } \left[1 + r^s e^{2i \gamma z_t} + \frac{ \gamma^2 }{ k_0^2 } (1 - r^p e^{2i \gamma z_t}) \right]\right) \label{ex2} \\ \left<|E^0_z({\bf r}_t)|^2\right> & = & \frac{\mu_0\omega \Theta(\omega,T) }{\pi^2} \Im\left( i\int_0^\infty\frac{K^3 dK}{ \gamma k_0^2 } (1 + r^p e^{2i \gamma z_t}) \right) \\ \left<|H^0_x({\bf r}_t)|^2\right> & = & \frac{\epsilon_0\omega \Theta(\omega,T) }{2\pi^2} \Im\left( i\int_0^\infty\frac{K dK}{ \gamma } \left[1 + r^p e^{2i \gamma z} + \frac{ \gamma^2 }{ k_0^2 } (1 - r^s e^{2i \gamma z}) \right]\right) \\ \left<|H^0_z({\bf r}_t)|^2\right> & = & \frac{\epsilon_0\omega\Theta(\omega,T)}{\pi^2} \Im\left( i\int_0^\infty\frac{K^3 dK}{ \gamma k_0^2 } (1 + r^s e^{2i \gamma z}) \right) \\ \left< E^0_x({\bf r}_t) H^{0*}_y({\bf r}_t) \right> &=& \frac{ \Theta(\omega,T) }{ 4\pi^2 } \left[ \int_0^\infty \frac{2K \gamma dK}{ |\gamma| } \Re\left( \frac{ \gamma }{ |\gamma| }(1 + r^s e^{2i \gamma z_t} - r^p e^{2i \gamma z_t}) \right)\right] \label{exhy} \end{eqnarray} We denote $\Theta(\omega,T) = \hbar \omega/[\exp[\hbar\omega/(k_BT)]-1]$ the mean thermal energy of an oscillator with angular frequency $\omega$, $\hbar$ and $k_B$ are the Planck and Boltzmann constants, and $T$ is the temperature. The integrals are somewhat similar to those in Eqs.(\ref{eq:HExy-Green}--\ref{eq:HHzz-Green}), being plane-wave expansions; we use the same notation as there. If the substrate and the particle dipole materials are known, the signal at the detector can be calculated from these formulas. If the signal is divided by the mean energy of an oscillator $\Theta(\omega,T)$, it only depends on the particle polarizability, the tip-sample distance and the surface optical properties. The polarizability can be extracted from reference experiments at another surface. Note that the dressed polarizability and the thermal electromagnetic field are closely related. Indeed, both involve the system's Green tensor taken at the tip position. As a consequence, if the thermal near field shows a resonance at a given frequency, it is likely that there will be one in the dressed polarizability at the same frequency. This entails a difficulty for the interpretation of the signal at the detector. Indeed, the electromagnetic field correlations are closely related to the electromagnetic local density of states (EM LDOS) \cite{Joulain:2005ih}, at least to a projected LDOS. Thus, $\langle |E_x|^2 \rangle$ is related to the parallel electric LDOS whereas $\langle |H_z|^2 \rangle$ is related to the perpendicular magnetic LDOS. Relating the detector signal to the LDOS would be very interesting, since it would give a way to detect this quantity like electronic tunneling microscopy does it for the electronic LDOS \cite{Tersoff:1985wm}. Unfortunately, it is not possible to find a simple and universal relation between the two quantities. We shall see, however, that in specific situations, one contribution may be the leading term, so that the projected EM LDOS and the detected signal become proportional. \subsubsection{Probing a polar material} \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig9.eps} \caption{Spectrum of thermal radiation scattered by a tip (TRSTM signal). The tip is described by a tungsten sphere above a SiC sample. The signal is detected in a direction making an angle of 30$^\circ$ with the vertical direction. (a) Signal and its contributions at height $z = 200\,{\rm nm}$. We take $R_t = z /2$ in panel (a--c). (b) Signal at height $z = 1\,\mu{\rm m}$. (c) Comparison of the signal with the LDOS for $z = 200\,{\rm nm}$, $z = 1\,\mu{\rm m}$ and $z = 1.6\,\mu{\rm m}$. (d) Signal spectrum for $R_{p} = 100\,{\rm nm}$ at two different distances $z = 200$ nm (left scale) and $z = 1\,\mu{\rm m}$ (right scale). } \label{sigdetWSiC} \end{center} \end{figure} We now study the spectrum of the signal detected by a probe above SiC. As we have seen in the previous section, this signal strongly depends on the dressed polarizability and therefore on the tip size and the tip distance. To illustrate this, we show, in Fig.~\ref{sigdetWSiC}, the spectral signal detected by a tungsten probes above SiC with radii 100 nm and $1\,\mu{\rm m}$. It is observed that the emission peak is rather narrow for the small tip. This peak appears around the phonon polariton resonance frequency that is 948 cm$^{-1}$ for SiC. We also note that for a small tungsten tip, the signal above SiC is dominated at short distance by electrical terms. Parallel and perpendicular field components contribute both significantly. These contributions show a slight spectral shift with respect to each other: this can be attributed to the different denominators in the dressed polarizability tensor of Eq.(\ref{alphaeffapprox}). For a 500 nm tip at $1\,\mu{\rm m}$ from the interface, we note that the main contribution comes from the parallel electric field. Magnetic fields are more important here because the magnetic polarizability is larger. These magnetic terms do not contribute a lot because near the polariton resonance in a polar material, the energy density is dominantly electric. However, the mixed term, which involves both magnetic and electric fields (and both polarizabilities), is the second contribution to the signal and is therefore not negligible. Fig.~\ref{sigdetWSiC}(c) shows various signal spectra for three tips at (center-surface) distances equal to the their diameter. They are compared with the LDOS at $200\,{\rm nm}$ and $1.6\,\mu{\rm m}$. At 200 nm distance, the tip gives a signal similar to the electromagnetic LDOS: a well-defined peak appears around the polariton frequency. Its width is very similar to the LDOS, but the position is slightly shifted. When the tip size increases, the peak tends to shift and to broaden. Recent SNOM experiments based on the measurement of the near-field thermal emission using a tungsten tip seem to confirm this shifting and broadening, suggesting that the approximation of the tip by a simple spherical dipole is valid to some extent \cite{Babuty:dYmzb9en}. Note also that this broadening, although less pronounced than for a large tip, also occurs when a small tip is retracted from the surface as it can be seen in Fig.~\ref{sigdetWSiC}(d). As shown in Fig.~\ref{sigdetWSiC}, the signal calculated with a small tip is very similar to the LDOS around the polariton resonance. At short distance, the signal is indeed dominated by the parallel electric contribution. This means that the signal mainly depends on $\alpha^{EE}_{xx}$ and on $|E^0_x|^2$. In the case of a thermal signal, this last quantity is proportional to the projected parallel electromagnetic LDOS. Therefore, a SNOM experiment detecting the thermal near field will measure the product of one of the dressed polarizabilities by a partial contribution to the EM LDOS. If the dressed polarizability has a flat spectral response (i.e. it varies only weakly with frequency), one can say that the signal at the detector is proportional to the projected EM LDOS. This is not strictly the case in the present situation where the polarizability can be increased by 10--20\% around the resonance (Fig.~\ref{aleff100200}): the resulting signal is the product of two peaks at approximately the same frequency. As the value of the EM LDOS in the peak spectral band is about 100 to 1000 times its value outside this band, its multiplication by the dressed polarizability will give a peak that is not exactly the projected EM LDOS but which is representative of the LDOS. Consider now the results for the larger tip (radius $1\,\mu{\rm m}$). The signal scattered to the detector by such tips has a broader spectrum and the frequency corresponding to the emission spectrum is shifted to lower frequencies [Fig.~\ref{sigdetWSiC}(b)]. When the tip size increases, polarizability is shifted to lower frequency and broadened [compare Figs.\ref{aleff100200} and \ref{aleff5001000}]. At larger distances, the EM LDOS is also broadened [Fig.~\ref{sigdetWSiC}(c)]. The resulting signal is now the product of two peaks which are rather similar in shape but with a slight frequency offset. This product shows a maximum mid-way between the peaks of the polarizability and the LDOS. Under these conditions, the signal at the detector is a peak related to the EM LDOS, but is not strictly speaking giving the EM LDOS. However, one can see from Fig.~\ref{sigdetWSiC}(c) that the detected signal and the LDOS have a very similar shape. \subsubsection{Probing a metallic sample} We now consider the example of a tip above a metal surface. Fig.~\ref{sigdetWAu} shows the signal scattered by a tungsten particle above a gold surface heated at 300 K. Note the significantly different spectral dependence and the reduced magnitude, compared to the polar dielectric of the previous section. For a small particle, the contribution to the signal is dominated by the magnetic term below 1000 cm$^{-1}$. Indeed, the magnetic energy is larger than the electrical energy in this spectral range where Au is highly reflecting. As seen in Fig.~\ref{aleff100200}, the magnetic polarizability has a rather flat response for small tips above 400 cm$^{-1}$. As a consequence, the signal at the detector is proportional to the parallel magnetic LDOS between 400 and 1000 cm$^{-1}$. Since the total EM LDOS is dominated by its magnetic contribution, the signal detected in this spectral range is close to the EM LDOS. On the contrary, above 1000 cm$^{-1}$, the signal is dominated by the perpendicular electric contribution but the parallel magnetic and the mixed term are of the same order of magnitude even if smaller. For such situation, it is not correct to state that the signal is proportional to the EM LDOS although it is related to it. For a large tip, there is no spectral range where one component clearly dominates the contribution to the signal (except at high frequencies, but there the validity of the dipole approximation becomes questionable). The detected signal cannot be considered as simply proportional to the EM LDOS. \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig10.eps} \caption{Spectrum of TRSTM signal, as in Fig.~\ref{sigdetWSiC}, but above a gold sample. The tip is described by a tungsten sphere. The signal is detected in a direction making an angle of 30$^\circ$ with the vertical (a) Total signal and different contributions for a 100 nm radius sphere at $z=200\,{\rm nm}$ above surface. (b) Total signal and different contributions for a 500 nm radius sphere at $z=1\,\mu{\rm m}$ above the surface.} \label{sigdetWAu} \end{center} \end{figure} \subsection{Modeling the tip by a dipolar particle} \label{s:check-dipole-approx} Considering the calculations made in this paper and the sharp tips in real experiments, one may wonder why a conical tip is well described by a spherical particle. As a matter of fact, TRSTM experiments are in excellent agreement with this theoretical model~\cite{Babuty:dYmzb9en}. This can be qualitatively understood with the following argument. To be specific, we consider the detection of the thermal near field above a sample and a sharp conical tip whose apex is touching the surface. The field intensity (autocorrelation function) typically decays like $1/z^3$ with the distance $z$ from the surface. A slice of the tip at height $z$ has a volume proportional to $z^2$ that contributes to the scattered field, provided the field fully penetrates into the material. Therefore, the signal arising from this slice is proportional to $1/z^3\times(z^2)^2 \propto z$, favoring the tip's shaft rather than the apex. Above a certain altitude, the field does not penetrate into the metallic tip, and the scattering volume of a slice is determined by its circumference, proportional to $z$. This leads to a scattered intensity $\propto 1/z$, indicating that most of the signal arises from an optimal height proportional to the skin depth and depending on the cone's opening angle. The skin depth for tungsten in the infrared around $10\,\mu{\rm m}$ is $\lambda/\sqrt{\epsilon}\sim 200\,{\rm nm}$. This means that even for tips touching the sample, the field scattered will arise from a distance of several hundreds of nanometers. This argument makes it plausible why the simple model of a particle with relatively large size permits to capture much of the phenomenon, as found by comparing to experimental data in Ref.\cite{Babuty:dYmzb9en}. The same argument also holds for the situation of a coherently excited surface mode, as illustrated by Fig.~\ref{Theoplasm} above: indeed, the local field intensity then also decays with distance, even though the dependence on $z$ differs. \section{Thermal emission of a nanoparticle} We now consider a small dipolar particle heated to a temperature $T_t$. This particle radiates an electromagnetic field in all space which can be detected in the far field. This configuration corresponds to recent experimental situations \cite{Jones:2012fx} where a heated tip is moved into the near field of a sample. Note that in this situation, the detector is also picking up the thermal emission coming from the sample, as calculated in the preceding section. We focus in the following on the radiation from the tip; it could be isolated by suitable imaging and modulation techniques. The tip signal has again contributions coming directly from the particle in straight line and undergoing one reflection at the interface. The signal at the detector reads: \begin{eqnarray} \label{ } \left<S^d(\omega)\right>&=&\frac{d\Omega}{32\pi^3}\Theta(\omega,T_t)\frac{\omega^3}{c^3}\left[\Im[\alpha^E_{xx}]\left(|1+r^s(\theta)e^{i \phi}|^2+\cos^2\theta|1-r^p(\theta)e^{i \phi}|^2\right)\right.\nonumber\\ &+&\left.\Im[\alpha^E_{zz}]|1+r^p(\theta)e^{i \phi}|^2\sin^2\theta\right] \end{eqnarray} where $T_t$ is the tip temperature. Here, the spectral dependence comes mainly from the dressed polarizability. The far field reflection coefficients $r^{s,p}( \theta )$ typically show no resonant features since surface modes appear in the evanescent sector. One thus expects that the detected spectra will follow polarizability variations that we have described in the preceding sections. In Fig.~\ref{Chff}, we plot the signal detected in the far field in a direction making an angle of 45$^\circ$ with the vertical direction. The tip is modeled by a tungsten sphere (100 nm radius) heated at 300 K and held at various distances from a vacuum-SiC interface. Note that the tip sample distance is much smaller than the wavelength so that $\phi\ll1$. When the particle-interface distance is 200 nm, the signal is peaked around the SiC surface resonance ($948\,{\rm cm}^{-1}$), similarly to the polarizability of a 100 nm-radius sphere. When the particle is retracted from the interface, the signal is reduced and the spectrum broadens. Above a certain distance, the dressing corrections to the polarizability become negligible, and for tungsten, the spectral dependence becomes flat. The spectral behaviour of the signal comes mainly from variations of the reflection coefficient. One notes that the signal is more important in the frequency range where SiC is known to be highly reflective i.e. between $850\,{\rm cm}^{-1}$ and $950\,{\rm cm}^{-1}$. A similar behaviour is observed for a micron-sized tip except that the signal is broader at the minimum distance. This can easily be explained by inspecting the dressed polarizabilities of a 500 nm-radius sphere above a SiC interface where this broadening is also observed (compare Figs.~\ref{aleff100200} and \ref{aleff5001000}). \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig11.eps} \caption{Signal at the detector at an angle $\theta = 45 ^0$ to the surface normal (log scale). The tip is described by a tungsten sphere heated at $T = 300$ K and located at various distances to the interface. (a) Signal in log-scale for $R_t = 100\,{\rm nm}$, for different tip-sample distances. Inset: signal for $R_t= 100\,{\rm nm}$ and $z= 200\,{\rm nm}$ in linear scale. (b) Signal for $R_t = 500 \,{\rm nm}$ and at distances $z = 1, 2, 3\,\mu{\rm m}$.} \label{Chff} \end{center} \end{figure} Note that for such situation, the detected signal can exhibit a peak very similar to the one observed in the energy density spectrum or in the EM LDOS. However, this peak is the signature of the dressed polarizability which exhibits a resonance at frequencies close to surface resonance. Even if such an experiment does not probe the EM LDOS, it can probe surface resonances if the tip is sufficiently close to the interface~\cite{Jones:2012fx}. \section{(Non)radiative cooling of a particle} When a particle is heated, it exchanges energy with the environment and cools down. In free space and assuming the surroundings at zero temperature, the spectral power lost by the particle (temperature $T_t$, modelled by a dipole) is \cite{Joulain:2008tp} \begin{equation} \label{ } P(\omega)=\frac{\omega^3}{\pi^2 \epsilon_0 c^3}\Im[\alpha(\omega)] \Theta(\omega,T_t) \end{equation} where $\Theta(\omega, T_t)$ is again the mean thermal energy of the dipole oscillator. When the particle is close to a surface, the transferred power depends on the electromagnetic mode density at the particle position (EM LDOS) and can be much larger than in free space because non-radiative channels (evanescent modes) open up. This phenomenon is very similar to the Purcell effect when an atom or a molecule has its spontaneous emission rate modified inside a cavity or when approaching a surface. The cooling rate depends on the EM LDOS~\cite{BenAbdallah:2011tz} in a similar way as the spontaneous emission rate. The cooling rate formula given by Mulet et al.\cite{Mulet:2001kp} should be corrected at very close distances to take into account the dressed polarizability. We thus find for the heat transferred from particle to sample \begin{equation} \label{eq:spectrum-cooling} P(\omega) = \frac{2}{\pi} \Theta(\omega,T_t) \sum_{i=x,y,z} \Im(\alpha^E_{ii}) \Im[ G^{EE}_{ii}({\bf r}_t,{\bf r}_t) ] \end{equation} This is easily generalized to particle and surroundings at temperatures $T_t, T_s > 0$, see Ref.\cite{Joulain:2008tp}. We emphasize that the heat exchange is the observable that shows the ``cleanest'' connection to the EM LDOS which is itself given by the imaginary part of the EM Green function. In Fig.~\ref{refr}, we plot the heat exchanged between a SiC particle (at 300 K) and a SiC substrate ($T_s = 0$) as a function of the distance. We also show the spectrum of the exchanged flux. We compare, in each plot, Eq.(\ref{eq:spectrum-cooling}) to the results of Ref.~\cite{Mulet:2001kp}. We observe that both expressions give similar results until at distances as low as 400 nm, the corrections in the dressed polarizability set in. The heat spectrum is peaked at a the particle plasmon resonance in the SiC particle, where $\epsilon( \omega ) = -2$, i.e., around $935\,{\rm cm}^{-1}$ [Fig.~\ref{refr}(c)]. When one enters in the near field [Fig.~\ref{refr}(b)], two phenomena occur: evanescent contributions to the EM LDOS become stronger, in particular those of the surface phonon polariton resonance where $\epsilon( \omega ) =-1$, around $948\,{\rm cm}^{-1}$, leading to a double-peaked spectrum. In addition, the corrections from the dressed polarizability shift the peaks in relative weight and position. Depending on the distance, these shifts can both enhance or reduce the heat transfer compared to the bare polarizability used in Ref.\cite{Mulet:2001kp}, see Fig.~\ref{refr}(a). (The dipolar model is no longer valid at the shortest distances shown, however.) For magnetic particles, similar corrections should be applied. \begin{figure} \begin{center} \includegraphics[width=12cm]{Fig12.eps} \end{center} \caption{(a) Total exchanged heat power between a 100nm radius SiC particle (300 K) above a SiC substrate ($T = 0$) versus distance. The two curves are based on the bare (``Mulet 2001'') and dressed (``full'') polarizabilities. (b) Spectrum of heat power at distance 200 nm. (c) Spectrum at distance $1\,\mu{\rm m}$, same particle size. } \label{refr} \end{figure} \section{Conclusion} We have shown how electromagnetic near fields in the infrared are modified by the interaction between a probing SNOM tip and the sample it is scanning. The tip's response (electric and magnetic polarizabilities) can be modified to account for this interaction. We have combined the effect of retardation to the one of the image dipole. We have analyzed the relation between the signal detected in the far field by an apertureless SNOM and the near fields, excited either coherently (surface plasmon resonance) or incoherently (thermal emission). The technique performs a local spectroscopy of the surface and for some cases the signal can even be proportional to the EM LDOS, a fundamental quantity. By comparing to experimental data, we demonstrated one salient feature of strong tip-sample interactions, namely interferences between direct and reflected rays that lead to oscillatory signals (standing waves) as a function of distance. We have also corrected the formula for near-field radiative heat transfer in cases where there is a strong particle-sample interaction. In the future, this theory could be extended to tips with more elongated shapes \cite{Huth:2010fg} such as conical ones, often used in experiment, in order to provide a more quantitative comparison. \begin{acknowledgments} The authors thank J.-J. Greffet for fruitful discussions. Y. De Wilde and A. Babuty thank the team of R. Colombelli and A. Bousseksou for fruitful collaborations on plasmonics. Y. De Wilde and A. Babuty acknowledge support from the Agence Nationale de la Recherche (Grant No. ANR-07-NANO-039 ``NanoFtir'') and from the R\'egion Ile-de-France in the framework of C'Nano IdF (nano-science competence center of Paris Region). K. Joulain and P. Ben-Abdallah acknowledge support from Agence Nationale de la Recherche (Grant No. ANR-10-BLAN-0928-01 ``Sources-TPV''). This work is supported by LABEX WIFI and LABEX INTERACTIFS (Laboratory of Excellence within the French Program ``Investments for the Future'') under reference ANR-10-IDEX-0001-02 PSL* and ANR-11-LABX-0017-01. C. Henkel enjoyed the hospitality of Universit\'e de Poitiers where part of this work was done. \end{acknowledgments} \bibliographystyle{vancouver}
\section{Introduction} The search for a consistent thermal field theory in the perturbative regime has led to realization that all the so-called {\it hard thermal loops} have to be taken into account. In momentum space, these are amplitudes with loop momenta of the order of the temperature, which is large compared with all the external momenta. In the case of gauge field theories, it has been shown that it is possible to construct a closed form expression for the effective Lagrangian, which generates all hard thermal loops \cite{Frenkel:1989br,Taylor:1990ia,Braaten:1990az}. In gauge theories the hard thermal loop amplitudes are related to each other through Ward identities. This property, together with the characteristic non-localities exhibited by the amplitudes, are key ingredients for the construction of a gauge invariant effective action. In principle, the same approach can be employed for hard thermal loops in a background of soft gravitational fields. It is known that, similarly to the gauge field amplitudes, the graviton thermal amplitudes satisfy, in the high temperature limit, simple Ward identities which reflect the symmetry under local coordinate transformations (in addition, these amplitudes are also related by Weyl identities which arise from the scale invariance) \cite{Brandt:1993dk}. Nevertheless, an explicit closed form expression for the one-loop order effective Lagrangian is only known in two special limits, when the background gravitational field is either time independent or spatial independent. Each of these limits (which physically corresponds to static or long wavelength plasma perturbations) yield two different local effective Lagrangians which are functionals of the background field \cite{Brandt:2009ht,Francisco:2013vg}. In the case of a general configuration of the background gravitational fields, so far only an implicit representation of the one-loop effective Lagrangian is known \cite{Brandt:1994mv}. In previous works it has been shown that the leading contributions of hard thermal loops in a static background can be obtained by evaluating them at zero external energy-momentum. This has been shown both at one-loop \cite{Frenkel:2009pi} and subsequently at two-loops \cite{Brandt:2012mn}. More recently, this result has also been generalized to all orders \cite{bfs2013}. This interesting property has prompted us to consider a more direct approach in seeking for the effective Lagrangian, by making use of the background field method \cite{peskin_scroeder} (for a pedagogical review article see also \cite{abbott82}). As an example of the usefulness of this method, the one-loop effective Lagrangian has been previously derived in a simple manner in the case of thermal scalar fields in a static background \cite{Brandt:2012ei}. The result, which has been known for many years \cite{Rebhan:1991yr}, has the same metric dependence as in Eq. \eqref{eq11}, differing only by a factor which counts the degrees of freedom of the fields. This form exhibits some interesting properties. First, it has a characteristic dependence on the local temperature \begin{equation}\label{Tloc} T_{\rm loc}\equiv \frac{T}{\sqrt{g_{00}}}, \end{equation} being proportional to $T^d_{\rm loc}$, where $d$ is the space-time dimension and $T$ is the asymptotic Minkowski space temperature. $T_{\rm loc}$ is the temperature measured by a standard local thermometer, such as a Carnot cycle \cite{Balazs2:1965,Ebert:1973}. This behavior is in agreement with the so-called {\it Tolman-Ehrenfest effect} which argues that in a system at thermal equilibrium in a stationary gravitational field, the temperature varies with the space-time metric according to the relation \eqref{Tloc}. This effect was originally discovered in the context of a classical fluid interacting with external static gravitational fields \cite{Tolman:1930zza,Tolman:1930a,Tolman:1930zza1}. Another important property of the one-loop static effective Lagrangian is its invariance under conformal transformations, for any value of $d$. This can be simply understood since, in order to behave like a density, the factor $T^d_{\rm loc}$ in the effective Lagrangian has to be multiplied by $\sqrt{|g|}$. Consequently, the resulting expression is invariant under the rescaling $g_{\mu\nu} \rightarrow \sigma g_{\mu\nu}$. The main purpose of the present work is to obtain the higher loop corrections to the effective Lagrangian of a QED plasma in a static gravitational background. The one-loop results above described, do not take into account the interactions between electrons and photons. In order to consider these effects, we will apply the same basic idea of the background field method to higher loop orders. As we will show, in a $d$-dimensional space-time, the interactions break the Weyl symmetry when $d\neq 4$. The two-loop contribution, given by Eq. \eqref{eq21}, is obtained computing the 1PI diagrams, with no external legs, in the background of static gravitational fields. We also compute a non-perturbative contribution to the effective Lagrangian which arises from the summation of all higher order infrared divergent 1PI diagrams. In this case, there are two different forms, given by Eqs. \eqref{Leven} and \eqref{Lodd}, depending whether the space-time dimension is even or odd, respectively. The above results have a very simple structure. They are equivalent, up to a factor of $\sqrt{|g|}$, to the pressure of a QED plasma at high temperature in $d$-dimensional Minkowski space-time, with $T$ replaced by the local temperature $T_{\rm loc}$. This shows that the Tolman-Ehrenfest effect is explicitly manifested even when the quantum corrections are taken into account. These and other related aspects of the effective Lagrangian are discussed further in the concluding section. \section{One-loop effective Lagrangian} In this section we will introduce our basic notation and method. Also, for completeness we derive the one-loop effective Lagrangian. Let us consider the Lagrangian for photons and electrons in a gravitational background \begin{align}\label{eq:lll} \mathcal{L} &= \sqrt{|g|} \left[ i \bar{\psi} g^{\mu \nu} \gamma_\mu(\partial_\nu -i e A_\nu) \psi -\frac{1}{4} g^{\mu \nu}g^{\alpha \beta} F_{\mu \alpha}F_{\nu \beta} \right. \nonumber \\ & - \left. \frac{1}{2} (g^{\mu \nu} \partial_\mu A_\nu )^2 + g^{\mu \nu} \partial_\mu \bar{C} \partial_\nu C \right], \end{align} where $g_{\mu\nu}$ is the metric tensor ($|g| = |\det{g_{\mu\nu}}|$), $F_{\mu\nu} = \partial_{\mu} A_\nu - \partial_{\nu} A_\mu$ is electromagnetic field tensor, $\psi$ is the fermion field and $C$ is the ghost field (we are employing the Feynman gauge condition for the gauge field $A_\mu$). As we have pointed out in the introduction, for the purpose of obtaining the static effective Lagrangian in the high temperature limit, we will neglect all the space-time derivatives of the metric as well as the fermion masses since these quantities would be suppressed by the much larger scale of temperature. The Feynman rules for photons, fermions and ghosts in a gravitational background can be readily obtained using the vierbein formalism, which allows us to write the Lagrangian in the following form \begin{align}\label{eq:ll123} \mathcal{L} &= \sqrt{|g|} \left[ i \bar{\psi} {\tilde{\gamma}}^a(\tilde{\partial}_a -i e \tilde{A}_a) \psi -\frac{1}{4} \tilde{F}_{a b}\tilde{F}^{ab} \right. \nonumber \\&-\left.\frac{1}{2} ( \tilde{\partial}_a \tilde{A}^a )^2 + \tilde{\partial}_a \bar{C} \tilde{\partial}^a C \right], \end{align} where $A_a = E^{\mu}_a A_\mu$ is the field in the local frame ($E^\mu_a$ is the vierbein) and the Dirac matrices $\tilde\gamma_a$ satisfy \begin{align}\label{eq::algebra} \{ \tilde{\gamma}_a, \tilde{\gamma}_b\} = 2 {E_a}^\mu {E_b}^\nu g_{\mu \nu} = 2 {E_a}^\mu {E_b}^\nu {e^c}_\mu {e^d}_\nu \eta_{c d} = 2\eta_{ab}, \end{align} From this Lagrangian one obtains the following Feynman rules for the effective propagators and vertices \begin{subequations}\label{frules} \begin{eqnarray}\label{fprop} \includegraphics[scale=0.8]{fermion_prog} & : \;\;\; & \;\;\; \displaystyle{{1\over \sqrt{|g|}\gamma^a p_a}} \end{eqnarray} \begin{eqnarray} \includegraphics[scale=0.8]{photon_prop} & : \;\;\;&\;\;\; \displaystyle{{\eta_{ab}\over \sqrt{|g|} p^c p_c}} \end{eqnarray} \begin{eqnarray} \includegraphics[scale=0.8]{ghost_prop} & : \;\;\;&\;\;\; -\displaystyle{{1\over \sqrt{|g|} p^a p_a}} \end{eqnarray} \begin{eqnarray} \begin{array}{c}\includegraphics[scale=0.8]{photon_fermion}\end{array} & : \;\;\;&\;\;\; e\sqrt{|g|} \gamma^a \end{eqnarray} \end{subequations} \begin{figure}[t!] \includegraphics[scale=0.8]{one_loop_fermion.pdf}\qquad\includegraphics[scale=0.8]{one_loop_photon}\qquad\includegraphics[scale=0.8]{one_loop_ghost} \caption{One-loop diagrams which contribute to the effective Lagrangian. The solid line represents a fermion and wavy and dashed lines denote respectively gauge and ghost particles, all in the gravitational background.} \label{fig1} \end{figure} The lowest order contributions to the effective Lagrangian are represented diagrammatically in Fig.~\ref{fig1}. Let us first consider the fermion loop contribution. Using the imaginary time formalism and performing the Dirac algebra in the vierbein basis, we obtain \begin{align}\label{LF} \mathcal{L}_1^F &=T \sum_{n} \int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \mathrm{tr} \log ( \beta \tilde{\gamma}^a \tilde{p}_a)\nonumber\\ &= \frac{T}{2} 2^{E(d/2)} \sum_{n}\int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \log (-\beta^2 g^{\mu \nu} {p}_\mu {p}_\nu ), \end{align} were $\beta=1/T$ and we are considering a $d$-dimensional space-time. The time component of the momentum is $p_0 = i \omega^F_n$, where $\omega^F_n = (2n+1)\pi T$ are the fermionic Matsubara frequencies and $n=0,\pm 1, \pm 2, \dots$. Here we are employing the minimal representation for the Dirac matrices, so that the trace of the identity is given by $\mathrm{tr} I =2^{E(d/2)}$, where $E(d/2)$ is the integer part of $d/2$. Similarly the respective contributions from the photon and ghost loops shown in Fig.~\ref{fig1} can be expressed as follows \begin{subequations}\label{LB} \begin{eqnarray} \mathcal{L}_1^P &=& -\frac{d}{2} T \sum_n \int\frac{d^{d-1} q}{(2\pi)^{d-1}} \log( -\beta^2 g^{\mu \nu} q_\mu q_\nu) \\ \mbox{and} &&\nonumber \\ \mathcal{L}_1^G &=& T \sum_n \int\frac{d^{d-1} q}{(2\pi)^{d-1}} \log( -\beta^2 g^{\mu \nu} q_\mu q_\nu), \end{eqnarray} \end{subequations} where $q_0 = i \omega^B_n$ and $\omega^B_n = 2 n \pi T$ is the bosonic Matsubara frequency , with $n=0,\pm 1, \pm 2, \dots$. In order to perform the sum/integrals in Eqs. \eqref{LF} and \eqref{LB} we now use a locally rest vierbein frame as defined in the appendix \ref{aa}. Proceeding in this way, Eq. \eqref{LF} yields \begin{align} \mathcal{L}_1^F &= \frac{T}{2}2^{E(d/2)} \sum_{n} \int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \log \left\{ -\beta^2 \left[ (p_0/\sqrt{g_{00}})^2 \right. \right. \nonumber \\ &+ \left. \left. g^{ij}(p_i+ (g^{-1})_{ik}g^{0k} p_0 ) (p_j+ (g^{-1})_{jl} g^{0l} p_0) \right] \right\}. \end{align} Making a change of variables in the $p_i$ integration, it is possible the factorize all the metric dependence and we obtain the following result \begin{align} \mathcal{L}_1^F &= 2^{E(d/2)} \frac{ \sqrt{(-1)^{d-1} \bold{g}^{-1} g_{00} }}{g_{00}^{d/2}} \,\, \frac{T}{2} \sum_{n} \int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \nonumber \\ &\times \log \left\{ -\beta^2 \left[ p_0^2 - |\vec{p}|^2 \right] \right\}. \label{eq:P_0f234aaa} \end{align} where $\bold{g}^{-1} = \det (g^{-1})_{ij}$. The sum/integral in the previous expression is the same as in flat space-time \cite{kapusta:book89}, here generalized to $d$ space-time dimensions. The temperature-independent part of \eqref{eq:P_0f234aaa} leads to a divergent integral. This divergent result gives just the zero-point energy of the vacuum, which can be subtracted off since it is an unobservable constant. On the other hand, the $T$-dependent part of \eqref{eq:P_0f234aaa} leads to a finite result. The metric dependence can be dealt with expanding the determinant of the metric in terms of co-factors (see appendix \ref{apb}). In this way, the fermionic contributions to the one-loop effective Lagrangian reduces to the following expression \begin{equation} \mathcal{L}_1^F = \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d \frac{ \Gamma(d) \zeta(d) 2^{E(d/2)}(1-2^{1-d}) }{(2\sqrt{\pi})^{d-1} \Gamma(\frac{d+1}{2})} , \label{eq:p0f} \end{equation} where $\Gamma$ and $\zeta$ are the Euler and Riemann functions, respectively. Proceeding similarly, the sum of the photon and ghost contributions in \eqref{LB} yields \begin{equation}\label{eq:p0b} \mathcal{L}_1^B = \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d (d-2) \frac{\Gamma(d) \zeta(d)}{(2 \sqrt{\pi})^{d-1} \Gamma(\frac{d+1}{2})}. \end{equation} Finally, adding together the results \eqref{eq:p0f} and \eqref{eq:p0b}, we obtain the one-loop effective Lagrangian in the form \begin{align}\label{eq11} \mathcal{L}_1 &= \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d \frac{\Gamma(d) \zeta(d)}{(2 \sqrt{\pi})^{d-1} \Gamma\left(\frac{d+1}{2}\right)} \nonumber \\ &\times \left[(d-2)+2^{E(d/2)}(1-2^{1-d}) \right]. \end{align} As expected for a density, Eq. \eqref{eq11} exhibits the factor $\sqrt{|g|}$. It also displays a temperature dependence in terms of the {\it local temperature} defined in Eq. \eqref{Tloc}, which is a simple consequence of the Tolman-Ehrenfest effect. The combination of these two factors leads to the invariance under the scale transformation $g_{\mu\nu} \rightarrow \sigma g_{\mu\nu}$. In the following sections we will investigate the effect of higher order corrections, when the thermal photons interact also with thermal fermions. \section{Effective Lagrangian at two-loop order} \begin{figure} \center \includegraphics[scale=1.0]{two_loops}\\ \caption{Two-loop contribution to the effective Lagrangian. The effective vertex and propagators are given by Eqs. \eqref{frules}}\label{2-loops} \end{figure} Let us now apply the technique illustrated in the simple one-loop calculation of the previous section, in order to obtain the two-loop order contribution to the effective Lagrangian. This can be obtained computing the diagram shown in Fig.~\ref{2-loops}. Using the effective Feynman rules given in \eqref{frules} we obtain \begin{align} \mathcal{L}_2 &= \frac{e^2T^2}{2} \sum_{m,\, n} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \frac{d^{d-1}q}{(2\pi)^{d-1}} \nonumber \\ & \times \mathrm{tr} \left( {\gamma}^\mu \frac{1}{{\gamma}^\nu p_\nu} {\gamma}_\mu \frac{1}{{\gamma}^\alpha p_\alpha} \right) \frac{1}{\sqrt{|g|}(p - q)^\gamma(p - q)_\gamma}, \end{align} where $q_0 = i (2n+1)\pi T$ and $p_0 = i (2m+1) \pi T$. In the vierbein basis the Dirac algebra yields \begin{align} \mathcal{L}_2 &= \frac{e^2T^2}{2\sqrt{|g|}} 2^{E(d/2)}(2-d) \sum_{m,\, n} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \frac{d^{d-1}q}{(2\pi)^{d-1}} \nonumber \\ &\times \frac{\tilde{p}^a \tilde{q}_a }{(\tilde{p}^b \tilde{p}_b) (\tilde{q}^c \tilde{q}_c)(\tilde{p}-\tilde{q})^d(\tilde{p}-\tilde{q})_d }, \end{align} which can be rewritten as \begin{align} \mathcal{L}_2 &=\frac{e^2 2^{E(d/2)}(2-d)}{4\sqrt{|g|}} \nonumber \\ &\times \left[2T^2 \sum_{m,\, l} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \frac{d^{d-1}k}{(2\pi)^{d-1}} \frac{1}{(\tilde{p}^a \tilde{p}_a) (\tilde{k}^b \tilde{k}_b)} \right. \nonumber \\ & \left. - T^2\sum_{m,\, n} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \frac{d^{d-1}q}{(2\pi)^{d-1}} \frac{1}{(\tilde{p}^a\tilde{p}_a)(\tilde{q}^b \tilde{q}_b)} \right] , \label{eq:p2fac} \end{align} where $k=p-q$ is the photon momentum ($k_0= i 2 l \pi$). There are two independent sum/integrals in Eq. \eqref{eq:p2fac}, namely \begin{equation} I_1 = T \sum_{m} \int \frac{d^{d-1}p}{(2 \pi)^{d-1}} \frac{1}{\tilde{p}_a \tilde{p}^a } \end{equation} and \begin{equation} I_2 = T \sum_{l} \int \frac{d^{d-1}k}{(2 \pi)^{d-1}} \frac{1}{\tilde{k}_b \tilde{k}^b}, \end{equation} which differ since $k_0= i 2\pi l T$ and $p_0 = i \pi (2m+1) T$. Using the definition of the locally rest vierbein \eqref{eq:vierbein-final}, the integral $I_1$ can be written as \begin{widetext} \begin{align} I_1 &=T \sum_{m} \int \frac{d^{d-1}p}{(2 \pi)^{d-1}} \frac{1}{ (p_0/\sqrt{g_{00}})^2 + g^{ij}[p_i - (g^{-1})_{ik} g^{0k} p_0][p_j - (g^{-1})_{jl} g^{0l} p_0]}. \end{align} \end{widetext} Proceeding as in the one-loop case, a change of variables in $p_i$ allows one to perform the thermal part of the resulting integral, which yields the result \begin{align} I_1 &= \frac{\sqrt{|g|}}{(g_{00})^{d/2-1}} T^{d-2}(2-2^{4-d}) \frac{\Gamma(d-2) \zeta(d-2)}{\Gamma(\frac{d-1}{2})(2\sqrt{\pi})^{d-1}}. \label{eq:i1-final} \end{align} Similarly, we obtain the following result for the integral $I_2$ \begin{align} I_2 &= -2 \frac{\sqrt{|g|}}{(g_{00})^{d/2-1}} T^{d-2} \frac{\Gamma(d-2) \zeta(d-2)}{\Gamma(\frac{d-1}{2})(2\sqrt{\pi})^{d-1}} . \label{eq:i2-final} \end{align} Substituting \eqref{eq:i1-final} and \eqref{eq:i2-final} in Eq. \eqref{eq:p2fac}, yields \begin{align}\label{eq21} \mathcal{L}_2 &= \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d \left[e^2 \left(\frac{T}{\sqrt{g_{00}}}\right)^{d-4} \right] 2^{E(d/2)-2}(2-d) \nonumber \\ &\times \left[\frac{\Gamma(d-2) \zeta(d-2)}{\Gamma(\frac{d-1}{2})(2\sqrt{\pi})^{d-1}}\right]^2 (2-2^{4-d})(6-2^{4-d}). \end{align} This result, which can be identified with the two-loop contribution to the pressure, in the high temperature limit, is similar to the flat space-time result, corrected by the factor $\sqrt{|g|}$ (as expected for a density). Also, the two-loop result exhibit the simple dependence on the local temperature, as defined in \eqref{Tloc}. \section{Non-perturbative contribution to the effective Lagrangian} \begin{figure} \includegraphics[scale=0.5]{rings} \caption{Ring diagrams.}\label{rings} \end{figure} Higher loop corrections to the effective Lagrangian may exhibit infrared divergences which arise from the dominant high temperature contribution of the zero mode. In order to deal with these divergences one has to sum an infinite series of diagrams which are individually divergent. Each successive order is obtained by inserting an extra one-loop static photon self-energy (the photon self-energy diagram is shown in Fig. \ref{pi-foton_g} of the appendix \ref{apc}). By power counting one can see that these zero mode contributions will produce infrared divergences in the high temperature limit, when the number of insertions is large enough. For instance, in four space-time dimensions diagrams with two or more self-energy insertions are infrared divergent. Figure \ref{rings} shows three such diagrams, corresponding to three-, four- and five-loops. Diagrams in this set are called {\it ring diagrams}. A diagram with $N$ self-energy insertions has the form \begin{widetext} \begin{align} R_N &= \frac{T}{2N} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \left[ (-1)^N\tilde{D}^{a_1 b_1} \tilde{\Pi}_{b_1 a_2} \cdots \tilde{\Pi}_{b_{N-1} a_{N}} \tilde{D}^{a_N b_N} \tilde{\Pi}_{b_N a_1} \right]_{\tilde{p}_0=0} , \end{align} where $D^{a b}$ is the photon propagator. Using the Eq. \eqref{eq:static-self-energy} for the static self-energy we obtain \begin{align} R_N&= \frac{T}{2N} \int \frac{d^{d-1}p}{(2\pi)^{d-1}} \left[ \frac{2m^2}{\sqrt{|g|} g^{ij} p_i p_j}\right]^N , \label{eq:jebun} \end{align} where $m$ is the thermal mass given by \eqref{Tmass}. Performing a change of variables, we obtain \begin{align} R_N &= \left( \frac{2 m^2}{\sqrt{|g|} }\right)^{(d-1)/2} T \sqrt{(-1)^{d-1} \bold{g}^{-1} } \int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \frac{1}{2N}\left[ \frac{1}{\eta^{ij} p_i p_j}\right]^N \nonumber \\ &= \sqrt{|g|} \left( \frac{2 m^2}{\sqrt{|g|} }\right)^{(d-1)/2} \frac{T}{\sqrt{g_{00}}} \int \frac{d^{d-1} p}{(2 \pi)^{d-1}} \frac{1}{2N}\left[ \frac{1}{\eta^{ij} p_i p_j}\right]^N, \label{eq:rnrn} \end{align} \end{widetext} where we have used Eq. \eqref{detg} (from now on $p_i$ are dimensionless variables). Simple power counting shows that these individual ring diagrams contributions are infrared divergent for $N\geq d/2$ or $N \geq (d-1)/2$ respectively when $d$ is even or odd (notice that the number of loops is $N+1$). Because of this different behavior, one has to consider separately the two cases. The sum of all the infrared divergent contributions can be written as \begin{widetext} \begin{align}\label{eq25aa} \mathcal{L}_{ring} &= \sum_{N=E(d/2)}^\infty R_N = \sqrt{|g|} \left( \frac{2 m^2}{\sqrt{|g|} }\right)^{(d-1)/2} \frac{T}{\sqrt{g_{00}}} \frac{1}{2^{d-2} \pi^{(d-1)/2} \Gamma \left( \frac{d-1}{2} \right) } \sum_{N=E(d/2)}^\infty \int_0^\infty dp \frac{p^{d-2}}{2N} \left[ \frac{-1}{p^2} \right]^N. \end{align} Using integration by parts, we can write \begin{eqnarray} \sum_{N=E(d/2)}^\infty \int_0^\infty dp \frac{p^{d-2}}{2N} \left[ \frac{-1}{p^2} \right]^N = \sum_{N=E(d/2)}^\infty \left[ \left. \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} \right|_0^\infty + \int_0^\infty dp \frac{p^{d-1}}{d-1} \frac{(-1)^N}{p^{2N+1}} \right] . \label{eq:intgeral} \end{eqnarray} One can readily verify that the surface term vanishes when $d$ is an even number. First, for $p\rightarrow \infty$ it is immediate that \begin{equation} \sum_{N=d/2}^\infty \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} \longrightarrow 0 . \end{equation} When $p\rightarrow 0$ we obtain \begin{align} \sum_{N=d/2}^\infty \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} = \frac{p^{d-1}}{2(d-1)}\log(1+1/p^2) + \sum_{N=1}^{d/2-1} \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} \longrightarrow 0. \end{align} Therefore, Eq. \eqref{eq:intgeral} with even values of $d$ reduces to \begin{align} \sum_{N=d/2}^\infty \int_0^\infty dp \frac{p^{d-2}}{2N} \left[ \frac{-1}{p^2} \right]^N = \frac{(-1)^{d/2}}{d-1} \int_0^\infty \frac{dp}{p^2} \frac{1}{1+1/p^2} = \frac{\pi (-1)^{d/2}}{2(d-1)}. \label{eq:sum-ring-even} \end{align} Substituting the Eq. \eqref{eq:sum-ring-even} into Eq. \eqref{eq25aa}, we obtain the following non-perturbative contribution for even space-time dimensions \begin{align}\label{Leven} \mathcal{L}_{ring}^{d-even} &= \sqrt{|g|} \frac{T}{\sqrt{g_{00}}} \frac{(-1)^{d/2}}{2^d \pi^{(d-3)/2} \Gamma\left( \frac{d+1}{2} \right) } \left[ \frac{2 m^2}{\sqrt{|g|}} \right]^{(d-1)/2} \nonumber \\ &= \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d \left[e^2 \left(\frac{T}{\sqrt{g_{00}}}\right)^{d-4} \right]^{(d-1)/2} \frac{(-1)^{d/2}}{2^d \pi^{(d-3)/2} \Gamma\left( \frac{d+1}{2} \right) } \left[ \frac{2^{E(d/2)+1}(1-2^{3-d}) \Gamma(d-1) \zeta(d-2)}{\Gamma\left(\frac{d-1}{2}\right) (2 \sqrt{\pi})^{d-1}} \right]^{(d-1)/2} , \end{align} which exhibits a non-analyticity in the coupling constant of the form $(e^2)^{(d-1)/2}$. Let us now consider the case when $d$ is odd. In this case, one can show that the surface term in \eqref{eq:intgeral} does not vanish. Indeed, although in the limit $p\rightarrow 0$ we obtain \begin{align}\label{eqodd1} \sum_{N=(d-1)/2}^\infty \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} = \frac{p^{d-1}}{2(d-1)}\log(1+1/p^2) + \sum_{N=1}^{(d-3)/2} \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} \longrightarrow 0, \end{align} when $p\rightarrow \infty$ we are left with the following finite contribution \begin{align}\label{eqodd2} \sum_{N=(d-1)/2}^\infty \frac{p^{d-1}}{2N(d-1)} \frac{(-1)^N}{p^{2N}} \longrightarrow \frac{(-1)^{(d-1)/2}}{(d-1)^2}. \end{align} Substituting Eqs. \eqref{eqodd1} and \eqref{eqodd2} into Eq. \eqref{eq:intgeral} we obtain \begin{align} \sum_{N=E(d/2)}^\infty \int_0^\infty dp \frac{p^{d-2}}{2N} \left[ \frac{-1}{p^2} \right]^N &= \frac{(-1)^{(d-1)/2}}{(d-1)^2} + \sum_{N=(d-1)/2}^\infty \int_0^\infty dp \frac{p^{d-1}}{d-1} \frac{(-1)^N}{p^{2N-1}} \nonumber \\ &= \frac{(-1)^{(d-1)/2}}{(d-1)} \left[ \frac{1}{d-1} + \int_0^\infty \frac{dp}{p} \frac{1}{1+1/p^2} \right]. \end{align} The resulting integral is logarithmically increasing at large momenta, where the approximations used for the ring diagrams are no longer valid. In order to regularize this behavior, we will employ a cut-off $\sqrt{-g^{ij} P_i P_j}$ in the momentum of the original integral in Eq. \eqref{eq:jebun}, where $P$ is naturally of the same order as the local temperature. In terms of the parameter $\mu = \sqrt{-\eta^{ij} P_i P_j}$, we then get \begin{align} \sum_{N=E(d/2)}^\infty \int_0^\infty dp \frac{p^{d-2}}{2N} \left[ \frac{-1}{p^2} \right]^N &= \frac{(-1)^{(d-1)/2}}{2(d-1)} \left[ \frac{2}{d-1} + \log \left( 1+ \frac{\mu^2 \sqrt{|g|}}{2 m^2} \right) \right], \end{align} which yields the following expression for the non-perturbative contribution to the effective Lagrangian \begin{align}\label{Lodd} \mathcal{L}_{ring}^{d-odd} &= \sqrt{|g|} \left( \frac{2 m^2}{\sqrt{|g|} }\right)^{(d-1)/2} \frac{T}{\sqrt{g_{00}}} \frac{(-1)^{(d-1)/2} }{2^{d} \pi^{(d-1)/2} \Gamma \left( \frac{d+1}{2} \right) } \left[ \frac{2}{d-1} + \log \left( 1+ \frac{\mu^2 \sqrt{|g|}}{2 m^2} \right) \right] \nonumber \\ &= \sqrt{|g|} \left(\frac{T}{\sqrt{g_{00}}}\right)^d \left[e^2 \left(\frac{T}{\sqrt{g_{00}}}\right)^{d-4} \right]^{(d-1)/2} \frac{(-1)^{(d-1)/2} }{2^{d} \pi^{(d-1)/2} \Gamma\left( \frac{d+1}{2} \right) } \left[ \frac{2^{E(d/2)+1}(1-2^{3-d}) \Gamma(d-1) \zeta(d-2)}{\Gamma\left(\frac{d-1}{2}\right) (2 \sqrt{\pi})^{d-1}} \right]^{(d-1)/2} \nonumber \\ & \times \left\{ \frac{2}{d-1} + \log \left[ 1+ \mu^2 \left(\frac{T}{\sqrt{g_{00}}}\right)^{-2} \left(e^2 \left(\frac{T}{\sqrt{g_{00}}}\right)^{d-4} \right)^{-1} \frac{\Gamma\left(\frac{d-1}{2}\right) (2 \sqrt{\pi})^{d-1}}{2^{E(d/2)+1}(1-2^{3-d}) \Gamma(d-1) \zeta(d-2)} \right] \right\}. \end{align} \end{widetext} This expression has a logarithmic non-analyticity in the coupling constant $e^2$. The same type of non-analyticities have been found previously in the context of scalar fields in flat backgrounds \cite{Brandt:2012mu}. As in the previous results for the one- and two-loop contributions to the effective action, the non-perturbative results in this section given by Eqs. \eqref{Leven} and \eqref{Lodd} can be expressed in terms of the local temperature as defined in Eq. \eqref{Tloc}. This confirms that the Tolman-Ehrenfest effect is explicitly manifested even in the extreme case when an infinite number of interactions are taken into account. \section{Discussion} In the present work, we have employed the equivalence between static and zero energy-momentum thermal amplitudes, which holds for the leading contributions at high temperature. Using this correspondence, we have obtained the effective Lagrangian of static gravitational fields interacting with a plasma of photons and electrons at high temperature, up to two-loops order. We have also obtained a non-perturbative contribution from the sum of the infinite set of ring diagrams. This generalizes the previous results for the static effective action which were known only at one-loop order \cite{Brandt:2012ei}. It is interesting to remark that the contributions generated by the static gravitational fields correspond to those obtained for the pressure of a QED plasma in Minkowski space-time in the following simple way: apart from an overall factor of $\sqrt{|g|}$ which is required by gauge invariance, the only modification involves the replacement of the Minkowski temperature by the local temperature \eqref{Tloc}. From a physical point of view, this universal behavior (which has also been derived using other approaches \cite{Balazs2:1965,Ebert:1973,Rovelli:2010mv,Haggard:2013fx}) can be traced back to the requirement of thermal equilibrium in a gravitational field. Indeed, the emergence of a local temperature, and consequently a temperature gradient, is unavoidable in thermal equilibrium to prevent heat (which interacts with gravity) to flow from regions of higher to those of lower gravitational potential. Another salient feature is that the conformal invariance of the effective Lagrangian, which is present at one-loop order for any space-time dimension $d$, is not in general satisfied by the higher loop corrections (this is also the case of the one-loop photon self-energy given by Eq. \eqref{eq:static-self-energy}). Both in the two-loop correction in Eq. \eqref{eq21} as well as in the non-perturbative contributions in Eqs. \eqref{Leven} and \eqref{Lodd} (which receives contributions of an infinite number of photon self-energy insertions), we find terms like \begin{align}\label{eq36} \left[ e^2 \left( \frac{T}{\sqrt{g_{00}}} \right)^{d-4} \right] \end{align} which will not be conformal invariant in general. The physical reason for this behaviour may be understood by noting that $e^2$ is a dimensionful quantity with canonical mass dimension $4-d$. We finally remark that the above leading results at high temperature were obtained by neglecting all masses compared with the temperature. In four dimensions, since $e^2$ is dimensionless, these results should therefore be scale invariant. In this case, because $\sqrt{|g|} T_{\rm loc}^4$ is invariant under scale transformations, we see that the modification $T\rightarrow T_{\rm loc}$ is the only possibility which is consistent with the Weyl symmetry. When $d\neq 4$, the leading thermal results are no longer scale invariant (see Eq. \eqref{eq36}), but the violation of the Weyl symmetry still occurs according to the simple prescription $T\rightarrow T_{\rm loc}$, which enforces a smooth behaviour when $d\rightarrow 4$. Based on the above physical considerations and explicit calculations we conclude that, to all orders, the simple correspondence $T\rightarrow T_{\rm loc}$ leads to the effective action of static gravitational fields interacting with a QED plasma at high temperature. Moreover, the terms involving $T_{\rm loc}$ ensure the invariance of this action under time-independent local coordinate transformations. Our treatment in arbitrary space-time dimensions was motivated by various unified field theories of gravitational, electromagnetic and other interactions, which have been often formulated in higher dimensions. But at present, we can indicate a direct physical application of the above results only when $d=4$. These results may then be useful to calculate, for example, the pressure in a plasma of electrons and photons surrounding a hot star. \acknowledgments We would like to thank FAPESP and CNPq (Brazil) for a grant. J. F. is indebted to Prof. J. C. Taylor for a helpful correspondence.
\section{Introduction} The discussion of matrix eigenvalue problems in terms of deflating subspaces and the solutions of matrix equations has received a great deal of attention, particularly in its wide range of important applications in control theory~\cite{Gohberg1982,Mehrmann1991,Zhou1996}. It is famous that the solutions of Riccati and matrix polynomial equations can be obtained by computing its corresponding deflating subspaces and invariant subspaces, respectively. For example, consider a matrix (unilateral) polynomial equation with degree $m$ \begin{align}\label{MM} A_m X^m+A_{m-1} X^{m-1}+\cdots+A_0=0, \end{align} where the complex coefficient matrices $A_0,\cdots,A_m$ and unknown matrix $X$ are of size $n\times n$. Let the \emph{companion} matrix pencil $\mathcal{M}-\lambda \mathcal{L}\in \mathbb{C}^{mn \times mn} \times \mathbb{C}^{mn \times mn}$ be defined as \begin{align}\label{MLL} \mathcal{M}= \begin{bmatrix} 0 & I & 0 & \cdots & 0 \\ 0 & 0 & I & & 0 \\ \vdots & \vdots & \ddots & 0 & \\ 0 & 0 & \cdots & 0 & I \\ -A_0 & -A_1 & \cdots & -A_{m-2} & -A_{m-1} \\ \end{bmatrix},\,\, \mathcal{L}= \begin{bmatrix} I & 0 &\cdots&\cdots& 0 \\ 0 & I & & &0\\ \vdots & &\ddots& & \\ 0 & 0 & & & A_m \\ \end{bmatrix}, \end{align} where $I$ and $0$ denote respecitively the identity and zero matrix with appropriate size. It is easy to show \begin{align}\label{ML} \mathcal{M} U =\mathcal{L} UX, \end{align} with the ``column matrix'' $U=\begin{bmatrix} I \\ X\\\vdots\\ X^{m-1}\end{bmatrix}$. According to the equality~\eqref{ML}, the information of solution $X$ of ~\eqref{MM} is embedded in the eigeninformation of the generalized eigenvalue problem~\eqref{MLL}. Our main point in this work is to connect deflating subspaces, or more specifically invariant subspaces, with solutions of the following linear matrix equations, \begin{subequations} \begin{itemize} \item[1.] Standard Sylvester matrix equation: \begin{align}\label{sylvester} AX+XB=C,\quad A\in\mathbb{C}^{m \times m},B\in\mathbb{C}^{n \times n},C,X\in\mathbb{C}^{m \times n}, \end{align} \item[2.] $\star$-Sylvester matrix equation: \begin{align}\label{CTS} AX+X^\star B=C,\quad A,B,C,X\in\mathbb{C}^{n \times n}, \end{align} where $\star=\top$ is the transport operator or $\star=H$ is the Hermitian operator. \end{itemize} \end{subequations} The study of the Sylvester equations of the form~\eqref{sylvester} has been widely discussed in theory and applications. Especially, ~\eqref{sylvester} plays an indispensable role in a variety of fields of control theory \cite{Lancaster1995}. On the other hand, the study of the $\star$-Sylvester equations stems from the treatment of completely integrable mechanical systems. When $\star=\top$, a formula for the solution of \eqref{CTS} in terms of generalized inverse and some necessary and sufficient conditions for the existence of the solution of \eqref{CTS} were proposed in \cite{Piao2007}. In recent years, an extensive amount of iterative methods based on the conjugate gradient method were studied and developed for solving the generalized $T$-Sylvester equation \[ A X B + C X^T D = E \ , \ \ \ A,B,C,D,E,X \in \mathbb{R}^{n \times n}. \] See, e.g., \cite{Su2010, Hajariann2013} and the references cited therein. However, for the matrix equation~\eqref{CTS} especially with $\star = H$, there are not many references in the literature and in particular, the developed method so far for solving~\eqref{CTS} is through the application of the generalized Schur form of the pencil $A-\lambda B^\star$~\cite{Chiang2012}. In this paper, the solutions of the standard and $\star$-Sylvester equations are studied in terms of the study of invariant subspace and deflating subspace methods. In particular, we are mainly interested in the square cases when $m = n$ for $\star$-Sylvester equations. Other relative works can be found in \cite{Hajariann2013,De2011}. The paper is organized as follows. In Section 2 we recall some properties of the generalized eigenvalue problems. In Section 3 we give an invariant subspace method for computing the solution of the standard Sylvester equation~\eqref{sylvester}. In Section 4 we show how the deflating subspace method can be applied to solve {$\star$-Sylvester matrix equation}~\eqref{CTS}. In Section 5 a quadratic convergence method is provided for finding the stabilizing solution of ~\eqref{CTS} and concluding remarks are given in Section 6. \section{Preliminaries} \label{sec_background} In this section we briefly review some properties of matrix pencil which are required in the statements and the proofs in the following sections. To facilitate our discussion, we use $\sigma(A)$ and $\sigma(A-\lambda B)$ to denote the spectrum of the matrix $A$ and the matrix pair $A-\lambda B$, respectively and the notion $\sim$ to denote the spectral equivalent condition, that is, $A-\lambda B \sim \widetilde{A} -\lambda \widetilde{B}$ implies that $\sigma(A-\lambda B) = \sigma(\widetilde{A} -\lambda \widetilde{B})$. Given a matrix pencil $A-\lambda B$, the pair $A-\lambda B$ is said to be \emph{regular} if $\det(A-\lambda B) \neq 0$ for some $\lambda\in \mathbb{C}$. One strategy to analyze the eigeninformation is to transform one matrix pencil to its simplified and equivalent form. That is, two matrix pencils $A-\lambda B$ and $\widetilde{A} -\lambda \widetilde{B}$ are said to be equivalent if and only if there exist two nonsingular matrices $P$ and $Q$ such that \begin{equation*} P(A-\lambda B) Q = \widetilde{A} -\lambda \widetilde{B}. \end{equation*} Like similarity discussed in the ordinary eigenvalue problem, the property of equivalence preserves eigenvalues and transforms eigenvectors in a similar way. This result can be easily understood through the following well-known result given by Weierstrass~\cite{Weierstrass1868} and Kronecter~\cite{Kronecker1890} and is also discussed in~\cite[Defintion 1.13]{Stewart1990}. \begin{theorem}\label{Weierstrass} [Weierstrass canonical form for a matrix pencil] Let $A-\lambda B$ be a regular pair. Then there exist nonsingular matrices $P$ and $Q$ such that \begin{equation}~\label{wJform} PAQ = \left[\begin{array}{cc} J & 0 \\0 & I\end{array}\right] \mbox{ and } PBQ = \left[\begin{array}{cc} I & 0 \\0 & N\end{array}\right], \end{equation} where $J$ and $N$ are in Jordan canonical form and diagonal entries of $N$ are zero (i.e., $N$ is nilpotent). Also, we can simplify this result by the notation of direct sum. That is, $A-\lambda B \sim (J\oplus I)-\lambda (I \oplus N)$. Sometimes the canonical form in~\eqref{wJform} is also called the \emph{Kronecker canonical form}. \end{theorem} From theorem~\ref{Weierstrass}, it is easy to see that if $\lambda$ is an eigenvalue of $A-\lambda B$ with eigenvector $\mathbf{x}$, then $\lambda$ is an eigenvalue of $PAQ - \lambda PBQ$ with eigenvector $Q^{-1}\mathbf{x}$. Note that for an $n\times n$ matrix $A$, the generalized eigenvectors of $A$ span the entire $\mathbb{R}^n$ space. This property is also true for every regular matrix pencil and is demonstrated as follows. For a detailed proof, the reader is referred to~\cite[Theorem 7.3]{Gohberg1982}. \begin{theorem}\label{lancaster} Given a matrix pair of $n\times n$ matrix $A$ and $B$, if the matrix pencil $A-\lambda B$ is regular, then its Jordan chains corresponding to all finite and infinite eigenvalues carry the full spectral information about the matrix pencil and consists of $n$ linearly independent vectors. \end{theorem} \section{Standard Sylvester equations} It is true that the standard Sylvester matrix equation~\eqref{sylvester} can be viewed as the application of Newton method to the nonsymmetric algebraic Riccati equation (NARE): \begin{equation}\label{nare} C+XA+DX-XBX = 0, \end{equation} where $X\in\mathbb{C}^{m\times n}$ is the unknown, and where the coefficients are $A\in\mathbb{C}^{n\times n}$, $B\in\mathbb{C}^{n\times m}$, $C\in\mathbb{C}^{m\times n}$, and $D\in\mathbb{C}^{m\times m}$, and the solution $X$ of~\eqref{nare} can be solved by considering the invariant subspace of the Hamiltonian-like matrix~\cite{GuoXULin2006} \begin{equation} \mathcal{H} = \left[\begin{array}{cc}-A & B \\C & D\end{array}\right]. \end{equation} But, unlike the NARE, the solvability of the standard Sylvester equation don't require rigorous constraints on the matrix $\mathcal{H}$ as given in~\cite{GuoXULin2006}. Instead, we want to show that the invariant subspace method for solving~\eqref{sylvester} can be developed by simply applying the following solvability condition of the Sylvester equation~\eqref{sylvester}~\cite{Horn1994}. \begin{theorem} The equation of~\eqref{sylvester} has a unique solution for each $C$ if and only if $\sigma(-A)\cap\sigma(B) = \phi$. \end{theorem} In this section, we want to establish the idea of invariant subspace. Its theory is an application of the following result~\cite{Chu1987}. \begin {theorem}\label{Eric0} Given two regular matrix pencils $A_i-\lambda B_i\in\mathbb{C}^{n_i\times n_i}$, $1\leq i \leq 2$. Consider the following equations with respect to $U,V\in\mathbb{C}^{n_1\times n_2}$ \begin{subequations}\label{CM} \begin{align} A_1 U &= V A_2,\\%\label{CM1} B_1 U &= V B_2 \end{align} \end{subequations} If $\sigma(A_1-\lambda B_1)\cap\sigma(A_2-\lambda B_2)=\phi$, then~\eqref{CM} has a unique solution $U=V=0$. \end {theorem} Note that Theorem~\ref{Eric0} yields a Corollary which is simple, but useful in our subsequent discussion. \begin{corollary}\label{Eric} Let $A\in\mathbb{C}^{n\times n}$ and $T\in\mathbb{C}^{k\times k}$. If $\sigma(A)\cap\sigma(T)=\phi$, then the equation \begin{align*} AU=U T \end{align*} have the unique solution $U=0\in\mathbb{C}^{n\times k}$. \end{corollary} Now we have enough tools for discussing the invariant subspace corresponding to~\eqref{sylvester}. \begin {theorem}\label{thm2} Let $A$, $B$, $C$ be the matrices given in~\eqref{sylvester}. If $\mathcal{M}=\begin{bmatrix} -A & C\\0 & B \end{bmatrix} \in \mathbb{C}^{(m+n)\times (m+n)}$ and if $\sigma(-A)\cap \sigma(B)=\phi$, let us write \begin{align* \mathcal{M}\begin{bmatrix} U \\ V \end{bmatrix}= \begin{bmatrix} U\\V \end{bmatrix} T, \end{align*} where $\begin{bmatrix} U\\V \end{bmatrix}$ is full rank. Then, we have \begin{enumerate} \item $V=0$ and $-A U=U T$ if $\sigma (T)=\sigma(-A)$. \item $V$ is nonsingular and $X=UV^{-1}$ is the solution of~\eqref{sylvester} if $\sigma (T)=\sigma(B)$. \end{enumerate} \end {theorem} \begin{proof} Since $\sigma(-A)\cap \sigma(B)=\phi$, it follows from Corollay~\ref{Eric} that the equation $BV = -VA$ has only one solution $V = 0$. This proves part 1. Consider two equations \begin{align*} \mathcal{M}\begin{bmatrix} U_1 \\ 0 \end{bmatrix} &= \begin{bmatrix} U_1\\0 \end{bmatrix} T_1,\, \mbox{with } \sigma(T_1)=\sigma(-A),\\ \mathcal{M}\begin{bmatrix} U_2 \\ V_2 \end{bmatrix} &= \begin{bmatrix} U_2\\V_2 \end{bmatrix} T_2,\, \mbox{with } \sigma(T_2)=\sigma(B). \end{align*} From Theorem~\ref{lancaster}, all column vectors of $\begin{bmatrix} U_1\\0 \end{bmatrix}$ and $\begin{bmatrix} U_2\\V_2 \end{bmatrix}$ are linearly independent. This implies that $\begin{bmatrix} U_1 & U_2\\0 & V_2 \end{bmatrix}$ is nonsingular, that is, $V_2$ is nonsingular. Observe further that \begin{subequations} \begin{align*} -A U_2 + C V_2 &= U_2 T_2,\\ B V_2 &= V_2T_2. \end{align*} \end{subequations} Upon substitution, we see that $A (U_2V_2^{-1}) + (U_2V_2^{-1}) B = C$ and the proof is complete. \end{proof} From Theorem~\ref{thm2}, we know that, in order to compute the solution of~\eqref{sylvester}, it is sufficient to compute a base for the invariant subspace associated with the eigenvalues of $B$. Some acceleration iterative methods like doubling algorithm \cite{GuoXULin2006} for finding the unique solution of ~\eqref{sylvester} are based on the relationship between ~\eqref{sylvester} and invariant subspace in Theorem~\ref{thm2}. We don't further discuss here. \section{$\star$-Sylvester equations} Before demonstrating the unique solvability conditions~\eqref{CTS}, we need to define that a subset $\Lambda = \{\lambda_1,\ldots,\lambda_n\}$ of complex numbers is said to be \emph{$\star$-reciprocal free} if and only if $\lambda_i\neq1/\lambda_j^\star$ for $1\leq i,j \leq n$. This definition also regards $0$ and $\infty$ as reciprocals of each other. Note that the necessary and sufficient conditions for unique solvability of~\eqref{CTS} are given in~\cite{Wimmer1994} by means of Roth's criterion. Consult also~\cite[Lemma 5.10]{Byers2006} and~\cite[Lemma 8]{Kressner2009}, where solvability conditions for the $\star$-Sylvester equations with $m = n$ were obtained, without considering the details of the solution process. That is, we have the following solvability conditions of~\eqref{CTS}. \begin{theorem}\label{solvable} Suppose that the pencil $A-\lambda B^\star$ is regular, the $\star$-Sylvester matrix equation~\eqref{CTS} is uniquely solvable if and only if \begin{itemize} \item[1.] For $\star = \top$, $\sigma(A^\top-\lambda B)\backslash \{1\}$ is reciprocal free, and whenever $1\in\sigma(A^\top-\lambda B)$, $1$ is simple; \item[2.] For $\star = H$, $\sigma(A^H-\lambda B)$ is reciprocal free, and $|\lambda| \neq 1$, whence $\lambda\in\sigma(A^H-\lambda B)$. \end{itemize} \end{theorem} The most straightforward way to find the solution of~\eqref{CTS} is through the analysis of its corresponding \emph{palindromic} eigenvalue problem \cite{Chu2010,NLA:NLA612,Mackey2006,doi:10.1137/050628350} \begin{equation}\label{eq:PEP} \mathcal{Q}(\lambda)x: = (\mathcal{Z}^\star-\lambda \mathcal{Z})x=0, \end{equation} where \begin{equation}\label{eq:Zmat} \mathcal{Z} =\begin{bmatrix} 0 & B \\ A & -C \end{bmatrix}\in\mathbb{C}^{2n\times 2n}, \end{equation} and the discussion of the deflating subspace of~\eqref{eq:PEP}. Again, the operator $(\cdot)^\star$ denotes either the transpose ($\top$), or the conjugate transpose ($H$) of a matrix. We interpret both cases in a unified way hereafter. The name ``palindromic'' stems from the invariant property of $\mathcal{Q(\lambda)}$ under reversing the order and taking (conjugate) transpose of the coefficient matrices, that is, \begin{equation*} \mbox{rev} \mathcal{Q}(\lambda) = \mathcal{Q}^\star(\lambda),\, \mbox{where } \mbox{rev} \mathcal{Q}(\lambda) : = (\mathcal{Z}-\lambda \mathcal{Z}^\star). \end{equation*} The corresponding eigenvalue problems of~\eqref{eq:PEP} were originally generated from the vibrational analysis of trail tracks for obtaining information on reducing noise between wheel and rail~\cite{Ipsen04,Hilliges2004}. Our task in this section is to identify eigenvectors of problem~\eqref{eq:PEP} and then associate these eigenvectors with the solution of~\eqref{CTS}. We begin this analyst by studying the eigeninformation of two matrices $A$ and $B$, where $A-\lambda B$ is a regular matrix pencil. \begin{lemma}\label{LemmAB} Let $A-\lambda B\in\mathbb{C}^{n\times n}$ be a regular matrix pencil. Assume that matrices $X_i,Y_i\in\mathbb{C}^{n\times n_i}$, $i=1,2$, are full rank and satisfy the following equations \begin{subequations}\label{eq:abxy} \begin{eqnarray} A X_i &=& Y_i R_i,\\ B X_i &=& Y_i S_i, \end{eqnarray} \end{subequations} where $R_i$ and $S_i$, $i=1,2$, are square matrices of size $n_i\times n_i$. Then \begin{itemize} \item [i)] $R_i-\lambda S_i\in\mathbb{C}^{n_i\times n_i}$ are regular matrix pencils for $i=1,2$. \item [ii)] if $\sigma(R_1-\lambda S_1)\cap \sigma(R_2-\lambda S_2)=\phi$, then the matrix $\begin{bmatrix} X_1 & X_2\end{bmatrix}\in\mathbb{C}^{n\times (n_1+n_2)}$ is full rank. \end{itemize} \end{lemma} \begin{proof} Fix $i\in \{1,2\}$. To show that $R_i-\lambda S_i$ is a regular pencil, it suffices to show that if $\lambda$ is an eigenvalue of $R_i-\lambda S_i$, then $\lambda$ is an eigenvalue of $A-\lambda B$. Let $\mathbf{x} \neq 0$ be an eigenvector of $R_i-\lambda S_i$ corresponding to the eigenvalue $\lambda$, that is, \begin{align}\label{eq:rs} \begin{array}{rc} (R_i - \lambda S_i) \mathbf{x} = 0, & \mbox{if }\lambda < \infty,\\ (S_i - 0 R_i) \mathbf{x} = 0, & \mbox{if }\lambda = \infty. \end{array} \end{align} We then pre-multiply both sides of~\eqref{eq:rs} by $X_i$ and obtain \begin{align*} \begin{array}{rc} (A - \lambda B) X_i\mathbf{x} = 0, & \mbox{if }\lambda < \infty,\\ (B - 0 A) X_i\mathbf{x} = 0, & \mbox{if }\lambda = \infty. \end{array} \end{align*} Since $X_i$ is full rank and $\mathbf{x}\neq 0$, we have $X_i \mathbf{x}\neq 0$. Hence, $\lambda$ is an eigenvalue of $(A, B)$. This proves part i). Next, by Theorem~\ref{Weierstrass} there exist nonsingular matrices $P_i,\,Q_i$, Jordan block matrices $J_i\in\mathbb{C}^{k_i\times k_i}$, and nilpotent matrices $N_i$, $i=1,2$ (exactly one of $N_1$ or $N_2$ exists since the regularity of $R_i-\lambda S_i$) such that \begin{subequations}\label{eq:pq} \begin{eqnarray} (P_1R_1Q_1, P_1S_1Q_1)&=&(\begin{bmatrix} J_1 & \\ & I \end{bmatrix}, \begin{bmatrix} I & \\ & N_1 \end{bmatrix}),\\ (P_2R_2Q_2, P_2 S_2Q_2)&=& (\begin{bmatrix} J_2 & \\ & I \end{bmatrix}, \begin{bmatrix} I & \\ & N_2 \end{bmatrix}). \end{eqnarray} \end{subequations} By~\eqref{eq:abxy} and~\eqref{eq:pq}, it can be seen that \begin{align*} A X_{i} Q_i &= Y_i P_i^{-1} \begin{bmatrix} J_i & 0\\ 0 & I \end{bmatrix},\\ B X_{i} Q_i &= Y_i P_i^{-1} \begin{bmatrix} I & 0 \\ 0 & N_i \end{bmatrix}. \end{align*} Let $X_iQ_i=\begin{bmatrix} X_{i,1} & X_{i,2} \end{bmatrix}$ and $Y_iP_i^{-1}=\begin{bmatrix} Y_{i,1} & Y_{i,2} \end{bmatrix}$ be two partitioned matrices with size $n_i \times (k_i+(n_i-k_i)) $ for $i=1,2$. It then follows from direct computation that the matrix pair $A-\lambda B$ satisfies \begin{align*} A \begin{bmatrix} X_{1,1} & X_{2,1} \end{bmatrix} &= B \begin{bmatrix} X_{1,1} & X_{2,1} \end{bmatrix} \begin{bmatrix} J_1 & 0\\ 0 & J_2 \end{bmatrix},\\ B \begin{bmatrix} X_{1,2} & X_{2,2} \end{bmatrix} &= A \begin{bmatrix} X_{1,2} & X_{2,2} \end{bmatrix} \begin{bmatrix} N_1 & 0 \\ 0 & N_2 \end{bmatrix}. \end{align*} Since $\sigma(R_1-\lambda S_1)\cap \sigma(R_2-\lambda S_2)=\phi$, we might assume without loss of generality that $N_2$ does not exist, i.e., $(P_2R_2Q_2, P_2 S_2Q_2)= (J_2, I )$. Since the condition $\sigma(R_1-\lambda S_1)\cap \sigma(R_2-\lambda S_2)=\phi$ holds, it then follows from Theorem~\ref{lancaster} that the matrix $\begin{bmatrix} X_1 & X_2\end{bmatrix}=\begin{bmatrix} X_{1,1}&X_{2,1}&X_{1,2}\end{bmatrix}$ is full rank. \end{proof} Armed with the property given in Lemma~\ref{LemmAB}, we can now attack the problem of determine how the deflating subspace is related to the solution of ~\eqref{CTS}. \begin {theorem}\label{deflatingthm} Corresponding to~\eqref{CTS}, let $\mathcal{Z}$ be a matrix defined by~\eqref{eq:Zmat}. If $\sigma(A^\star-\lambda B)$ is reciprocal free, and $U_i,\,V_i\in\mathbb{C}^{n\times n}$, $i = 1,2$, are matrices satisfying \begin{subequations}\label{eq:cond1} \begin{align} \mathcal{Z}^\star \begin{bmatrix} U_1 \\ V_1 \end{bmatrix}= \begin{bmatrix} U_2 \\ V_2 \end{bmatrix} T_1^\star,\label{d1}\\ \mathcal{Z} \begin{bmatrix} U_1 \\ V_1 \end{bmatrix}= \begin{bmatrix} U_2 \\ V_2 \end{bmatrix} T_2,\label{d2} \end{align} \end{subequations} for some matrices $T_1, T_2 \in \mathbb{C}^{n\times n}$. Then, \begin{itemize} \item [i)] $V_1=U_2=0$ if $T_1^\star-\lambda T_2\sim B^\star-\lambda A$; \item [ii)] $V_1$ is nonsingular if $T_1^\star-\lambda T_2\sim A^\star-\lambda B$. Moreover, $X=U_1V_1^{-1}=-U_2^{-\star} V_2^\star$ solves the $\star$-Sylvester matrix equation~\eqref{CTS} if $B$ is a nonsingular. \end{itemize} \end {theorem} \begin{proof} \begin{itemize} \item[i)] It follows from \eqref{d1} and \eqref{d2} that $A^\star V_1 = U_2 T_1^\star$ and $B V_1 =U_2 T_2$. Since $\sigma(A^\star-\lambda B)\cap\sigma(T_1^\star-\lambda T_2)=\phi$, we have $V_1 = U_2 = 0$ by Theorem~\ref{Eric0}. \item[ii)] Notice that, since $T_1^\star-\lambda T_2\sim A^\star-\lambda B$, there exist nonsingular matrices $U$ and $V$ such that \begin{equation}\label{eq:ccond2} B^\star U = V T_2^\star \mbox{ and } A U = V T_1. \end{equation} Thus, it can be seen that \begin{subequations}\label{eq:k} \begin{align} \mathcal{Z}^\star\begin{bmatrix} U \\ 0 \end{bmatrix} &=\begin{bmatrix} 0 \\ V\end{bmatrix} T_2^\star,\label{space1}\\ \mathcal{Z}\begin{bmatrix} U \\ 0 \end{bmatrix} &=\begin{bmatrix} 0 \\ V\end{bmatrix} T_1. \label{space2} \end{align} \end{subequations} Hence, by~\eqref{eq:cond1} and~\eqref{eq:k}, we have \begin{subequations}\label{eq:cond2} \begin{align} \mathcal{Z}^\star\begin{bmatrix} U & U_1\\ 0& V_1 \end{bmatrix} &=\begin{bmatrix} 0 & U_2\\ V & V_2 \end{bmatrix} \begin{bmatrix} T_2 & 0\\ 0 & T_1^\star\end{bmatrix},\\ \mathcal{Z}\begin{bmatrix} U & U_1\\ 0& V_1 \end{bmatrix} &=\begin{bmatrix} 0 & U_2\\ V & V_2 \end{bmatrix} \begin{bmatrix} T_1^\star & 0\\ 0 & T_2\end{bmatrix}. \label{space22} \end{align} \end{subequations} Since $\sigma(\mathcal{Z}^\star - \lambda \mathcal{Z}) = \sigma(A^\star - \lambda B) \cup \sigma(B^\star - \lambda A)$, by Lemma~\ref{LemmAB}, the matrix $\begin{bmatrix} U & U_1\\ 0& V_1 \end{bmatrix}$ is nonsingular. Thus, $V_1$ is nonsingular. To show that $X = U_1V_1^{-1}$ is the solution of~\eqref{CTS}, we first claim that $U_2$ is nonsingular. From~\eqref{eq:ccond2}, we see that $T_2$ is invertible since matrices $B$, $U$, and $V$ are nonsingular. From~\eqref{space22}, we know that $ B V_1=U_2T_2 $. Since $B$, $T_2$ and $V_1$ are nonsingular, this implies that $U_2$ is invertible. Let $\widehat{T}_1=U_2 T_1^\star V_1^{-1}$, $\widehat{T}_2=U_2 T_2 V_1^{-1}$, $X = U_1V_1^{-1}$, and $Y = V_2U_2^{-1}$. It follows from~\eqref{eq:cond2} that \begin{align*} A^\star &= \widehat{T}_1, \quad B^\star X-C^\star = Y \widehat{T}_1,\\ B &= \widehat{T}_2, \quad AX-C =Y\widehat{T}_2. \end{align*} This implies that $ A (-Y^\star)+X^\star B =C \mbox{ and } AX+(-Y)B =C, $ i.e., \begin{align}\label{eqxyy} A (X+Y^\star) - (X+Y^\star)^\star B=0. \end{align} Since $\sigma(A^\star-\lambda B)$ is reciprocal free, it follows that $\sigma(A^\star+\lambda B)$ is reciprocal free. Thus, we have $X=-Y^\star$ and $AX + X^\star B = C$ by the uniqueness of the solution of~\eqref{eqxyy}. \end{itemize} \end{proof} Theorem~\ref{deflatingthm} shows that if $A^\star-\lambda B$ is reciprocal free and $B$ is nonsingular, we can solve~\eqref{CTS} by the deflating subspace method. Also, the reciprocal free condition implies that either $A$ or $B$ is nonsingular. Thus, we might assume without loss of generality that $B$ is nonsingular. Otherwise, we can replace $\mathcal{Z}^\star$ of \eqref{d1} with $\mathcal{Z}$ of \eqref{d2} in Theorem~\ref{deflatingthm}. In the proof of Theorem~\ref{deflatingthm}, we know that if $B$ is nonsingular, then $T_2$ is invertible. We then are able to transform the formulae defined in~\eqref{eq:cond1} into the {palindromic} eigenvalue problem as follows. \begin{corollary}\label{CTSsol} Suppose that $\sigma(A^\star - \lambda B)$ is reciprocal free and the matrix $B$ is nonsingular. If there exists a full rank matrix $\left[\begin{array}{c}U \\V\end{array}\right]$ such that \begin{equation} \mathcal{Z}^\star \left[\begin{array}{c}U \\V\end{array}\right] = \mathcal{Z} \left[\begin{array}{c}U \\V\end{array}\right] T, \end{equation} for some matrix $T$ with $\sigma(T) = \sigma(A^\star-\lambda B)$, then $V$ is nonsingular and $X= UV^{-1}$ is the unique solution of the $\star$-Sylvester matrix equation~\eqref{CTS}. \end{corollary} From Corollary~\ref{CTSsol}, it can be seen that the solution of~\eqref{CTS} can be obtained by solving the palindromic eigenvalue problem~\eqref{eq:PEP}. We refer the reader to~\cite{Chu2008,Chu2010, Hilliges2004} for more details. It should also be noted that the assumption of the existence of a full rank matrix $\begin{bmatrix} U \\ V\end{bmatrix}$ is always true, since the eigenvalues of $\mathcal{Z}^\star -\lambda\mathcal{Z}$ are composed of $\sigma(A^\star - \lambda B) \cup \sigma(B^\star - \lambda A)$. \section{An efficient iterative method} In this section, we want to discuss how to apply the palindromic doubling algorithm (abbreviated as PDA) \cite{Li2011} to find the stabilizing solution of ~\eqref{CTS}. The stabilizing solution $X$ of algebraic Riccati equations has been an extremely active area of the design of feedback controller \cite{Lancaster1995}. Our interest in the stabilizing solution $X$ of ~\eqref{CTS} originates from the solution of the $\star$-Riccati matrix equation \begin{align}\label{starRic} XAX^\star+XB+CX^\star+D=0 \end{align} from an application related to the palindromic eigenvalue problem \cite{Hilliges2004,Chu2010}. Finding the solution of the $\star$-Riccati matrix equation is a difficult treatment. The application of Newton's method is a reasonable possibility and leads to the iterative process \begin{eqnarray}\label{NW} (C+X_kA)X_{k+1}^\star+X_{k+1}(B+AX_k^\star)=X_k A X_k^\star-D, \end{eqnarray} which is $\star$-Sylvester matrix equation with respect to $X_{k+1}$ for nonnegative integer $k$. To guarantee the convergence of the Newton's method \eqref{NW}, under some mild assumptions on coefficient matrices of $\star$-Riccati matrix equation~\eqref{NW}, one can choose the initial value $X_0$ such that the spectrum set of $\sigma(B+AX_k A-\lambda (C^\star+A^\star X_k^\star))$ lies in unit circle \cite{Chiang2012} for nonnegative integer $k$ so that the Newton's method will quadratically converge to the stabilizing solution. We focus on how efficiently solve the $\star$-Sylvester matrix equation in this paper. According to above discussion, the definition of stabilizing solution $X$ of ~\eqref{CTS} is stated as follows, \begin{Definition}\label{stab} If $\sigma(A^\star-\lambda B)$ lies in the unit circle in ~\eqref{CTS}, then the solution $X$ (if exist) of ~\eqref{CTS} is called a stabilizing solution. \end{Definition} For a matrix pencil $A^\star-\lambda B$, if $\sigma(A^\star-\lambda B)$ lies in the unit circle, it is clear that $\sigma(A^\star-\lambda B)$ is $\star$-reciprocal free. It follows that the stabilizing solution always exists and is unique. Our approach in the next subsection is to develop an efficient numerical algorithm for finding the stabilizing solution of \eqref{CTS}. To this end, we start with a review of the so-called palindromic doubling algorithm. This method has been applied to obtain the stabilizing solutions of the generalized continuous-time (and the generalized discrete-time) algebraic Riccati equations~\cite{Li2011}. \subsection{Palindromic Doubling Algorithm} Given a matrix pencil $\mathcal{A}-\lambda \mathcal{B}$ and assume $-1\not\in \sigma (\mathcal{A}-\lambda\mathcal{B})$, since $ \mathcal{B} (\mathcal{A}+ \mathcal{B})^{-1} (\mathcal{A}+ \mathcal{B}) = \mathcal{B} = (\mathcal{A}+ \mathcal{B}) (\mathcal{A}+ \mathcal{B})^{-1} \mathcal{B}$, it is easy to see that \begin{eqnarray}\label{swap} \mathcal{B}(\mathcal{A}+\mathcal{B})^{-1} \mathcal{A} &=& \mathcal{B} -\mathcal{B}(\mathcal{A}+\mathcal{B})^{-1} \mathcal{B} = \mathcal{A}(\mathcal{A}+\mathcal{B})^{-1} \mathcal{B}\label{ZZstar}. \end{eqnarray} We now consider the \emph{doubling transformation} $\mathcal{A}-\lambda \mathcal{B}\rightarrow \widehat{\mathcal{A}}-\lambda \widehat{\mathcal{B}}$ by \begin{subequations}\label{doublingtrans} \begin{align} \widehat{\mathcal{A}} = \mathcal{A}(\mathcal{A}+\mathcal{B})^{-1}\mathcal{A},\\ \widehat{\mathcal{B}} = \mathcal{B}(\mathcal{A}+\mathcal{B})^{-1}\mathcal{B}. \end{align} \end{subequations} The following theorem is to show that such transformation will keep the eigenspace unchanged and double the original eigenvalues. \begin{theorem}\label{doublingtransthm} The matrix pair $\widehat{\mathcal{A}}-\lambda \widehat{\mathcal{B}}$ has the doubling property, i.e., if \begin{align}\label{CAB} \mathcal{A} \begin{bmatrix} U\\V\end{bmatrix}=\mathcal{B} \begin{bmatrix} U\\V\end{bmatrix} {T}, \end{align} where $U,V\in\mathbb{C}^{2n\times n}$ and $ T\in\mathbb{C}^{n\times n}$, then \begin{align}\label{CAB2} \widehat{\mathcal{A}} \begin{bmatrix} U\\V\end{bmatrix}=\widehat{\mathcal{B}} \begin{bmatrix} U\\V\end{bmatrix} {T}^2. \end{align} \end{theorem} \begin{proof} Pre-multiplying the both sides of \eqref{CAB} by $\mathcal{A}(\mathcal{A}+\mathcal{B})^{-1}$ and applying \eqref{swap} and \eqref{doublingtrans}, we obtain \eqref{CAB2}, which completes the proof. \end{proof} To see how the doubling transformation can be applied to obtain the stabilizing solution of \eqref{CTS}, write the matrix $Z$ in~\eqref{eq:Zmat} as \begin{subequations}\label{PDA} \begin{align}\label{DA1} \mathcal{Z}=\mathcal{H}+\mathcal{K}, \end{align} where \begin{equation}\label{DA2} \mathcal{H}=\frac{1}{2}(\mathcal{Z}^\star+\mathcal{Z})=\mathcal{H}^\star,\quad \mathcal{K}=\frac{1}{2}(-\mathcal{Z}^\star+\mathcal{Z})=-\mathcal{K}^\star. \end{equation} \end{subequations} Note that matrices $\mathcal{H}$ and $\mathcal{K}$ are the Hermitian(symmetric) part and skew-Hermitian(skew-symmetric) part of $ \mathcal{Z}$ with $\star=H(\top)$, respectively. Since $\mathcal{Z}$ is an upper anti-triangular block matrix, we proceed to express $\mathcal{K}$ as \begin{align}\label{matK} \mathcal{K} =\begin{bmatrix} 0 & K_{12}\\ -(K_{12})^\star&-K_{22}\end{bmatrix}, \end{align} and define an initial matrix $\mathcal{H}_0$ for the PDA as \begin{align}\label{matH} \mathcal{H}_0=\begin{bmatrix} 0 & H_{12}^{(0)}\\ (H_{12}^{(0)})^\star& -H_{22}^{(0)}\end{bmatrix}, \end{align} where $H_{12}^{(0)}=\frac{A^\star+B}{2}$ and $H_{22}^{(0)}=\frac{C^\star+C}{2}$, $K_{12}=\frac{-A^\star+B}{2}$ and $K_{22}=\frac{-C^\star+C}{2}$. Based on the above notation, we then generalize the PDA technique given in~\cite{Li2011} for obtaining the stabilizing solution of~\eqref{CTS} as follows. \vspace*{0.3cm} \noindent{\textbf{Algorithm 1: the PDA}} \\ Given matrices $\mathcal{K}$ and $\mathcal{H}_0$ as defined by~\eqref{matK} and~\eqref{matH},\\ \textbf{for} $k=0,1,\ldots$, \textbf{compute until convergence} \\ \begin{subequations}\label{Ntmalg} \begin{eqnarray} T_k&=&(H_{12}^{(k)})^{-1} K_{12},\\ H_{12}^{(k+1)}&=& \frac{1}{2}(H_{12}^{(k)}+K_{12}T_k),\\ H_{22}^{(k+1)}&=&\frac{1}{2}(H_{22}^{(k)}+T_k^\star H_{22}^{(k)}T_k +K_{22}T_k-T_k^\star K_{22}). \end{eqnarray} \end{subequations} \textbf{end for}\\ \textbf{End of algorithm} \vspace*{0.3cm}\\ Of particular interest is that the stabilizing solution $X$ of \eqref{CTS} can be represented by \begin{equation*} X=\lim\limits_{k\rightarrow\infty} X_k, \end{equation*} where \begin{align}\label{eq:xk} X_k=(H_{12}^{(k)}+K_{12})^{-\star}(H_{22}^{(k)}+K_{22})^\star. \end{align} We prove this result in the following theorem. \begin{theorem}\label{PDA Solvent} Suppose that $\sigma(A-\lambda B^\star)$ lies in the unit circle and let $X$ be the stabilizing solution of~\eqref{CTS}. Then, all iterations given in Algorithm~1 are well-defined, and the sequences $\{Z_{12}^{(k)}, Z_{21}^{(k)}\}$ (See the below definition in (33)) and $\{X_k\}$ in~\eqref{eq:xk} satisfy \begin{eqnarray*} Z_{12}^{(k)} & \rightarrow& -A^\star+B, \mbox{ quadratically as } k\rightarrow\infty,\\ Z_{21}^{(k)} &\rightarrow& 0, \mbox{ quadratically as } k\rightarrow\infty,\\ X_k &\rightarrow& X, \mbox{quadratically as }k\rightarrow\infty, \end{eqnarray*} with convergence rate $\rho$ define by $\rho = \max\limits_{\tau\in\sigma(A-\lambda B^\star)} |\tau|<1$. \end{theorem} \begin{proof} Let $\{\mathcal{H}_k\}$ be a sequence given by \begin{align*} \mathcal{H}_k=\begin{bmatrix} 0 & H_{12}^{(k)}\\ (H_{12}^{(k)})^\star& -H_{22}^{(k)}\end{bmatrix}. \end{align*} Following from a direct computation of the inverse of $\mathcal{H}_k$ (if exist) and the fact that \begin{equation*} \mathcal{H}_k^{-1} = \begin{bmatrix} (H_{12}^{(k)})^{-\star} H_{22}^{(k)} (H_{12}^{(k)})^{-1} & (H_{12}^{(k)})^{-\star} \\ (H_{12}^{(k)})^{-1} & 0 \end{bmatrix}, \end{equation*} we obtain \begin{align}\label{HK} \mathcal{H}_{k+1}=\frac{1}{2}(\mathcal{H}_k+\mathcal{K}\mathcal{H}_k^{-1}\mathcal{K}). \end{align} Now consider a sequence of iterations given by \begin{align}\label{eq:Zk} \mathcal{Z}_k=\mathcal{H}_k+\mathcal{K}\equiv \begin{bmatrix} 0 & Z_{12}^{(k)}\\ Z_{21}^{(k)} & -Z_{22}^{(k)}\end{bmatrix}, \end{align} for $k=0,1,\cdots$. It follows from~\eqref{HK} that \begin{align*} \mathcal{Z}_{k+1}&=\mathcal{H}_{k+1}+\mathcal{K}=\frac{1}{2}(\mathcal{H}_k+\mathcal{K}+\mathcal{K}\mathcal{H}_k^{-1}\mathcal{K} +\mathcal{K})\\ &=\frac{1}{2}(\mathcal{H}_k+\mathcal{K})\mathcal{H}_k^{-1}(\mathcal{H}_k+\mathcal{K})=\mathcal{Z}_k(\mathcal{Z}_k+\mathcal{Z}_k^\star)^{-1}\mathcal{Z}_k, \end{align*} which is exactly the doubling transformation in \eqref{doublingtrans} with $\mathcal{A}\rightarrow\mathcal{Z}_k^\star$ and $\mathcal{B}\rightarrow\mathcal{Z}_k$. Furthermore, the anti-diagonal block matrix of $\mathcal{Z}_k$ satisfies \begin{align*} Z_{12}^{(k+1)}&=Z_{12}^{(k)}(Z_{12}^{(k)}+(Z_{21}^{(k)})^\star)^{-1}Z_{12}^{(k)},\\ Z_{21}^{(k+1)}&=Z_{21}^{(k)}(Z_{21}^{(k)}+(Z_{12}^{(k)})^\star)^{-1}Z_{21}^{(k)}, \end{align*} i.e., the matrix pencil $(Z_{21}^{(k+1)})^\star-\lambda Z_{12}^{(k+1)}$ is a doubling transformation of the matrix pencil $(Z_{21}^{(k)})^\star-\lambda Z_{12}^{(k)}$. Thus, $\sigma((Z_{21}^{(k)})^\star-\lambda Z_{12}^{(k)})$ lies in the unit circle for each $k$, since $(Z_{21}^{(0)})^\star-\lambda Z_{12}^{(0)}=A^\star-\lambda B$. We thus conclude that $Z_{12}^{(k)}$ is invertible for all $k$ by the regularity of $A^\star-\lambda B$. On the other hand, according to Theorem~\ref{deflatingthm}, the initial value $\mathcal{Z}_0=\mathcal{Z}$ gives rise to the fact that \begin{equation}\label{init} \mathcal{Z}_0^\star \left[\begin{array}{c} U \\ V\end{array}\right] = \mathcal{Z}_0 \left[\begin{array}{c}U \\V\end{array}\right] T, \end{equation} for some matrix $T$ with $\sigma(T) = \sigma(A^\star-\lambda B)$ and a nonsingular matrix $V$. This implies that the intersection of the spectrum set $\mathcal{Z}_0^\star-\lambda \mathcal{Z}_0$ and the unit circle is empty. It follows that $-1\not\in\sigma(\mathcal{Z}_k^\star-\lambda \mathcal{Z}_k)$ for each $k$ from Theorem~\ref{doublingtransthm}. Thus, $\mathcal{Z}_k^\star+\mathcal{Z}_k$ is invertible and hence $\mathcal{H}_k=\frac{1}{2}(\mathcal{Z}_k^\star+\mathcal{Z}_k)$ is also a nonsingular matrix for all $k$. That is, all iterations in Algorithm~1 are well defined. It follows from Theorem~\ref{doublingtransthm} that \begin{equation}\label{doubspac} \mathcal{Z}_k^\star \left[\begin{array}{c} U \\ V\end{array}\right] = \mathcal{Z}_k \left[\begin{array}{c}U \\V\end{array}\right] T^{2^k}, \end{equation} that is, \begin{subequations} \begin{align} ({Z}_{21}^{(k)})^{\star} V &={Z}_{12}^{(k)} V T^{2^k},\label{eq:sub1}\\ ({Z}_{12}^{(k)})^{\star} U-({Z}_{22}^{(k)})^{\star} V &= (({Z}_{21}^{(k)})^{\star} U-({Z}_{22}^{(k)})^{\star} V )T^{2^k}.\label{eq:sub2} \end{align} \end{subequations} By equalities~\eqref{eq:Zk} and~\eqref{eq:sub1}, we have \begin{align*} (H_{12}^{(k)}-K_{12}) V &=(H_{12}^{(k)}+K_{12}) V T^{2^k}, \end{align*} that is, \begin{align*} H_{12}^{(k)} - K_{12} =2 K_{12} V T^{2^k}(I-T^{2^k})^{-1} V^{-1}. \end{align*} This implies that $\lim\limits_{k\rightarrow\infty}H_{12}^{(k)}=K_{12} $. Thus we conclude that \begin{align*} \lim\limits_{k\rightarrow\infty}Z_{21}^{(k)}=\lim\limits_{k\rightarrow\infty}(H_{12}^{(k)}-K_{12}^{(0)})^\star=0,\,\, \lim\limits_{k\rightarrow\infty}Z_{12}^{(k)}=\lim\limits_{k\rightarrow\infty}(H_{12}^{(k)}+K_{12}^{(0)})=-A^\star+B. \end{align*} Note that $Z_{21}^{(k)}$ and $Z_{12}^{(k)}$ converge quadratically to $0$ and $-A^\star+B$ with rate $\rho(T)=\rho$, respectively. On the other hand, it follows from~\eqref{eq:sub2} and~\eqref{eq:Zk} that \begin{align*} (H_{12}^{(k)})^\star U-H_{22}^{(k)} V=-(K_{12}^\star U+K_{22}V)(I+T^{2^k})(I-T^{2^k})^{-1}. \end{align*} Thus, the right hand side of \eqref{eq:sub2} can be expressed as \begin{align}\label{eq:sub3} &({Z}_{21}^{(k)} U-{Z}_{22}^{(k)} V )T^{2^k}=((H_{12}^{(k)})^\star U-H_{22}^{(k)} V)T^{2^k}-(K_{12}^\star U+K_{22}V)T^{2^k}\nonumber\\ &=-(K_{12}^\star U+K_{22}V)(I+(I+T^{2^k})(I-T^{2^k})^{-1})T^{2^k}. \end{align} Let $X = UV^{-1}$. It follows from Theorem~\ref{doublingtransthm} that $X$ is a solution of~\eqref{doubspac}. Pre-multiplying and post-multiplying \eqref{eq:sub2} by $({Z}_{12}^{(k)})^{-\star}$ and $V^{-1}$, respectively, we have \begin{align*} X-X_k=({Z}_{12}^{(k)})^{-\star}({Z}_{21}^{(k)} U-{Z}_{22}^{(k)} V )T^{2^k}V^{-1}, \end{align*} where \begin{equation*} X_k= (H_{12}^{(k)}+K_{12})^{-\star}(H_{22}^{(k)}+K_{22})^\star = ({Z}_{12}^{(k)})^{-\star}({Z}_{22}^{(k)})^{\star}. \end{equation*} From~\eqref{eq:sub3} we have $\limsup\limits_{k\rightarrow\infty}\sqrt[2^k]{\| X-X_k\|}\leq \max\limits_{\tau\in\sigma(A-\lambda B^\star)} |\tau|=\rho$, which completes the proof. \end{proof} From Theorem~\ref{PDA Solvent}, it is natural to modify the definition $X_k$ in~\eqref{eq:xk} by solving the linear system \begin{align*} (-A+B^\star) X_k=Z_{22}^{(k)},\,\, k=1,2,\cdots \end{align*} Besides, the capacity of Algorithm~1 for solving~\eqref{CTS} with eigenvalues of $\sigma(A^\star -\lambda B)$ lying on the unit circle is something worthy of our discussion. Note that for the case $\star={H}$, there does not exist any unimodular eigenvalue lying on the unit circle under the condition of the uniquely solvable solution of \eqref{CTS}. However, from Theorem~\ref{solvable} we know that if $1$ is a simple eigenvalue of $\sigma(A^\top -\lambda B)$ and $\sigma(A^\top-\lambda B)\backslash \{1\}$ is reciprocal free, then the equation~\eqref{CTS} with $\star= \top$ is still solvable. In this critical case, we call the solution as the \emph{almost} stabilizing solution~\cite{Lancaster1995}. It is interesting to know that Algorithm~1 can also be applied to solve the critical case of~\eqref{CTS} with no difficulty. For the convergence analysis of this algorithm, we first consider the following analysis of the eigenstructure of the unimodular eigenvalues of $\mathcal{Z}^\top- \lambda \mathcal{Z}$ (i.e., $\lambda=1$). \begin{lemma}\label{Lem} Suppose that $1\in\sigma(A^\top-\lambda B)$ and $1$ is simple. Then $\mbox{nullity}(\mathcal{Z}^\top-\mathcal{Z})=2$. \end{lemma} \begin{proof} Considering two orthogonal matrices $Q_1$ and $Z_1$, let $Z_1 A Q_1^\top := \widehat{A}= [\hat{a}_{ij}]$ and $ Q_1 B Z_1^\top := \widehat{B} = [\hat{b}_{ij}]$ be the QZ or generalized Schur decomposition of $A^\top$ and $B$ so that $\widehat{A}^\top$ and $\widehat{B}$ are upper-triangular. Corresponding to this decomposition, the matrix pencil $Z^\top - Z$ can be expressed as \begin{equation} \mathcal{Z}^\top - \mathcal{Z} = \begin{bmatrix} Q_1^\top & 0\\ 0 &Z_1^\top \end{bmatrix} \left[\begin{array}{cc} 0 & \widehat{A}^\top-\widehat{B} \\ \widehat{B}^\top-\widehat{A} & \widehat{C}-\widehat{C}^\top \end{array}\right] \begin{bmatrix} Q_1 & 0\\ 0 & Z_1 \end{bmatrix} \end{equation} with $\widehat{C} := Z_1 C Z_1^\top$. Thus, in order to discuss the dimension of the null space of the matrix $\mathcal{Z}^\top - \mathcal{Z}$, we can assume without loss of generality that $A^\top$ and $B$ are upper triangular matrices and $a_{11}=b_{11}$. It follows that there exist a nonzero vector $\mathbf{x}_0$ and an unit vector $\mathbf{e}_1$ such that vectors $\mathbf{x}_0$ and $\mathbf{e}_1$ are in the null space of matrices $B^\top-A$ and $A^\top-B$, respectively. Since the first row of $B^\top-A$ is a zero row vector and the other rows of $B^\top-A$ are linearly independent, there exists a vector $\mathbf{x}_1$ satisfying $(B^\top-A)\mathbf{x}_1=(C^\top-C)\mathbf{e}_1$. Recall also from~\eqref{eq:PEP} that \begin{align*} \mathcal{Z}^\top-\mathcal{Z}=\begin{bmatrix} 0 & A^\top-B\\B^\top-A & C-C^\top\end{bmatrix}. \end{align*} We see that $\mathbf{v}_1=\begin{bmatrix} \mathbf{x}_1 \\ \mathbf{e}_1\end{bmatrix}$ and $\mathbf{v}_2=\begin{bmatrix} \mathbf{x}_0\\ \mathbf{0}\end{bmatrix}$ are two linearly independent vectors in the null space of $\mathcal{Z}^\top-\mathcal{Z}$, that is, $\mbox{nullity}(\mathcal{Z}^\top-\mathcal{Z})=2$. \end{proof} Lemma~\ref{Lem} tells that the eigenvalue $``1''$ of $\mathcal{Z}^\top-\lambda\mathcal{Z}$ has partial multiplicity one (two $1\times 1$ Jordan blocks) if the eigenvalue $``1''$ of $A^\top-\lambda B$ is simple. Based on Lemma~\ref{Lem}, it can be shown that Algorithm~1 can be applied to solve this problem with a linear rate of convergence. Because the analysis of the convergence property of Algorithm 1 is an analogous result for the palindromic generalized eigenvalue problem in~\cite{Li2011}[Theorem 3.1], we omit it here. \begin{theorem}\label{DAconvthm} In the critical case, all sequences generated in Algorithm~1 for finding the almost stabilizing solution $X$ of~\eqref{CTS} are well defined. Moreover, \begin{eqnarray*} Z_{12}^{(k)}&\rightarrow& -A^T+B,\,\mbox{linearly}\,\, \mbox{as } k\rightarrow\infty, \\ Z_{21}^{(k)} &\rightarrow& 0,\,\mbox{linearly}\,\, \mbox{as } k\rightarrow\infty, \\ X_k &\rightarrow& X,\,\mbox{linearly}\,\, \mbox{as } k\rightarrow\infty, \end{eqnarray*} with convergence rate at least $1/2$. \end{theorem} \subsection{A numerical example} In this subsection, we use a numerical example to illustrate the efficiency of Algorithm~1 with the assumption of Theorem~\ref{PDA Solvent} and to demonstrate the numerical behavior of Algorithm~1 in the critical case. All computations were performed in MATLAB/version 2011 on a PC with an Intel Core i7-4770 3.40 GHZ processor and 32 GB main memory. For evaluating the performance, we define the error (ERR), relative error (RERR) and relative normalized residual (RES) as follows \begin{eqnarray*} \mbox{ERR} &\equiv& \| Z_{21}^{(k)}\|_F,\,\,\mbox{RERR}\equiv \frac{\| X_k-X\|_F}{\|X\|_F}, \\ \mbox{RES}&\equiv& \frac{\| AX_k+X_k^\star B-C\|_F}{\|A\|_F\|X_k\|_F+\|B\|_F\|X_k\|_F+\|C\|_F}, \end{eqnarray*} respectively, where $X$ is the (almost) stabilizing solution of \eqref{CTS}. All iterations are terminated whenever the errors or relative errors or the relative normalized residual residuals are less than $n^2\mathbf{u}$, where $\mathbf{u}=2^{-52}\cong 2.22e-16$ is the machine zero. \begin{example}~\label{ex1} Let $\widehat{A}^\top, \widehat{B} \in \mathbb{R}^{n \times n}$ be two real lower-triangular matrices with given diagonal elements (specified by $a,b \in \mathbb{R}^n$) and random strictly lower-triangular elements. They are then reshuffled by the orthogonal matrices $Q,Z \in \mathbb{R}^{n \times n}$ to form $(A,B) = (Q \widehat{A} Z, Q \widehat{B} Z)$, that is, in \texttt{MATLAB} commands, we define \begin{eqnarray*} \widehat{A} &=&triu(randn(n),-1)+diag(a), \,\, \widehat{B}=tril(randn(n),-1)+diag(b), \\ C&=&AX+X^\top B, \end{eqnarray*} where \[a=[temp_1.*temp_2,1-\epsilon] \mbox{ and } b=[temp_1;1]\] with \begin{align*} temp_1=rand(n-1,1),\,\,temp_2=rand(n-1,1). \end{align*} where $X=randn(n)$ is given as the (almost) stabilizing solution of \eqref{CTS} and $0\leq \epsilon <1$. The setup of $a$ and $b$ guarantee that $\sigma(A^T-\lambda B)$ lies within the unit circle if $\epsilon\neq 0$, and ``1'' will be a simple eigenvalue of $\sigma(A^T-\lambda B)$ if $\epsilon=0$. Let $n=10$, Table~\ref{table1} contains the iteration numbers (refer as ``ITs''), ERR, RERR and RES for various numbers of $\epsilon$. We see that Algorithm~1 quadratically converges to the stabilizing solution $X$ if $\epsilon=1e-1,1e-2$, and all measurements for errors almost have the same accuracy for every tests. Since the limit of the sequence $\{Z_{12}^k\}$ is $-A^\top + B$, the condition of $\{Z_{12}^k\}$ becomes ill-conditioning as $\epsilon$ approaches zero. Note that the sequence $\{X_k\}$ generated by Algorithm~1 converges well to the stabilizing solution before the sequence $\{Z_{12}^k\}$ tends to a singular matrix. However, the number of iterations required in the computation dramatically increases, once $\epsilon$ approaches zero and finally loses the property of quadratic convergence. Also, it can be observed that forward errors of the almost stable eigenvalues are approximately equal to $\sqrt{\mathbf{u}}$. Similar phenomena can also be seen in~\cite{Li2011} for the study of the palindromic generalized eigenvalue problem. \begin{table}[htbn] \begin{center} \begin{tabular}[c]{c|cccc} & & \\[-2.4ex] $\epsilon$ & \multicolumn{1}{c}{ITs} & \multicolumn{1}{c}{ERR} & \multicolumn{1}{c}{RERR} & \multicolumn{1}{c}{RES} \\ \hline & & & \\[-2.4ex] 1e-1 & 5.00e+00 & 9.2527e-17&3.9015e-18 & 8.1211e-16 \\ 1e-2 & 7.00e+00 & 9.8965e-15& 5.8186e-15 & 8.0061e-16 \\ 1e-4 & 1.60e+01 & 1.7175e-13& 1.1288e-13 & 7.2224e-14 \\ 1e-8 & 2.40e+01 & 2.8457e-10& 1.3213e-10 & 1.0476e-10 \\ 0 & 3.80e+01 & 6.6106e-8& 3.9109e-8 & 3.4419e-8 \\ \end{tabular} \caption{Results for Example~\ref{ex1}} \label{table1} \end{center} \end{table} \end{example} \section{Conclusion} One common procedure to solve the Sylvester equations is by means of the Schur decomposition. In this paper, we present the invariant subspace method and, more generally, the deflating subspace method to solve the Sylvester equations. Our methods are based on the analysis of the eigeninformation of two square matrices defined in Section 3 and 4. We carry out a thorough discussion to address the various eigeninformation encountered in the subspace methods. These ideas can then be implemented into a doubling algorithm for solving the (almost) stabilizing solution of \eqref{CTS} in Section~5. On the other hand, it follows from Theorem~\ref{solvable} that~\eqref{CTS} still has an unique solution, even if some eigenvalues of $A-\lambda B^\star$ lie beyond the unit circle. How to apply Algorithm~1 for solving this problem is under investigation and will be reported elsewhere. \section*{Acknowledgement} The authors wish to thank editor and two anonymous referees for many interesting and valuable suggestions on the manuscript.
\section{Conclusion}\label{conclusion} With the motivation of exploring high-Reynolds-number rotating turbulent flows, we developed a highly-efficient parallel DNS code for Taylor-Couette flows. The incompressible Navier-Stokes equations in cylindrical coordinates are solved in primitive variables by using a projection method proposed by Hugues and Randriamampianina~\cite{Hugues_ijnmf1998}, which is second-order accurate in both pressure and velocity. This method leads at each time step to the solution of five linear differential equations, either of Poisson or of Helmholtz type, which simplifies significantly the programming of the code. For the spatial discretization, we used a combination of Fourier spectral in axial and azimuthal directions and high-order finite differences in the radial direction, which allow the use of tailored stretched grids. The computing cost scales linearly with the number of grid points in each direction. In order to reach higher Reynolds numbers and to take full advantage of the modern HPC facilities, the code was parallelized by a hybrid MPI-OpenMP strategy, combining the simplicity of a MPI-based one-dimensional ``slab'' domain decomposition in Fourier space with efficient exploitation of the remaining coarse-grained parallelism by OpenMP threading. Compared to a flat MPI-parallelization, the hybrid code maps more naturally to the current multi-node, multi-core architectures, keeps the number of MPI tasks in the well-manageable regime of a few thousand, and, most importantly, reduces inter-node communications, which improves the overall efficiency and scalability. The strong scaling study which was performed with scientifically relevant setups demonstrates the scalability of the code up to more than 20\,000 processor cores. This allows to perform simulations with much higher resolutions than previously possible. With the current HPC technology, this code pushes the achievable $Re$ to the order of magnitude of $\mathcal{O}(10^5)$ in DNS of Taylor-Couette flow, which therefore opens up the possibility to study quasi-Keplerian flows at experimentally relevant parameters. The new code was shown to be very accurate in various regimes: laminar Couette flow, wavy vortices, transitional and turbulent flow at high Reynolds number. With the high efficiency of the hybrid parallel scheme, this code possesses great potential to explore the turbulent TCf in a much broader parameter space. \section*{Acknowledgments} We thank Florian Merz (IBM) for optimizing the global transposition routine. L. Shi and B. Hof acknowledge research funding by Deutsche Forschungsgemeinschaft (DFG) under Grant No. SFB963/1 (project A8). Computations were performed on the HPC system "Hydra" of the Max-Planck-Society at RZG. \bibliographystyle{elsarticle-num-names} \section{Introduction}\label{sec1} Rotating fluid flows with radially increasing angular momentum are known to be linearly stable because of the inviscid Rayleigh criterion~\cite{Rayleigh_prsla1917}. A particularly important application is astrophysical Keplerian flow, with angular velocity profile decreasing radially as $\Omega \propto r^{-3/2}$. Whether Keplerian flows become turbulent because of nonlinear instabilities or remain laminar even at extreme Reynolds numbers, has great implications for accretion processes in weakly-ionized astrophysical disks~\cite{Balbus_nature2011}. This question has been recently investigated with experiments of fluid flows between rotating cylinders (Taylor--Couette flow, TCf), which can in principle approximate Keplerian profiles. Experiments conducted by Ji and co-workers~\cite{JiGoodman_nature2006, SchartmanGoodman_aa2012,edlund2014} found no hydrodynamic turbulence in quasi-Keplerian TCf in the range $Re\sim\mathcal{O}(10^5-10^6)$, whereas similar studies~\cite{PaolettiLathrop_prl2011,PaolettiLathrop_aa2012} report strongly turbulent flows that could account for the observed accretion rates in astrophysical disks. This discrepancy may arise from the axial boundary conditions: numerical simulations of the experimental setups show that top and bottom endwalls confining the fluid strongly disrupt Keplerian velocity profiles and causes turbulence to arise already at $Re\sim\mathcal{O}(10^3)$ \cite{Avila_prl2012}. Hence, the interpretation and extrapolation of experimental data remains controversial because of the prominent role played by axial endwalls. Numerical simulations with axially periodic boundaries resolve this problem and allow to directly probe the stability of Keplerian flows. Very recently, Ostilla-M\'onico \emph{et al.} \cite{ostilla2014} have carried out such simulations at $Re=8.1\times 10^4$ and observed that the turbulence decays to laminar flow. However, the role of initial conditions and numerical dissipation in the decay process are unknown and require more detailed studies. Probing the stability of Keplerian flows requires achieving yet larger Reynolds numbers $Re\sim\mathcal{O}(10^5)$, while still resolving the dissipation scales. We note that as the key question is concerned about the existence of turbulence, modeling strategies such as Reynolds-averaged equations (RANS) and Large-Eddy Simulation (LES) are precluded and one has to resort to direct numerical simulation (DNS) of the Navier--Stokes equations (see~\cite{Pope_2000} for details about these simulation techniques). Starting with the study of homogeneous isotropic turbulence conducted by Orszag and Patterson~\cite{OrszagPatterson_prl1972}, DNS has been proven as a very powerful approach to explore the physics of turbulent flows (see Ref.~\cite{MoinMahesh_arfm1998,Jimenez_jot2003}). It has been widely used in fundamental research on both transitional and fully-developed turbulence in boundary layers over a flat plate (e.g. \cite{Spalart_jfm1988,Schlatter_jfm2010}), channel (e.g. \cite{KimMoinMoser_jfm1987,HoyasJimenez_pof2006}), pipe (e.g. \cite{eggels1994,WuAdrian_jfm2012}) and Taylor--Couette flows (e.g. \cite{CoMa96,dong2007,BrauckmannEckhardt_jfm2012}). Distinguished from RANS and LES, a carefully performed DNS resolves all temporal and spatial scales relevant to turbulence and thus provides data of high fidelity. Its advantage is also its main drawback: resolving the physics of turbulence implies a scaling of the computational complexity as $\mathcal{O}(Re^3)$~\cite{Pope_2000}. In this paper we develop a highly efficient DNS code for TCf with axially periodic boundary conditions using a hybrid two-level parallelization strategy. It enables DNS to be performed up to $Re\sim\mathcal{O}(10^5)$, and thus provides access to a broad range in the parameter space of TCf, including quasi-Keplerian flows at experimentally relevant Reynolds numbers. Generally, finite differences \cite{ostilla2014} or spectral-element methods \cite{dong2007} can be used to perform DNS of Taylor--Couette flow at large Reynolds numbers. However, the most efficient and accurate method for discretizing partial differential equations with periodic boundary conditions is the spectral Fourier--Galerkin method, so we use this in the axial and azimuthal directions. Many authors have also used the spectral Galerkin-method in the non-periodic radial direction by employing Chebyshev, Legendre or Jacobi polynomials ~\cite{Moser_jcp1983,marcus1984,canuto2007}. The two latter render however a computational complexity of $\mathcal{O}(M^2)$, where $M$ is the degree of the approximation, due to the lack of fast transformations between physical and spectral spaces. This makes computations too expensive at large Reynolds numbers. In contrast, with the Chebyshev method the fast cosine transform allows it to keep the cost at $\mathcal{O}(M\log(M))$. However, in order to use accurate quadratures the projection basis must be different from the basis used to discretize the Navier--Stokes equations (Petrov-Galerkin method) ~\cite{Moser_jcp1983,Meseguer_EPJ2007}. On the other hand, if the spectral method is used directly at a collocation grid (in physical space) the resulting differentiation matrices are dense. Hence the solution of the Poisson equations, for example with the diagonalization method \cite{marcus1984}, requires $\mathcal{O}(M^2)$ operations. A common drawback of all the aforementioned spectral methods is that the density of collocation nodes towards the boundaries scales as $\mathcal{O}(M^2)$. Although this allows to properly resolve boundary layers with relatively low resolutions, at large Reynolds numbers the clustering is excessive and the required resolution is often given by the spacing of nodes far from the boundaries. Moreover, this clustering poses a severe restriction on the time step of $\Delta t=\mathcal{O}(M^{-2})$ because of the CFL condition. Although transformations of the node distribution have been proposed~\cite{KasloffTalEzer_jcp1993}, these result in the loss of the spectral convergence. Finally, it becomes impractical to use the Chebyshev method for large resolutions $M\gtrsim 600$, as needed in the simulation of turbulence at large $Re$. For these reasons we use the high-order finite-difference (FD) method in the radial direction, which makes the stretching of grid nodes straightforward. The incompressible Navier--Stokes equations in primitive variables are integrated in time with a second-order $\mathcal{O}(\Delta t^2)$ time-splitting scheme proposed by Hugues \& Randriamampianina \cite{Hugues_ijnmf1998}, who tested it in two dimensions in combination with a Chebyshev-Chebyshev discretization. The scheme is semi-implicit and is second-order accurate also for the pressure, rendering a very small $\mathcal{O}(\Delta t^3)$ slip-velocity error at the boundary while fulfilling the incompressibility constraint. It is straightforward to implement: it avoids staggered grids and requires the solution of five equations of Poisson or Helmholtz type. Raspo \emph{et al.}~\cite{RaspoBontoux_ComptFluids2002}, and later Avila \emph{et al.}~\cite{avila2008}, subsequently extended the scheme to three-dimensional Taylor--Couette flow with no-slip axial boundary conditions, and the Fourier--Galerkin method in the azimuthal direction. These codes have been extensively used for the simulations of centrifugal \cite{czarny2003,czarny2004} and endwall-driven instabilities in Taylor--Couette flow \cite{avila2008,Avila_prl2012}. Mercader \emph{et al.} \cite{mercader2010} have also extended the scheme to convection in a cylindrical container. Here we combine the scheme of Hugues \& Randriamampianina \cite{Hugues_ijnmf1998} with the finite-difference method in the radial direction and the Fourier--Galerkin method in the azimuthal and axial periodic directions. The nonlinear advective term is computed in physical space with the pseudospectral method. The code is parallelized here by combining the Message Passing Interface (MPI) and the Open Multiprocessing (OpenMP) paradigms. The Fourier-Galerkin method leads to mode-decoupled linear equations, which makes the one-dimensional MPI parallelization rather straightforward to implement. OpenMP threading within MPI tasks allows to efficiently use modern high performance computing (HPC) architectures and mitigates the overhead induced by MPI All-to-all inter-task communications which are typical of spectral methods. The paper is structured as follows. In \S\ref{sec2}, we formulate the Taylor--Couette problem and then present the numerical method in \S\ref{sec3}. In \S\ref{sec:parallel_implementation} we describe the parallelization strategy employed in the code and its implementation. The accuracy and performance of the code are discussed in \S\ref{sec:validation} and \S\ref{sec:benchmarks}, respectively, before the conclusions in \S\ref{conclusion}. \section{Governing equations and geometry}\label{sec2} \begin{figure}[!h] \centering \includegraphics[width=0.4\textwidth]{tCf_geometry.eps} \caption{Schematic of the Taylor-Couette system in cylindrical coordinates. The inner and outer cylinder rotate independently with speeds $\Omega_i$ and $\Omega_o$, respectively. No-slip boundary conditions at the cylinder are used together with axially periodic boundary conditions. The fluid between the cylinders (hatched region) moves by the shear force due to the fluid viscosity.} \label{fig:TCgeometry} \end{figure} We solve the equations governing the motion of an incompressible fluid of kinematic viscosity $\nu$ and constant density $\rho$ \begin{equation} \partial_{t}\textbf{u}+\textbf{u}\cdot\nabla\textbf{u} =-\frac{1}{\rho}\nabla p^h+\nu\Delta\textbf{u},\quad \nabla\cdot\textbf{u}=0, \label{eq:NS} \end{equation} where $\textbf{u}(\textbf{r},t)$ is the velocity field and $p^h(\textbf{r},t)$ is the hydrodynamic pressure. Here cylindrical coordinates $\textbf{r}=(r,\theta,z)$ are used. The geometry of the system is shown in Fig.~\ref{fig:TCgeometry} and consists of fluid confined between two concentric cylinders. The inner (outer) cylinder has radius $r_{i}$ $(r_o)$ and rotates at a speed of $\Omega_i$ $(\Omega_o)$. The Reynolds number in the inner and outer cylinder is defined as $Re_{i,o}=\Omega_{i,o}r_{i,o}d/\nu$, where $d=r_o-r_i$ is the gap between the cylinders. The geometry is fully specified by two dimensionless parameters: the radii-ratio $\eta=r_i/r_o$ and the length-to-gap aspect-ratio $\Gamma=L_z/d$, where $L_z$ is the axial length of the cylinders. At the cylinders no-slip boundary conditions are applied, whereas in the axial direction periodic boundary conditions are imposed to avoid endwall effects. This approximates the case of very long cylinders. In the azimuthal direction periodic boundary conditions occur naturally. However, it is often computationally convenient to simulate only an angular section $L_\theta\le 2\pi$ of the cylinders, and periodic boundary conditions are then used for $\theta \in [0,L_\theta]$. This is justified provided that the correlation length of the turbulent flow in the azimuthal direction is shorter than $r_iL_\theta$ \cite{BrauckmannEckhardt_jfm2012}. Henceforth, all variables will be rendered dimensionless using $d$, $\tau=d^2/\nu$, and $\nu^2/d^2$ as units for space, time, and the reduced pressure $p=p^h/\rho$, respectively. The Navier-Stokes equations \eqref{eq:NS} for this scaling become \begin{equation} \begin{aligned} \partial_{t}\textbf{u}+\textbf{u}\cdot\nabla\textbf{u} &=-\nabla p+\Delta\textbf{u}\\ \nabla\cdot\textbf{u}&=0. \end{aligned} \label{eq:NS_dimless} \end{equation} In cylindrical coordinates the equations read \begin{equation} \begin{aligned} (\partial_t+\textbf{u}\cdot\nabla)u_r-u_{\theta}^2/r &= -\partial_rp+\Delta u_r-u_r/r^2-2\partial_{\theta}u_{\theta}/r^2\\ (\partial_t+\textbf{u}\cdot\nabla)u_{\theta}+u_{\theta}u_r/r &= -\partial_{\theta}p/r+\Delta u_{\theta}-u_{\theta}/r^2 +2\partial_{\theta}u_r/r^2\\ (\partial_t+\textbf{u}\cdot\nabla)u_z &= -\partial_zp+\Delta u_z,\\ u_r/r+\partial_ru_r+\partial_{\theta}u_{\theta}/r+\partial_zu_z &=0.\\ \end{aligned} \label{eq:NS_cyl} \end{equation} with $\nabla = (\partial_r, \partial_{\theta}/r,\partial_z)$ and $\Delta = \partial_r/r+\partial_{rr}^2+\partial_{\theta\theta}^2/r^2+\partial_{zz}^2$. Note that the Reynolds numbers enter the system through the boundary conditions \begin{equation} \begin{aligned} u_{\theta}(r_{i,o},\theta,z)&=Re_{i,o},\\ u_{r,z}(r_{i,o},\theta,z)&=0,\\ \textbf{u}(r,\theta,z)&=\textbf{u}(r,\theta + L_\theta,z),\\ \textbf{u}(r,\theta,z)&=\textbf{u}(r,\theta,z+\Gamma). \end{aligned} \label{eq:nsBC} \end{equation} By taking the divergence of the first equation and then applying the incompressibility condition, we obtain a Poisson equation for the pressure, \begin{equation} \Delta p = -\nabla\cdot\textbf{N}(\textbf{u}),\quad \text{where} \quad \textbf{N}(\textbf{u}) = \textbf{u}\cdot\nabla\textbf{u}, \label{eq:pp} \end{equation} with consistent boundary conditions~\cite{GreshoSani_inmf1987} \begin{equation} \partial_np|_{r=r_{i,o}} = \textbf{n}\cdot [-\partial_t\textbf{u}-\textbf{N}(\textbf{u})+\Delta\textbf{u}]. \label{eq:ppBC} \end{equation} As explained in \S\ref{sec3:numericalMethod}, this equation will be solved for the pressure prediction. \section{Numerical Formulation}\label{sec3} The governing equations~\eqref{eq:NS_cyl} are solved for the primitive variables ($\textbf{u},p$). We discretize the equations with a combination of the Fourier--Galerkin method with the finite-difference method (FD) in space, whereas time is advanced with the semi-implicit fractional-step method of Hugues and Randriamampianina~\cite{Hugues_ijnmf1998}, who employ second-order-accurate backward differences with linear (second-order) extrapolation for the nonlinear term. The pseudospectral technique with 3/2-dealiasing is applied to compute the nonlinear term $\textbf{N}(\textbf{u})$ in physical space~\cite{OrszagPatterson_prl1972}. \subsection{Spatial discretization} In the periodic axial and azimuthal directions, the velocity field and pressure are approximated as \begin{equation} \begin{aligned} \textbf{u}(r,\theta,z)=\sum_{l=-L}^{L}\sum_{n=-N}^{N}\hat{\textbf{u}}^{ln}(r) e^{i(lk_z z+nk_{\theta}\theta)},\\ p(r,\theta,z)=\sum_{l=-L}^{L}\sum_{n=-N}^{N}\hat{p}^{ln}(r) e^{i(lk_zz+nk_{\theta}\theta)}, \end{aligned} \label{eq:fourier} \end{equation} where $k_z$ is the minimum (fundamental) axial wavenumber and fixes the axial non-dimensional length $\Gamma=2\pi/k_z$ of the computational domain. Similarly, $L_\theta=2\pi/k_\theta$ is the azimuthal arc degree; $k_{\theta}=1$ corresponds to the natural periodic boundary condition in the azimuthal direction, whereas $k_\theta=4$ corresponds to one quarter of an annulus. The hat symbol $\hat{}$ in \eqref{eq:fourier} denotes quantities in Fourier space and the tuple $(L,N)$ determines the spectral numerical resolution. By substituting~\eqref{eq:fourier} into~\eqref{eq:NS_cyl} and projecting the result onto a basis $e^{-i(lk_z z+nk_{\theta} \theta)}$ ($l = -L,\dots,L; n=-N,\dots,N$), we obtain the mode-decoupled Navier-Stokes equations. For each Fourier mode $(l,n)$, they read \begin{equation} \begin{aligned} \partial_t\hat{u}_r+\hat{N}_r &= -\partial_r\hat{p}+\hat{\Delta}\hat{u}_r- \hat{u}_r/r^2-2ink_{\theta}\hat{u}_{\theta}/r^2,\\ \partial_t\hat{u}_{\theta}+\hat{N}_{\theta} &= -ink_{\theta}\hat{p}/r+ \hat{\Delta}\hat{u}_{\theta}-\hat{u}_{\theta}/r^2-2ink_{\theta}\hat{u}_r/r^2,\\ \partial_t\hat{u}_z+\hat{N}_z &= -ilk_z\hat{p}+\hat{\Delta}\hat{u}_z.\\ \end{aligned} \label{eq:modalNS} \end{equation} Here $\hat{\Delta}=\partial_r/r+\partial_{rr}-n^2k_{\theta}^2/r^2-l^2k_z^2$, and the superscripts $(l,n)$ have been omitted for clarity. Note that the nonlinear term couples Fourier modes and it is thus computed in physical space with the pseudospectral method. Details of the implementation and parallelization of the nonlinear term are given in \S\ref{sec:parallel_implementation}. Equations \eqref{eq:modalNS} couple the radial and azimuthal velocities. By applying the following change of variables \cite{orszag1983} \begin{equation*} \begin{aligned} \hat{u}_+&=\hat{u}_r+i\hat{u}_{\theta},\\ \hat{u}_-&=\hat{u}_r-i\hat{u}_{\theta}, \end{aligned} \end{equation*} to equation~\eqref{eq:modalNS}, we obtain the decoupled equations \begin{equation} \begin{aligned} \partial_t\hat{u}_+(r)+\hat{N}_+(r) &= -\partial_r\hat{p}(r)+nk_{\theta}\hat{p}(r)/r +(\hat{\Delta}-1/r^2-2nk_{\theta}/r^2)\hat{u}_+,\\ \partial_t\hat{u}_-(r)+\hat{N}_-(r) &= -\partial_r\hat{p}(r)-nk_{\theta}\hat{p}(r)/r +(\hat{\Delta}-1/r^2+2nk_{\theta}/r^2)\hat{u}_-,\\ \partial_t\hat{u}_z(r)+\hat{N}_z(r) &= -ilk_z\hat{p}(r)+\hat{\Delta}\hat{u}_z, \end{aligned} \label{eq:modalNSdecoup} \end{equation} where $\hat{N}_\pm=\hat{N}_r\pm i\hat{N}_{\theta}$. We use a standard high-order, central finite-difference method to approximate the radial derivatives in equations~\eqref{eq:modalNSdecoup} (see Ref.~\cite{Fornberg_cambridge}). The radial nodes are distributed as ~\cite{KasloffTalEzer_jcp1993} \begin{equation} r_j= \dfrac{1+\eta}{2(1-\eta)}+\dfrac{\sin^{-1}(-\alpha \cos(\pi j/M))}{2\, \sin^{-1}\alpha},\qquad j=0,\dots,M. \label{eq:nodeDist} \end{equation} For $\alpha=1$ the grid is uniform, whereas for $\alpha\rightarrow 0$ the Chebyshev collocation points are obtained. Here stencils of $n_s=9$ points, corresponding to a scheme of formally order 8 was found to give the best compromise in our tests. Note that we reduce the stencil length gradually towards the boundaries in order to keep the FD-matrices banded. We show in \S\ref{sec:validation} that due to the clustering of nodes near the walls with typical values of $\alpha=0.5$ this reduction of the order of accuracy does not produce a larger error at the boundaries. With $(L,N)$ Fourier modes and $M$ radial nodes, the number of grid points in physical space is $(n_r,n_{\theta},n_z)=(M,2N+1,2L+1)$ in the radial, azimuthal and axial directions, respectively. Note that we dealiase the nonlinear term by computing it on a grid of $(M,3N+1,3L+1)$ points. \subsection{Temporal scheme}\label{sec3:numericalMethod} A stiffly stable temporal scheme based on a backward differentiation formula with extrapolation for the nonlinear term is adopted (see Ref.~\cite{Hugues_ijnmf1998,KarniadakisOrszag_jcp1991}). It reads \begin{equation} \frac{3\textbf{u}^{i+1}-4\textbf{u}^{i}+\textbf{u}^{i-1}}{2\Delta t} + 2\textbf{N}^i(\textbf{u})-\textbf{N}^{i-1}(\textbf{u}) = -\nabla p^{i+1} + \Delta{\textbf{u}^{i+1}}. \label{eq:timeScheme} \end{equation} In the literature this is often referred to as Adams-Bashforth backward-difference method of second order (AB2BD2). The viscous terms are discretized implicitly, whereas the nonlinear terms are treated explicitly. At each time step, equation~\eqref{eq:timeScheme} is solved through a fractional step method proposed by Hugues and Randriamampianina~\cite{Hugues_ijnmf1998}. The method is summarized below. Here $(\hat{\textbf{u}}^i,\hat{p}^i)$ denote the spectral coefficients at the $i^{th}$ time step. \begin{enumerate}[\itshape 1\upshape)] \item Obtain spectral coefficients of the nonlinear term, $\hat{\textbf{N}}^{i}(\textbf{u})$, using the 3/2-dealiasing rule \begin{itemize} \item Do matrix-vector multiplication to calculate $\partial_r\hat{\textbf{u}}^i$ (FD method) \item Compute dot product in Fourier space to calculate $\partial_{\theta}\hat{\textbf{u}}^i$ and $\partial_z\hat{\textbf{u}}^i$ \item Perform Fourier transform of $\partial_{r,\theta,z}\hat{\textbf{u}}^i$ and $\hat{\textbf{u}}^i$ to obtain the velocity field and all its derivatives in physical space; \item Calculate $\textbf{N}^{i}(\textbf{u})=\textbf{u}^i\cdot\nabla\textbf{u}^i$; \item Perform inverse Fourier transform to obtain the spectral coefficients $\hat{\textbf{N}}^{i}(\textbf{u})$. \end{itemize} \item Obtain the pressure prediction, $\hat{p}^*$: solve the Poisson equation \begin{align} \Delta \hat{p}^* = \nabla\cdot[-2\hat{\textbf{N}}^i(\textbf{u})+\hat{\textbf{N}}^{i-1}(\textbf{u})], \label{eq:pressure} \end{align} with consistent Neumann boundary conditions~\eqref{eq:ppBC}. \item Obtain the velocity prediction, $\hat{\textbf{u}}^*$: solve the three Helmholtz equations \begin{align} \frac{3\hat{\textbf{u}}^*-4\hat{\textbf{u}}^{i}+\hat{\textbf{u}}^{i-1}}{2\Delta t} + 2\hat{\textbf{N}}^i(\textbf{u})-\hat{\textbf{N}}^{i-1}(\textbf{u}) = -\nabla \hat{p}^* + \Delta\hat{\textbf{u}}^* \label{eq:vel} \end{align} with Dirichlet boundary conditions~\eqref{eq:nsBC}. \item Correct via an intermediate variable $\phi= 2\Delta t(\hat{p}^{i+1}-\hat{p}^*)/3$. The incompressibility condition $\nabla\cdot \hat{\textbf{u}}^{i+1}=0$ leads to a Poisson equation for $\phi$ with homogeneous Neumann boundary conditions (see Ref.~\cite{GreshoSani_inmf1987,Hugues_ijnmf1998}) \begin{equation} \begin{aligned} &\Delta \phi = \nabla\cdot\hat{\textbf{u}}^*,\\ &\partial_r\phi|_{r=r_{i,o}} =0 \label{eq:correction} \end{aligned} \end{equation} \item Compute pressure and velocity correction, $\hat{p}^{i+1}$ and $\hat{\textbf{u}}^{i+1}$: \begin{equation} \begin{aligned} \hat{p}^{i+1}&=\hat{p}^*+3\phi/(2\Delta t)\\ \hat{\textbf{u}}^{i+1}&=\hat{\textbf{u}}^*-\nabla\phi \end{aligned} \label{eq:update} \end{equation} \item Go back to step 1 \end{enumerate} The Navier-Stokes equations are thus advanced in time by solving five systems of linear equations \eqref{eq:pressure}-\eqref{eq:correction}, of Poisson or Helmholtz type, for each Fourier mode. This method accounts for a divergence-free velocity field and a small slip at the wall of the order of $\mathcal{O}(\Delta t^3)$ in the tangential velocities, $u_z$ and $u_\theta$. We note that the method was originally developed and tested \cite{Hugues_ijnmf1998} for the two-dimensional Navier-Stokes equation discretized on a Chebyshev-Chebyshev collocation grid, and the Poisson and Helmholtz equations were solved using the double diagonalization method, thus rendering quadratic computational complexity in each direction. Here, the FD-discretized Poisson and Helmholtz equations render banded matrices which are solved with the LU-method. The decompositions are precomputed at the beginning of a simulation and at each time step only backward and forward substitutions need to be computed, resulting in an operation count of $\mathcal{O}(M)$ for the solution of each system. Note that for the axially and azimuthally invariant Fourier mode, $n=l=0$, the Poisson equations \eqref{eq:pressure} and \eqref{eq:correction} are singular: their solution is defined up to a constant because of the Neumann boundary conditions. Here a Dirichlet homogeneous boundary condition was employed at the outer cylinder to select a particular solution. \section{Parallelization scheme and its implementation}\label{sec:parallel_implementation} A hybrid MPI-OpenMP parallelization strategy is adopted for the implementation of the code. Since the linear equations \eqref{eq:pressure}--\eqref{eq:correction} are mode-independent, it is convenient to employ an MPI-based, one-dimensional domain decomposition (also known as ``slab'' decomposition, Fig.~\ref{fig:mpiDist}): The Fourier coefficients $(\hat{u}_+,\hat{u}_-,\hat{u}_z,\hat{p})$ corresponding to different modes are distributed across the MPI tasks, which allows to solve equations \eqref{eq:pressure}--\eqref{eq:correction} concurrently, without inter-task communications. Each of the $N_\mathrm{tasks}$ MPI tasks operates on data corresponding to a number of $m_{\theta}\cdot m_z/N_\mathrm{tasks}$ modes, where $(m_r,m_{\theta},m_z)=(M,N+1,2L)$ are the dimensions of variables in Fourier space. OpenMP threading inside each MPI task allows to efficiently exploit the remaining coarse-grained parallelism (see below). \begin{figure}[!ht] \centering \includegraphics[width=1.0\textwidth]{mpi_alltoall_v2.eps} \caption{Schematic of the MPI-based, one-dimensional ``slab'' domain decomposition and the global transposition by using the function {\tt MPI\_Alltoall()}. Mode-independent spectral coefficients in Fourier space are distributed among different MPI tasks. Each variable has a dimension of $(m_r,m_{\theta},m_z)$.} \label{fig:mpiDist} \end{figure} We compute the nonlinear term (step \emph{1} in Section~\ref{sec3:numericalMethod}) by performing global matrix transpositions (Fig.~\ref{fig:mpiDist}) of the discretized fields $\partial_r\hat{\textbf{u}}$ and $\hat{\textbf{u}}$ such that for each radial point the complete spectrum of Fourier modes is localized in one MPI task. This requires a collective communication operation of type "all-to-all" but allows to most efficiently compute the Fourier transformations and the derivatives with respect to the spectral coordinates, namely $\theta$ and $z$. Finally, inverse transpositions are performed for the resulting array $\hat{\textbf{N}}$. In our applications, typically $m_{\theta}\cdot m_z \gg m_r$ applies, i.e.\ there are many more Fourier modes than radial grid points. Hence, the number of MPI tasks in our slab decomposition is bounded by $N_\mathrm{tasks} \le m_r$ (cf.~Fig..~\ref{fig:mpiDist}) and consequently the achievable parallel speedup with respect to the serial code would be at most $m_r$. However, OpenMP threads allow to parallelize over the $m_{\theta} \cdot m_z/N_\mathrm{tasks}$ modes within a MPI task, while retaining the one-dimensional MPI domain decomposition, which is conceptually straightforward to implement. Similarly, we can exploit concurrency in the nonlinear part if $N_\mathrm{tasks} < m_r$ applies. In addition, the Fourier transformations and the individual partial derivatives required for evaluating $\textbf{u}\cdot\nabla\textbf{u}$ are computed concurrently and the transposition of $\partial_r\hat{\textbf{u}}$ is overlapped with the computation of $\textbf{u}$, $\partial_{\theta}\textbf{u}$, and $\partial_{z}\textbf{u}$. Theoretically, this strategy allows to utilize a number of $\min(m_r,m_{\theta}\cdot m_z)\cdot N_\mathrm{threads}$ processor cores where $N_\mathrm{threads}$ is the maximum number of threads a shared-memory compute node provides. Current high performance computing (HPC) platforms feature at least 16 cores with 32 logical threads per node (e.g. Intel Xeon E5 Sandy-Bridge or Ivy-Bridge processors), and thread-based concurrency on the node-level is expected to increase substantially in the near future, in particular with the many-core processors and GPU-accelerated nodes \cite{dongarra_hpc2011}. In practice, we achieve the best parallel efficiencies when MPI tasks are mapped to the individual "sockets" (i.e.\ CPUs or NUMA domains) of a compute node and the number of OpenMP threads equals the number of physical cores per socket. Due to the smaller number of MPI tasks per node (compared with a plain MPI parallelization) the amount of inter-node communications is reduced in the global transposition. This transposition, which is implemented by {\tt MPI\_Alltoall} collective communication and task-local transpositions, ultimately limits the overall parallel scalability of the code at high task counts (see\ Section~\ref{sec:benchmarks}). The code is implemented in FORTRAN~90 and has been ported to a number of major HPC architectures, including IBM Power and BlueGene, as well as compute clusters based on x86\_64 processors and high-performance interconnects such as InfiniBand. We employ vendor-optimized BLAS and LAPACK routines for the matrix-vector multiplication (BLAS level-2 routine {\tt DGEMV}) and the linear solvers (LAPACK routines {\tt DGBTRF}, {\tt DGBTRS} taken e.g.\ from the Intel Math Kernel Library, MKL, or IBM ESSL), respectively, and utilize the MKL or the FFTW library \cite{FFTW} for performing the Fourier transformations in the nonlinear part of the code. For data output we employ the parallel HDF5 libraries which enable collective output of the MPI-distributed data into a single file in a transparent and efficient way. This facilitates data handling, post-processing and visualization, \textit{e.g.} with VisIT or Paraview (cf.\ Fig.~\ref{fig:highReRun}). \section{Numerical Accuracy and Code Validation}\label{sec:validation} The code has been tested\footnotemark[1] over a wide range of Reynolds numbers $Re\in[50,100\,000]$. A number of specific test cases will be given in the following. \footnotetext[1]{The platform for the tests is a small departmental cluster with two Intel Xeon E5640 four-core processors in each node. The Intel Compiler (v12.0) and the Intel MKL library (v10.3) were employed. The experiments described in the last part in this section (fully turbulent flow) were done on the same platform as described in \S6.} \subsection{Laminar flow} We firstly computed the laminar velocity profile, which is also known as circular Couette flow. It can be expressed as $\textbf{U}=(0,U_{\theta}(r),0)$, where $U_{\theta}(r)=C_1r+C_2/r$ with $ C_1= (Re_o-\eta Re_i)/(1+\eta)$ and $C_2= \eta(Re_i-\eta Re_o)/((1-\eta)(1-\eta^2))$, and corresponds to pure rotary shear flow. The tests were performed at $Re_i=50, Re_o=200$ and at $\eta=0.5$. \begin{figure}[!ht] \centering \includegraphics[width=0.7\textwidth]{lamComp} \\ \includegraphics[width=0.7\textwidth]{errVsR_v2}\\ \caption{Laminar Couette flow at $Re_i=50, Re_o=200$ and $\eta=0.5$ with Chebyshev points ($\alpha\rightarrow 0$ in equation \eqref{eq:nodeDist}). Top: numerically obtained streamwise velocity profile (blue squares) and pressure (green circles) for $n_r=32$. The dashed lines show the corresponding curves for the exact Couette solution. Bottom: local relative error $\epsilon_u$ for $n_r=64$ as a function of $r$ and several stencil lengths $n_s$.} \label{fig:lamProfile} \end{figure} A non-uniform grid according to formula (\ref{eq:nodeDist}) was used in the radial direction. Fig.~\ref{fig:lamProfile}(top) shows the numerical velocity and pressure profiles for $\alpha\rightarrow 0$ (Chebyshev points) and $n_r=32$, which match well with the theoretical curves (dashed lines). The distributions of the relative error $\epsilon_u(r)=|\frac{u_{\theta}-U_{\theta}}{U_{\theta}}|$ along the radial direction are shown in Fig.~\ref{fig:lamProfile}(bottom) for $n_r=64$ and different stencil lengths $n_s$. In the FD method, the stencil length is the number of consecutive points used to approximate the derivatives. As $n_s$ is increased, the relative error decreases until approaching the machine precision. We found that a stencil of 9 points gives a good compromise between computing time and accuracy, so $n_s=9$ is kept for the following tests. To measure the global error, we integrated the local error $\epsilon_u$ over the radial direction, $\mathcal{E}_u=\int_{r_i}^{r_o} \epsilon_u r dr$. This is shown in Fig.~\ref{fig:relErrE} as a function of $n_r$ and $\alpha$. In the left panel, $\mathcal{E}_u$ scales as a power law with $n_r$ for both $\alpha=0$ and $\alpha =0.5$. The power exponent is about $-11$, which is better than expected from the 9-point-stencil FD scheme. The right panel shows that the error is minimized for $\alpha\simeq 0.5$ and that below 0.5 the errors are almost at the same level. \begin{figure}[!ht] \centering \includegraphics[width=0.49\textwidth]{gErrVsNr_ns9both.eps} \includegraphics[width=0.47\textwidth]{gErrVsAlpha_nr32ns9.eps} \caption{Global relative error $\mathcal{E}_u$ as a function of $n_r$ (left) and $\alpha$ for $n_r=32$ (right). The dashed line in the left panel is a power fit with an exponent of -11. The stencil length is $n_s=9$ in both panels.} \label{fig:relErrE} \end{figure} \subsection{Hydrodynamic instability and three-dimensional time-dependent flow} As the Reynolds number of the inner cylinder increases beyond a certain value, laminar Couette flow gives way to Taylor vortices, and subsequently to wavy vortex flow. Following Jones~\cite{Jones_jfm1985} we computed the onset of wavy vortex flow for $\eta=0.56$, $\Gamma=2.2$ and $k_\theta=1$. The simulations were initialized with Taylor vortex flow, which was disturbed with a perturbation of azimuthal wavenumber $n$. We then measured the exponential growth/decay rate of the kinetic energy of the disturbed mode, which vanishes at the critical point. In agreement with Jones we found that the dominant mode has $n=1$. The critical Reynolds number, obtained with $(n_r,n_\theta,n_z)=(64,48, 128)$ and $\Delta t = 2\times 10^{-5}$, was determined to $Re_i^{c}\in [408.09,408.1]$, which was reproduced by using the Petrov--Galerkin code of Meseguer \emph{et al.}~\cite{Meseguer_EPJ2007}. We note that Jones quotes a slightly lower value (399, Table~1 of~\cite{Jones_jfm1985}). We attribute this discrepancy of about 2\% to the limited axial and radial resolutions, which Jones could use in the eighties. Nonlinear time-dependent wavy vortex flow was computed at $Re_i=458.1, Re_o=0$, $\eta=0.868$ and is shown in Fig.~\ref{fig:taylorVortices}. The axial length was chosen as $\Gamma=2.4$ and $k_{\theta}=6$ to compare to the experimental observations of King \emph{et al.}~\cite{King_jfm1984} and numerical simulations of Marcus~\cite{marcus1984}. Wavy Taylor vortices are a relative equilibrium: they consist of a constant pattern rotating as a solid at a constant wave speed. Marcus~\cite{marcus1984} notes: `\emph{A test that is more sensitive than the comparison of torques is the comparison of the numerically computed wave speed with the experimentally observed wave speed}'. We performed this test with spatial resolution $(n_r,n_\theta,n_z)=(32, 32, 32)$ and time-step size $\Delta t=2\times 10^{-5}$. The wave speed normalized by the rotation speed of the inner cylinder was accurately computed with a rigorous method based on Brent's minimization algorithm~\cite{Brent_Dover}. Our result, namely a wave speed of $c=0.34432$, with the pattern rotating at about one-third of the speed of the inner cylinder, agrees to all decimal places given in~\cite{King_jfm1984}. The same result was reproduced with higher resolutions (as in \S\ref{sec:slip}) and on various HPC platforms. \begin{figure}[!ht] \centering \includegraphics[height=0.3\textwidth]{contour_uthm}\\[5pt] \caption{Contour plots of the streamwise velocity in the middle $(\theta,z)$ plane for wavy Taylor vortices. The outer cylinder is stationary, whereas the inner cylinder rotates with $Re_i=458.1$. The geometrical parameters are $\eta=0.868$ and $\Gamma=2.4$ and only one sixth of the circle ($k_\theta=6$) was used in the simulations and is displayed. Here $(n_r,n_\theta,n_z)=(32, 32, 32)$.} \label{fig:taylorVortices} \end{figure} The choice $k_\theta=6$, as in \cite{marcus1984}, automatically fixes the symmetry of the observed wavy mode. To remove this constraint, we did an extra set of simulations with $k_{\theta}=1$, {\em i.e.} the whole annulus in the azimuthal direction. After initial transients the flow stabilized to a wavy state, whose wavenumber and speed depended on the disturbance used to initialize the run. When a disturbance with $n=6$ was used, exactly the same result as above was obtained. We also tried $n=5$ an $n=7$ and obtained new wavy-vortex-flow sates, all of which had very similar wavespeed. The same results were reproduced by doubling the axial length of the domain, i.e. $\Gamma=4.2$, and for much longer cylinders ($\Gamma=14.4$). Finally, we note that for other disturbances, which simultaneously excited several modes, we could also obtain chaotic flow states. Overall, the results are in agreement with the experiments of Coles \cite{coles1965}, who demonstrated the coexistence of several flow states at identical parameter values depending on the history of the flow. An investigation of the basin of attraction of each of the possible flow states would be very interesting and may be pursued in the future. We expect extreme multiplicity, much beyond what Coles observed. \subsection{Velocity slip at the cylinders}\label{sec:slip} We further examined the tangential velocity slip at the cylinders. In the projection scheme we employed, the incompressibility constraint $\nabla\cdot \textbf{u}=0$ is discretely fulfilled by construction, in that the Poisson equation for $\phi$ in \S\ref{sec3:numericalMethod} is derived by applying the divergence-free condition. However, the velocities at the inner and outer cylinders slip by a predicted amount of $|\nabla \phi|=\mathcal{O}(\Delta t^3)$ after the correction step~\cite{Hugues_ijnmf1998,RaspoBontoux_ComptFluids2002}. We have evaluated the L2-norm of the tangential velocity slip at the inner cylinder, $\int_{\theta}\int_{z}\sqrt{((u_{\theta}-Re_i)^2-u_z^2)|_{r=r_{i}}}d\theta dz$. In Fig.~\ref{fig:slipVel} the relative velocity slip, i.e. slip velocity normalized with $Re_i$, is shown as a function of $\Delta t$ for several radial resolutions $n_r$. For the lowest resolution $n_r=32$ the curve rapidly levels off, indicating that spatial-discretization errors dominate over temporal errors. Note that with the largest time-step size allowed for stability and lowest resolution we already obtain five digits in the accuracy of $c$. As $n_r$ is increased the slip velocity decreases and its scaling gradually approaches a power law, here with an exponent of approximately 2.5. The reason why this is slightly below the predicted value of $3$, as observed by Raspo \emph{et al}. \cite{RaspoBontoux_ComptFluids2002} for simple cosine and sine flows, is unclear. The behaviour seen in Fig.~\ref{fig:slipVel} suggests that very high spatial resolutions may be needed to observe the asymptotic scaling. Nevertheless we stress that the even with the coarsest resolution and time step, the relative slip error is of about $10^{-6}$. We repeated this study for fully turbulent flow at $Re_i=8000$ (see \S\ref{sec:turb} for further tests with the same setup). The results were very similar, and in fact, even with $\Delta t =2\times 10^{-7}$, which is close to the stability limit ($\Delta t_\text{lim}\approx 3\times 10^{-7}$), the relative error was $10^{-9}$. We conclude that in typical simulations the dominating source of error comes from the spatial discretization, which determines the slip error observed in practice. A time step moderately smaller than permitted by stability yields very accurate results in the solution and very small slip velocities. \begin{figure}[!ht] \centering \includegraphics[width=0.7\textwidth]{slipVel_v2} \caption{Normalized velocity slip at the inner cylinder versus time-step size $\Delta t$ for different $n_r$. The parameters are the same as in Figure~\ref{fig:taylorVortices}, corresponding to wavy vortex flow. The spatial resolution in the axial and azimuthal directions is $(n_\theta,n_z)=(32,32)$} \label{fig:slipVel} \end{figure} \subsection{Localized turbulence at moderate Re}\label{sec:stripe} Localized turbulence, interspersed in the surrounding laminar flow, is a typical feature of transitional Reynolds numbers in shear flows. The turbulent stripe pattern found in the counter-rotating Taylor-Couette flow in the narrow-gap limit is an example. We computed this pattern for exact counter-rotation ($Re_i=-Re_o=680$) and $\eta=0.993$. The time-step size was $\Delta t = 2\times 10^{-5}$ and the domain size in the axial direction was $\Gamma=50$, whereas $k_{\theta}= 179$. Folowing Barkley and Tuckermann~\cite{barkley2005} the $\theta$-direction in our computational domain is tilted with an angle of $24^{\circ}$ to the streamwise direction (for details see~\cite{ShiHof_prl2013}). We tested the probability distributions of the splitting times of turbulent stripes reported in~\cite{ShiHof_prl2013}, which were obtained by using the spectral Petrov-Galerkin code of Meseguer \emph{et al.}~\cite{Meseguer_EPJ2007}. We here used $n_r=32$ in the radial direction, whereas Shi \emph{et al.}~\cite{ShiHof_prl2013} used modified Chebyshev polynomials of degree up to 26. In both cases the azimuthal and axial resolutions are $n_\theta=48$ and $n_z=640$, respectively. The exponential distributions of splitting times obtained by both codes are statistically equivalent (see the inset in Fig.~\ref{fig:resolComp}): our computed characteristic time of the exponential distributions is well within the $95\%$ confidence interval reported in~\cite{ShiHof_prl2013}. \begin{figure}[!ht] \centering \includegraphics[width=0.75\textwidth]{resolCompare_v1} \caption{Probability distributions of the splitting time of a single turbulent stripe at $Re_i=680,Re_o=-680$ and at $\eta=0.993$. Circles correspond to the data set obtained with the spectral Petrov--Galerkin code of Meseguer \emph{et al.}~\cite{Meseguer_EPJ2007}, whereas squares correspond tot the data set obtained with the present code with $(n_r,n_\theta,n_z)=(32,48,640)$. Inset: characteristic splitting time estimated with the sample mean. The error bar shows the $95\%$ confidential interval.} \label{fig:resolComp} \end{figure} \subsection{Fully turbulent flow at high Re}\label{sec:turb} The robustness of the code was further validated at high Reynolds numbers in the linearly unstable regime, where the flow is fully turbulent. Here we computed the global torque exerted by the fluid on the inner and outer cylinders, which characterizes the turbulence intensity and the transport of angular momentum~\cite{BrauckmannEckhardt_jfm2012}. The tests were done at $Re_i=8000$ and stationary outer cylinder with $\eta=0.5$, $\Gamma = 2\pi/k_z=\pi$ and $k_{\theta}=2$. The time-step size is $\Delta t = 2\times 10^{-7}$. As is shown in Fig.~\ref{fig:nuVSresol}(top), the quasi-Nusselt number $Nu_{\omega}$~\cite{BrauckmannEckhardt_jfm2012}, which is the torque normalized by the torque of the laminar flow, converges to $8.815$ at the resolution $(n_r,n_{\theta},n_z)=(128,192,320)$. This value agrees very well to the value of $8.816$ recently reported by Brauckmann and Eckhardt~\cite{BrauckmannEckhardt_jfm2012}, who also used the spectral Petrov-Galerkin code of Meseguer \emph{et al.}~\cite{Meseguer_EPJ2007}. The temporal fluctuation of the quasi-Nusselt number obtained with the highest resolution is shown in the bottom figure. At this $Re$, we also examined the influence of the radial node distribution by varying the parameter $\alpha$ in equation~\eqref{eq:nodeDist}. Three runs with $\alpha=0,0.5,0.99$ were done, and all three rendered $Nu_{\omega}=8.81\pm 0.05$. The maximum time-step size for stability was found to be $\Delta t_\text{lim} \approx 3\times 10^{-7}$ for all three values of $\alpha$. This is explained by the fact that the CFL number is dominated by the azimuthal direction. Thus at this Reynolds number the Chebyshev node distribution does not impose a restriction in the time-step yet. \begin{figure}[!ht] \centering \includegraphics[width=0.75\textwidth]{nuConvg_Re8000_v3.eps}\\ \includegraphics[width=0.75\textwidth]{tseries_nu_Re8000.eps} \caption{(Top) Quasi-Nusselt number at $Re_i=8000$, $Re_o=0$, $\eta=0.50$, $\Gamma=2\pi/k_z=\pi$ and $k_\theta=2$, as a function of the azimuthal and axial resolutions for $n_r=128$. The dashed line corresponds to the value ($Nu_{\omega}\simeq 8.816$) reported in~\cite{BrauckmannEckhardt_jfm2012}. The error bars indicate the $95\%$ confidential interval. (Bottom) The temporal fluctuation of the quasi-Nusselt number at the highest resolution $(n_r,n_\theta,n_z)= (128,192,320)$.} \label{fig:nuVSresol} \end{figure} Another test run was performed at $Re_i=10^5, Re_o=79685$ and at $\eta=0.71$. We used $k_\theta=16$ and axial length $\Gamma=0.5$, with a spatial resolution $(n_r,n_\theta,n_z)=(1152, 384, 384)$ and time-step size $\Delta t = 10^{-9}$. The initial condition at $t=0$ is taken from the optimal initial perturbation, which gives the maximal transient energy growth~\cite{Maretzke_jfm2014} supplemented with very small three-dimensional noise. Fig.~\ref{fig:highReRun} shows the 3D contour plot of the streamwise vorticity, $\omega_{\theta}=\partial_z u_r-\partial_r u_z$, at $t=5\times10^{-4}$ ($\simeq 3.3$ cylinder rotations) which illustrates a transiently turbulent flow state. The research is still ongoing and will be disseminated in future publications. We expect that the results will contribute to clarify the role of pure hydrodynamic turbulence in astrophysical disks~\cite{Balbus_nature2011}. \begin{figure}[!ht] \centering \includegraphics[width=0.9\textwidth]{wth_Re1e5.eps} \caption{Isosurfaces of the streamwise vorticity in the quasi-Keplerian regime at $Re_i=10^5, Re_o=79685$, $\eta=0.71$, $\Gamma=0.5$ and $k_\theta=16$. The resolution is $(n_r,n_\theta,n_z)=(1152, 384, 384)$.} \label{fig:highReRun} \end{figure} \section{Computational efficiency}\label{sec:benchmarks} \subsection{Benchmark setup} In this section we report benchmarks results using up to 20\,480 processor cores of an IBM iDataPlex compute cluster with Intel Ivy-Bridge processors and a fully nonblocking InfiniBand (FDR~14) fat-tree interconnect. Each shared-memory compute node hosts two Intel Xeon E5-2680v2 ten-core processors (CPUs) with a clock frequency of 2.8~GHz. We employ Intel compilers (version 14.0), and the Intel Math Kernel Library (MKL 11.1) for BLAS, LAPACK, and FFT functionality. We have performed two strong scaling studies, which show the scaling of the runtime with increasing number of CPU cores for a fixed problem size. Two different, representative setups were considered: \begin{enumerate}[\itshape a\upshape)] \ite a ``SMALL'' setup with $(n_r,n_{\theta},n_z)=(32,384,640)$. This setup is used to investigate localized turbulence at the transitional stage ($Re\sim \mathcal{O}(10^2)$), where the structures inside the turbulence are relatively large. The probability distributions of the splitting time of localized turbulent stripe in \S\ref{sec:stripe} are obtained with this resolution. \ite a ``LARGE'' setup with $(n_r,n_{\theta},n_z)=(2048,384,2048)$. This resolution is representative of our ongoing studies of hydrodynamic turbulence in Taylor-Couette flows with quasi-Keplerian velocity profiles at Reynolds numbers up to $\mathcal{O}(10^5)$. \end{enumerate} For the LARGE setup, a weak scaling study, i.e.\ increasing the problem size along with the number of CPU cores, is presented in addition. \subsection{Benchmark results and discussion} Fig.~\ref{fig:scaling} provides an overview of the strong scalability of the hybrid code. Different colors and symbols are used to distinguish runs with different numbers of MPI tasks ($N_\mathrm{tasks}$) and OpenMP threads ($N_\mathrm{threads}$). The total number of processor cores is given by $N_\mathrm{cores}=N_\mathrm{tasks}\cdot N_\mathrm{threads}$. For both setups we achieve scalability up to the maximum number of cores our parallelization scheme admits on this computing platform, \textit{i.e.}, \ $N_\mathrm{cores}=32\cdot 20=640$ for the SMALL setup, and $N_\mathrm{cores}=2048\cdot 10=20480$ for the LARGE setup. The latter number is constrained by the maximum number of nodes that can be used by a single job. For the SMALL setup, Fig.~\ref{fig:scaling} (upper left) shows that up to a number of 10 threads per MPI task the run times for a given number of cores are virtually the same, independent of the distribution of the resources to MPI tasks and OpenMP threads. This indicates that the efficiency of our coarse-grained OpenMP parallelization is almost the same as the explicit, MPI-based domain decomposition. Moreover, as the results for the LARGE setup (Fig.~\ref{fig:scaling}, upper right) show, it can even be more efficient to use less than the maximum of $n_r$ MPI tasks for a given number of cores and utilize the resources with OpenMP threads (compare the cyan and the red symbols at moderate core counts). This is due to the fact that a lower number of MPI tasks per node reduces the amount of inter-node MPI communication (specifically the {\tt MPI\_Alltoall} communication pattern for the global transpositions). Notably, for the LARGE setup, the hybrid code shows nearly perfect scaling between 1280 and 2560 cores and continues to scale up to more than 20\,000 processor cores (1024 nodes), albeit with a marginally efficient speedup of 8 (corresponding to a parallel efficiency of slighly more than 50\%) when compared with the baseline run at 1280 cores. A floating-point performance of about 20 GFlop/s per compute node is reached which is roughly 5\% of the nominal peak performance. The memory footprint remains below 10 GB per node, making the code well prepared for future CPU architectures with scarce memory resources \cite{dongarra_hpc2011}. \begin{figure}[!h] \begin{tabular}{@{\extracolsep{-15pt}}lr} \includegraphics[width=0.48\textwidth]{scaling_SMALL.eps} \hfill \includegraphics[width=0.48\textwidth]{scaling_LARGE.eps} \\ \includegraphics[width=0.48\textwidth]{speedup_SMALL.eps} \hfill \includegraphics[width=0.48\textwidth]{speedup_LARGE.eps} \end{tabular} \caption{Runtime per time step (upper row) and relative speedups (bottom row) for the SMALL setup (left panels) and for the LARGE setup (right panels) as a function of the number of cores, $N_\mathrm{cores}=N_\mathrm{tasks}\cdot N_\mathrm{threads}$. Different colors and symbols are used to distinguish runs with different numbers of MPI tasks ($N_\mathrm{tasks}$) and OpenMP threads ($N_\mathrm{threads}$), respectively. The slope of an ideal scaling curve is indicated by dotted lines.} \label{fig:scaling} \end{figure} \bigskip The details on the absolute run times and the parallel efficiencies of the whole code (the bottom row) as well as the individual parts of the algorithm (cf.\ Section~\ref{sec3:numericalMethod}) are listed in Table~\ref{tab:efficiency}. The first column, which corresponds to a plain MPI-parallelization using the maximum number of tasks ($N_\mathrm{tasks}=n_r$) for the given setup, is assigned an efficiency of 100\%, by definition. \begin{table}[!h]\footnotesize \begin{center} \begin{tabular}{|l|lr|lr|lr|lr|lr|} \hline\hline\multicolumn{11}{|c|}{SMALL setup $(32,384,640)$}\\\hline\hline \hline \multicolumn{1}{|r|}{cores ($N_\mathrm{threads}$)} & \multicolumn{2}{|c|}{$32 (1)$}& \multicolumn{2}{|c|}{$64 (2)$}& \multicolumn{2}{|c|}{$160 (5)$}& \multicolumn{2}{|c|}{$320 (10)$}& \multicolumn{2}{|c|}{$640 (20)$}\\ \hline & $T_1$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_2$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_5$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_{10}$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_{20}$ [s] & \multicolumn{1}{l|}{$\eta$}\\ \hline nonlinear \hfill (\emph{1}) & 0.696 & 100\% & 0.388 & 90\% & 0.161 & 86\% & 0.101 & 69\% & 0.093 & 37\% \\ predictor-corrector \hfill (\emph{2,3,4}) & 0.152 & 100\% & 0.080 & 95\% & 0.033 & 92\% & 0.017 & 89\% & 0.001 & 79\% \\ \hline complete step & 0.864 & 100\% & 0.472 & 92\% & 0.200 & 86\% & 0.120 & 72\% & 0.104 & 42\% \\ \hline \hline\hline\multicolumn{11}{|c|}{LARGE setup $(2048,384,2048)$}\\\hline\hline \hline \multicolumn{1}{|r|}{cores ($N_\mathrm{threads}$)} & \multicolumn{2}{|c|}{$2048 (1)$}& \multicolumn{2}{|c|}{$4096 (2)$}& \multicolumn{2}{|c|}{$10240 (5)$}& \multicolumn{2}{|c|}{$20480 (10)$}& \multicolumn{2}{|c|}{} \\ \hline & $T_1$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_2$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_5$ [s] & \multicolumn{1}{l|}{$\eta$}& $T_{10}$ [s] & \multicolumn{1}{l|}{$\eta$}& & \\ \hline nonlinear \hfill (\emph{1}) & 3.875 & 100\%& 1.862 & 104\%& 0.870 & 89\%& 0.695 & 70\%& & \\ predictor-corrector \hfill (\emph{2,3,4}) & 0.407 & 100\%& 0.240 & 85\%& 0.097 & 84\%& 0.048 & 85\%& & \\ \hline complete step & 4.325 & 100\%& 2.134 & 101\%& 0.977 & 87\%& 0.751 & 58\%& & \\ \hline \end{tabular} \caption{Runtime per time step, $T_n$ and parallel efficiency $\eta$ of the OpenMP parallelization as a function of the number $N_\mathrm{threads}$ of OpenMP threads per MPI task, using the maximum number of 32 MPI tasks for the SMALL setup, and 2048 MPI tasks for the LARGE setup, respectively. Parallel efficiency is conventionally defined as $\eta:= T_1/(n\cdot T_n)$ with $n=N_\mathrm{threads}$. Different rows show the contributions of the individual algorithmic steps (numbering in brackets chosen according to Section~\ref{sec3:numericalMethod}) to the total runtime of a complete time step (the bottom row).}\label{tab:efficiency} \end{center} \end{table} For the SMALL setup (the upper part of Table~\ref{tab:efficiency}) we observe good OpenMP efficiency up to 10 threads (which are pinned to the 10 physical cores of a single CPU socket) per MPI-task for the pressure and velocity predictor steps, the corrector step, and also the matrix-vector multiplication in the nonlinear part. When using all 20 cores of a shared-memory node with a single MPI task (cf.\ the magenta curve in Fig.~\ref{fig:scaling}, left) one notices a degradation in OpenMP efficiency which is due to memory-bandwidth limitations, NUMA effects, and limited parallelism in the nonlinear part. The overall parallel efficiency (the bottom row) can be considered as very good up to 320 cores, but gets increasingly bounded by the global transposition ({\tt MPI\_Alltoall} communication) in the nonlinear part. \begin{figure}[!ht] \centering \includegraphics[width=0.8\textwidth]{weakscaling_LARGE.eps} \caption{Weak scaling of the total runtime (solid line) and the contribution of the global transposition (dashed) for the LARGE setup. The number of cores is related to the number of radial zones as $N_\mathrm{cores}=10\cdot n_r$. In all cases, $N_\mathrm{threads}=10$ was used.} \label{fig:weakscaling} \end{figure} For the LARGE setup (the lower part of Table~\ref{tab:efficiency}), although the highly scalable linear parts (predictor-corrector, steps 2-4) and the matrix-vector multiplications contribute only a minor part to the total runtime, the code maintains an excellent OpenMP efficiency up to more than 10\,000 cores (87\%). At very high core counts the {\tt MPI\_Alltoall} communication dominates the total runtime and becomes the major bottleneck for overall scalability. This is also apparent in the weak scaling analysis (cf.\ Fig.~\ref{fig:weakscaling}). The global transposition exhibits good weak scalability up to 10\,240 cores (512 nodes), and its contribution to the total runtime remains subdominant but it seriously impedes the scalability up to the maximum of 20\,480 cores (1024 nodes). Note that parts of the global transposition (roughly a third, in terms of runtime) are performed concurrently with computations and are thus not accounted for separately in Fig.~\ref{fig:weakscaling}. Using an adapted setup of $(n_r,n_\theta,n_z)=(1792,384,512)$, we were able to run the code on the largest, fully interconnected partition with 1792 nodes (35\,840 cores) of the high-performance computer of the Max-Planck-Society, "Hydra", resulting in a run time of $0.3$~s per time step. Computing times of this order enable us to perform highly resolved simulations (\textit{e.g.}\ of Keplerian flows which require on the order of a million time steps) within a couple of days. \bigskip
\section{Introduction} Let $C \subset \PP^2$ be a projective curve of degree $d > 2$ such that all its singularities are nodes, and let $\nu\colon \hatC \to C$ be the normalization mapping. For a point $p \notin C$ consider the projection $\pr_p: C \to \PP^1 = p^\perp$ from $p$; here $p^\perp \subset \mathbb P^2$ is the set of lines passing through $p$, and $\pr _p$ assigns to $x$ the line joining $p$ and $x$. The composition $\pr _p \circ \nu: \hat C \to \mathbb P^1$ will be called a generalized projection; its branch locus coincides with $p^\perp \cap C^*$, where $C^* \subset (\PP^2)^*$ is the curve dual to $C$. The generalized projection has simple ramification if and only if $p^\perp$ is transversal to $C^*$. Suppose that $\pi\colon C'\to p^\perp$, where $C'$ is a smooth projective curve, is a holomorphic mapping with simple ramification and such that the branch locus of $\pi$ coincides with $p^\perp \cap C^*$. We are looking for a criterion for $\pi$ to be isomorphic to the generalized projection $\pr_p\circ\nu$. \subsubsection*{Problem background} Consider first the case when $C$ is smooth. An easy dimension count shows that if $d > 3$ then the branch locus of a projection is not arbitrary. Namely, it follows from the Riemann--Hurwitz formula that a degree $d$ map from a curve of degree $d$ has $d(d-1)$ critical values, provided the ramification is simple. So, its branch locus is a point of $\Sym^{d(d-1)}\PP^1 = \PP^{d(d-1)}$. The space of projective curves $C \subset \PP^2$ of degree $d$ has dimension $\binom{d+2}{2}-1 = d(d+3)/2$. For a point $p \in \PP^2$ there is a $3$-dimensional group of projective automorphisms $\PP^2 \to \PP^2$ preserving $p$ and all the lines containing $p$. So, the space of all projections has the dimension $m_d \mathrel{:=} d(d+3)/2-3 = (d^2+3d-6)/2 \ll d(d-1)$ for large $d$. Consider now small $d$. For $d=1$ and $d=2$ the situation is trivial. For $d=3$ one has $m_3 = 6 = 3(3-1)$, and indeed, it is easy to see that any $6$ pairwise distinct points $a_1, \dots, a_6 \in \PP^1$ are branch points of a suitable generic projection of a cubic curve. The case $d=4$ was studied in detail by R.Vakil in \cite{Vakil}. Here $m_4 = 11$, $4(4-1) = 12$, so branch loci of generic projections lie in a hypersurface $V \subset \PP^{12}$. For a point $a \in V$ there are $255$ pairs $(C',\pi)$ where $C'$ is a smooth curve and $a$ is the branch locus of the mapping $\pi: C' \to \PP^1$ of degree $4$. For $120$ pairs $C'$ is a smooth plane projective curve and $\pi$ is isomorphic to a projection. For the remaining $135$ pairs the curve $C'$ is genus $3$ hyperelliptic, hence not plane. For $d > 4$ it is easy to derive from~\cite[Proposition 1]{EGH} that if $C'$ is a smooth plane projective curve then each degree $d$ mapping $\pi: C' \to \PP^1$ is isomorphic to a projection. We do not, though, assume $C'$ to be a plane curve, so a situation similar to the $d=4$ case may arise; cf.\ Proposition~\ref{prop:example}. For $C$ nodal a degree $d$ map $\hatC \to \PP^1$ does not need to be equivalent to a projection. \subsubsection*{Main results} To make our problem more tractable we impose some generality hypotheses on the curves and mappings in question. To wit, we will assume that the mappings $\pi\colon\hatC\to\PP^1$ have simple ramification (see Definition~\ref{def:moderate_covering} below) and that the nodal curve $C\subset\PP^2$ is general enough in the sense of Definition~\ref{def:general_curve}. If $\nu\colon\hatC\to C$ is the normalization, then for the general point $p\in\PP^2\setminus C$ the generalized projection $\pr _p\circ\nu \colon \hatC\to\PP^1$ has simple ramification. We will say that the holomorphic mappings $\ph_1\colon C_1\to\PP^1$ and $\ph_2\colon C_2\to\PP^1$ are \emph{equivalent} if there exists a holomorphic isomorphism $\psi\colon C_1\to C_2$ such that $\ph_2\circ\psi=\ph_1$. \begin{theorem}\label{th.main} Suppose that $C\subset \PP^2$ is a nodal curve of degree~$>2$ that is general enough in the sense of Definition~\ref{def:general_curve}. Suppose that a point $p\in\PP^2\setminus C$ is such that the composition $\pr_p\circ\nu\colon\hatC\to\PP^1$, where $\nu\colon\hatC\to C$ is the normalization and $\pr_p\colon C \to p^\perp = \PP^1$ is the projection from $p$, has simple ramification. Suppose that $C'$ is a smooth projective curve and $\pi\colon C'\to p^\perp$ is a holomorphic mapping with simple ramification such that the branch locus of $\pi$ coincides with $p^\perp \cap C^*$. If $\deg C=3$, assume in addition that $C$ has a node or $\deg\pi\ne4$. Then the following two conditions are equivalent: \textup{(a)} $\pi$ is equivalent to $\pr _p \circ \nu$: there exists an isomorphism $\ph\colon C'\to\hat C$ such that $(\pr_p\circ\nu)\circ \ph=\pi$. \textup{(b)} There exist a smooth projective surface $X$, a finite holomorphic mapping $f\colon X\to(\PP^2)^*$ that is ramified exactly over $C^*$, and an isomorphism $\ph\colon C'\to f^{-1}(p^\perp)$ such that $\left.f\right|_{f^{-1}(p^\perp)}\circ \ph=\pi$. \end{theorem} \begin{note} The implication $(\mathrm a)\Rightarrow(\mathrm b)$ holds without the genericity hypotheses, for arbitrary nodal curves and arbitrary generalized projections. See the proof of Proposition~\ref{gen(a)=>(b)} below. \end{note} \begin{note} If $C$ is a smooth cubic and $\deg f = 4$ then the theorem is wrong; see Remark~\ref{chisini:example}. \end{note} \begin{note} The condition of smoothness of $X$ in the theorem cannot be omitted, at least for $\deg f$ equal to $3$ or $4$: in Section \ref{sec:example} we show that for $d=3$ and $4$ there exists a general enough smooth curve $C\subset\PP^2$ of degree $d$, a line $p^\perp \subset (\PP^2)^*$ (for some $p \in \PP^2$) transversal to~$C^*$, a smooth projective curve $C'$, and a morphism $\pi\colon C'\to p^\perp$ simply ramified over $p^\perp \cap C^*$ such that condition~(b) of Theorem \ref{th.main} with the smoothness omitted holds but $\pi$ is not equivalent to the projection $\pr _p$ (Proposition~\ref{prop:example}). We do not know similar counterexamples with $d > 4$. \end{note} In other words, if we are given a simply ramified covering of $\PP^1$ with the same branch locus as that of a (generalized) projection from a given point~$p$, then this covering is the generalized projection if and only if it can be extended to a finite ramified covering of $(\PP^2)^*$ branched over $C^*$, with a smooth surface upstairs. Theorem~\ref{th.main} may be restated in topological terms: we will show that if $L \subset (\PP^2)^*$ is a projective line then the mapping $\pi\colon C'\to L$ is equivalent to the generalized projection $\pr_p\circ\nu$ if and only if the covering $\pi^{-1}(L\setminus C^*)\to L\setminus C^*$ can be extended to a covering $X_0\to(\PP^2)^*\setminus C^*$ such that its fiber monodromy satisfies some extra conditions ($C^*$ must have no bad nodes or bad cusps in the sense of Definitions~\ref{def:good.nodes} and~\ref{def:bad.cusp}); see Theorem~\ref{th.main2}. Note that we make (almost) no assumption about genus of the curve $C'$ or degree of the morphism~$f$. The $(\mathrm a) \Rightarrow (\mathrm b)$ implication in Theorem~\ref{th.main} is easy (see Section~\ref{(a)=>(b)}). If $\deg C\ge 7$, the $(\mathrm b) \Rightarrow (\mathrm a)$ implication follows very easily from Theorem~10 of Victor Kulikov's paper~\cite{Kulikov1999} (see Proposition~\ref{Chisini_argument} below), but for the remaining cases $3\le\deg C\leq 6$, where the Chisini conjecture for coverings ramified over $C^*$ is not completely proved, it requires more work. We give an argument that works uniformly in all cases, including the case $\deg C\ge 7$; our proof is quite different from Kulikov's one. As a by-product, we establish the Chisini conjecture for coverings of $\PP^2$ ramified over curves dual to general enough nodal curves of degrees $4$, $5$, and $6$ (and of degree $3$ provided that either the curve does have a node or degree of the covering is not $4$). This extends Theorem~10 from~\cite{Kulikov1999}. The paper is organized as follows. In Section~\ref{(a)=>(b)} we prove the easy part of Theorem~\ref{th.main}. In Section~\ref{sec:plastering} we study coverings of the projective plane that are ramified over a curve having only nodes and standard cusps as singularities, and in Section~\ref{sec:PL} we prove the difficult part of Theorem~\ref{th.main}. Finally, in Section~\ref{sec:example} we show that, at least if $\deg C = 3$ and $\deg C = 4$, the smoothness condition in Theorem~\ref{th.main} cannot be omitted. The authors would like to thank Maxim Kazarian, Victor Kulikov, Alexander Kuznetsov, Sergey Lando, Sergey Natanzon, Ossip Schwarzman, Andrey Soldatenkov, and Victor Vassiliev for useful discussions. We are grateful to Igor Khavkine, Francesco Polizzi, Damian R\"ossler, Will Sawin, and particularly Rita Pardini for consultations at \url{http://mathoverflow.net}. The first named author wishes to thank the Technion --- Israel Institute of Technology, where the final version of the paper was written, for the warm hospitality. Our special thanks go to Boris Shapiro, at whose instigation this research was started. Without his constant encouragement this paper would never have been written. \section*{Notation, conventions, and definitions} The base field for all the algebraic varieties will be the field \C of complex numbers. Except for the case when we mention explicitly Zariski topology, all topological spaces will be assumed Hausdorff and locally simply connected, and all the topological terms will refer to the classical topology of complex algebraic varieties. We will say that a singular point of a curve is a \emph{node} if it is locally analytically isomorphic to the singularity of the plane curve defined by the equation $xy=0$, and that a singular point is a \emph{standard cusp} (or simply a cusp) if it is locally analytically isomorphic to the singularity of the plane curve defined by the equation $y^2+x^3=0$. If $f\colon X\to Y$ is a finite holomorphic map of smooth varieties of equal dimension, then by \emph{ramification locus} (a.k.a.\ the set of critical points) of $f$ we mean the closed subset \[ R=\{x\in X\mid\text{derivative of $f$ is degenerate at $x$}\}; \] by \emph{branch locus} (a.k.a.\ the set of critical values) of $f$ we mean the subset $f(R)\subset Y$. Let $f\colon W\to V$ be a covering of topological spaces, and take $v_0\in V$. Then the action of the group $\pi_1(V,v_0)$ on the set $f^{-1}(v_0)$ via loop lifting will be called the \emph{fiber monodromy} of $f$. If $f\colon Y \to X$ is a finite morphism of algebraic varieties of equal dimension (i.e., a proper holomorphic map with finite fibers) and $B \subset X$ is its branch locus, then the restriction of $f$ to $f^{-1}(X \setminus B) \subset Y$ is a finite-sheeted covering of topological spaces. Take $x_0 \in X \setminus B$. Then by fiber monodromy of $f$ we will mean fiber monodromy of the covering $f^{-1}(X \setminus B) \to Y\setminus B$ with respect to the point~$x_0$. Suppose now that $X = \PP^1$, so the branch locus $B\subset \PP^1$ is a finite set. If $p \in B$ and $\gamma_p$ is the conjugacy class in $\pi_1(\PP^1\setminus B)$ corresponding to a small loop around $p$, then the image of $\gamma_p$ under the fiber monodromy homomorphism is a conjugacy class in the symmetric group~$S_d$, where $d$ is the number of points in $f^{-1}(x_0)$. This conjugacy class (or the corresponding partition of $d$) is called the \emph{cyclic type} of the point~$p$. \begin{definition}\label{def:moderate_covering} Suppose that $C$ is a smooth projective curve. We will say that a finite morphism $f\colon C\to\PP^1$ has \emph{simple ramification} if each branch point $\xi\in\PP^1$ has the cyclic type of a transposition. \end{definition} If $C \subset \PP^2$ is a nodal curve and $\nu\colon\hatC\to C$ is the normalization, then the generalized projection of $\hatC$ from a general point of $\PP^2$ has simple ramification. \begin{definition}\label{def:general_curve} Suppose that $C\subset\PP^2$ is an irreducible projective curve such that each singular point of $C$ is a node. Let us say that $C$ is \emph{general enough} if the following conditions are satisfied. -- all the inflexion points of $C$ are simple; -- no line is tangent to $C$ at more than two points; -- if a line is tangent to $C$ at an inflexion point, it is not tangent to $C$ elsewhere. \end{definition} Here, we assume that a line is tangent to $C$ if it is either tangent to $C$ at a smooth point or tangent to a branch of $C$ at a node. By inflexion point we mean either a smooth point $x\in C$ for which the local intersection index of $C$ with its tangent at $x$ is greater than $2$, or a node $x\in C$ for which the local intersection index of at least one of the two (limiting) tangent lines to branches a $x$ with the corresponding branch is greater than~$2$; we say that an inflexion point is \emph{simple} if the intersection index in question equals exactly~$3$. For a projective subspace $\alpha \subset \PP^n$ we denote by $\alpha^\perp \subset (\PP^n)^*$ the set of hyperplanes in $\PP^n$ containing $\alpha$, where $(\PP^n)^*$ is the dual projective space. If $\dim \alpha = k$ then $\alpha^\perp \subset (\PP^n)^*$ is a projective subspace of dimension $n-1-k$; in particular, if $\alpha$ is a point then $\alpha^\perp$ is a hyperplane (denoted by $H_\alpha$ in SGA7, see \cite{SGA7.2}). If $X \subset \PP^n$ is a projective variety, then by $X^*\subset{(\PP^n)}^*$ we will denote its projective dual, i.e., the closure of the set of hyperplanes tangent to $X$ at smooth points. We also will be using the notation $\alpha^\perp$ and $X^*$ when $\alpha, X \subset (\PP^n)^*$, where the canonical isomorphism $((\PP^n)^*)^* = \PP^n$ is assumed. If $X\subset\PP^2$ is a projective curve, we say that a line $L\subset\PP^2$ is transversal to $X$ if $L$ does not pass through singular points of $X$ and is not tangent to $X$ at any non-singular point. If $D_1$ and $D_2$ are Cartier divisors on a projective surface, then $(D_1,D_2)$ stands for their intersection index. \section{Proof of the $(\mathrm a) \Rightarrow (\mathrm b)$ implication in Theorem \ref{th.main}}\label{(a)=>(b)} This part of Theorem~\ref{th.main} follows immediately from \begin{proposition}\label{gen(a)=>(b)} Suppose that $C\subset\PP^2$ is a nodal curve, $\nu\colon \hat C\to C$ is the normalization, $p\in\PP^2\setminus C$, and $\pr _p\colon C\to p^\perp$ is the projection from $p$. Then there exists a smooth projective surface $X_C$, a finite regular mapping $f_C\colon X_C\to(\PP^2)^*$ having the dual curve $C^* \subset(\PP^2)^*$ as its branch locus, and an isomorphism $\ph\colon \hatC\to f_C^{-1}(p^\perp)$ such that $f_C\circ \ph=\pr_p\circ\nu$. \end{proposition} \begin{proof} Put \begin{equation}\label{def_of_X_C} X_C=\{(x,t)\in \hatC\times(\PP^2)^*\mid \nu(x)\in t^\perp\}. \end{equation} The surface $X_C$, being the projectivization of the vector bundle $\nu^*\T_{\PP^2}(-1)$, is smooth. Define the mapping $f_C\colon X_C\to (\PP^2)^*$ by the formula \begin{equation}\label{def:f_C} f(x,t) = t. \end{equation} For $t\in(\PP^2)^*$, \[ f_C^{-1}(t)=\{(x,t)\mid \nu(x)\in t^\perp\}; \] $f$ is a finite morphism of degree~$d$ since for the general $t\in(\PP^2)^*$ the line $t^\perp$ intersects $C$ at $d$ smooth points. Its branch locus is the set of $t$ such that $t^\perp$ is either tangent to $C$ at a smooth point or tangent to a branch of $C$ at a node, so the branch locus of $f$ coincides with~$C^*$. If $p\in \PP^2\setminus C$ then \[ f_C^{-1}(p^\perp)=\{(x,t)\in \hatC\times(\PP^2)^*\mid \nu(x), p \in t^\perp\}. \] It is clear that the mapping $\ph\colon \hatC\to f^{-1}(p^\perp)$ sending $x$ to the pair $(x,\overline{px})$ is the required isomorphism. \end{proof} \section{Plastering of generic coverings over complements to curves}\label{sec:plastering} In this section we prepare for the proof of the implication $\mathrm{(b)}\Rightarrow\mathrm{(a)}$ in Theorem~\ref{th.main}. We start with a simple general fact. Suppose that $D\subset\PP^2$ is an arbitrary projective curve, $X_0$ is a smooth affine surface and $f_0\colon X_0\to \PP^2\setminus D$ is a finite (topological) covering. \begin{proposition}\label{prop:same_mono} Suppose that, in the above setting, there exists a line $L_0\subset\PP^2$ transversal to $D$ such that the induced covering $f_0^{-1}(L_0)\to L_0\setminus D$ has simple ramification. Then for any line $L\subset\PP^2$ transversal to $D$ the induced covering $f_0^{-1}(L)\to L\setminus D$ also has simple ramification. \end{proposition} Following \cite{Kulikov1999}, we will say that coverings satisfying hypotheses of the proposition are \emph{generic}. \begin{proof}[Proof of the proposition] The set \[ U=\{t\in(\PP^2)^*\mid \text{$t^\perp$ is transversal to~$D$}\} \] is arcwise connected; let $\gamma: [0,1] \to U$ be a path such that $\gamma(0) = L_0$, $\gamma(1) = L$. Let also \[ \X=\{(x,t)\in(\PP^2\setminus D)\times U\mid x\in t^\perp\}; \] the natural projection $p\colon \X\to U$ is a locally trivial fiber bundle with the fiber homeomorphic to $\PP^1$ punctured at $\deg D$ points. The pullback covering $p^* f_0\colon \tilde \X\to \X$ restricted to $p^{-1}(t) \subset \X$ is isomorphic to the covering~$f_0^{-1}(t^\perp)\to t^\perp$. The pullback bundle $\gamma^* p: \overline X \to [0,1]$ is trivial, so the proposition follows. \end{proof} If $f_0\colon X_0\to \PP^2\setminus D$ is a covering, then $X_0$ carries a unique structure of complex variety for which the mapping $f$ is a finite unramified morphism. Suppose now that the covering is generic and all the singularities of the curve $D$ are nodes or standard cusps. It follows from a theorem of Grauert and Remmert~\cite[Expos\'e~XII, Th\'eor\`eme 5.4]{SGA1} that in this situation $f_0$ extends uniquely to a finite mapping $f\colon X\to\PP^2$ with normal~$X$. A well-known GAGA-style result~\cite[Expos\'e~12, Corollaire 4.6]{SGA1} implies that the resulting complex space $X$ will be a projective surface. We are going now to describe explicitly the singularities of surface $X$ as well as the structure of the mapping $f$ near ramification locus. In the sequel, $\Delta$ will denote the unit disk $\{z\colon \left|z\right|<1\} \subset \C$, and $\Delta^*$ will stand for the punctured disk $\Delta \setminus \{0\}$. Suppose that $q$ is a node of $D$; choose a neighborhood $U\ni q$ and an isomorphism $U\to \Delta^2$ such that $D\cap U$ is mapped to the set $\{(x,y) \in D^2\mid xy=0\}$. Fix a point $q_0\in U\setminus D$. The covering $f_0^{-1}(U\setminus D)\to U\setminus D$ induces a fiber monodromy action of the group $\pi_1(U\setminus D,q_0) = \Z \oplus \Z$ on the set $f^{-1}(q_0)$. The two generators of $\pi_1(U\setminus D,q_0)$ are the small loops around two branches of $D$ at $q$; since $f_0$ is a generic covering, Proposition~\ref{prop:same_mono} shows that the generators act on $f^{-1}(q_0)$ by transpositions. These transpositions commute, so they can be either disjoint or equal. \begin{definition}\label{def:good.nodes} In the above setting, we will say that a node $q\in D$ is \emph{good} (with respect to the covering $f_0$) if the two generators of $\pi_1(U\setminus D,q_0)$ act by disjoint transpositions. We will say that a node $q\in D$ is \emph{bad} (with respect to $f_0$) if they act by equal transpositions. \end{definition} Suppose now that $q$ is a (standard) cusp of the curve $D$. Choose a neighborhood $U\ni q$ and an isomorphism $U\to \Delta^2$ such that $D\cap U$ is mapped to the set $\{(x,y)\in D^2\mid x^3+y^2=0\}$. Fix a point $q_0\in U\setminus D$. The covering $X\setminus f^{-1}(D)\to\PP^2\setminus D$ induces a fiber monodromy action of the group $\pi_1(U\setminus D,q_0)$ on the set $f^{-1}(q_0)$. It is well known that $\pi_1(U\setminus D,q_0)$ is isomorphic to the Artin braid group on $3$ strings $B_3=\langle u,v\mid uvu=vuv\rangle$, where the generators $u$ and $v$ correspond to the two loops around two intersection points $\ell\cap D$, where $\ell$ is a line close to $q$. Since $f_0$ is a generic covering, the elements $u$ and $v$ of $\pi_1(U\setminus D,q_0)$ act on $f^{-1}(q_0)$ by transpositions. These transpositions (call them $U$ and $V$) satisfy the relations $UVU = VUV$, so they are either non-commuting or equal. \begin{definition}\label{def:bad.cusp} In the above setting, we will say that the cusp $q\in D$ is \emph{good} (with respect to the covering $f_0$) if the transpositions $U$ and $V$ are non-commuting. We will say that the cusp $q\in D$ is \emph{bad} (with respect to $f_0$) if $U = V$. \end{definition} \begin{proposition}[cf.\ {\cite[Section 1.3]{Kulikov1999}}]\label{not.smooth.over.bad} The local behavior of the mapping $f$ near points of the preimage $f^{-1}(q)$, $q \in D$, is as follows: \begin{enumerate} \item\label{It:Smooth} If $q$ is a smooth point of $D$ then $f^{-1}(q)$ consists of $d-1$ points, and $X$ smooth at all of them. At exactly one of them $f$ is ramified, and near this point $f$ is locally isomorphic to $f(x,y) = (x,y^2)$. \item\label{It:GoodNode} If $q$ is a good node then $f^{-1}(q)$ consists of $d-2$ points, and $X$ is smooth at all of them. At exactly two of them $f$ is ramified, with the same local behavior as in Case \ref{It:Smooth}. \item\label{It:BadNode} If $q$ is a bad node then $f^{-1}(q)$ consists of $d-1$ points. All of them except one are smooth. The remaining point is singular, locally isomorphic to $\{(x,y,z) \in \Delta^3 \mid z^2 = xy\}$ \textup(Du Val's $A_1$\textup); the mapping $f$ in the same coordinates is $f(x,y,z) = (x,y)$. \item\label{It:GoodCusp} If $q$ is a good cusp then $f^{-1}(q)$ consists of $d-2$ points, and $X$ is smooth at all of them. At exactly one of them $f$ is ramified, and the pair $(X,f)$ near this point is locally isomorphic to $X = \{(x,y,z) \in \Delta^3 \mid z^3 + 3xz + 2y = 0\}$, $f(x,y,z) = (x,y)$. \item\label{It:BadCusp} If $q$ is a bad cusp then $f^{-1}(q)$ consists of $d-1$ points. All of them except one are smooth, the remaining point is singular, locally isomorphic to $\{(x,y,z) \in \Delta^3 \mid z^2=x^3+y^2\}$ \textup(Du Val's $A_2$\textup); the mapping $f$ in the same coordinates is $f(x,y,z) = (x,y)$. \end{enumerate} \end{proposition} \begin{proof} A direct computation shows that the branch loci $B$ of the local models described above are as follows: $B = \{(x,0) \mid x \in \Delta\}$ for case \ref{It:Smooth}, $B = \{(x,y) \in \Delta^2 \mid xy=0\}$ for the cases \ref{It:GoodNode} and \ref{It:BadNode}, and $B = \{(x,y) \in \Delta^2 \mid x^3+y^2=0\}$ for the cases \ref{It:GoodCusp} and \ref{It:BadCusp}. Thus, in all the cases the branch loci are locally biholomorphic to the curve $D$ in a neighbourhood of $q$. By the uniqueness statement of the Grauert--Remmert theorem it is enough to check that the monodromy action of $\pi_1(\PP^2 \setminus D)$ on the fiber $f_0^{-1}(q) \subset X$ is isomorphic to the action of $\pi_1(\Delta^2 \setminus B)$ on the fiber over $0$ in the local model. If $a \in f^{-1}(q)$ is a smooth point such that $f$ has the normal form $f(z) = z^k$ in it, then the monodromy action is a cycle of length $k$; this proves the statement in cases \ref{It:Smooth} and \ref{It:GoodNode}. In case \ref{It:BadNode} the branch locus is a union of two lines; the group $\pi_1(\Delta^2 \setminus B)$ is generated by two loops circling around these lines; apparently, they both act by the same transposition of the two preimages of the origin. To cover cases \ref{It:GoodCusp} and \ref{It:BadCusp} notice that $\pi_1(\Delta^2 \setminus B) = B_3$ in both of them. Any homomorphism $B_3 \to S_3$ maps the standard generators of the group into permutations $U$ and $V$ satisfying $UVU = VUV$; it is easy to see that such $U$ and $V$ must be transpositions, either equal or non-commuting. So, there exist only two actions, up to conjugation, of $B_3$ on a set of three elements; this proves the statement in cases \ref{It:GoodCusp} and \ref{It:BadCusp}. \end{proof} Suppose that $C\subset \PP^2$ is a nodal projective curve that is general enough in the sense of Definition~\ref{def:general_curve} and that $L\subset(\PP^2)^*$ is a line transversal to $C^*$. Suppose in addition that $\pi\colon C'\to L$, where $C'$ is a smooth projective curve, is a holomorphic mapping with simple ramification and with branch locus $L\cap C^*$. Choose a point $t_0\in L\setminus C^*$. \begin{corollary}\label{cor:factors=>exists_X} If the fiber monodromy homomorphism $\pi_1(L\setminus C^*,t_0)\to\Perm(\pi^{-1}(t_0))$ factors through the homomorphism $\pi_1(L\setminus C^*,t_0)\to \pi_1(\PP^2\setminus C^*,t_0)$ induced by the embedding $L\hookrightarrow (\PP^2)^*$, then there exists a unique pair $(X,f)$, where $X$ is a normal projective surface and $f\colon X\to (\PP^2)^*$ is a finite holomorphic mapping such that $\left.f\right|_{f^{-1}(L)}=\pi$. Over each node of $C$, there is at most one singular point of $X$, and it must be a Du~Val $A_1$-singularity. Over each cusp of $C^*$, there is at most one singular point of $X$, and it must be a Du~Val $A_2$-singularity. The other points of $X$ are smooth. \end{corollary} \begin{proof} Since the monodromy $\pi_1(L\setminus C^*,t_0)\to\Perm(\pi^{-1}(t_0))$ factors through the homomorphism $\pi_1(L\setminus C^*,t_0)\to \pi_1(\PP^2\setminus C^*,t_0)$, there exists a topological covering $X_0\to (\PP^2)^*\setminus C^*$ of degree $\deg\pi$ extending the covering $\pi^{-1}(L\setminus C^*)\to L\setminus C^*$. The restriction of $f_0$ to the line transversal to $C^*$ has simple ramification, so $f_0$ is a generic covering by Proposition \ref{prop:same_mono}. The rest follows from Proposition~\ref{not.smooth.over.bad} with $D = C^*$. \end{proof} \section{Smooth surfaces simply ramified over duals to general nodal curves}\label{sec:PL} In this section we prove the $(\mathrm b) \Rightarrow (\mathrm a)$ implication in Theorem~\ref{th.main}. Throughout this section we will be working in the following setting. $C\subset\PP^2$ is a nodal curve that is general enough in the sense of Definition~\ref{def:general_curve}, $C^*\subset(\PP^2)^*$ is the dual curve, $X$ is a smooth projective surface, $f\colon X\to(\PP^2)^*$ is a finite holomorphic mapping with simple ramification and branch locus~$C^*$. Recall that the smooth surface $X_C$ and the mapping $f_C\colon X_C\to(\PP^2)^*$ were defined by equations~\eqref{def_of_X_C} and~\eqref{def:f_C}, respectively. For the point $p \in \PP^2$ denote by $\pr_p\colon C \to p^\perp = \PP^1$ the projection of $C$ from $p$. Also denote by $\nu\colon\hatC\to C$ the normalization mapping. Suppose that $\Phi\colon X\to X_C$ is an isomorphism such that $f_C\circ\Phi=f$. \begin{proposition}\label{final_step} Let $p$ be such that the composition $\pr_p\colon C \to p^\perp$ has simple ramification. Then for $C' \mathrel{{:}{=}} f^{-1}(p^\perp)$ there exists an isomorphism $\ph\colon \hatC\to C'$ such that $f \circ \ph=\pr_p\circ\nu$. \end{proposition} \begin{proof} The required isomorphism $\ph$ is just the restriction of $\Phi^{-1}$ to $f_C^{-1}(p^\perp)$. \end{proof} \begin{proposition}\label{Chisini_argument} The $(\mathrm b) \Rightarrow (\mathrm a)$ implication in Theorem~\ref{th.main} holds in each of the following cases. \textup{(1)} $\deg C\ge7$; \textup{(2)} $\deg C=6$ and $C$ has at least one node; \textup{(3)} $\deg C=5$ and $C$ has at least three nodes. \end{proposition} \begin{proof} Theorem 10 from~\cite{Kulikov1999} says that if $C$ satisfies the hypothesis of the proposition, then a finite mapping $f\colon X\to(\PP^2)^*$ with simple ramification over $C^*$ is unique. Thus, the mapping $f\colon X\to (\PP^2)^*$ is equivalent to $f_C\colon X_C\to(\PP^2)^*$, and Proposition~\ref{final_step} applies. \end{proof} The rest of this section is devoted to the proof of the $(\mathrm b) \Rightarrow (\mathrm a)$ implication in Theorem~\ref{th.main} in the remaining cases. This proof is independent of the results of \cite{Kulikov1999} and works for all the cases in Theorem~\ref{th.main}, including those covered by Proposition~\ref{Chisini_argument}. It follows from the hypotheses on $C$ that the dual curve $C^*\subset(\PP^2)^*$ has no inflexion points, all the singularities of $C^*$ are nodes and standard cusps, and no line is tangent to $C^*$ at three points (bitangents of $C^*$ correspond to nodes of~$C$). \begin{proposition}\label{smooth_f^-1} Suppose that $L\subset(\PP^2)^*$ is a line such that the point $L^\perp$ does not lie on $C$. Then the curve $f^{-1}(L)\subset X$ is smooth. \end{proposition} \begin{proof} By virtue of projective duality one has $C=(C^*)^*$. Hence, for a line $L \subset (\PP^2)^*$, one has $L^\perp \in C$ if and only if $L$ either is tangent to $C^*$ at a smooth point, or is tangent to a branch of $C^*$ at a node (we call such $L$ a tangent to $C$ at the node), or is the (limiting) tangent to $C$ at a cusp. In the third case $L^\perp$ is an inflexion point of $C$, in the second case it lies on a double tangent to $C$, and in the first case neither takes place. Now the result follows immediately from cases \ref{It:Smooth}, \ref{It:GoodCusp} of Proposition~\ref{not.smooth.over.bad}, where one puts $D=C^*$. \end{proof} \begin{proposition}\label{singular_f^-1} Suppose that a line $L\subset(\PP^2)^*$ is tangent to $C^*$ (at a smooth point, a node, or a cusp). Then the curve $f^{-1}(L)$ has singular points only over the points of tangency; moreover, there is exactly one singular point of $f^{-1}(L)$ over each point of tangency, and this singular point is a node. \end{proposition} \begin{proof} Since the curve $(C^*)^*=C$ has no cusps, the curve $C^*$ has no inflexion points; now everything follows from Proposition~\ref{not.smooth.over.bad}. \end{proof} Let the line $L\subset(\PP^2)^*$, where $L^\perp\notin C$, vary. Proposition~\ref{smooth_f^-1} shows that the curve $f^{-1}(L)$ is smooth for such $L$, so variation of $L$ induces a monodromy action of $\pi_1(\PP^2 \setminus C)$ on $H^1(f^{-1}(L),\Z)$. More formally, consider the incidence variety \begin{equation}\label{inc.var} I=\{(p,x)\in \PP^2\times X\colon f(x)\in p^\perp\} \end{equation} and denote by $\pr_1: I \to \PP^2$ and $\pr_2: I \to X$ the projections. For each $p\in\PP^2\setminus C$ the fiber $\pr_1^{-1}(p)$ is isomorphic to $f^{-1}(p^\perp)$. It follows from Proposition~\ref{smooth_f^-1} that the derivative of $\pr_1$ has maximal rank everywhere on $\pr_1^{-1}(\PP^2\setminus C)$, whence $\pr_1$ restricts to a locally trivial fibration over the preimage of $\PP^2\setminus C$. So, for a point $q \in \PP^2\setminus C$ the fundamental group $\pi_1(\PP^2\setminus C,q)$ acts on $H^1(f^{-1}(q^\perp),\Z)$; this is the monodromy action in question. \begin{proposition}\label{finite_monodromy} The image of the monodromy action of $\pi_1(\PP^2 \setminus C,q)$ on $H^1(f^{-1}(q^\perp),\Z)$ is cyclic of order dividing $d=\deg C$. \end{proposition} \begin{proof} Since $C$ is a nodal projective curve, the group $\pi_1(\PP^2\setminus C)$ is abelian (see~\cite{Del}), whence it is isomorphic to $H_1(\PP^2\setminus C,\Z)$; the latter is obviously cyclic of order~$d$. \end{proof} Recall the main result of local Picard--Lefschetz theory for Lefschetz pencils with one-dimensional fiber (\cite[Expos\'e XIV, 3.2.5]{SGA7.2} or \cite[\S\S\,5 and~6]{Lamotke}). \begin{proposition}\label{PL} Suppose that \X is a $2$-dimensional complex manifold and $p\colon \X\to\Delta$ is a proper surjective holomorphic mapping with the following properties. \textup{(i)} Over $\Delta^*$, the mapping $p$ has no critical points. \textup{(ii)} In $p^{-1}(0)$, the mapping $p$ has only one critical point~$w$ and the Hessian of $p$ at $w$ is non-degenerate. \textup{(iii)} All fibers of $p$ are connected. Fix a point $z_0\in \Delta^*$ and put $C=p^{-1}(z_0)$ \textup(in view of~\textup{(i)}, $C$ is a compact Riemann surface\textup). Then \textup{(a)} The curve $C_0=p^{-1}(0)$ is smooth everywhere except for a node at~$w$. \textup{(b)} The curve $C$ contains an embedded circle~$S$ such that $C_0$ is homeomorphic to the quotient space $C/S$. \textup{(c)} The monodromy operator on $H^1(C,\Z)$ corresponding to the generator of $\pi_1(\Delta^*)$ is defined by the formula \begin{equation}\label{eq:PL} x\mapsto x-(x,c)c, \end{equation} where $c\in H^1(C,\Z)$ is the Poincar\'e dual of the fundamental class of $S$. \end{proposition} (The circle $S$, as well as its Poincar\'e dual cohomology class, is called \emph{vanishing cycle}.) Put $C_0=f^{-1}(\ell)$, where $\ell\subset (\PP^2)^*$ is a general line, and denote the genus of $C_0$ by~$g$. \begin{proposition}\label{tangent:generalities} Suppose that $L\subset (\PP^2)^*$ is a line such that $L^\perp \in C$. Then the curve $f^{-1}(L)$ is connected, and $L$ can be tangent to $C^*$ only at one or two points. \end{proposition} \begin{proof} The connectedness of $f^{-1}(L)$ (for arbitrary $L$) follows from~\cite[Proposition~1]{FH}. A line $L$ cannot be tangent to $C^*$ at more than two points since all the singularities of the curve $C=(C^*)^*$ are nodes. \end{proof} \begin{proposition}\label{f(-1)(tangent)} If a line $L\subset (\PP^2)^*$ is tangent to $C^*$ at one point, then the curve $f^{-1}(L)$ is homeomorphic to the quotient space $C_0/S$, where $S\subset C_0$ is homeomorphic to the circle and $S$ is homologous to zero in $C_0$. \end{proposition} \begin{proof} Take a point $t \in L \setminus C^*$, so that $L^\perp \in t^\perp \subset \PP^2$, and $t^\perp$ is transversal to $C$. Then consider the incidence variety \begin{equation}\label{eq:Lefschetz_pencil} \tilde X=\{(x, a)\in X\times t^\perp\colon f(x)\in a^\perp\} \end{equation} (the surface $\tilde X$ is just the blow-up of $X$ at $f^{-1}(t)$). The existence of $S$ follows now from Proposition~\ref{PL} applied to the natural projection $\tilde X \to t^\perp$ restricted on the preimage of a small disk $\Delta \subset t^\perp$ centered at $L^\perp$. We are left only to check that the vanishing cycle is zero-homologous. Choosing $z_0\in \Delta\setminus \{L^\perp\}$ and putting $C_0=f^{-1}(L)$, observe that $\Delta \setminus L^\perp \subset \PP^2 \setminus C$, so the monodromy representation \begin{equation}\label{local_mono} \pi_1(\Delta\setminus \{L^\perp\})\to \Aut(H^1(C_0,\Z)) \end{equation} factors through the monodromy representation \begin{equation}\label{global_mono} \pi_1(\PP^2\setminus C)\to \Aut(H^1(C_0,\Z)). \end{equation} The image of the homomorphism~\eqref{global_mono} is finite by Proposition~\ref{finite_monodromy}, whence the image of the generator of $\pi_1(\Delta\setminus\{L^\perp\})$ under the homomorphism~\eqref{local_mono} has finite order~$k$. The $k$th power of the monodromy operator \eqref{eq:PL} is \[ x\mapsto x - k\cdot(x,c)c, \] which is identity if and only if $(c,x)=0$ for all $x\in H^1(C_0,\Z)$. So, $c=0$ and the curve $S$ is homologous to zero. \end{proof} \begin{proposition}\label{unique_section} Suppose that $\deg C>2$ and $L\subset(\PP^2)^*$ is a line tangent to $C^*$ at exactly one point~$q$. Then only the following two cases are possible: \textup{(a)} $f^{-1}(L)=Y_1\cup Y_2$, where $Y_1$ and $Y_2$ are smooth projective curves intersecting transversally at a point lying over~$q$, where the restriction $\left.f\right|_{Y_1}\colon Y_1\to L$ is an isomorphism, the restriction $\left.f\right|_{Y_2}\colon Y_2\to L$ has degree~$>1$, and $(Y_1,Y_1)=0$; \textup{(b)} $C$ is a smooth cubic curve and $f$ is equivalent to a projection of the Veronese surface $v_2(\PP^2)\subset\PP^4$ \textup(in particular, $\deg f=4$\textup). In this case both $Y_1$ and $Y_2$ is a smooth rational curve, and each of the restrictios $\left.f\right|_{Y_1}\colon Y_j\to L$ has degree~$2$. \end{proposition} \begin{proof} Propositions~\ref{f(-1)(tangent)} and~\ref{tangent:generalities} show that $f^{-1}(L)=Y$ is a connected curve homeomorphic to the quotient space $Y'/S$, where $Y'$ is a sphere with handles, and $S\subset Y'$ is a zero-homologous circle. Then $Y'/S$ is homeomorphic to the wedge sum of two spheres with handles, so the curve $Y$ has exactly two components. Since the divisor $Y=Y_1+Y_2$ is the inverse image of the line~$L\subset (\PP^2)^*$ and $f(Y_1)=f(Y_2)=L$ (which follows from the finiteness of the morphism~$f$), one has, for $j=1$ or~$2$, \begin{equation}\label{eq:pos} (Y_j,Y_1+Y_2)=\deg(\left.f\right|_{Y_j})>0. \end{equation} The curves $Y_1$ and $Y_2$ intersect transversally at one point, whence $(Y_1,Y_2)=1$ and therefore \begin{equation}\label{both_non-neg} (Y_1,Y_1)\ge0,\quad (Y_2,Y_2)\ge0. \end{equation} Denote by $V\subset H^2(X,\R)$ the subspace generated by fundamental classes of $Y_1$ and $Y_2$. Consider two cases. \emph{Case 1}: $\dim V=1$. In this case the classes of $Y_1$ and $Y_2$ are proportional: $Y_2 = rY_1$. Observe that the number $r$ must be positive because $(Y_1,f^{-1}L)>0$, $(Y_2,f^{-1}L)>0$ for a general line $L\subset(\PP^2)^*$. Since \[ 1=(Y_1,Y_2)=r(Y_1,Y_1)=r^{-1}(Y_2,Y_2), \] and both $(Y_1,Y_1)$ and $(Y_2,Y_2)$ are integers, one has $r=1$, so that $Y_1 = Y_2$ in cohomology, and $(Y_1,Y_1)=(Y_2,Y_2)=1$. Since homology classes of $C$ and $C'$ are equal and $C+C'=f^{-1}(L)$ is an ample divisor, the divisors $C$ and $C'$ are also ample. Now $\deg f=(Y_1+Y_2,Y_1+Y_2)=4$. Since fundamental classes of the curves $Y_1$ and $Y_2$ coincide, their genera are the same; denote this number by $g$. \begin{lemma}\label{2g=g} The curves $Y_1$ and $f^{-1}(L)$ have the same genus. \end{lemma} \begin{proof} Observe that the surface $\tilde X$ defined by the equation~\eqref{eq:Lefschetz_pencil} with the natural projection $q\colon \tilde X\to t^\perp$ is a Lefschetz pencil. Now a standard argument (see, for example, \cite[Expos\'e XVIII, Theorems 5.6.8 and~5.6.2]{SGA7.2}) shows that image of the natural injection $H^1(X,\C)\hookrightarrow H^1(f^{-1}(L),\C)$, where the line $L\subset(\PP^2)^*$ is transversal to $C^*$ and passes through~$t$, coincides with the invariant subspace \[ H^1(f^{-1}(L))^{\pi_1(L\setminus C^*)}. \] Put $f^{-1}(L)=H$; since the monodromy group is generated by the operators~\eqref{local_mono} and all the vanishing cycles~$c$ are zero in view of Proposition~\ref{f(-1)(tangent)}, we conclude that the restriction $H^1(X,\C)\hookrightarrow H^1(H,\C)$ is an isomorphism, whence the homomorphism $H^1(X,\Ooo_X)\to H^1(H,\Ooo_H)$ from the exact sequence \[ 0\to\Ooo_X(-H)\to\Ooo_X\to\Ooo_H\to 0 \] is also an isomorphism; the Kodairra vanishing theorem and Serre's duality show then that this is equivalent to the equation $\dim \lmod K_X\rmod = \dim \lmod K_X+H\rmod$; since $\dim \lmod H\rmod > 0$, this is equivalent to the relation $\lmod K_X+H\rmod=\varnothing$ (cf.~\cite[Proposition 6.1]{Lvovski}). Since $H=Y_1+Y_2$ and $Y_2$ is an effective divisor, it follows now that the linear system $\lmod K_X+Y_1\rmod$ is also empty; since the divisor $Y_1$ is ample, the above argument in the reverse order shows that $H^1(X,\Ooo_X) \cong H^1(Y_1,\Ooo_{Y_1})$. Thus, genera of the curves $f^{-1}(L)=H$ and $Y_1$ are both equal to $\dim H^1(X,\Ooo_X)$. \end{proof} If a line $L'\subset(\PP^2)^*$ is transversal to $C^*$, then, according to Proposition~\ref{PL}b, $Y_1\cup Y_2$ is homeomorphic to $Y'=f^{-1}(L')$ with a circle contracted to a point. Hence, the genus of $Y'$ equals $2g$. Since $2g=g$ by Lemma~\ref{2g=g}, we infer that $g=0$. Thus, $Y_1$ and $Y_2$ are smooth rational ample curves with self-intersection indices~$1$ on the smooth projective surface~$X$. It is well known (see, for example, \cite[Proposition 2.3]{Lvovski2}) that if a smooth rational curve $Y$ on a smooth projective surface $X$ is ample and $(Y,Y)=1$, then there exists an isomorphism from $X$ to $\PP^2$ mapping $Y$ to a line. Thus, $X\cong \PP^2$ and $\Ooo_X(f^{-1}(L))\cong \Ooo_{\PP^2}(2)$, so the mapping $f\colon X\to\PP^2$ is isomorphic to a projection $\pr_\Lambda\colon v_2(\PP^2)\to\PP^2$, where $v_2(\PP^2)\subset \PP^5$ is the quadratic Veronese surface and $\Lambda\subset\PP^5$ is a $2$-plane disjoint from $v_2(\PP^2)$. Since the curve $C^*$ is the branch locus of the projection $\pr_\Lambda$, its dual curve $C\subset\PP^2$ coincides with the intersection $(v_2(\PP^2))^*\cap\Lambda^\perp$ (cf. the proof of Proposition~\ref{prop:example} below), which is clearly a smooth plane cubic since $(v_2(\PP^2))^*$ is the variety of degenerate symmetric $3\times3$-matrices. \emph{Case 2}: $\dim V = 2$. By the Hodge index theorem, \[ \begin{vmatrix} (Y_1,Y_1)&(Y_1,Y_2)\\ (Y_1,Y_2)&(Y_2,Y_2) \end{vmatrix} <0, \] whence $(Y_1,Y_1)(Y_2,Y_2)<1$. So at least one of the factors must be zero; assume without loss of generality that $(Y_1,Y_1)=0$, then \[ \deg(\left.f\right|_{Y_1})=(Y_1,Y_1+Y_2)=1 \] Thus, the restriction of $f$ to $Y_1$ is an isomorphism from $Y_1$ to the line $L$. Let us prove that degree of the restriction $\left.f\right|_{Y_2}\colon Y_2\to f(Y_2)$ is greater than one. Indeed, $L$ is tangent to $C^*$ at only one point, so $Y$ has only two components. Thus, if $Y_2$ maps with degree~$1$ onto~$L$, then \[ (Y_2,Y_2)=(Y_2,Y_1+Y_2)-(Y_2,Y_1)=1-1=0 \] and $\deg f=(Y_1+Y_1,Y_2+Y_2)=2(Y_1,Y_2)=2$, so all nodes and/or cusps of $C^*$ must be bad in the sense of Definition~\ref{def:bad.cusp}, which contradicts the smoothness of $X$ in view of Proposition~\ref{not.smooth.over.bad}. Thus, $C^*$ must be smooth. Since its dual curve $(C^*)^*=C$ has no cusps, the smooth curve~$C^*$ has no inflexion points. It is well known (and follows, for example, from the Pl\"ucker formulas~\cite[Section 9.1, Theorem 1]{BK}) that the only smooth plane curve without inflexions is the conic, whence $C$ is a conic, which contradicts the hypothesis. \end{proof} \begin{note} The trick with Hodge index theorem is essentially contained in Van de Ven's paper~\cite{VandeVen} (see the proof of Theorem~I). A similar argument allows one to give a proof of the main result of the paper~\cite{Zak} that is valid in arbitrary characteristic (see~\cite{Lvovski2}). \end{note} \begin{proposition}\label{X_C=X} Suppose that $C\subset\PP^2$ is a smooth or nodal curve of degree~$>2$ that is general enough in the sense of Definition~\ref{def:general_curve}. Let also $X$ be a smooth projective surface and $f\colon X\to(\PP^2)^*$ be a finite morphism with branch locus $C^*$ and such that the ramification is simple. If $\deg C=3$, assume in addition that $C$ has a node or $\deg f\ne 4$. Then there exists an isomorphism $\Phi\colon X\to X_C$ such that $f_C\circ\Phi=f$, where $X_C$ and $f_C$ are defined by equations~\eqref{def_of_X_C} and~\eqref{def:f_C}, respectively. \end{proposition} We know two proofs of this proposition, one ``topological'' and one algebraic geometric. For the expositon in this paper, we have chosen the latter since its rigorous version appears to be shorter. \begin{proof} As before, denote the normalization of $C$ by $\nu\colon \hatC\to C$. Let $\gamma\colon \hatC\to C^*$ be the Gauss mapping, which attaches to a point $x\in C$ the tangent to $C^*$ at $\nu(x)$, where by tangent we mean the limiting tangent if $\nu(x)$ is a cusp of $C^*$ and the limiting tangent at the branch corresponding to $x$ if $\nu(x)$ is a node of~$C^*$. If follows from projective duality that the definition of the surface $X_C$ (see~\eqref{def_of_X_C}) may be rewritten as \[ X_C=\{(x,t)\in \hatC\times(\PP^2)^*\mid t\in \gamma(x)^\perp\}. \] Now put \[ \Zcal=\{(x,y)\in \hatC\times X\mid f(y)\in \gamma(x)^\perp\}. \] \begin{lemma}\label{lemma:component} There exists a component $\Zcal_1\subset\Zcal$ consisting of components of fibers that have intersection index $1$ with $D=f^{-1}(L)$ \textup(or, equivalently, project isomorphically onto tangents to~$C^*$; in the proof of Proposition~\ref{unique_section} these components were called~$Y_1$\textup). \end{lemma} \begin{proof}[Proof of the lemma] We will be using the language of schemes. Denote the natural projection $\Zcal\to\hatC$ by $\pi_1$ and the morphism $(x,y)\mapsto f(y)$ by $\pi_2 \colon\Zcal\to(\PP^2)^*$. Pick a line $L\subset(\PP^2)^*$ transversal to $C^*$ and denote by $D\subset\Zcal$ the closed subscheme $\Pi_2^{-1}L$. Proposition~\ref{unique_section} shows that for almost all (in the sense of Zariski topology) closed points $x\in\hatC$, the fiber $\pi_1^{-1}(x)$ consists of two components $Y_1$ and $Y_2$ and $D\cap \pi_1^{-1}(x)$ is a reduced scheme of length $\deg C^*$; moreover, one of the points of $D\cap \pi_1^{-1}(x)$ lies in $Y_1$ and the remaining $\deg C^*-1$ lie in $Y_2$. Let $K$ stand for the function field of \hatC and $\bar K$ for its algebraic closure; put $\eta=\Spec K$, $\bar\eta=\Spec\bar K$, and let $\Zcal_\eta$ and $\Zcal_{\bar\eta}$ be the generic and generic geometric fibers of $\pi_1$ respectively. Then it is apparently evident that $\Zcal_{\bar\eta}$ consists of two components and the pullback of $D$ is also a reduced scheme of length $\deg C^*$ such that one of its points lies on one of the components of $\Zcal_{\bar\eta}$, while the remaining $\deg C^*-1$ points lie on the other component. To prove this assertion rigorously, we invoke EGAIV. To wit, consider the coherent sheaf $\F=\Ooo_\Zcal\oplus\Ooo_D$ on \Zcal. The primary type (``type primaire'') \cite[Remarque 9.8.9]{EGAIV-3} (roughly speaking, it is the collection of lengths of fibers of \F at the generic points of components of its support and at the generic points of various intersections of the closures of these points) of pullbacks of \F to fibers over almost all (in the sense of Zariski topology) closed points is the same; since the set of points with the property ``the pullback of \F to the geometric fiber over the point in question has a given \emph{type primaire}'' is constructible due to~\cite[9.8.9.1]{EGAIV-3}, it follows that \emph{type primaire} of the pullback of \F to $\Zcal_{\bar\eta}$ is the same as that of its pullbacks to almost all fibers over closed points, whence the assertion. Now observe that the Galois group $G=\mathrm{Gal}(\overline K/K)$ acts on the set of the two components of $\Zcal_{\bar\eta}$. Since the pullback of $D$ to $\Zcal_{\bar\eta}$ is $K$-rational, it is $G$-invariant. On the other hand, one of these component contains only one point of $D$, while the other contains $\deg C^*-1>1$ such points (if $\deg C^*-1=1$, then $C^*$ and $C$ are conics, which contradicts the hypothesis). Thus, both these components are $G$-invariant, whence $\Zcal_\eta$ contains a component containing only one point from the pullback of~$D$. This proves the lemma. \end{proof} Returning to the proof of the proposition, observe that self-intersection index of almost all of the fibers of the projection $\Zcal_1\to C^*$ as curves on $X$ is zero due to Proposition~\ref{unique_section}, so they are disjoint. Thus, the natural mapping $\Phi_1\colon\Zcal_1\to X$ (induced by projection to the second factor) is generically one to one; since $X$ is smooth, Zariski's main theorem implies that $\Phi_1$ is an isomorphism. On the other hand, if we define the morphism $\Phi_2\colon \Zcal_1\to X_C$ by the formula~$(x,y)\mapsto (x,f(y))$, then it is easy to see that $\Phi_2$ is also generically bijective, whence isomorphic. Now we may put $\Phi=\Phi_2\circ\Phi_1^{-1}$. \end{proof} Now the $(\mathrm b) \Rightarrow (\mathrm a)$ implication in Theorem~\ref{th.main} follows immediately from Propositions~\ref{X_C=X} and~\ref{final_step}. \begin{note}\label{chisini:example} If $\deg C=3$ and $\deg f=4$, Theorem~\ref{th.main} does not hold. Indeed, if $f\colon v_2(\PP^2)\to\PP^2$ is a general projection of the quadratic Veronese surface, then the branch locus of $f$ is the curve $C^*\subset \PP^2$ that is dual to a smooth cubic~$C$. If $L\subset \PP^2$ is a general line, then the restriction $\pi=\left.f\right|_{f^{-1}(L)}\colon f^{-1}(L)\to L$ is a generic covering of degree $4$ ramified over $L\cap C^*$, that is, over the branch locus of the projection $\pr_{L^\perp}\colon C\to L$. By the very construction, the mapping $\pi$ can be extended to the mapping $f\colon v_2(\PP^2)\to\PP^2$ ramified over~$C^*$, but $\deg f=4\ne3$, so $f$ is not equivalent to a projection of the plane cubic. \end{note} The following corollary of Theorem \ref{th.main} is essentially its ``topological'' reformulation: \begin{corollary}\label{th.main2} Suppose that $C\subset \PP^2$ is a nodal curve of degree~$>2$ that is general enough in the sense of Definition~\ref{def:general_curve}. Suppose that a point $p\in\PP^2\setminus C$ is such that the composition $\pr _p\circ\nu\colon\hatC\to\PP^1$, where $\nu\colon\hatC\to C$ is the normalization and $\pr _p$ is the projection from $p$, has simple ramification. Denote by $L\subset(\PP^2)^*$ the line in the dual plane corresponding to the point $p\in\PP^2$, and suppose that $C'$ is a smooth projective curve and $\pi\colon C'\to L$ is a holomorphic mapping with simple ramification and such that the branch locus of $\pi$ coincides with $L\cap C^*$. If $\deg C=3$, assume in addition that $C$ has a node or $\deg f\ne4$. Then the following two conditions are equivalent. \textup{(a)} There exists an isomorphism $\ph\colon C'\to\hat C$ such that $(\pr _p\circ\nu)\circ \ph=\pi$. \textup{(b)} The covering $\pi^{-1}(L\setminus C^*)\to L\setminus C^*$ can be extended to a covering $X_0\to(\PP^2)^*\setminus C^*$ with respect to which all nodes and cusps of the curve~$C^*$ are good in the sense of Definitions~\ref{def:good.nodes} and~\ref{def:bad.cusp}. \end{corollary} \begin{proof} Immediate from Theorem~\ref{th.main} and Proposition~\ref{not.smooth.over.bad}. \end{proof} Our proof of Theorem~\ref{th.main} implies the following corollary: \begin{corollary}\label{chisini_duals} Suppose that $C\subset\PP^2$ is a nodal curve that is general enough in the sense of Definition~\ref{def:general_curve}. If $C$ is not a smooth cubic, then any two generic coverings $X\to(\PP^2)^*$, ramified over $C^*$ and with smooth $X$, are equivalent. \end{corollary} If $\deg C \ge 7$, or $\deg C = 6$ and $C$ has at least one node, or $\deg C = 5$ and $C$ has at least three nodes, then this corollary is implied by \cite[Theorem~10]{Kulikov1999}. It is well known (see, for example,~\cite{Kulikov1999}) that this assertion is wrong if $C$ is a smooth cubic. A~relevant counterexample is essentially contained in our proof of Proposition~\ref{unique_section} (Case~1). \section{The smoothness condition is (sometimes) essential}\label{sec:example} In this section we show that the smoothness condition in Theorem \ref{th.main} is not trivial, at least for the case of smooth curves of degree $d=3$ or $4$ and mappings of degree~$d$. To wit, we prove the following. \begin{proposition}\label{prop:example} Suppose that $d=3$ or $4$. Then there exist a smooth curve $C\subset\PP^2$, $\deg C=d$, a line $p^\perp\subset(\PP^2)^*$ (where $p \in \PP^2$ is a point) transversal to~$C$, a smooth projective curve $C'$, and a holomorphic mapping $\pi\colon C'\to p^\perp$ with simple ramification such that $\deg\pi=d$ and $\pi$ is ramified exactly over $C^*\cap p^\perp$ with the following properties. \textup{(1)} For a point $t_0\in p^\perp \setminus C^*$, the fiber monodromy homomorphism $\pi_1(p^\perp\setminus C^*,t_0)\to \Perm (\pi^{-1}(t_0))$ can be factored through the homomorphism $\pi_1(p^\perp\setminus C^*,t_0)\to \pi_1((\PP^2)^*\setminus C^*,t_0)$ induced by the embedding $p^\perp \setminus C^*\hookrightarrow (\PP^2)^*\setminus C^*$. \textup{(2)} The mapping $\pi\colon C'\to p^\perp$ is \emph{not} equivalent to the projection $\pr _p\colon C\to p^\perp$. \end{proposition} We will need the following lemma. \begin{lemma}\label{self-dual:3,4} If $d=3$ or $4$, then there exists a surface with isolated singularities $X\subset\PP^3$, $\deg X=d$, such that its dual $X^*\subset(\PP^3)^*$ is isomorphic to $X$ and the general hyperplane section of $X$ is general enough in the sense of Definition~\ref{def:general_curve}. \end{lemma} \begin{proof}[Proof of the proposition modulo Lemma~\ref{self-dual:3,4}] Let $X \subset \PP^3$ be a surface from Lemma~\ref{self-dual:3,4}. For a general point $r\in\PP^3$, the projection $\pr_r\colon X\to \PP^2$, where $\PP^2$ is a projective plane (incidentally, $\PP=(r^\perp)^*$), is simply ramified. If $B\subset\PP^2$ is the branch locus of $\pr_r$, then a line $\ell\subset\PP^2$ is tangent to $B$ if and only if the plane $\Pi_r$ spanned by $r$ and $\ell$ is tangent to $X$. Thus, the dual curve $B^*\subset(\PP^2)^*$ is projectively isomorphic to $X^*\cap r^\perp =C$; if $p$ is general, Lemma~\ref{self-dual:3,4} asserts that $C$ is a general enough smooth plane curve. Take $p \in \PP^3$ general and let $C'=X\cap\Pi_p$; denote the restriction of $\pr_p$ to~$C'$ by $\pi\colon C'\to L$. We see (taking into account that $B=C^*$) that $\pi$ extends to a finite morphism $\pr_p\colon X\to (\PP^2)^*$ ramified over $C^*$, where the surface $X$ is singular. Thus, $\pi$ satisfies Condition (b) of Theorem \ref{th.main} with the smoothness omitted. On the other hand, $\pi$ cannot be equivalent to the projection $\pr_r$. Indeed, if such an equivalence existed, then by Proposition ~\ref{gen(a)=>(b)} the restriction of $\pi$ to the preimage of the complement of the branch locus could be extended to a finite morphism $X_C\to\PP^2$ ramified over $C^*$ with smooth $X_C$. However, if such an extension to a finite mapping $X\to(\PP^2)^*$ ramified over $C^*$ and with normal (in particular, smooth) $X$ exists, this extension is unique by the theorem of Grauert and Remmert~\cite[Expos\'e~XII, Th\'eor\`eme 5.4]{SGA1} we cited before. Since we already have such an extension with singular $X$, we arrive at the desired contradiction. \end{proof} It remains to prove Lemma~\ref{self-dual:3,4}. If $d=3$, we let $X$ be the projective closure of surface in $\mathbb A^3$ defined by the equation $x_1x_2x_3=1$ (this surface was considered in~\cite{ACT}). A direct computation shows that $X$ contains only three singular points (of the type $A_2$) and that $X$ is projectively isomorphic to~$X^*$. Since any smooth plane cubic is automatically general enough in the sense of Definition~\ref{def:general_curve}, the surface $X$ is as required. To treat the case $d=4$ we need to recall the definition and some properties of Kummer surfaces. Suppose that $J$ is a principally polarized Abelian variety of dimension~$2$ that is not the product of two elliptic curves and $\Theta\subset J$ is a theta divisor of polarization. Then the complete linear system $\lmod 2\Theta\rmod$ is base point free and defines a finite morphism $J\to\PP^3$; the image of this morphism is a surface of degree $4$ with exactly $16$ singular points of the type $A_1$; the surfaces obtained by this construction are called Kummer surfaces (see, for example, \cite[Section 10.3]{Dolgachev}). Any Kummer surface $X\subset\PP^3$ is projectively isomorphic to its dual $X^*\subset(\PP^3)^*$ (see \cite[Theorem 10.3.19]{Dolgachev}). Thus, to finish the proof of Lemma~\ref{self-dual:3,4}, it remains to show that there exists a Kummer surface such that its general plane section is general enough in the sense of Definition~\ref{def:general_curve}; since such sections are plane quartics, it suffices to establish that a general plane section has only simple inflexion points. The following lemma and its proof were pointed out to us by Rita Pardini. \begin{lemma}[R. Pardini]\label{Pardini} Each non-hyperelliptic curve of genus $3$ \textup(smooth and projective\textup) is isomorphic to a plane section of an appropriate Kummer surface in~$\PP^3$. \end{lemma} \begin{proof}[Proof of the lemma] Suppose that $C$ is such a curve, and let $\sigma\colon C_1\to C$ be an unramified covering of degree~$2$. Denote by $(J,\Theta)$ the Prym variety corresponding to this covering~\cite[Section 12.2]{BiLa}. According to \cite[Proposition~12.5a]{BiLa}, the Abel--Prym map embeds $C'$ in $J$ as an element of the linear system~$2\Theta$. Multiplication by $-1$ on $J$ restricts on $C_1$ to the involution induced by $\sigma$, hence the image of $C_1$ via the map given by $\lmod 2\Theta\rmod$ is a plane section of the Kummer surface associated to $(J,\Theta)$, and this section is isomorphic to~$C$. \end{proof} Now if we pick a smooth plane curve $C_0\subset\PP^2$ of degree $4$ such that all its Weierstrass points have index one, or, equivalently, all its inflexion points are simple, Lemma~\ref{Pardini} shows that $C_0$ is a plane section of a Kummer surface $X\subset\PP^3$; since the property ``all the Weierstrass points have index one'' is open, a generic plane section of $X$ has only simple inflexion points, which completes the proof of Lemma~\ref{self-dual:3,4}. \bibliographystyle{amsalpha}
\section{Introduction} The Ising model needs no introduction, being perhaps the most studied example, since its formulation by Lenz~\cite{Len20}, of a system undergoing a phase transition. The transition it exhibits was found to be of rather broad relevance, though of course its features do not exhaust the range of possible behaviors in statistical mechanics. The model has provided the testing ground for a large variety of techniques, which partially compensate for the lack of exactly solvable models above two dimensions. Our goal here is to present a tool for addressing the question of continuity of the model's phase transition. Doing so, we advance the technique of the model's \emph{random current} representation which was developed in \cite{Aiz82} starting from the Griffiths-Hurst-Sherman switching lemma (which was earlier used in~\cite{GHS70} for the GHS inequality). In this representation the onset of the Ising model's symmetry breaking is presented as a percolation transition in a system of random currents with constrained sources. The perspective that this picture offers has already shown itself to be of value in yielding a range of results for the model's critical behavior (c.f.~\cite{Aiz82,AF86,ABF87,Sak07}). The incremental step taken here is to consider directly the limiting shift invariant infinite systems of random currents. This allows to add to the available tools arguments based on the `uniqueness of infinite cluster' principle, which is of relevance to the question of continuity of the state at the model's critical temperature. Among the specific cases for which the results presented here answer a long open question is the continuity at the critical point of the spontaneous magnetization of the standard three dimensional Ising model. To highlight this it may be noted that what is proven here for Ising spin systems is not valid for the broader class of the $Q$-state Potts models, with counter-examples existing for $Q>4$ (Ising case corresponding to $Q=2$). \subsection{Notation} We focus here on the $d$-dimensional version of the ferromagnetic Ising model, on the transitive graph ${\mathbb Z}^d$. More generally, the model may be formulated on a graph, whose vertex set and edge set we denote by $\mathbb{G}$ and $E\subset \mathbb{G}^2$ correspondingly. Associated with the sites are $\pm1$ valued spin variables, whose configuration is denoted $\sigma=(\sigma_x:x\in \mathbb{G})$. For a general ferromagnetic pair interaction, the system's Hamiltonian defined for finite subsets $\Lambda \subset \mathbb{G}$ and boundary conditions $\tau \in \{-1,0,1\}^ {\mathbb{G} \backslash \Lambda}$ is given by the function \be \label{eq:H} H_\Lambda^\tau(\sigma)~:=~ -\sum_{x\in \Lambda} h \sigma_x \ -\ \sum_{\{x,y\}\subset \Lambda: x\neq y}J_{x,y}\sigma_x\sigma_y \ -\ \sum_{x\in \Lambda: y \in \mathbb{G} \backslash \Lambda}J_{x,y}\sigma_x\tau_y \,, \end{equation} for any $\sigma\in\{-1,1\}^{\Lambda}$, where $(J_{x,y})_{x,y\in\mathbb{Z}^d}$ is a family of nonnegative {\em coupling constants}, and $h$ is the magnetic field. For $\beta \in (0, \infty)$, finite volume Gibbs states with boundary conditions $\tau$ are given by probability measures on the spaces of configurations in finite subsets $\Lambda \subset \mathbb{G}$ under which the expected values of functions $f:\{-1,1\}^{\Lambda}\rightarrow \mathbb{R}$ are $$\langle f\rangle_{\Lambda,\beta,h}^\tau=\sum_{\sigma\in\{-1,1\}^{\Lambda}}f(\sigma)\frac{e^{-\beta H^\tau_\Lambda(\sigma)}}{Z^\tau(\Lambda,\beta,h)}\, ,$$ where the sum is normalized by the partition function $Z^\tau(\Lambda,\beta,h)$ so that $\langle 1\rangle_{\Lambda,\beta,h}^\tau=1$. Of particular interest is the following pair of boundary conditions (b.c.): \begin{itemize} [noitemsep,nolistsep] \item {\em free b.c.}: $\tau_x = 0 $ for all $x\in \mathbb{G}\backslash \Lambda$ (or alternatively, the last term in \eqref{eq:H} is omitted). \item {\em plus b.c.}: $\tau_x = 1$ for all $x\in \mathbb{G} \backslash \Lambda$. \end{itemize} The corresponding measures, or expectation value functionals, are denoted $ \langle \cdots \rangle_{\Lambda,\beta,h}^0$ and $ \langle \cdots \rangle_{\Lambda,\beta,h}^+$ (with ``$ \cdots $'' a place holder for functions of the spin configurations). For each of these two boundary conditions, the finite volume Gibbs states are known to converge to the corresponding infinite-volume Gibbs measures (for a discussion of the concept see e.g. \cite{Georgii}). By default the volume subscript will be omitted when it refers to the full graph, i.e. $\Lambda = \mathbb{G}$. For simplicity we focus here on the prototypical example of $\mathbb{G}= \mathbb{Z}^d$, and interactions which are: \vspace{-.4cm} \begin{enumerate} [noitemsep] \item[\bf C1] translation invariant: $J_{x,y}=J_{0,y-x}$, \item[\bf C2] ferromagnetic: $J_{x,y}\ge 0$, \item [\bf C3] locally finite: $|J|:=\sum_{x\in\mathbb{Z}^d}J_{0,x}<\infty$. \item [\bf C4] aperiodic: for any $x\in \mathbb{Z}^d$, there exist $0=x_0,x_1,\dots,x_{m-1},x_m=x$ such that $J_{x_0,x_1}J_{x_1,x_2},\dots , J_{x_{m-1},x_m}>0$. \end{enumerate} The last condition is benign since under {\bf C1} the lattice can be divided into sub-lattices with {\bf C4} holding for each sub-lattice, and the arguments presented below can be adapted to such setup \footnote{Among the essential features of the graphs $\mathbb{Z}^d$ is their transitivity and sub-exponential growth. Our arguments can be extended, however not to graphs of positive Cheeger constant, such as regular trees and more generally nonamenable Cayley graphs. Nor are the statements proven below valid at such generality (the percolation aspect of this distinction is discussed in \cite{SNP00} and references therein).}. \\ Of particular interest is the model's phase transition, which in the $(\beta,h)$ plane occurs along the $h=0$ line and is reflected in the nonvanishing of the symmetry breaking order parameter: \be m^*(\beta) \ := \ \langle\sigma_0\rangle_{\beta}^+ \, . \end{equation} For temperatures ($T\equiv \beta^{-1}$) at which $m^*(\beta) >0$, the mean magnetization at nonzero magnetic field $h$ changes discontinuously at $h=0$. The discontinuity is symptomatic of the co-existence of two distinct Gibbs equilibrium states : \begin{eqnarray} \langle \cdots \rangle_{\beta}^+ &= & \lim_{h\searrow 0} \langle \cdots \rangle_{\beta,h} \notag \\ \mbox{} \\ \langle \cdots \rangle_{\beta}^- &= & \lim_{h\nearrow 0} \langle \cdots \rangle_{\beta,h} \notag \end{eqnarray} which carry the residual magnetizations: \be \langle \sigma_0 \rangle_{\beta}^\pm \ = \ \pm m^*(\beta) \, , \end{equation} with $m^*(\beta)$ customarily referred to as the {\em spontaneous magnetization}. \\ Property $\bf C3$ guarantees that at small $\beta$ (in particular, for $\beta< |J|^{-1}$, see e.g.~\cite{Fisher,Dob70,Aiz82}) $m^*(\beta)=0$, and there is no symmetry breaking. However, in dimensions $d>1$ each such model exhibits a phase transition at some $\beta_c \in (|J|^{-1} , \infty)$, with $ m^*(\beta)>0$ for $\beta> \beta_c$~\cite{Pei36}. For $d=1$ such a transition occurs if $J_{x-y} \ge 1/|x-y|^\alpha$ with $\alpha \in (1,2)$ \cite{Dys69} and also for the boundary value $\alpha = 2$~\cite{ACCN88}, in which case $m^*(\beta)$ is discontinuous at $\beta_c$~\cite{Tho69,ACCN88}). \\ \subsection{The continuity question} The main result presented here addresses the following two related questions concerning the continuity of the correlation functions at this transition: \begin{enumerate} \item[Q1:] Is $m^*(\beta)$ continuous at $\beta_c$, or equivalently is\footnote{Being the limit of a decreasing sequence of (finite volume) continuous functions, $m^*(\beta)$ is upper semicontinuous, and hence $m^*(\beta) = \lim_{\varepsilon \searrow 0} \ m^*(\beta + \varepsilon)$ for all $\beta \ge 0$ \cite{LML72}.}: \be \label{eq:cont_m} \lim_{\beta \searrow \beta_c} \ m^*(\beta) \ = \ 0 \, ? \end{equation} \item[Q2:] Is the Gibbs state $ \langle \cdots \rangle^+_\beta$ continuous at $\beta_c$? \end{enumerate} By the arguments of \cite{Leb77, Leb72} for ferromagnetic Ising models in the class discussed here the answers to the two questions is the same. (For completeness the argument is summarized here in Appendix~\ref{app:continuity}.) As it is often the case, questions which at the level of Statistical Mechanics concern continuity of Gibbs states have Thermodynamic level manifestation in terms of differentiability properties of the \emph{pressure} \be \label{eq:pressure} P(\beta, h) \ := \ \lim_{L \to \infty} \frac{1} {|\Lambda_L|} \log Z^\tau(\Lambda_L,\beta,h), \end{equation} where $\Lambda_L=[-L,L]^d$ (with $\frac{-1}{\beta} P(\beta, h)$ also referred to as the \emph{free energy}). On general grounds, the limit \eqref{eq:pressure} is known to: i) exist, ii) be independent of the boundary conditions, and iii) yield a jointly convex function of $(\beta, \beta h) $. For ferromagnetic Ising models it is known that for $h\neq0$ the function $P(\beta, h) $ is analytic in $\beta$ and $h$~\cite{LeeYang} and thus the continuity questions are limited to the line $h=0$. J.L. Lebowitz~\cite{Leb77} proved that for any $\beta$ the differentiability of the free energy with respect to $\beta$ (in the class of models described above) is equivalent to the continuity of the expectation values of even spin functions (i.e. functions which are invariant under global spin flip) averaged over any of the translation invariant Gibbs states, or also equivalently to the relation \be \label{eq:f=pm} \langle \cdots \rangle^0_{\beta_c} \ = \ \frac{1}{2} \left[ \langle \cdots \rangle^+_{\beta_c} + \langle \cdots \rangle^-_{\beta_c} \right], \end{equation} where $ \langle \cdots \rangle^-_{\beta_c}$ is the Gibbs measure with {\em minus b.c.}, or simply the image of $ \langle \cdots \rangle^+_{\beta_c}$ under global spin flip. Furthermore, it was pointed out there that by convexity the differentiability condition can fail at only a countable set of $\beta$. For finite range models T.~Bodineau~\cite{Bod06} reduced that to the one point set consisting of just the critical point $\beta_c$, and furthermore showed that the magnetization is also continuous at all $\beta\neq \beta_c$. (Some further related results are mentioned below.) It is generally expected that for all Ising models of the kind described above ([{\bf C1 - C4}]) equation \eqref{eq:f=pm} is met also at $\beta_c$. When condition \eqref{eq:f=pm} holds that it is not due to most general thermodynamic reason, as it fails for the ferromagnetic nearest-neighbor Potts models with $Q$ large enough: $Q>4$ in $d=2$ dimensions (see \cite{Bax73} for exact results, and \cite{KS,LMR,DST13,DC_pf} for partial rigorous results) and $Q>2$ for $d\ge3$ (see \cite{KS,BCC} for partial mathematical results in this direction). The failure of \eqref{eq:f=pm} implies that \eqref{eq:cont_m} also fails, i.e $m^*(\beta_c) \neq 0$ (as is explained in Appendix~\ref{app:continuity}). However the converse does not seem to be tautologically true, as is indicated\footnote{It is expected, but not proven, that in this borderline case \eqref{eq:f=pm} is satisfied, while $m^*(\beta_c)\neq 0$.} by the special case of the one-dimensional model with $J_{x,y} = 1/|x-y|^2$, c.f. \cite{Tho69,ACCN88}). \subsection{Statement of the main results} Relevant to the continuity of the spontaneous magnetization is the Long Range Order (LRO) parameter which is defined by: \be M_{LRO}(\beta)^2 \ := \ \lim_{n\to \infty} \frac{1}{|\Lambda_n|} \sum_{x\in \Lambda_n} \langle\sigma_0\sigma_x\rangle_{\beta}^0 \, , \end{equation} where $\Lambda_n=[-n,n]^d$, for the model on ${\mathbb Z}^d$ (the limit existing by monotonicity arguments), or the LRO parameter's variant \be \widetilde M_{LRO}(\beta)^2 \ := \ \inf_{B\subset {\mathbb Z}^d, |B|<\infty} \frac{1}{|B|^2} \sum_{x,y \in B} \langle \sigma_x \sigma_y \rangle_{\beta}^0 \ \equiv \ \inf_{B\subset {\mathbb Z}^d, |B|<\infty} \left \langle \left[\frac{1}{|B|} \sum_{x\in B} \sigma_x \right]^2 \right \rangle_{\beta}^0 \, \end{equation} which satisfies \be \inf_{x\in \mathbb{Z}^d} \langle \sigma_0 \sigma_x \rangle^0_\beta \ \le \ \widetilde M_{LRO}(\beta)^2 \ \le \ M_{LRO}(\beta)^2 \,. \end{equation} It may be noted that whereas $m^*(\beta_c)$ provides direct information about the states at $\beta > \beta_c$, the monotonicity arguments of~\cite{Leb77} imply that $M_{LRO}(\beta_c)$ provides direct information about the states at $\beta < \beta_c$ and, furthermore, the following relation holds. \begin{proposition} For any translation invariant ferromagnetic Ising model on ${\mathbb Z}^d$: at all $\beta \ge 0$ \be M_{LRO}(\beta) \ \le \ m^*(\beta) \end{equation} with equality holding at values of $\beta$ at which $P(\beta,0)$ is continuously differentiable. \end{proposition} Our main general result is: \begin{theorem}\label{thm:continuity} For any ferromagnetic Ising model on $\mathbb{Z}^d$ whose coupling constants $(J_{x,y})_{x,y\in\mathbb{Z}^d}$ satisfy the conditions {\bf C1-C4}: if \be \label{eq:cont_cond} \widetilde M_{LRO}(\beta) \ = \ 0 \ \end{equation} then also \be m^*(\beta_c) \ =\ 0 \, , \end{equation} and the system has only one Gibbs state at $\beta_c$. \end{theorem} \noindent (In which case one may add that the Gibbs states $\langle \cdot \rangle^{\#}$ are continuous in $\beta$ at $\beta_c$, regardless of the choice of the boundary conditions, see Appendix~\ref{app:continuity}.) \noindent{\bf Remarks:} \begin{enumerate} \item[i.] The reference to $ \widetilde M_{LRO}(\beta)$ in \eqref{eq:cont_cond} instead of $ M_{LRO}(\beta)$ allows to base the estimate on the spin-spin correlations along just one of the principal axes. Kaufman and Onsager found such simplification helpful in calculations, by which they showed that $\widetilde M_{LRO}(\beta_c)=0$ for the {\it nearest neighbor} Ising model in two dimensions~\cite{KO49}. \item[ii.] Theorem~\ref{thm:continuity} carries interesting implication also for the case that the magnetization is discontinuous at $\beta_c$ (as in the above mentioned $1D$ model with the borderline $1/r^2$ interaction). It shows that even in such case the Ising model does not display coexistence at $\beta_c$ of two distinct phases: a `high temperature phase' with rapid decay of correlations and a `low temperature phase' exhibiting spontaneous magnetization, $m^*(\beta_c)>0$. Such phase coexistence would be characteristic of a `regular' first order phase transitions, as is found in the afore mentioned Q-state Potts models at $Q>4$. It may be of relevance to note here that it has already been known that for the Ising model the transition is `continuous' in the sense that the magnetic susceptibility $\chi(\beta) := \sum_x \langle \sigma_0 \sigma_x \rangle_{\beta} $, which is finite for all $\beta < \beta_c$~(\cite{ABF87}), diverges as $\beta \nearrow \beta_c$~(by the argument of \cite{GJ74}). That is now strengthened to the observation that $m^*(\beta_c)>0$ requires also $\lim_{\beta \nearrow \beta_c} M_{LRO} \ > \ 0$. \item[iii.] For the specific example of the one-dimensional Ising model with $J_{x,y}=1/|x-y|^2$ , for which $m^*(\beta_c)>0$ (\cite{Tho69,ACCN88}), Theorem~\ref{thm:continuity} combined with the monotonicity of the two point function (\cite{Sch,MessMSole}) implies that for all $x\in {\mathbb Z}^d$ \be \lim_{\beta \nearrow \beta_c} \langle\sigma_0\sigma_x\rangle_{\beta}^0 \ > \ m^*(\beta_c)^2 \ > \ 0 \,. \end{equation} \end{enumerate} \medskip Theorem~\ref{thm:continuity} is proven here by extending the random current representation of~\cite{Aiz82}, which is based on the switching lemma of~\cite{GHS70}, to infinite domains and establishing uniqueness of the infinite cluster for the resulting system of (duplicated) random currents on $\mathbb{Z}^d$. This provides a useful tool for studying the implications of condition \eqref{eq:cont_cond}. \subsection{Applications to reflection positive models} Among the few available tools for establishing the validity of \eqref{eq:cont_cond} for nonsolvable models, and the only one which applies in the intermediate but important dimension $d=3$, is the Gaussian domination bound of \cite{FSS76,FILS} which applies to reflection positive interactions (see e.g. \cite{Bis09} for a review on this crucial notion). For the Ising model on a torus, where due to the period boundary condition the correlation function $F_{L,\beta}(x,y) = \langle \sigma_x\sigma_y\rangle_{\mathbb{T}_L,\beta} $ depends only on $x-y$, we denote the Fourier transform by \be \widehat F_{L,\beta}(p) \ :=\ \sum_{x\in \mathbb{T}_L}e^{i p \cdot x}F_{L,\beta}(0,x) \end{equation} where $p$ ranges over $\mathbb{T}_L^\star=\left( \frac{2\pi}{L} {\mathbb Z} \right)^d \cap (-\pi,\pi]^d$, and $ u\cdot v $ denotes the scalar product between $u$ and $v$. \begin{proposition} [Gaussian domination bound~\cite{FSS76,FILS}] If the interaction $J_{x,y}$ is reflection positive, then or any $p\in\mathbb{T}_L^\star\setminus\{0\}$: \be \label{eq:89} \widehat F_{L,\beta}(p)\le \frac{1}{2\beta E(p)} \, , \\[2ex] \end{equation} where \be E(p)\ :=\ \sum_{x\in\mathbb{Z}^d}\left(1-e^{i p\cdot x} \right)J_{0,x} \ = \ 2 \sum_{x\in\mathbb{Z}^d} \sin^2\big(\tfrac{p\cdot x}{2}\big) \, J_{0,x} \, . \end{equation} \end{proposition} Here $E(p)$ is the energy function of modes of momentum $p$. The bound \eqref{eq:89} allows to prove that \eqref{eq:cont_cond} holds for reflection-positive models provided \be \label{eq:Ep_condition} \int_{[-\pi,\pi]^d}\frac{{\rm d}p}{(2\pi)^d}\frac{1}{E(p)}\ < \ \infty ~. \end{equation} \\ The relation \eqref{eq:Ep_condition} is also the condition for transience of the random-walk associated with the weights $(J_{x,y})_{x,y\in\mathbb{Z}^d}$, which is the Markov process defined by the transition probabilities \be \mathbb{P}(X_{n+1}=y|X_n=x)=\frac{J_{x,y}}{\sum_{z\in \mathbb{Z}^d}J_{x,z} }\quad\text{ for all }x,y\in\mathbb{Z}^d. \end{equation} This yields the following conclusion (derived below in Section~\ref{sec:3}): \begin{corollary}\label{cor:main} If the random-walk associated to $(J_{x,y})_{x,y\in\mathbb{Z}^d}$ is transient and the model is reflection-positive, then the magnetization of the Ising model is continuous at $\beta_c$. \end{corollary} For specific applications, let us quote from~\cite{FSS76,FILS} (c.f. also \cite{AF86}) that the following Ising models are reflection-positive (with respect to hyperplanes passing through vertices): \vspace{-.2cm} \begin{enumerate} \item (nearest neighbor interactions) $J_{x,y} = \delta_{\|x-y\|_1,1}$\, , \item (exponential decay) \hspace{1.8cm} $J_{x,y}=\exp(-\mu \|x-y\|_1)$ for $\mu>0$\, , \item (power-law potentials) \hspace{1.2 cm} $J_{x,y}=\|x-y\|_1^{-\alpha}$ for $\alpha > d$\, , \\ \end{enumerate} where $\|x\|_1=\sum_{i=1}^d|x_i|$ for $x=(x_1,\dots,x_d)$. Furthermore, Ising models whose couplings are linear combinations with positive coefficients of the couplings mentioned above are also reflection-positive, and in one dimension the existence of a phase transition requires both $\sum_{n\in \mathbb N} J_n = \infty$ and $\sum_{n\in \mathbb N} n \, J_n < \infty$, which corresponds to $\alpha \in (1, 2]$ (\cite{Dys69,ACCN88}).\\ In the above examples $E(p)$ vanishing for $p \to 0$ at the rates: $E(p)\approx |p|^2$ in cases (1) and (2), and $E(p) \approx |p|^{\min\{2,\alpha-d\}}$ in case (3). Thus, verifying the condition \eqref{eq:Ep_condition} we conclude: \begin{corollary}\label{example} The magnetization of the following models is continuous as a function of $\beta$ for \vspace{-.4cm} \begin{itemize}[noitemsep] \item any reflection positive ferromagnetic Ising model in $d>2$ dimensions, \item any such one dimensional model whose interaction includes long range term(s) with $1< \alpha <2$, and no other powers. \end{itemize} \end{corollary} It may be added that while \eqref{eq:Ep_condition} is not satisfied in two dimensions, Theorem~\ref{thm:continuity} is of relevance also for this case, since the condition \eqref{eq:cont_cond} can be established for the nearest neighbor model on the square, hexagonal and triangular lattices through the Fortuin-Kasteleyn random cluster representation (see \cite{Gri06} Theorem 6.72). \\ The planar case shows that transience of the associated random walk (namely \eqref{eq:Ep_condition}) is not necessary for the continuity of $m^*(\beta)$. On the other hand, for one-dimensional long range models with $J_{x,y} = 1/|x-y|^\alpha$ the criterion provided by \eqref{eq:Ep_condition} is sharp, since it holds just up to the value $\alpha=2$ at which the spontaneous magnetization is known to be discontinuous at $\beta_c$. \\ \subsection{Past results on continuity at $\beta_c$} The past results on the continuity of the spontaneous magnetization at $\beta_c$ are naturally split into two distinct classes: {\it i.} the special low dimensional case of $d=2$, and {\it ii.} high dimensions (as described below). For the standard nearest neighbor model, the only dimension which has been left out is the one of seemingly most physical interest: $d=3$. Thus, aside from the general principle, for the nearest-neighbor case the novelty in our results is mainly limited to that case (plus certain long range models that fall between these two ranges). The earliest results have been derived for the nearest neighbor model in $d=2$ dimensions, for which the spontaneous magnetization was computed by Yang~\cite{Yan52}, using the methods of~\cite{Ons44,K49}. Prior to that, Kaufman and Onsager~\cite{KO49} showed that $\widetilde M_{LRO}(\beta_c)=0$. More recently, continuity of $m^*(\beta)$ at $\beta_c$ was given a short proof in~\cite{Wer09} and a proof using discrete holomorphicity and the Russo-Seymour-Welsh theory was exposed in \cite{DHN10}. For high dimensions, the continuity of spontaneous magnetization in the nearest neighbor model was established in~\cite{AF86} for $d\ge 4$ (formally $d > 3 \frac{1}{2}$) through reflection positivity bounds combined with differential inequalities. The method used there yields also information on the critical exponent $\delta$ with which $m^*(\beta) \approx |\beta_c-\beta|_{-}^{1/\delta}$, and related results for reflection positive long range interactions due to which the effective dimension (as expressed through the `bubble diagram') is lowered. \\ Results which do not require reflection positivity were derived through the general method of `lace expansion'~\cite{Sl06} whose adaptation to Ising systems' random currents was accomplished by A. Sakai~\cite{Sak07}. The lace expansion is restricted to models above the upper critical dimension, $d> d_c$, which in the presence of long range interaction (with $\alpha > d$) is lowered to $d_c = \min \{ 4, 2(\alpha -d)\}$ \cite{HHS08} (see also \cite{CS12} for tight estimates on the two-point function). For optimal dimensional dependence of the results, the lace expansion requires also a sufficiently spread-out short range interaction (possibly for only technical reasons). However this method allows to deduce mean field behavior of $m^*(\beta)$ in high dimensions for a collection of non-solvable, and not necessarily reflection positive, models. \\ \section{The Random Current representation and its percolation properties} We shall use the following notation when discussing dependent percolation on the graph ${\mathbb Z}^d$. \paragraph{Notation.} For any subset $G \subset {\mathbb Z}^d$, we let $\mathcal{P}_2(G)=\{\{x,y\}:x,y\in G\}$. For a configuration of ``bond variables'' $\omega\in\{0,1\}^{\mathcal{P}_2(\mathbb{Z}^d)}$, an edge $\{x,y\}$ for which $\omega_{x,y}=1$ is said to be {\em open}, and otherwise it is {\it closed}. Two vertices $x$ and $y$ are said to be {\em connected} if there exist $x=x_0,\dots,x_m=y$ such that $\omega_{x_i,x_{i+1}}=1$ for every $0\le i<m$. The statement that $x$ and $y$ are connected is denoted by $x\lr{\omega} y$ (and when $\omega$ is deemed clear from the context, we drop it from the notation). The sites of $\mathbb{Z}^d$ are partitioned into maximal connected components of $\omega$, which are called {\em clusters}. We will often encounter also integer valued bond functions, i.e. elements $\omega \in \{0,1,2,...\}^{\mathcal{P}_2(G)} \ =: \ \Omega_G$. The associated percolation would refer to the projection $\widehat\cdot: \Omega_G \longrightarrow\{0,1\}^{\mathcal{P}_2(G)}$ defined by $$\widehat{n}_{x,y}=\begin{cases}1&\text{ if ${n}_{x,y}>0$,}\\ 0&\text{ otherwise.}\end{cases} $$ The `lattice shifts', by vectors $x\in \mathbb{Z}^d$, of configurations $\omega\in\{0,1\}^{\mathcal{P}_2(\mathbb{Z}^d)}$, or $\omega \in \Omega_{\mathbb{Z}^d}$, are the mappings $\tau_x$ defined by $\tau_x(\omega)_{a,b}=\omega_{a+x,b+x}$ for all $a,b\in\mathbb{Z}^d$.\\ The indicator function on a configuration space $\Omega$ corresponding to a condition $E$ will be denoted by $\indf{E}\equiv \indf{E}(\omega)$. The argument $(\omega)$ will be omitted when its deemed to be clear within the context. \\ \subsection{The random current representation} \begin{definition} A {\em current} $\mathbf{n}$ on $G\subset \mathbb{Z}^d$ (also called a current configuration) is a function from $\mathcal{P}_2(G)$ to $\{0,1,2,...\}$. A {\em source} of $\mathbf{n}=(\mathbf{n}_{x,y}:\{x,y\}\in \mathcal{P}_2(G))$ is a vertex $x$ for which $\sum_{y\in G}{\mathbf{n}}_{x,y}$ is odd. The set of sources of $\mathbf{n}$ is denoted by $\partial\mathbf{n}$, and the collection of current configurations on $G$ is $\Omega_G$. \end{definition} \subsubsection*{Random current representation for free boundary conditions} The partition function of a finite graph $G$ is: \begin{equation}\label{eq:4}Z^0(G,\beta)=\sum_{\sigma\in\{-1,1\}^G}\prod_{\{x,y\}\subset G}e^{\beta J_{x,y}\sigma_x\sigma_y}.\end{equation} Expanding $e^{\beta J_{x,y}\sigma_x\sigma_y}$ for each $\{x,y\}$ into $$e^{\beta J_{x,y}\sigma_x\sigma_y}=\sum_{{\mathbf{n}}_{x,y}=0}^\infty\frac{(\sigma_x\sigma_y)^{{\mathbf{n}}_{x,y}} (\beta J_{x,y})^{{\mathbf{n}}_{x,y}}}{{\mathbf{n}}_{x,y}!}$$ and substituting this relation in \eqref{eq:4}, one gets \begin{equation*} Z^0(G,\beta)=\sum_{\mathbf{n}\in\Omega_G}w_\beta(\mathbf{n})\sum_{\sigma\in\{-1,1\}^G}\prod_{x\in G}\sigma_x^{\sum_{y\in G} {\mathbf{n}}_{x,y}}, \end{equation*} where $$w_\beta(\mathbf{n}):=\prod_{\{x,y\}\subset G}\frac{(\beta J_{x,y})^{{\mathbf{n}}_{x,y}}}{{\mathbf{n}}_{x,y}!}.$$ Now, $$\sum_{\sigma\in\{-1,1\}^G}\prod_{x\in G}\sigma_x^{\sum_{y\in G} {\mathbf{n}}_{x,y}}=\begin{cases}0 &\text{ if $\sum_{y\in G} {\mathbf{n}}_{x,y}$ is odd for some $x\in G$,}\\ 2^{|G|}&\text{ otherwise.}\end{cases}$$ Above, $|G|$ denotes the number of sites of $G$. Thus, the definition of a current's source enables one to write \be Z^0(G,\beta)=2^{|G|}\sum_{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\emptyset}w_\beta(\mathbf{n}) ~. \end{equation} Similar expansions for the correlation functions involve currents with sources. For instance, \begin{equation*}\sum_{\sigma\in\{-1,1\}^G}\sigma_x\sigma_ye^{-\beta H^0_G(\sigma)}=2^{|G|}\sum_{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\{x,y\}}w_\beta(\mathbf{n}),\end{equation*} which gives \begin{equation}\label{eq:7}\langle\sigma_x\sigma_y\rangle_{G,\beta}^0=\frac{\displaystyle\sum_{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\{x,y\}}w_\beta(\mathbf{n})}{\displaystyle\sum_{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\emptyset}w_\beta(\mathbf{n})}.\end{equation} \subsubsection*{Random current representation for $+$ boundary conditions} For a finite subset $G\subset {\mathbb Z}^d$, the equilibrium state at $+$ boundary conditions is obtained by freezing all spins in the complementary set to the value $+1$. This may be conveniently represented by adding an additional vertex $\delta\notin \mathbb{Z}^d$, to which we refer as the {\em ghost spin} site\footnote{The notion is related to Griffith's {\em ghost spin}, which was added by R.B. Griffiths \cite{Grif_ghost} as a tool for the extension of correlation inequalities to states under an external field.}, and setting the coupling between it and sites $x\in G$ to $J_{x,\delta}=J_{x,\delta}(G):=\sum_{y\notin G}J_{x,y}$. For notational convenience we adapt the convention that the ghost site is not to be listed in the configurations source set $\partial \mathbf{n}$ regardless of the parity of the flux into $\delta$ (which can be determined from the parity of $|\partial \mathbf{n}|$).\\ For $+$ boundary conditions, a development similar to the above yields: \begin{equation}\label{eq:8}\langle\sigma_x\sigma_y\rangle_{G,\beta}^+=\frac{\displaystyle\sum_{\mathbf{n}\in\Omega_{G\cup\{\delta\}}:\,\partial\mathbf{n}=\{x,y\}}w_\beta(\mathbf{n})}{\displaystyle\sum_{\mathbf{n}\in\Omega_{G\cup\{\delta\}}:\,\partial\mathbf{n}=\emptyset}w_\beta(\mathbf{n})}\, . \end{equation} Observe that \eqref{eq:7} differs from \eqref{eq:8} in that the summation is over all currents on $G\cup\{\delta\}$ instead of $G$. Also note that $J_{x,\delta}$ depends on $G$. \\ \subsubsection*{Switching lemma} As mentioned in the introduction, the {\it random current} perspective on the Ising model's phase transition is driven by the observation that the onset of long range order coincides with a percolation transition in a dual system of currents. This point of view, as an intuitive guide to diagrammatic bounds which under certain conditions provide `hard information' on the critical model's scaling limits, was developed in \cite{Aiz82} and a number of subsequent works. Among the first tools which facilitate cancellations in this representation is the following graph-theoretic switching lemma, which was originally introduced in \cite{GHS70} and applied there for the Griffiths-Hurst-Sherman (GHS) inequality \begin{lemma}[Switching lemma\footnote{In allowing $G$ to be a strict subset of $ H$, Lemma~\ref{switching} forms a minor extension of the statement found in \cite{GHS70}. An allusion to it, and its other applications, was made in the explanation of Lemma 6.3 in \cite{AG83}, where this extension was applied. The difference in the proof is rather trivial. }] \label{switching} For any nested pair of finite sets $G\subset H$, pair of sites $x,y \in G$ and $A\subset H$, and a function $F:\Omega_H\rightarrow \mathbb{R}$: \begin{align*}\sum_{\substack{\mathbf{n}_1\in\Omega_G:\,\partial \mathbf{n}_1=\{x,y\}\\ \mathbf{n}_2\in\Omega_H:\,\partial \mathbf{n}_2=A }}&F(\mathbf{n}_1+\mathbf{n}_2)w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)\\ &=\sum_{\substack{\mathbf{n}_1\in\Omega_G:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_H:\,\partial \mathbf{n}_2=A\Delta\{x,y\}}}F(\mathbf{n}_1+\mathbf{n}_2)w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)\indf{ x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\mathrm{\ in\ }G }.\end{align*} with $A\Delta B$ denoting the symmetric difference $(A\setminus B)\cup(B\setminus A)$ between $A$ and $B$. \end{lemma} The essential graph-theoretic argument is given in \cite{GHS70}, and in the random current notation which is employed below in \cite{Aiz82}. Since the proof provides an introduction to the notation let us repeat it here. \begin{proof} In the argument, a current on the subgraph corresponding to $G$ is also viewed as a current on $H$ which vanishes on pairs $\{ x,y \}$ not contained in $G$. The switching is performed within collections of pairs of currents $\{ \mathbf{n}_1,\, \mathbf{n}_2 \}$ of a specified value for the sum $\mathbf{m}:=\mathbf{n}_1+\mathbf{n}_2$. It is therefore convenient to take as the summation variables the current pairs $\mathbf{m}$ and $\mathbf{n}=\mathbf{n}_1 \le \mathbf{m}$ (with $\mathbf{n} \le \mathbf{m}$ defined as the natural {\it partial order} relation). One obtains \begin{align*}\sum_{\substack{\mathbf{n}_1\in\Omega_G:\,\partial \mathbf{n}_1=\{x,y\}\\ \mathbf{n}_2\in\Omega_H:\,\partial \mathbf{n}_2=A }}&F(\mathbf{n}_1+\mathbf{n}_2)\, w_\beta(\mathbf{n}_1) \, w_\beta(\mathbf{n}_2)\\ &=\sum_{\substack{\mathbf{m}\in\Omega_H:\,\partial \mathbf{m}=A\Delta\{x,y\}}} F(\mathbf{m}) \, w_\beta(\mathbf{m})\sum_{\substack{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\{x,y\}\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}},\end{align*} and \begin{align*} \sum_{\substack{\mathbf{n}_1\in\Omega_G:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_H:\,\partial \mathbf{n}_2=A\Delta\{x,y\} }}&F(\mathbf{n}_1+\mathbf{n}_2) \, w_\beta(\mathbf{n}_1) \, w_\beta(\mathbf{n}_2)\, \indf { x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\text{ in }G }\\ &=\sum_{\substack{\mathbf{m}\in\Omega_H:\,\partial \mathbf{m}=A\Delta\{x,y\}}}F(\mathbf{m}) \, w_\beta(\mathbf{m}) \, \indf{x\lr{\widehat\mathbf{m}}y\text{ in }G }\sum_{\substack{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\emptyset\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}},\end{align*} where $\binom{\mathbf{m}}{\mathbf{n}}=\prod_{\{x,y\}\subset G}\binom{\mathbf{m}_{x,y}}{{\mathbf{n}}_{x,y}}$ and where we used the fact that $$w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)=\prod_{\{x,y\}\subset G\cup\{\delta\}}\left[\frac{(\beta J_{x,y})^{{\mathbf{n}}_{x,y}}}{{\mathbf{n}}_{x,y}!}\right]\, \left[\frac{(\beta J_{x,y})^{\mathbf{m}_{x,y}}}{\mathbf{m}_{x,y}!}\right]\ =\ w_\beta(\mathbf{m})\binom{\mathbf{m}}{\mathbf{n}}.$$ The claim follows if the relation below is proved for every current $\mathbf{m}\in\Omega_H$: \begin{equation}\label{eq:10} \sum_{\substack{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\{x,y\}\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}}\ = \ \indf{ x\lr{\widehat\mathbf{m}}y\text{ in }G }\sum_{\substack{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\emptyset\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}}. \end{equation} First, assume that $x$ and $y$ are not connected in $G$ by $\mathbf{m}$. The right-hand side is trivially zero. Moreover, there is no current $\mathbf{n}$ on $G$ which is smaller than $\mathbf{m}$ and which connects $x$ to $y$. The left-hand side is thus 0 and \eqref{eq:10} is proved in this case. Let us now assume that $x$ and $y$ are connected in $G$ by $\mathbf{m}$. Associate to $\mathbf{m}$ the graph $\mathcal{M}$ with vertex set $G$, and $\mathbf{m}_{a,b}$ edges between $a$ and $b$. For a subgraph $\mathcal{N}$ of $\mathcal{M}$, let $\partial\mathcal{N}$ be the set of vertices belonging to an odd number of edges. Since $x$ and $y$ are connected in $G$ by $\mathbf{m}$, there exists a subgraph $\mathcal{K}$ of $\mathcal{M}$ with $\partial\mathcal{K}=\{x,y\}$. The involution $\mathcal{N}\mapsto \mathcal{N}\Delta\mathcal{K}$ provides a bijection between the set of subgraphs of $\mathcal{M}$ with $\partial\mathcal{N}=\emptyset$, and the set of subgraphs of $\mathcal{M}$ with $\partial\mathcal{N}=\{x,y\}$. Therefore, these two sets have the same cardinality. Since the summations in $$\sum_{\substack{\mathbf{n}\in\Omega_G:\,\partial\mathbf{n}=\emptyset\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}}\quad\text{and}\quad\sum_{\substack{n\in\Omega_G:\,\partial\mathbf{n}=\{x,y\}\\ \mathbf{n}\le\mathbf{m}}}\binom{\mathbf{m}}{\mathbf{n}}$$ are over currents $\mathbf{n}$ in $G$, these sums correspond to the cardinality of the two sets mentioned above. In particular, they are equal and the statement follows. \end{proof} \subsection{Infinite-volume random current representation} Next, we formulate the infinite volume limit of the random current representation. This allows a more effective use of the asymptotic translation invariance, and enables us to deploy the uniqueness of the infinite cluster argument, which can be established in this context. Let ${\rm P}^0_{G,\beta}$ be the law on currents on $G$ defined by \be {\rm P}^0_{G,\beta}[\mathbf{n}]:=\frac{w_\beta(\mathbf{n}) \, \, \indf{\partial\mathbf{n}=\emptyset} }{\displaystyle\sum_{\mathbf{m}\in\Omega_G:\,\partial\mathbf{m}=\emptyset}w_\beta(\mathbf{m})}\quad, \qquad \forall \mathbf{n}\in\Omega_G~. \end{equation} This induces a measure on $\Omega_G:=\{0,1,2,...\}^{\mathcal{P}_2(G)}$ that we denote by $\widehat {\rm P}^0_{G,\beta}$. One may also define a law on currents on $G\cup\{\delta\}$ which induces a measure on $\Omega_{G\cup\{\delta\}} :=\{0,1,2,...\}^{\mathcal{P}_2(G\cup\{\delta\})}$ denoted by $\widehat {\rm P}^+_{G,\beta}$. Let $\Lambda_L=[-L,L]^d$ be the box of size $L$. \begin{theorem}\label{def:current infinite} Let $\beta>0$. There exist two laws $\widehat{\rm P}^+_\beta$ and $\widehat{\rm P}^0_\beta$ on $\Omega_{\mathbb{Z}^d}$ such that \begin{itemize} \item[\rm \bf R1] (Convergence) For any event $\mathcal{A}$ depending on finitely many edges, $$\lim_{L\rightarrow \infty}\widehat{\rm P}^+_{\Lambda_L,\beta}[\mathcal{A}]=\widehat{\rm P}^+_\beta[\mathcal{A}]\quad\text{and}\quad \lim_{L\rightarrow \infty}\widehat{\rm P}^0_{\Lambda_L,\beta}[\mathcal{A}]=\widehat{\rm P}^0_\beta[\mathcal{A}].$$ \item[\rm \bf R2] (Invariance under translations) $\widehat{\rm P}^+_\beta$ and $\widehat{\rm P}^0_\beta$ are invariant under the shifts $\tau_x$, $x\in \mathbb{Z}^d$. \item[\rm \bf R3] (Ergodicity) $\widehat{\rm P}^+_\beta$ and $\widehat{\rm P}^0_\beta$ are ergodic with respect to the group of shifts $(\tau_x)_{x\in\mathbb{Z}^d}$. \end{itemize}\end{theorem} \begin{proof} Except for a minor difference in the very last step the proof is identical for the $(+)$ and the free $(f)$ boundary conditions. Let us therefore use the symbol $\#$ as a marker for either of the two. \paragraph{Proof of R1} (Convergence) To prove convergence of the finite volume probability measures, let us first note that the distribution of the random currents simplifies into a product measure when conditioned on the parity variables ${\bf r}(\omega) = ({\bf r}_{x,y})_{x,y \subset G}$, with: \be {\bf r}_{x,y} (\omega)\ := \ (-1)^{{\bf n}_{x,y} (\omega)} \, . \end{equation} The conditional distribution of $\mathbf{n}$, given ${\bf r}(\omega)$, is simply the product measure of independent Poisson processes of mean values $\beta J_{x,y}$ conditioned on the corresponding parity. Thus, for a proof of convergence it suffices to establish convergence of the law of the parity variables ${\bf r}(\omega)$. For a set of bonds (i.e. graph edges) $E\subset\mathcal{P}_2(\mathbb{Z}^d)$, define the events \begin{eqnarray} \mathcal{C}_E & =& \left \{ \ \omega \ : \ {\bf r}_{x,y}(\omega) = 1 \quad \forall \{x,y\} \in E \ \right \} \, , \\ \mathcal{C}^{(0)}_E & =& \left \{ \ \omega \ : \ {\bf n}_{x,y}(\omega) = 0 \quad \forall \{x,y\} \in E \ \right \}\, . \notag \end{eqnarray} Let us prove that for any finite subset $E$ of edges of $\mathbb{Z}^d$, $\widehat{\rm P}^\#_{\Lambda_L,\beta}[\mathcal{C}_E]$ converges as $L$ tends to infinity. To facilitate a unified treatment of the two boundary conditions we denote \be \Omega^\#_L \ = \ \begin{cases} \Omega_{\Lambda_L }& \mbox{for $\# = 0$} \, , \\[2ex] \Omega_{\Lambda_L\cup\{\delta\} }& \mbox{for $\# = +$} \, . \end{cases} \, \end{equation} For $L$ large enough (so that $\Lambda_L \supset E$) we have: \begin{eqnarray} \label{eq:deletion} \widehat{\rm P}^\#_{\Lambda_L,\beta}[\mathcal{C}_E] \ & = & \frac{\displaystyle\sum_{\mathbf{n}\in\Omega^\#_L:\,\partial \mathbf{n}=\emptyset}w_\beta(\mathbf{n})\, \indf{\mathcal{C}_E}}{\displaystyle\sum_{\mathbf{n}\in\Omega^\#_L:\,\partial \mathbf{n}=\emptyset}w_\beta(\mathbf{n}) } \notag \\ & = & \frac{\displaystyle\sum_{\mathbf{n}\in\Omega^\#_L:\,\partial \mathbf{n}=\emptyset}w_\beta(\mathbf{n})\, \indf{\mathcal{C}^{(0)}_E}}{\displaystyle\sum_{\mathbf{n}\in\Omega^\#_L:\,\partial \mathbf{n}=\emptyset}w_\beta(\mathbf{n}) } \, \prod_{x,y\in E} \cosh (\beta J_{x,y}) \notag \\[2ex] & = & \frac{ Z^\#(\Lambda_L \setminus E, \beta) } { Z^\#(\Lambda_L,\beta)} \, \prod_{x,y\in E} \cosh (\beta J_{x,y}) \, \end{eqnarray} Above, $\Lambda_L\setminus E$ designates the graph obtained by removing the edges of $E$ but keeping all the vertices of $\Lambda_L$. The above ratio can be expressed in terms of an expectation value of a finite term: \be \label{eq:ProbE} \widehat{\rm P}^\#_{\Lambda_L,\beta}[\mathcal{C}_E] \ = \ \big \langle e^{ -\beta K_E } \big \rangle^\#_{\Lambda_L, \beta} \prod_{x,y\in E} \cosh (\beta J_{x,y})\end{equation} with the finite volume collection of energy terms \be \label{eq:K} K_E(\omega) := \sum_{x,y\in E} J_{x,y} \sigma_x \sigma_y \, . \end{equation} The convergence of the above expression follows now directly from the convergence of correlation functions as $L$ tends to infinity. The events $\mathcal{C}_E$ with $E$ ranging over finite sets of edges span (by inclusion-exclusion) the algebra of events expressible in terms of finite collections of the binary variables of ${\bf r}(\omega)$. This fact, and the above observation that the probability distribution of the random current $\mathbf{n}$ conditioned on ${\bf r}(\omega)$ does not depend on $L$, implies the existence of $\widehat{\rm P}^\#_\beta$. \paragraph{Proof of R2} (Translation invariance) Fix $x\in \mathbb{Z}^d$. The limit of the probability of the event $\mathcal{C}_E$, where $E$ is a finite set of edges, is the same if the sequence $(\Lambda_L)_{L\ge0}$ is replaced by the sequence $(x+\Lambda_L)_{L\ge 0}$. (Simply use \eqref{eq:ProbE} and the convergence of $\langle \cdots \rangle_{{x+\Lambda_L},\beta}^\#$ to $\langle \cdots \rangle_{\beta}^\#$.) This immediately implies that $\widehat{\rm P}^\#_{\beta}$ is invariant under translations. \paragraph{Proof of R3} (Ergodicity) Since every translationally invariant event can be approximated by events depending on a finite number of edges, it is sufficient to prove that for any events $A$ and $B$ depending on a finite number of edges, \be \label{eq:ergcond} \lim_{\|x\|_1\rightarrow \infty}\widehat{\rm P}^\#_\beta[A\cap \tau_xB]\ =\ \widehat{\rm P}^\#_\beta[A]\ \widehat{\rm P}^\#_\beta[B]. \end{equation} In view of the conditional independence of $\mathbf{n}$ given the parity variables ${\bf r}$, the requirement can be further simplify to the proof that for any two finite sets $E$ and $F$ of edges, \be \lim_{\|x\|_1\rightarrow\infty}\widehat{\rm P}^\#_\beta[\mathcal{C}_{E\cup(x+F)}]\ =\ \widehat{\rm P}^\#_\beta[\mathcal{C}_E] \ \widehat{\rm P}^\#_\beta[\mathcal{C}_F] ~. \end{equation} Using the expression \eqref{eq:ProbE}, for $x$ large enough so that $E\cap (x+F) =\emptyset$: \begin{eqnarray} \label{eq:ProbE2} \frac { \widehat{\rm P}^\#_{\beta}[\mathcal{C}_{E\cup(x+F)}] } {\widehat{\rm P}^\#_\beta[\mathcal{C}_E] \ \widehat{\rm P}^\#_\beta[\mathcal{C}_F] } & = & \frac {\big \langle e^{ -\beta K_E } e^{ -\beta K_{x+F} } \big \rangle^\#_{ \beta } } { \big \langle e^{-\beta K_E } \big \rangle^\#_{\beta } \ \big \langle e^{-\beta K_F } \big \rangle^\#_{\beta } } \end{eqnarray} Ergodicity of the random current states can therefore be presented as an implication of the statement that this ratio tends to 1. This condition holds as a consequence of the mixing property of the states $ \big \langle \cdots \big \rangle^{\#}_\beta$ when restricted to functions which are invariant under global spin flip (i.e. that $f(-\sigma)=f(\sigma)$ for every spin configuration $\sigma$). For completeness we enclose the proof of the statement, which may be part of the folklore among experts, in Appendix A. \end{proof} \begin{remark} The relation of the ergodicity of $\widehat{\rm P}_\beta^+$ and $\widehat{\rm P}_\beta^0$ to the partial ergodicity of $\langle \cdots \rangle_\beta^+$ or $\langle \cdots \rangle_\beta^0$ (i.e. ergodicity of only the restriction to the $\sigma$ algebra of even event) can be compared to a similar relation in the random-cluster representation of the $Q$-state Potts models, with wired and free boundary conditions (see \cite{Gri06} for more details on these models). The restriction is needed since in the presence of symmetry breaking (at $\beta>\beta_c$) the state $\langle \cdots \rangle_\beta^0$ is not even mixing on functions which are odd with respect to the global spin flip. \end{remark} \subsection{Percolation properties of the sum of random currents} Define $\mathbb P_\beta$ to be the law of $\widehat{\mathbf{n}_1+\mathbf{n}_2}$, where $\mathbf{n}_1$ and $\mathbf{n}_2$ are two independent currents with laws ${\rm P}^0_{\beta}$ and ${\rm P}^+_{\beta}$. We also set $\mathbb E_\beta$ for the expectation with respect to $\mathbb P_\beta$. Properties {\bf R2} and {\bf R3} of Theorem~\ref{def:current infinite} imply immediately that $\mathbb P_\beta$ is invariant and ergodic with respect to shifts. We now prove that there cannot be more than one infinite cluster. This claim will be crucial in the proof of Theorem~\ref{thm:continuity}: it will replace the use of the FKG inequality, which is not available for the random current representation. \begin{theorem}\label{thm:percolation} For any translation invariant ferromagnetic Ising model on ${\mathbb Z}^d$ satisfying $\mathbf{C1}-\mathbf{C4}$, there exists at most one infinite cluster $\mathbb P_\beta$-almost surely (at any $\beta \ge 0$). \end{theorem} The following lemma is an equivalent of the insertion tolerance valid for many spin models. \begin{lemma}\label{lem:insertion} Let $\widehat\Phi_N: \{0,1\}^{\mathcal{P}_2(\mathbb{Z}^d)}\longrightarrow \{0,1\}^{\mathcal{P}_2(\mathbb{Z}^d)}$ be the map opening all edges $\{x,y\}$ in $\Lambda_{N}$ with $J_{x,y}>0$. Let $N>0$, then there exists $c=c(N,J,\beta)>0$ such that for any event $\mathcal{E}$, $$\mathbb P_\beta[\widehat\Phi_N(\mathcal{E})] \ \ge\ c\ \mathbb P_\beta[\mathcal{E}].$$ \end{lemma} \begin{proof} It is sufficient to consider events $\mathcal{E}$ depending on a finite number of edges. Let $\mathbb P_{\Lambda_n,\beta}$ be the law of $\widehat{\mathbf{n}_1+\mathbf{n}_2}$, where $\mathbf{n}_1$ and $\mathbf{n}_2$ are two independent currents with respective laws ${\rm P}_{\Lambda_n,\beta}^0$ and ${\rm P}_{\Lambda_n,\beta}^+$. Property {\bf R1} of Theorem~\ref{def:current infinite} shows that $\mathbb P_{\Lambda_n,\beta}$ converges weakly to $\mathbb P_\beta$. This reduces the proof to showing the existence of $c=c(N,J,\beta)>0$ on $\Lambda_n$, with a value which does not depend on $n>N$. Consider the transformation $\Phi_N:(\Omega_{\Lambda_n})^2\longrightarrow(\Omega_{\Lambda_n})^2$ defined by $$\Phi_N(\mathbf{n}_1,\mathbf{n}_2)\{x,y\}=\begin{cases}(0,2)&\text{ if $(\mathbf{n}_1\{x,y\},\mathbf{n}_2\{x,y\})=(0,0)$,} \\ &\text{ $J_{x,y}>0$, and $x,y\in\Lambda_{N}$,}\\ (\mathbf{n}_1\{x,y\},\mathbf{n}_2\{x,y\})&\text{ otherwise},\end{cases}$$ where exceptionally $\mathbf{m}\{x,y\}$ denotes $\mathbf{m}_{\{x,y\}}$ for ease of notation. Fix $\omega\in\{0,1\}^{\Lambda_n}$ and let $\Omega_2=\{(\mathbf{n}_1,\mathbf{n}_2)\in(\Omega_{\Lambda_n})^2:\widehat{\mathbf{n}_1+\mathbf{n}_2}=\omega\}$. The set $\Phi_N(\Omega_2)$ is obtained from $\Omega_2$ by changing the value of the current $\mathbf{n}_2$ on edges $\{x,y\}$ with $\omega_{\{x,y\}}=0$ from 0 to 2. Therefore, \begin{multline}\mathbb P_{\Lambda_n,\beta}[\widehat\Phi_N(\mathcal{E})] \ =\ \sum_{\omega'\in\widehat\Phi(\mathcal{E})}\mathbb P_{\Lambda_n,\beta}[\omega']=\sum_{\omega\in\mathcal{E}}\frac{1}{{\rm Card}[\widehat\Phi_N^{-1}(\widehat\Phi_N(\omega))]}\mathbb P_{\Lambda_n,\beta}[\widehat\Phi_N(\omega)] \\[2ex] \ge \ 2^{-{\rm Card}(\mathcal{P}_2(\Lambda_N))}\sum_{\omega\in\mathcal{E}}\mathbb P_{\Lambda_n,\beta}[\widehat\Phi_N(\omega)] \ =\ 2^{-{\rm Card}(\mathcal{P}_2(\Lambda_N))}\sum_{\omega\in\mathcal{E}}{\rm P}_{\Lambda_n,\beta}^0\otimes{\rm P}_{\Lambda_n,\beta}^+[\Phi_N(\Omega_2)] \\[2ex] \ge\ 2^{-{\rm Card}(\mathcal{P}_2(\Lambda_N))}\sum_{\omega\in\mathcal{E}}\Big(\prod_{\{x,y\}\subset\Lambda_N:\,\omega_{x,y}=0}\frac{(\beta J_{x,y})^2}{2}\Big){\rm P}_{\Lambda_n,\beta}^0\otimes{\rm P}_{\Lambda_n,\beta}^+[\Omega_2] \\[2ex] \mbox{ } \qquad \quad \ge \ c\ \sum_{\omega\in\mathcal{E}}{\rm P}_{\Lambda_n,\beta}^0\otimes{\rm P}_{\Lambda_n,\beta}^+[\Omega_2]\ =\ c\ \mathbb P_{\Lambda_n,\beta}[\mathcal{E}] \, , \hfill \end{multline} where $c=c(N,J,\beta)>0$ does not depend on $n$. In the first inequality, we used the fact that the number of pre-images of each configuration is smaller than 2 to the power the number of pairs of points in $\Lambda_{N}$ (since one has to decide whether edges of $\Lambda_N$ were open or closed before the transformation). \end{proof} \begin{proof}[Proof of Theorem~\ref{thm:percolation}] For $\ell\in\mathbb{N}\cup\{\infty\}$, let $\mathcal{E}_\ell$ be the event that there exist exactly $\ell$ disjoint infinite clusters. We must prove that $\mathbb P_\beta[\mathcal{E}_\ell]=0$ for $\ell\ge 2$. The proof is based on a variation of the Burton-Keane argument \cite{BK89}. Let $k>0$ such that for any vertex $y$ satisfying $\|y\|_1=1$, there exist $0=x_0,\dots,x_m=y$ with $J_{x_0,x_1}\dots J_{x_{m-1},x_m}>0$ and with $x_i\in\Lambda_k$ for every $i\le m$ (the existence of this $k$ is guaranteed by the aperiodicity condition). \paragraph{Proof of $\mathbb P_\beta[\mathcal{E}_\ell]=0$ for $2\le \ell<\infty$.} Let $\ell\ge 2$. Let $\mathcal{F}_n$ be the event that the $\ell$ infinite clusters intersect $\Lambda_n$. Fix $N>k$ large enough so that $\mathbb P_\beta[\mathcal{F}_N]\ge\tfrac12\mathbb P_\beta[\mathcal{E}_\ell]$. Lemma~\ref{lem:insertion} implies that $\mathbb P_\beta[\widehat\Phi_{2N}(\mathcal{F}_N)]\ge \tfrac c2\mathbb P_\beta[\mathcal{E}_\ell]$. Any configuration in $\widehat\Phi_{2N}(\mathcal{F}_N)$ contains exactly one infinite cluster since all the vertices in $\Lambda_N$ are connected. Therefore, $$\mathbb P_\beta[\mathcal{E}_1]\ge \tfrac c2\mathbb P_\beta[\mathcal{E}_\ell].$$ Ergodicity implies that $\mathbb P_\beta[\mathcal{E}_\ell]$ and $\mathbb P_\beta[\mathcal{E}_1]$ are equal to $0$ or $1$, therefore $\mathbb P_\beta[\mathcal{E}_\ell]=0$. \paragraph{Proof of $\mathbb P_\beta[\mathcal{E}_\infty]=0$.} Assume that $\mathbb P_\beta[\mathcal{E}_\infty]>0$ and consider $N>2k$ large enough so that $$\mathbb P_\beta[\text{three distinct infinite clusters intersect the box $\Lambda_{N/2}$}]>0.$$ Lemma~\ref{lem:insertion} (applied to $\widehat\Phi_{N}$) implies that $\mathbb P_\beta[{\rm CT}_0]>0$, where ${\rm CT}_0$ is the following event: \begin{itemize}[noitemsep] \item all vertices in $\Lambda_{N/2}$ are connected to each other in $\Lambda_N$, \item If $\mathcal{C}$ is the cluster of 0, then $\mathcal{C}\cap(\mathbb{Z}^d\setminus\Lambda_N)$ contains at least three distinct infinite connected components.\end{itemize} A vertex $x\in (2N+1)\mathbb{Z}^d$ is called a coarse-trifurcation if $\tau_x{\rm CT}_0=:{\rm CT}_x$ occurs. By invariance under translation, $\mathbb P_\beta[{\rm CT}_x]=\mathbb P_\beta[{\rm CT}_0]$. Fix $n\gg N$. The set $T$ of {\em coarse-trifurcations} in $\Lambda_n$ has a natural structure of forest $\mathcal{F}$ constructed inductively as follows. \begin{itemize} \item[Step 1] At time 0, all the vertices in $T$ are {\em unexplored}. \item[Step 2] If there does not exist any unexplored vertex in $T$ left, the algorithm terminates. Otherwise, pick an unexplored vertex $t\in T$ and mark it explored (by this we mean that it is not considered as an unexplored vertex anymore). Go to Step 3. \item[Step 3] Consider the cluster $\mathcal{C}_t$ of vertices $x\in \mathbb{Z}^d$ connected to $t$ in $\Lambda_n$. This cluster decomposes into $k\ge 3$ disjoint connected components of $\mathbb{Z}^d\setminus (t+\Lambda_N)$ denoted $\mathcal{C}_t^{(1)},\dots,\mathcal{C}_t^{(k)}$. For $i=1,\dots, k$, do the following: \begin{itemize} \item if there exist two vertices $x\in\mathcal{C}_t^{(i)}$ and $y\notin \Lambda_n$ such that $\{x,y\}$ is open, and there exists an open path in $\Lambda_n$ going from $x$ to $t$ and not passing at distance $N$ from a coarse-trifurcation in $\mathcal{C}_t^{(i)}\cap (T\setminus\{t\})$, then add the vertex $y$ to $\mathcal{F}$ together with the edge $\{t,y\}$. \item if there is no such vertex, then there must be a coarse-trifurcation $s\in\mathcal{C}_t^{(i)}$ connected by an open path not passing at distance $N$ from a trifurcation in $\mathcal{C}_t^{(i)}\cap (T\setminus\{t,s\})$. If $s$ is not already a vertex of $\mathcal{F}$, add it. Then, add the edge $\{t,s\}$. \end{itemize} \item[Step 4] Go to Step 2. \end{itemize} The graph obtained is a forest (due to the structure of coarse-trifurcations). Each coarse-trifurcation corresponds to a vertex of the forest of degree at least three. Thus, the number of coarse-trifurcations must be smaller than the number of leaves. Let $N$ be the number of leaves, we find \begin{equation}\label{eq:198}\mathbb P_\beta[{\rm CT}_0]\frac{(2n+1)^d}{(2N+1)^d}\le \mathbb E_\beta[N].\end{equation} Yet, leaves are vertices outside $\Lambda_n$ which are connected by an open edge to a vertex in $\Lambda_n$, therefore $$\mathbb E_\beta[N]\le 2d(2n+1)^{d-1}\sum_{k=0}^n \sum_{x \in \mathbb{Z}^d: \, \|x\|_1 \ge k}J_{0,x} ~.$$ Since $|J|<\infty$, we find that $$0<\frac {\mathbb P_\beta[{\rm CT}_0]}{(2N+1)^d}\le \frac{\mathbb E_\beta[N]}{(2n+1)^d}\longrightarrow 0\quad\text{as $n\rightarrow \infty$. }$$ This contradicts $\mathbb P_\beta[{\rm CT}_0]>0$ and therefore $\mathbb P_\beta[\mathcal{E}_\infty]$ must be zero. The claim follows.\end{proof} \section{Proofs of the main results} \subsection{A bound on the percolation probability} Let us start with a crucial relation which justifies the consideration of $\mathbb P_\beta$. \begin{theorem}\label{thm:percolation} For $\beta$ at which $\widetilde M_{LRO}(\beta) =0$, also $\mathbb P_{\beta}\left[0\leftrightarrow\infty\right] = 0.$ \end{theorem} \begin{proof} Let $L>0$ and let $x,y\in\Lambda_L$. The switching lemma (Lemma~\ref{switching}) implies \begin{multline} \label{eq:2}{\rm P}^0_{\Lambda_L,\beta}\otimes {\rm P}^+_{\Lambda_L,\beta}\left[x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\text{ in }\Lambda_L\right] \\[1ex] :=\frac{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)\indf{x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\text{ in }\Lambda_L}}{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)} \notag \\[2ex] \end{multline} \begin{align} &=\frac{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\{x,y\}\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\} }}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)}{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)}.\end{align} The representations of spin-spin correlations \eqref{eq:7} and \eqref{eq:8} then imply that \begin{equation}\label{eq:6}{\rm P}^0_{\Lambda_L,\beta}\otimes {\rm P}^+_{\Lambda_L,\beta}\left[x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\text{ in }\Lambda_L\right]=\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^+\le \langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0. \end{equation} The right-hand side converges to $\langle\sigma_x\sigma_y\rangle_{\beta}^0$ as $L$ tends to infinity. Since the event on the left-hand side can be expressed in terms of $\widehat\mathbf{n}_1$ and $\widehat\mathbf{n}_2$, the convergence of $\widehat{\rm P}^+_{\Lambda_L,\beta}$ and $\widehat{\rm P}^0_{\Lambda_L,\beta}$ to $\widehat{\rm P}^+_{\beta}$ and $\widehat{\rm P}^0_{\beta}$ provided by Theorem~\ref{def:current infinite} implies that the left-hand side converges to $\mathbb P_{\beta}\left[x\leftrightarrow y\right]$. (The percolation event does not depend on finitely many edges, but justifying passing to the limit is straightforward by first considering the events that 0 is connected to distance $N$.) Therefore, \begin{equation}\label{eq:111} \mathbb P_{\beta}\left[x\leftrightarrow y\right]\ \le \ \langle\sigma_x\sigma_y\rangle_{\beta}^0\,. \end{equation} Let now $B \subset {\mathbb Z}^d$ be a finite subset. The Cauchy-Schwarz inequality applied to the random variable $X=\sum_{x\in B} \indf{ x\leftrightarrow \infty }$ leads to \begin{align*} \big(\,|B| \, \mathbb P_{\beta}[0\leftrightarrow\infty]\,\big)^2 :=\mathbb E_\beta[X]^2&\le \mathbb E_\beta[X^2]=:\sum_{x,y\in B}\mathbb P_{\beta}[x,y\leftrightarrow\infty]. \end{align*} The uniqueness of the infinite cluster thus implies \begin{align} \label{eq:133} \big(\,|B| \, \mathbb P_{\beta}[0\leftrightarrow\infty]\,\big)^2 &\le \sum_{x,y\in B}\mathbb P_{\beta}[x,y\leftrightarrow\infty]\le\sum_{x,y\in B}\mathbb P_{\beta}[x\leftrightarrow y].\end{align} Combining this relation with \eqref{eq:111} and optimizing over $B$ we get: \begin{align} \label{eq:567} \mathbb P_{\beta}[0\leftrightarrow\infty ]\, ^2 \ & \le \ \inf_{B\in {\mathbb Z}^d, |B|< \infty} \frac{1}{|B|^2}\sum_{x,y\in B}\langle\sigma_x\sigma_y\rangle_{\beta}^0 \ = \ \widetilde M_{LRO}(\beta)^2~, \end{align} which proves the claim. \end{proof} \begin{remark} In the last step of \eqref{eq:133} one can see uniqueness of the infinite cluster used as a substitute for the classical percolation argument which utilizes the FKG inequality, which we do not have for random currents.\end{remark} \subsection{Proof of Theorem~\ref{thm:continuity} } Fix a pair of vertices $x$ and $y$. Applying the switching lemma (Lemma~\ref{switching}) again, we find that for $L>0$: \begin{align}&\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^+-\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0=\frac{\displaystyle\sum_{\mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\}}w_\beta(\mathbf{n}_2)}{\displaystyle\sum_{\mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}w_\beta(\mathbf{n}_2)}-\frac{\displaystyle\sum_{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\{x,y\}}w_\beta(\mathbf{n}_1)}{\displaystyle\sum_{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset}w_\beta(\mathbf{n}_1)} \notag \\[2ex] &=\frac{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\}}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)-\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\{x,y\}\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)}{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)} \notag \end{align} \begin{align} &=\frac{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\}}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)\left(1-\indf{x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}y\text{ in }\Lambda_L}\right)}{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)}\,. \end{align} Yet, any configuration $\mathbf{n}_2$ with sources at $x$ and $y$ such that $x$ and $y$ are not connected in $\Lambda_L$ necessarily satisfies that $x$ and $y$ are connected to $\delta$. Therefore, \begin{equation}\label{eq:57} \langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^+-\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0\le\frac{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\}}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2) \indf{ x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}\delta }}{\displaystyle\sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)}.\end{equation} We shall now estimate the sum in the numerator by comparing it to the corresponding sum in which the source condition of $(\mathbf{n}_1,\mathbf{n}_2)$ is changed to $\partial \mathbf{n}_1 = \partial \mathbf{n}_2 =\emptyset $. Fix a sequence of vertices $x=x_0,\dots,x_m=y$ with $J_{x_i,x_{i+1}}>0$ for any $0\le i<m$. For any $L$ large enough so that $x_i\in\Lambda_L$ for all $i\le m$, consider the one-to-many mapping which assigns to each $\omega$ a modified current configuration $\mathbf{n}_2$ with the change limited to $\mathbf{n}_2$ along the set of bonds $e_j= \{ x_i, x_{i+1}\}$, $j=0,\ldots,m-1$, at which the parity of all these variables is flipped, and the value of the new one is at least $1$ at each bond. Under this mapping, the image of each pair $ (\mathbf{n}_1,\mathbf{n}_2)$ that contributes in the numerator of \eqref{eq:57} lies in the set for which the connection event $x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}\delta $ remains satisfied, but the source set of $\mathbf{n}_2$ is reset to $\partial \mathbf{n}_2 =\emptyset $. Classifying the current pairs according to the values of all the unaffected variables of $\{ (\mathbf{n}_1,\mathbf{n}_2)\}$, and the parity of $\mathbf{n}_2$ along the set of bonds $e_0,\dots,e_{m-1}$, it is easy to see that under this one-to-many map the measure of each set is multiplied by a factor which is larger than or equal to \be \Gamma_{x,y} \ = \ \prod_{j=1}^m \min \left \{ \frac{ \sinh (\beta J_{e_j}) }{ \cosh (\beta J_{e_j})}, \frac{ \cosh (\beta J_{e_j}) - 1}{ \sinh (\beta J_{e_j})} \right \} \,. \end{equation} (i.e. the original measure multiplied by $\Gamma_x $ is dominated by the measure of the image set.) This allows us to conclude: \begin{multline} \label{eq:202} \sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1\ =\ \emptyset \\ \mathbf{n}_2\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2=\{x,y\}}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2)\indf{ x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}}\delta } \ \le \ \\ \le \ \Gamma_{x,y}^{-1} \sum_{\substack{\mathbf{n}_1\in\Omega_{\Lambda_L}:\,\partial \mathbf{n}_1=\emptyset\\ \mathbf{n}_2'\in\Omega_{\Lambda_L\cup\{\delta\}}:\,\partial \mathbf{n}_2'=\emptyset}}w_\beta(\mathbf{n}_1)w_\beta(\mathbf{n}_2')\,\indf{ x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2'}}\delta }\, . \end{multline} Inserting this in \eqref{eq:57}, we find that \begin{align} \label{eq:489} \langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^+-\langle\sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0 & \ \le \ \Gamma_{x,y}^{-1} \ {\rm P}^0_{\Lambda_L,\beta}\otimes {\rm P}^+_{\Lambda_L,\beta}[x\lr{\widehat{\mathbf{n}_1+\mathbf{n}_2}} \delta]\, . \end{align} Taking the limit $L\to \infty$ (which exists by Theorem~\ref{def:current infinite}) we obtain \begin{equation} \label{eq:99} 0 \ \le \ \langle\sigma_x\sigma_y\rangle_{\beta}^+-\langle\sigma_x\sigma_y\rangle_{\beta}^0 \ \le\ \ \Gamma_{x,y}^{-1} \ \mathbb P_\beta\left[x\leftrightarrow\infty\right]\,. \end{equation} (The percolation event on the right does not depend on finitely many edges, but justifying passing to the limit is straightforward by first considering the events that $x$ is connected to distance $N$.) \medbreak We now consider $\beta$ for which \eqref{eq:cont_cond} holds. Applying Theorem~\ref{thm:percolation} we conclude that $ \mathbb P_\beta\left[x\leftrightarrow\infty\right] =0$, and hence for any $x,y\in \mathbb{Z}^d$: $\langle\sigma_x\sigma_y\rangle_{\beta_c}^+ = \langle\sigma_x\sigma_y\rangle_{\beta_c}^0$. Thus, using the FKG inequality \cite{FKG71} and \eqref{eq:99}: \be 0\le \langle\sigma_0\rangle_{\beta_c}^+\langle\sigma_x\rangle_{\beta_c}^+\le\langle\sigma_0\sigma_y\rangle_{\beta_c}^+ = \langle\sigma_0\sigma_x\rangle_{\beta_c}^0 \end{equation} for any $y\in\mathbb{Z}^d$. The assumption that $\langle\sigma_0\sigma_x\rangle_{\beta_c}^0$ averages to zero over translations leads to $\langle\sigma_0\rangle_{\beta_c}^+=0$, i.e. $m^*(\beta_c)=0$. The full statement of continuity of the Gibbs state then follows by the known general result which is presented as Proposition~\ref{prop:GScont} in Appendix~\ref{app:continuity}, and the observation that for translation invariant models $m^*(\beta_c)=0$ implies that $\langle \sigma_x\rangle_{\beta_c}^+ =0$ for all sites $x\in \mathbb{Z}^d$. \\\ \qed \subsection{Proof of Corollary~\ref{cor:main}}\label{sec:3} We include the proof of Corollary~\ref{cor:main} only for completeness, as the argument is not new. The Gaussian domination bound \eqref{eq:89} (also known as the {\it infrared} bound) can be equivalently formulated as the statement that for any function $(v_x)\in\mathbb{C}^{\mathbb{T}_L}$, the correlation function $F_{L,\beta}(x,y) = \langle \sigma_x\sigma_y\rangle_{\mathbb{T}_L,\beta} $ satisfies: \be \label{eq:finitebound} \sum_{x,y\in \mathbb{T}_L}v_x\overline{v_y}F_{L,\beta}(x,y)\le \ \frac{1}{2\beta} \sum_{x,y\in \mathbb{T}_L}v_x\overline{v_y}G_L(x,y) \ + \ \frac{1}{L^d} \widehat F_{L,\beta}(0) \left| \sum_{x\in \mathbb{T}_L} v_x \right|^2 ~, \end{equation} where \be \label{eq:GL} G_L(x,y)=\sum_{p \ \in \ \mathbb{T}_L^\star\setminus\{0\}} \frac{1}{L^d} \ \ \ \frac{e^{i \, p \cdot (x-y) }}{E(p)} ~ . \end{equation} The sum in \eqref{eq:GL} (with the weights $\frac{1}{L^d} $) forms a Riemann approximation. Under the assumed Condition \eqref{eq:Ep_condition} ($L^1$ integrability of $1/E(p)$) standard approximation arguments allow to conclude the pointwise convergence \be \lim_{L\rightarrow \infty}G_L(x,y) \ =\ \int_{[-\pi,\pi]^d}\frac{{\rm d}p}{(2\pi)^d}\frac{e^{i \, p \cdot (x-y) }} {E(p)} \ =:\ G(x,y) ~ , \end{equation} to a function satisfying: \be \label{eq:63} \lim_{N\rightarrow \infty}\frac {1}{|\Lambda_N|^2}\sum_{x,y\in \Lambda_N}G(x,y)\ =\ 0 \, . \end{equation} Furthermore, for $\beta < \beta_c$ and any fixed function $v$ of bounded support the zero momentum term in \eqref{eq:finitebound} can be omitted since \begin{eqnarray} \label{eq:67} 0\ \le \ \lim_{L\to \infty} \frac{1}{L^d} \widehat F_{L,\beta}(0) & = & \lim_{L\to \infty} \frac{1}{|\mathbb{T}_L|} \sum_{x\in \mathbb{T}_L} \langle\sigma_0\sigma_x\rangle_{\mathbb{T}_L,\beta} \notag \\ & \le & \lim_{L\to \infty} \frac{1}{|\Lambda_L|} \sum_{x\in \Lambda_L} \langle\sigma_0\sigma_x\rangle_{\Lambda_L,\beta}^+ \ = \ 0 \, , \end{eqnarray} where use is made of the fact that $\sum_{x\in \mathbb{Z}^d } \langle\sigma_0\sigma_x\rangle_{\beta}^+ < \infty$ for any $\beta < \beta_c$ (\cite{ABF87}). To apply the above to the free boundary condition correlation function $\langle \sigma_x\sigma_y\rangle_{\Lambda_L,\beta}^0 $ let us first note that, by the Griffith inequality \cite{Gri67}, the latter are monotone increasing functions of $\beta$ and also monotone increasing in $L$. Standard semicontinuity arguments which are applicable under such monotonicity assumptions allow to conclude that for each $x,y\in \mathbb{Z}^d$ \be \label{eq:77} \langle\sigma_x\sigma_y\rangle_{\beta_c}^0 \ = \ \lim_{\beta \nearrow \beta_c} \ \lim_{L\to \infty} \langle\sigma_x\sigma_y\rangle_{\Lambda_L, \beta}^0 \ \end{equation} Combining \eqref{eq:77} with \eqref{eq:finitebound} for the function $v_x = \frac{1}{|\Lambda_n|} {\bf 1}[ x\in\Lambda_n] $, and using \eqref{eq:67}, we find that for each fixed $n$: \begin{eqnarray} \label{eq:65} \frac{1}{|\Lambda_n|^2} \sum_{x,y\in\Lambda_n}\langle\sigma_x\sigma_y\rangle_{\beta_c}^0 & = & \lim_{\beta \nearrow \beta_c} \ \lim_{L\to \infty} \frac{1}{|\Lambda_n|^2} \sum_{x,y\in\Lambda_n}\langle\sigma_x\sigma_y\rangle_{\Lambda_L, \beta}^0 \ \notag \\[2ex] & \le & \lim_{\beta \nearrow \beta_c} \ \lim_{L\to \infty} \frac{1}{|\Lambda_n|^2} \sum_{x,y\in\Lambda_n}\langle\sigma_x\sigma_y\rangle_{\mathbb{T}_L, \beta} \notag \\[3ex] & \le & \frac1{2\beta_c} \ \frac{1}{|\Lambda_n|^2} \sum_{x,y\in \Lambda_n}G(x,y) \ ~. \end{eqnarray} where another use is made of the Griffith inequality \cite{Gri67}, by which $\langle\sigma_x\sigma_y\rangle_{\Lambda_L, \beta}^0\le \langle\sigma_x\sigma_y\rangle_{\mathbb{T}_L, \beta}$. Incorporating now \eqref{eq:63} in \eqref{eq:65}, we see that if Condition \eqref{eq:Ep_condition} holds, then $\widehat M_{LRO}(\beta_c) = 0$. Thus, by Theorem~\ref{thm:continuity} the spontaneous magnetization $m^*(\beta)$ vanishes continuously at $\beta_c$, as claimed in Corollary~\ref{cor:main}. \qed \vspace{1cm}
\section{Introduction} The problem of the origin of spin glitches in neutron stars has remained unsolved since the first glitch was observed in the Vela pulsar in 1969. This difficult problem is of considerable interest, as an understanding of the glitch phenomenon would offer insights into both the dynamical and ground-state properties of matter above nuclear density. To make progress on the glitch problem, it is crucial to identify which components of the liquid interior corotate with the crust, which store and release the angular momentum that drives glitches, and which produce the observed post-glitch response that occurs over days to years (for examples, see \citealt{lohsen75,lyne87,cordes_etal88,sl96,wang_etal00,wang_etal01,espinoza_etal11}). A more complete understanding of the processes that regulate post-glitch response would be very useful for elucidating the dynamical properties of the neutron star interior. Fortunately, it is possible to address this problem without knowledge of the basic instability that causes glitches, and progress is possible with minimal assumptions. The neutron star interior was predicted long ago to contain superfluid neutrons and superconducting protons (\citealt{migdal59,gk65}; for recent reviews, see \citealt{dh03} and \citealt{ch08}). As in all rotating neutral superfluids, the neutron superfluid is threaded by an array of quantized vortices whose arrangement determines the angular momentum of the superfluid, and whose motion determines the torque exerted on the charged components of the star and the crust. In laboratory liquid helium, vortices {\em pin} to bumps on the bottom of the vessel, trapping the system in metastable rotational states. As the vessel is spun down, jumps in its spin rate are observed \citep{tt80}, much like the glitches observed in neutron stars, as vortices unpin and the superfluid transfers angular momentum to the vessel. Vortex pinning is predicted to occur throughout much of a neutron star, with the vortex lines pinning to sites along their length. In the inner crust, where a neutron superfluid coexists with the ionic lattice, vortices interact with nuclei with energies of $\sim 1$ MeV per nucleus \citep{alpar77,eb88,dp06,abbv07}. This interaction pins the vortex lattice to the nuclear lattice, as confirmed by the simulations of \citet{link09}. In the outer core, the neutron flow around the vortices that thread the rotating superfluid entrains a proton mass current, strongly magnetizing the vortices as shown in the seminal work of \citet{als84}. Here the protons are expected to form a type II superconductor, with the magnetic flux arranged in flux tubes that are frozen to the highly-conductive charged component of the fluid. The proton superconductor rotates as a rigid body, without forming vortices. The vortices of the neutron superfluid pin to the flux tubes, primarily through a magnetic interaction, with pinning energies as high as $\sim 100$ MeV per vortex-flux tube junction \citep{srinivasan_etal90,jones91,mendell91a,chau_etal92,rzc98,link12b}. While pinning energies remain rather uncertain, the conclusion that vortices pin to inner-crust nuclei and outer-core flux tubes is on solid ground. The nature of the glitch instability is a problem under active research, and many possibilities have been suggested. As originally suggested by \citet{ai75}, glitches could represent sudden unpinning of many vortices; as liberated vortices move under dissipation, a spin-up torque is exerted on the crust plus charged components. In this connection, glitches could arise through a vortex avalanche from a critical state \citep{cheng_etal88,mpw08}. Other possibilities include increased frictional coupling from sudden heating of the crust \citep{greenstein79a,le96} such as by a starquake, crust failure under superfluid stresses \citep{ruderman91c}, or interactions of magnetized flux tubes in the outer core with the London current near the core boundary \citep{sc99}. \citet{pizzochero11} has proposed that glitches result from the motion of a vortex sheet into a region where pinning is not sustainable. Alternatively, glitches might represent transitions between different states of superfluid turbulence \citep{peralta_etal06,pm09,ga09}, though these studies do not account for vortex pinning which is likely to play an important role in the dynamics. Recent work has shown that pinning is hydrodynamically unstable in both the inner crust \citep{link12a} and the outer core \citep{link12b}; further work is need to see how the instability saturates, and if the system turns turbulent. Vortex pinning decreases the rotational coupling between the portions of the superfluid in which there is pinning and the rest of the star. If vortices remained perfectly pinned between glitches, the decoupling would be complete, and the superfluid would exert no torque on the charged components of the star (the crust, electrons, and core proton fluid). At finite temperature, though, vortices move through thermal activation over their pinning barriers, with an associated coupling time that is generally far longer than without pinning. Post-glitch response has been ascribed to internal torques exerted on the crust as the pinned superfluid responds to the glitch through thermally-activated vortex motion, as in the ``vortex creep theory'' of \citet{alpar_etal84a}, hereafter AAPS, that has been substantially developed and used to fit pulsar timing data \citep{alpar_etal84b,alpar_etal85,alpar_etal86,alpar_etal88,alpar_etal89,alpar_etal93,alpar_etal94,alpar_etal96}. In vortex creep theory there are two dynamical regimes: the ``non-linear creep'' regime introduced by AAPS, and the ``linear creep'' regime introduced by \citet{alpar_etal89}. In the ``non-linear creep'' regime, the system is always near the threshold for vortex unpinning. Post-glitch response is determined by two timescales, the {\em decoupling time} $t_d$ and the {\em recoupling time} $t_r$, and post-glitch response depends non-linearly on the size of the glitch.\footnote{In AAPS, $t_d$ is denoted $t_o$, and is called the ``offset time", while $t_r$ is denoted $\tau$, and is called the ``relaxation time''.} Typical response in the ``nonlinear" creep regime is depicted in Figure \ref{response_resid}; the system remains out of equilibrium for a time $t_d$, before recovering over a time $t_r$, with a total response time of $t_d+t_r$. By contrast, in the ``linear creep" regime post-glitch response depends linearly on the size of the glitch, and consists of simple exponential recovery over one timescale rather than two. The ``linear creep" regime is generally associated with regions in which pinning is very weak, termed ``superweak pinning" by AAPS. The "non-linear creep regime" does admit linear response in the limit $t_r>>t_d$. The important dynamical difference between the two regimes is that the total recovery time can be very short in the ``linear creep" regime, but always exceeds $t_d$ in the ``non-linear" regime. As vortex creep theory became more developed, terms were introduced in the crust response to account for internal torques from different parts of the star, vortex depletion regions, and crust cracking; see, {\it e.g.}, \citet{alpar_etal96}. Prompt exponential response is introduced by assuming the existence of regions that are in the ``linear creep'' regime. Fits to data introduce many free parameters, and do not offer conclusions with which the theory can be refuted. In particular, the ``linear creep" regime, which gives simple exponential response independent of glitch magnitude, is not well-constrained by observations as it is observationally indistinguishable from vortex motion in a drag regime in which there is no pinning at all (see Section \ref{dragregime}). Inspired by the work of AAPS, the purpose of this paper is to present a comprehensive theory of thermally-activated vortex motion in the inner crust and the outer core with which to interpret existing and future glitch data and, in particular, to make {\em falsifiable predictions}. The theory developed here, which I call ``thermally-activated vortex slippage'' or ``vortex slippage'', to distinguish it from the ``vortex creep theory'' of AAPS, begins with a derivation of the vortex velocity from the vorticity equations of motion, an ingredient that is lacking from the vortex creep theory of AAPS and its extensions. The activation energy for vortex slippage is obtained from a detailed study of the mechanics and energetics of vortex motion. A key conclusion of this paper is that the ``linear creep" regime invoked in vortex creep theory is not realized for physically reasonable parameters. The ``linear creep" regime relies on vortex motion anti-parallel to the Magnus force to be nearly equal to vortex motion parallel to the Magnus force. Here I show that the anti-parallel motion is so strongly suppressed by the large vortex self-energy as to be negligible. Fits to post-glitch data that invoke this regime \citep{alpar_etal93,alpar_etal94,alpar_etal96} should therefore be reassessed. Without a ``linear creep'' regime of vortex response, the problem of vortex motion through thermal activation becomes strongly constrained; effectively, only the ``non-linear" regime of thermally-activated vortex motion remains. A second important conclusion is that ``superweak pinning", which plays a key role in vortex creep theory, probably does not occur but is eliminated by thermal fluctuations. The other conclusions of this paper are in {\em qualitative} agreement with the ``non-linear" regime studied by AAPS. In particular: \begin{enumerate} \item If pinning occurs, post-glitch response through thermally-activated vortex slippage consists of two distinct phases. First, the pinned superfluid is {\em decoupled} from the crust over a timescale $t_d$, during which the external torque acts on a lower moment of inertia, and the crust spins down at a greater rate. After a time $t_d$, a region in which there is pinning will recouple over a timescale $t_r$. The total recovery time from the glitch is $t_d+t_r$. Post-glitch response generally depends {\em non-linearly} on the initial conditions. \item For a glitch in the spin rate of magnitude $\Delta\nu$, the decoupling time is \begin{equation} t_d\simeq 7\,\left(\frac{t_{sd}}{10^4\mbox{ yr}}\right) \left(\frac{\Delta\nu/\nu}{10^{-6}}\right) \mbox{ days,} \end{equation} where $t_{sd}$ is the spin-down age. This timescale represents a {\em lower limit} for the duration of post-glitch response. \item If the recoupling time $t_r$ satisfies $t_r<t_d$, as can happen in a large glitch, and a portion $\Delta I_s$ of the core's moment of inertia participates in the response, the fractional change in $\dot\nu$ is \begin{equation} \frac{\delta\dot\nu}{\vert\dot\nu\vert}= -\frac{\Delta I_s}{I} \frac{1}{1+({\rm e}^{t_d/t_r}-1){\rm e}^{-t/t_r}}, \end{equation} depicted by the dashed curve in Figure \ref{response_dresid}. Identification of this characteristic non-exponential response through analyses of existing and future timing data would give evidence that thermally-activated vortex slippage plays a key role in post-glitch response. \item If a glitch is small enough that $t_d<<t_r$ is satisfied, the glitch recovers as ${\rm e}^{-t/t_r}$. \end{enumerate} Post-glitch response in the vortex slippage theory of this paper is mathematically identical to the ``non-linear creep" regime of AAPS, but with crucial quantitative differences. In this paper, the scaling of the recoupling time with the pinning energy $E_p$ is found to be $t_r\propto E_p^{0.9}$ for the inner crust and $t_r\propto E_p^{3/5}$ for the outer core. By contrast, AAPS conclude that $t_r$ is {\em independent} of $E_p$ in the ``non-linear" regime, and exponentially dependent on $E_p$ in the ``linear regime". To summarize, the vortex theory of AAPS and its extensions has two regimes of vortex motion, the ``non-linear" and ``linear" regimes, while the vortex slippage theory given here has only one, the ``non-linear" regime. The vortex slippage theory presented here predicts a minimum post-glitch response time of $t_d$, which is determined by the size of the glitch. Vortex creep theory invokes a ``linear creep" regime that is not realized for physically reasonable parameters, and that gives very short recovery times without a waiting time $t_d$ before the system recouples. By contrast, in the vortex slippage theory developed in this paper, the total post-glitch response time can only exceed $t_d$, giving strong constraints on the theory. This paper presents the theory of thermally-activated vortex slippage, with preliminary comparison with data; detailed comparisons with data and constraints of the theory will be given in a forthcoming publication. In Section \ref{overview}, the implications of post-glitch response are described in terms of a simple but quite general model of the coupling between the crust and the liquid interior. In Section \ref{dragregime}, the response of a neutron star without pinning is reviewed. In Section \ref{pinning}, pinning of vortices to inner-crust nuclei and outer-core flux tubes is described. In Section \ref{slippage}, the theory of thermally-activated vortex slippage is developed. In Section \ref{AE}, the mechanics and energetics of vortex pinning is discussed. In Section \ref{relaxation}, the relaxation dynamics are calculated and described. In Section \ref{timescales}, the coupling timescales are estimated for the inner crust and the outer core. In Section \ref{data}, preliminary comparisons of the calculations with observed post-glitch response are made. In Section \ref{comparison}, the results are compared with the previous work of AAPS and its extensions. Appendix \ref{transition} determines the conditions under which vortex pinning disappears, and vortex motion enters the drag regime. Appendix \ref{antiparallel} shows that vortex motion against the Magnus force is strongly suppressed. Appendix \ref{random_pinning} shows that vortex pinning is unlikely to occur in the denser regions of the inner crust, including the pasta region, so that the ``superweak pinning'' regime proposed by AAPS, a regime that plays a key role in vortex creep theory, is probably ruled out. \section{Post-glitch Response: General Considerations} \label{overview} Post-glitch response shows bewildering variety, with no clear systematics \citep{lss00,wang_etal00}. Most glitches are consistent with simple steps in the spin rate, with a recovery fraction $Q<<1$ over timescales of hundreds to thousands of days, as shown schematically in Figure \ref{typical_glitch}. In many pulsars, the Vela pulsar for example, the spin-down rate of the star does not recover to its post-glitch value before the next glitch occurs \citep{cordes_etal88,lyne_etal96,dodson_etal02,dodson_etal07}, indicating that rotational equilibrium is never reached; the internal torque on the star never becomes constant. The post-glitch response of such pulsars is problematic to study theoretically, since the glitch cannot be described as a perturbation about a state of rotational equilibrium. Consider the stellar crust plus charges, component $C$ of moment of inertia $I_c$, coupled to the superfluid interior, component $S$ of moment of inertia $I_s$ comprising the neutron superfluids of both the inner crust and the core. Let a glitch occur at $t=0$. To account for the fact that most glitches do not recover completely ($Q<1$), I introduce a third component, component $L$, which imparts angular momentum $J_l$ at the time of the glitch. Unlike $C$ and $S$, this ``loose screw'' is never in rotational equilibrium; it couples to $C$ only at the time of the glitch, but remains otherwise completely decoupled. As described in Section \ref{timescales}, much of the neutron star interior can remain decoupled by glitches if there is pinning; decoupled components in either the crust or the core could represent the loose screw that drives glitches. Assuming axisymmetry for simplicity, the rotational dynamics of the system is given by \begin{equation} I_c\dot{\Omega}_c+\int dI_s\, \dot{\Omega}_s=N_{\rm ext}+J_l\,\delta(t), \label{jcons} \end{equation} where $N_{\rm ext}$ is the external spin-down torque on $C$. Sufficiently long after the glitch, the system will relax to {\em spin-down equilibrium}, with $\dot{\Omega}_c=\dot{\Omega}_s=\dot{\Omega}_0$, where $\dot{\Omega}_0$ is the equilibrium spin-down rate of both $C$ and $S$. The external torque is given by $N_{\rm ext}=I\dot{\Omega}_0$, where $I\equiv I_c+I_s$. (Note that $I$ excludes component $L$). Defining the {\em lag} $\omega(r,t)\equiv\Omega_s(r,t)-\Omega_c(t)$, where $r$ is the polar radius, Equation [\ref{jcons}] can be rewritten as \begin{equation} \dot{\Omega}_c=-\frac{1}{I}\int dI_s\, \dot{\omega}(r,t)+\dot{\Omega}_0 +\frac{J_l}{I}\delta(t). \label{z7} \end{equation} The change in spin rate across the glitch is \begin{equation} \Delta\Omega_c\equiv\Omega_c(0+)-\Omega_c(0-)=-\frac{1}{I}\int dI_s\, \left[\omega(r,0+)-\omega(r,0-)\right] +\frac{J_l}{I}, \label{one} \end{equation} and the solution to Equation [\ref{z7}] for $t>0$ is \begin{equation} \Omega_c(t)-\Omega_c(0+)= -\frac{1}{I}\int dI_s\,\left[\omega(r,t)-\omega(r,0+)\right]+\dot{\Omega}_0\,t, \label{two} \end{equation} assuming constant $N_{\rm ext}$. The {\em spin rate residual} is the difference between the post-glitch spin rate and the value it would have had if the glitch had not occurred: \begin{equation} \delta\Omega_c(t)\equiv \Omega_c(t) -\left[\Omega_c(0-)+\dot{\Omega}_0t\right]. \label{residdef} \end{equation} Adding Equations [\ref{one}] and [\ref{two}], and using the definition of the spin residual, gives \begin{equation} \delta\Omega_c(t)=-\frac{1}{I}\int dI_s\, \left[\omega(r,t)-\omega(r,0-)\right] +\frac{J_l}{I}. \label{z3} \end{equation} Of general interest is coupling between $C$ and $S$ that depends only on the local lag $\omega(r,t)$: \begin{equation} \dot{\Omega}_s(r,t)=f(\omega(r,t)), \label{coupling} \end{equation} where the coupling could be linear or non-linear in $\omega$. The state of spin-down equilibrium, $\omega_0(r)$, is given by the solution to \begin{equation} \dot{\Omega}_0=f(\omega_0(r)). \end{equation} Let $S$ and $C$ be in spin-down equilibrium just before the glitch. Then \begin{equation} \delta\Omega_c(t)=-\frac{1}{I}\int dI_s\, \left[\omega(r,t)-\omega_0(r)\right] +\frac{J_l}{I}. \label{resid} \end{equation} At late times after the glitch, $\omega(r,t)$ recovers to $\omega_0(r)$, and the residual becomes \begin{equation} \lim_{t\rightarrow\infty}\delta\Omega_c(t)=\frac{J_l}{I}=(1-Q)\Delta\Omega_c. \label{Jl} \end{equation} From Equation [\ref{jcons}], $I_c\Delta\Omega_c=J_l$, so that \begin{equation} Q=\frac{I_s}{I}. \end{equation} That many observed glitches have $Q<<1$ with recovery of the spin-down rate to its pre-glitch value suggests, in the context of this simple model, that glitches are driven by an out-of-equilibrium component, and that only a small fraction of the superfluid is associated with post-glitch response.\footnote{ More generally, we might expect that some fraction $f$ of the angular momentum imparted to $C$ at the time of the glitch comes from $L$, with the remainder coming from $S$. For this situation, $Q$ lies in the range \begin{equation} \frac{I_s}{I}\le Q\le 1. \label{Qrange} \end{equation} where $Q=I_s/I$ corresponds to $f=1$, while $Q=1$ corresponds to $f=0$. Henceforth I take $f=1$, a restriction that does not affect the conclusions.} \section{Dynamical Response without Vortex Pinning} \label{dragregime} The dynamics of a superfluid is determined by the motion of the vortices, which is in turn determined by the dissipative forces on the rotating superfluid. In regions of the star where there is no vortex pinning, vortex motion is determined by drag forces. Here I review the dynamics in this regime. Suppose a straight vortex is moving at velocity $\vbf_v$ with respect to its environment. The environment could be the nuclear lattice of the inner crust, or the proton-electron fluid of the outer core. The motion of a straight vortex is given by balance between the Magnus force on the vortex and the drag force: \begin{equation} \rho_s\kappabf\times\left(\vbf_v-\vsbf\right)-\eta \vbf_v = 0, \label{drag} \end{equation} where $\vsbf$ is the flow velocity of the ambient superfluid in the rest frame of the environment, $\rho_s$ is the superfluid mass density, $\kappa=h/2m_n$ is the vorticity quantum where $m_n$ is the neutron mass$, \kappabf$ is directed along the vortex, and $\eta$ is the drag coefficient.\footnote{ In the low-temperature environment of a neutron star, a vortex has negligible effective mass; its velocity is determined entirely by $\vsbf$ and local forces \citep{bc83}.} In cylindrical coordinates $(r,\phi,z)$, with $z$ and $\kappabf$ along the rotation axis, the vortex velocity is given by the solution of Equation [\ref{drag}]; see \citet{be89} and \citet{eb92}: \begin{equation} \vbf_v=v_s\left(\frac{1}{2}\sin 2\theta_d\,\hat{r}+\cos^2\theta_d\,\hat{\phi}\right)= v_s\cos\theta_d\left(\sin\theta_d\,\hat{r}+\cos\theta_d\,\hat{\phi}\right) =r\omega\cos\theta_d\,\hat{n}, \label{vv} \end{equation} where the {\em dissipation angle} $\theta_d$ is given by \begin{equation} \tan\theta_d\equiv\frac{\eta}{\rho_s\kappa}. \label{theta} \end{equation} The vortex moves in direction $\hat{n}$, at an angle $\theta_d$ with respect to $\vsbf$ with a component away from the rotation axis for positive lag $\omega$. An important feature of vortex dynamics is that local forces determine the vortex velocity, not the vortex acceleration. The dissipation angle is simply related to the relaxation time of the system. Assuming axisymmetry, the angular acceleration of the superfluid follows from vorticity conservation (AAPS; \citealt{leb93}) \begin{equation} \dot{\Omega}_s(r,t)=-\frac{1}{r}\left(2\Omega_s(r,t)+ r\frac{\partial}{\partial r}\Omega_s(r,t)\right) \vbf_v\cdot\hat{r}, \label{sfeom1} \end{equation} and is determined by the radial component of the vortex velocity which vanishes in the limit of zero dissipation ($\theta_d=0$). Assuming uniform superfluid rotation $\partial\Omega_s/\partial r=0$ for simplicity, the equation of motion for the angular velocity lag $\omega\equiv\Omega_s-\Omega_c$ is, from Equation [\ref{jcons}], \begin{equation} \dot{\omega}=\frac{I}{I_c}(\vert\dot{\Omega}_0\vert+\dot{\Omega}_s) - \frac{J_l}{I_c}\delta(t). \label{lagdot} \end{equation} Linearizing Equation [\ref{sfeom1}], assuming constant $\theta_d$, and using Equation [\ref{vv}], gives \begin{equation} \dot{\Omega}_s=-\omega\,\Omega_s\sin 2\theta_d, \end{equation} where $\Omega_s$ is the unperturbed rotational velocity of the superfluid. Assuming spin-down equilibrium with $\omega(0-)=\omega_0$ just before the glitch, the solution is \begin{equation} \omega(t)=\omega_0+(\omega(0+)-\omega_0){\rm e}^{-t/t_r} \qquad t_r=\frac{I_c}{I}\frac{1}{\Omega_s\sin 2\theta_d}. \label{tr_drag} \end{equation} From Equation [\ref{lagdot}], the equilibrium lag is \begin{equation} \omega_0=\frac{I}{I_c}t_r\vert\dot{\Omega}_0\vert. \end{equation} In the $C+S+L$ model of Section \ref{overview}, with a spin glitch driven by angular momentum from components $L$, the response for uniform rotation follows from Equations [\ref{resid}], [\ref{Jl}], and [\ref{jcons}]: \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}=1-Q(1-{\rm e}^{-t/t_r}) \qquad Q=\frac{I_s}{I}. \label{friction} \end{equation} In this linear treatment, the timescale of relaxation is independent of the size of the glitch. As described below, vortices are expected to pin to nuclei in the inner crust and to flux tubes in the outer core (if the outer core is a type II superconductor). Pinning, though, is not expected to occur everywhere in the star. Let us therefore first consider calculated values for the drag strength without pinning, beginning with the inner crust. If vortices in the inner crust move in the drag regime, the dominant drag at low lag arises from scattering of the moving vortex against the electron-phonon system of the lattice. \citet{jones90a} finds $3\times 10^{-5}\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$} \sin 2\theta_d\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$} 10^{-6}$, with a typical relaxation time of \begin{equation} t_r=0.2\,\frac{I}{I_c}\left(\frac{\nu}{10\mbox{ Hz}}\right)^{-1} \left(\frac{\sin 2\theta_d}{10^{-6}}\right)^{-1}\mbox{ days}, \end{equation} where $\nu$ is the stellar spin rate. Relaxation through this process is shorter than observed, but the timescale increases significantly as the interaction energy between a nucleus and a vortex is reduced. \citet{be89} studied the scattering of electrons with the electron cloud induced around a vortex through its interaction with nuclei. They found $\sin 2\theta_d\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$} 10^{-10}$, corresponding to a relaxation time of 2000 d for $\nu=10$ Hz. \citet{feibelman71} considered the scattering of electrons with the magnetic moments of thermally-excited neutrons in the vortex core. The results are strongly sensitive to the neutron pairing gap and to the temperature. From these calculations, using a range of plausible pairing gaps, $\sin 2\theta_d$ can range from $\sim 10^{-8}$ to effectively zero. In the outer core, entrainment of the neutron and proton mass currents when both species are superfluid endows a vortex with a magnetic field of $B_v\sim 10^{14}$ G \citep{als84}. Electron scattering with the strongly-magnetized vortex cores gives $\sin 2\theta_d \sim 10^{-3}$ \citep{as88}. Hence, if there is no pinning in the core, electron scattering is so effective that the neutral and charged fluids of the core corotate over timescales $\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 10^3\,\Omega_c^{-1}$. The core liquid is thus typically considered to corotate with the crust. \citet{ss95} and \citet{ssct95} have obtained the opposite conclusion; ignoring the natal magnetic field of the neutron star, \citet{ss95} and \citet{ssct95} argue that magnetic flux tubes of the outer core form clusters around vortices, which could increase the coupling time to observable timescales of days or longer. They suggest that post-glitch response has a core component. The effects of the natal magnetic field on these conclusions have not been studied, and could prove to be important. The point of view taken in this paper is that post-glitch response could involve both the inner-crust and core superfluids, and both possibilities will be considered. \section{Vortex Pinning} \label{pinning} \subsection{Pinning of Vortices to Nuclei in the Inner Crust} As originally pointed out by \citet{pines71} and elaborated upon by \citet{ai75} and \citet{alpar77}, vortices of the inner crust interact with nuclei. The main contribution to this interaction is the density dependence of the superfluid condensation energy per particle. When the increase of energy associated with the destruction of superfluidity within the vortex core exceeds the corresponding increase inside the nucleus, the energy cost is minimized if the vortex overlaps with the nucleus, giving an attractive interaction. If the inequality goes the other way, the energy is minimized if the vortices can get as far from nuclei as possible, in which case vortices could pin to the interstices of the nuclear lattice. Pinning is predicted to be strongest in the denser regions of the inner crust. Recent pinning calculations generally agree on the magnitude and density dependence of $E_p$, but not its sign. \citet{dp06} find that the vortex-nucleus interaction energy is $E_p\simeq 3$ MeV, and attractive, for baryon densities $\rho_b$ in the narrow range $3\times 10^{13}<\rho_b<5\times 10^{13}$ \hbox{\rm\hskip.35em g cm}$^{-3}$. \citet{abbv07} find a similar interaction strength in this density range, though repulsive. Both calculations show a sharp drop in $\vert E_p\vert$ to nearly zero above $\rho_b\simeq 6\times 10^{13}$ \hbox{\rm\hskip.35em g cm}$^{-3}$. These calculations did not address the spatial dependence of the interaction, and so only estimates of the pinning force based on the relevant length scales could be obtained. The length scale of the interaction is of order the coherence length $\sim\xi_n$ of the neutron superfluid, and so the force per pinning site is $F_p\sim E_p/\xi_n$ for pinning to nuclei. This estimate should apply for a repulsive interaction as well. A vortex is an object with finite rigidity (see below), and in order to move through the lattice it must be brought close to nuclei that repel the vortex over a length scale $\xi_n$. Hence, nuclear and interstitial pinning should be very similar.\footnote{For a repulsive interaction, \citet{dp06} estimate the interaction force per site to be $F_p\sim E_p/a$, where $a$ is the nuclear spacing. Since $a$ is much larger than $\xi_n$ in regions where they find the interaction to be repulsive, they conclude that interstitial pinning is relatively unimportant, contrary to the argument given here.} \citet{eb88}, using Ginzburg-Landau theory, obtained $E_p\simeq 10$ MeV; Ginzburg-Landau theory is however inadequate for this problem, since the superfluid coherence length is comparable to the size of a nucleus. Above a baryon density $\rho_b \simeq 6\times 10^{13}$ \hbox{\rm\hskip.35em g cm}$^{-3}$, the pinning situation is unclear. The coherence length exceeds the nuclear spacing \citep{schwenk_etal03,cao_etal06}, which weakens pinning. At $\rho_b\simeq 10^{14}$ \hbox{\rm\hskip.35em g cm}$^{-3}$, the nuclei acquire non-spherical pasta shapes \citep{rpw83}. Pinning to nuclear pasta is probably weaker than pinning to spherical nuclei. Hence, the base of the inner crust could represent a physically distinct region with relatively weak pinning or no pinning. Pinning to nuclear pasta has been studied by \citet{mochizuki1995self} and \cite{mochizuki1997exotic}, who suggest that a vortex can induce nuclear rod formation along its length, becoming self-pinned. Pinning without idealized geometries has not been studied; a first attempt at this problem is given in Appendix \ref{random_pinning}. If a vortex could bend to intersect nuclei separated by $a$, the pinning force per unit length would be \begin{equation} f_p= \frac{F_p}{a}\simeq\frac{E_p}{a\xi_n}, \end{equation} typically of order $10^{17}$ dyn cm$^{-1}$ in the inner crust. This force is reduced by the vortex's large self-energy, or {\em tension}, associated with the kinetic energy of the flow about the vortex \citep{thomson1880,fetter67}: \begin{equation} T_v=\frac{1}{2}\int d^2r\, \rho_s \left(\frac{\kappa}{r}\right)^2 =\frac{\rho_s\kappa^2}{4\pi}\Lambda \qquad \qquad \Lambda\equiv\ln\frac{l_v}{\xi_n}, \label{tension} \end{equation} where $\rho_s$ is the density of free superfluid neutrons. A lower cut-off at the vortex core radius and an upper cut-off at the inter-vortex separation $l_v$ have been introduced; the logarithmic factor $\Lambda$ is typically $\simeq 3$, as assumed henceforth. Because the vortex is stiff, it cannot bend to intersect every nucleus, but bends over a length scale $l_p>a$, as shown in Figure \ref{nuclear_pinning}. Throughout much of the inner crust, most of the neutron superfluid is non-dissipatively {\em entrained} by the nuclei and does not participate in the superfluid flow. \citet{chamel05,chamel12} finds that $\sim 90$\% of the neutron mass is entrained in the denser regions of the inner crust. The effects of entrainment can be treated by taking \begin{equation} \rho_s=f_c\rho_n \end{equation} where $\rho_n$ is the total mass density in neutrons, both bound to nuclei and unbound, $\rho_s$ is the density of superfluid {\em conduction neutrons} that are not entrained by nuclei, and $f_c\sim 0.1$ is the fraction of of the neutron fluid comprising superfluid conduction neutrons. Accounting for nuclear entrainment, the typical vortex self energy is \begin{equation} T_v\simeq 0.6\, f_c\, \rho_{n,14}\mbox{ MeV fm$^{-1}$}, \label{tension1} \end{equation} where $\rho_{n,14}$ is the total superfluid mass density in units of $10^{14}$ \hbox{\rm\hskip.35em g cm}$^{-3}$. Tension makes the vortex difficult to bend over the typical nuclear spacing $a\simeq 50$ fm of the inner crust. Dynamical simulations of a vortex in a random potential at zero temperature by \citet{link09}, accounting for vortex tension, show that pinning is inevitable below a critical value of the superfluid flow speed with respect to the lattice, but that the pinning force per unit length $f_p$ is reduced by a factor $a/l_p$, where $l_p$ is the characteristic bending length of a vortex, typically $10-30a$ (if nuclear entrainment is neglected). A variational estimate (\citealt{link12a}; see also \citealt{lc02}) gives for the pinning force per unit length \begin{equation} f_p=\frac{F_p}{l_p}\simeq \frac{E_p}{a\xi_n}\left(\frac{2E_p}{3aT_v}\right)^{1/2} \qquad \mbox{where} \qquad \frac{l_p}{a}=\left(\frac{3aT_v}{2E_p}\right)^{1/2}, \label{lp} \end{equation} typically of order $10^{16}$ dyn cm$^{-1}$. A similar number was obtained by \citet{gp12}. The critical velocity difference $v_s$ in the rest frame of the solid above which pinning becomes unstable follows by equating the Magnus force per unit length to the pinning force per unit length. In terms of the critical lag $\omega_{\rm crit}=v_{s,{\rm crit}}/r$ at polar radius $r$, accounting for entrainment by nuclei, the Magnus force is \begin{equation} f_c\rho_n\kappa r\omega_{\rm crit} =f_p. \end{equation} Combining with Equations [\ref{tension1}] and [\ref{lp}] gives \begin{equation} \omega_{\rm crit}=\frac{E_p}{rf_c\rho_s\kappa\xi_n l_p}\sim 4\, \rho_{n,14}^{-3/2}\,E_p(\mbox{MeV})^{3/2}\left(\frac{f_c}{0.1}\right)^{-3/2} \left(\frac{\xi_n}{10\mbox{ fm}}\right)^{-1} \left(\frac{r}{\mbox{10 km}}\right)^{-1} \, \mbox{ \hbox{\rm\hskip.35em rad s}$^{-1}$}. \label{omc_crust} \end{equation} Note that entrainment of the neutron superfluid by nuclei increases $\omega_{\rm crit}$ by reducing both the tension and the Magnus force. For small $E_p$, thermal motion of the vortex precludes pinning. As shown in Appendix \ref{transition}, pinning disappears for \begin{equation} E_p<0.04\mbox{ MeV}\, \left(\frac{kT}{\mbox{10 keV}}\right)^{3/2} \left(\frac{a}{\mbox{50 fm}}\right) \left(\frac{\rho_{n,14}}{0.5}\right)^{1/3} \left(\frac{f_c}{0.1}\right)^{1/3}. \label{Epbound} \end{equation} Without entrainment, pinning vanishes for $E_p\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$} 0.1$ MeV. As discussed in Appendix \ref{random_pinning}, pinning to nuclei is likely to disappear above a baryon density of $\rho_b\simeq 6\times 10^{13}$. Above this density, the coherence length $\xi_n$ quickly begins to exceed the size of the Wigner-Seitz cell \citep{schwenk_etal03,cao_etal06}, and vortices interact primarily with spatial variations in the number density of nuclei. The effective pinning energy becomes so low that thermal fluctuations preclude pinning unless the interaction energy per nucleus $E_p$ exceeds $\simeq 0.4$ MeV. Recent pinning calculations show a sharp drop in $E_p$ to nearly zero above $\rho_b\simeq 6\times 10^{13}$ \hbox{\rm\hskip.35em g cm}$^{-3}$\ \citep{dp06,abbv07}. These effects appear to eliminate ``superweak pinning'' proposed by AAPS to exist in this density regime; typical values of $E_p$ for ``superweak pinning'' used in vortex creep theory are $\simeq 0.3$ MeV ({\it e.g.}, \citealt{alpar_etal89}). For $\rho_b\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 6\times 10^{13}$ \hbox{\rm\hskip.35em g cm}$^{-3}$\ in the inner crust, pinning probably does not occur at all, and vortex motion enters the drag regime. \subsection{Pinning of Vortices to Flux Tubes in the Outer Core} The protons of the outer core are predicted to form a type II superconductor. As the protons condensed when the star was young, the very high electrical conductivity of the relativistically-degenerate electrons prevented Meissner expulsion of the core's natal magnetic field \citep{bpp69a}, and the field formed a dense tangle of magnetic flux tubes with which vortices interact. This primarily magnetic interaction pins the vortices to the magnetic tangle which is frozen to the superconducting fluid. The flux tube tangle will be treated as completely immobile under the stresses exerted by the vortex array. Overlap of a vortex line and a flux tube is energetically favored because the volume of uncondensed fluid is minimized by such overlap; the interaction energy is $\sim 0.1$ MeV per junction \citep{srinivasan_etal90}. A far larger contribution to the interaction energy is the magnetic interaction between the two structures. The magnetic field in a flux tube is $B_\Phi\sim 10^{15}$ G \citep{als84}. The magnetic interaction energy between a vortex and a flux tube $E_p$ is of order $B_v B_\Phi V$, where $V$ is the overlap volume, and has been estimated to be $\sim 100$ MeV by a number of authors \citep{jones91,mendell91a,chau_etal92,rzc98,link12b}. The angle-averaged interaction energy for the intersection of a vortex with a flux tube is \citep{link12b} \begin{equation} E_p \simeq 10^2\, \left(\frac{m_p^*/m_p}{0.5}\right)^{-1/2} \left(\frac{\vert\delta m_p^*\vert/m_p}{0.5}\right) \left(\frac{x_p}{0.05}\right)^{1/2} \left(\frac{\rho_{n,14}}{4}\right)^{1/2} \mbox{ MeV}, \label{Ep} \end{equation} where $m_p$ is the bare proton mass, $m_p^*\equiv m_p+\delta m_p^*\sim m_p/2$ \citep{sjoberg,ch06} is its effective mass, and $x_p\simeq 0.05$ is the proton mass fraction. The range of the interaction between a vortex and flux tubes is of order the characteristic radius of a flux tube, or the London length, given by \citep{als84}, \begin{equation} \Lambda_*=50\, \left(\frac{m_p^*/m_p}{0.5}\right)^{1/2} \left(\frac{x_p}{0.05}\right)^{-1/2} \left(\frac{\rho_{n,14}}{4}\right)^{-1/2}\mbox{ fm.} \label{Lambda} \end{equation} The pinning force is $F_p\sim E_p/\Lambda_*$, about 1 MeV fm$^{-1}$. For a single vortex immersed in a tangle of flux tubes, the average length between intersections will equal the average distance between flux tubes \citep{link12b}, \begin{equation} l_p\sim l_\Phi\simeq 5\times 10^3\, B_{12}^{-1/2}\mbox{ fm}, \label{lphi} \end{equation} where $B_{12}$ is the magnetic field in units of $10^{12}$ G. Note that the relevant field is the average {\em internal} field of the star, which can be significantly larger than the dipolar component \citep{braithwaite09}. Because a vortex must cross a ``fence'' of flux tubes in order to move, the pinning spacing is not increased by the vortex self energy, as is the case for pinning to nuclei in the inner crust. In the outer core, entrainment between the protons and neutrons, though essential for pinning, has only a small effect on the free neutron density \citep{ch06,link12b}; $f_c$ is effectively unity in evaluating both the tension and the Magnus force. The critical angular velocity difference $\omega_{\rm crit}$ between the neutron superfluid and charged component to which the flux tubes are frozen, is given by \begin{equation} \rho_n\kappa r\omega_{\rm crit} l_\Phi=\frac{E_p}{\Lambda_*}, \label{critical} \end{equation} giving \citep{link12b} \begin{equation} \omega_{\rm crit} \sim 10^{-1}\, \left(\frac{x_p}{0.05}\right) \frac{\vert\delta m_p^*\vert}{m_p^*}\left(\frac{E_p}{100\mbox{ MeV}}\right) \left(\frac{r}{\mbox{10 km}}\right)^{-1} \,B_{12}^{1/2}\mbox{ \hbox{\rm\hskip.35em rad s}$^{-1}$}. \label{omc_core} \end{equation} As shown in Appendix \ref{transition}, thermal motion eliminates pinning when \begin{equation} E_p<0.3\mbox{ MeV}\, \left(\frac{kT}{\mbox{10 keV}}\right)^{3/2}B_{12}^{-1/2}. \end{equation} The large value of $E_p$ of Equation [\ref{Ep}] suggests that pinning occurs wherever type II protons coexist with superfluid neutrons. \subsection{Dissipation with Vortex Pinning} \label{dissipation_with_pinning} Pinning causes differential rotation between superfluid and the crust to develop as the crust is spun down by external torque. If a vortex segment unpins, it will be dragged though its environment at relatively high velocity, and the dissipation will be generally stronger than the processes discussed in Section \ref{dragregime} for small velocities. In the inner crust, pinning of vortices to nuclei sustains velocity differences of $\sim r\omega_{\rm crit}\sim 10^6$ \hbox{\rm\hskip.35em cm s}$^{-1}$. The dominant contribution to vortex drag on an unpinned segment is the excitation of Kelvin phonons on the vortex \citep{eb92,jones92}, giving $\sin 2\theta_d$ as large as $\sim 0.1$. For $\omega\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$} 10^{-2}$ \hbox{\rm\hskip.35em rad s}$^{-1}$, Kelvin phonon production is expected to be greatly reduced \citep{jones92}, with a corresponding reduction in the drag. The reduction has not been calculated. In the outer core, pinning of vortices to flux tubes produces velocity differences of $\sim r\omega_{\rm crit}\sim 10^5$ \hbox{\rm\hskip.35em cm s}$^{-1}$. The strength of vortex drag has not been calculated in this regime, but it is reasonable to expect the dissipation to be stronger than the low velocity limit considered by \citet{als84}, for which $\sin 2\theta_d\sim 10^{-3}$. The motion of vortex segments with respect to flux tubes will also excite Kelvin phonons on the vortices, and produce wave excitations on flux tubes; both processes will contribute to the dissipation, and have not been calculated. Dissipation due to the forcing of vortices through flux tubes at super-critical velocities has been studied by \citet{link03}. In the estimates below, $\sin 2\theta_d>10^{-2}$ will be assumed for the inner crust, and $\sin 2\theta_d>10^{-3}$ will be assumed for the outer core. The coupling timescales calculated in Section \ref{timescales} depend very weakly on these choices. \section{Thermally-activated Vortex Slippage} \label{slippage} As a result of vortex pinning to nuclei in the inner crust, and to flux tubes in the outer core, the lag $\omega$ approaches its local critical value as the charged component is spun down by the electromagnetic torque. Before $\omega_{\rm crit}$ is reached, though, vortices will slip over their pinning barriers through hermal activation. I now calculate the vortex slippage velocity. The superfluid circulation speed around a vortex is inversely proportional to the distance from the vortex. The ratio of the pinning interaction length to the pinning spacing is $\sim 0.2$ in the inner crust, and $\sim 0.1$ in the outer core. A pinned vortex will henceforth be treated as a thin string with tension in classical continuum mechanics. Consider pinning of a vortex along axis $P_1$; see Figure \ref{trajectories}. Vortex motion occurs as follows. First, thermal fluctuations excite a vortex segment of length $L_p$ to a saddle-point configuration at $S$ that lies out of the range $r_0$ of the pinning potential. The saddle-point configuration is shown in Figure \ref{saddlepoint} for an idealized geometry of a linear array of pinning centers spaced by $l_p$. The saddle-point configuration will be described in detail below. In this unstable configuration, the stress on the endpoints of the segment is increased by a factor $L_p/l_p$, typically greater than unity. The segment will then ``unzip'' under the Magnus force until the vortex can reach another prospective pinning axis a distance $d$ away; $d=a$ for the inner crust and $l_\Phi$ for the outer core. During this stage, the vortex segment is out of range of the pinning potential, and drifts at velocity $\vbf_v$ in direction $\hat{n}$, given by Equation [\ref{vv}], toward axis $P_2$. For pinning at $P_2$ to occur, the vortex segment must unzip to length $L_f>L_p$, so that pinning to one site along $P_2$ is stable. This configuration is shown in Figure \ref{repinned}. The condition for mechanical equilibrium is \begin{equation} 2T_v \sin\left[\tan^{-1}\left(\frac{2d}{L_f}\right)\right]\,\hat{n}- \fbf_{\rm mag}L_f=\Fbf_p. \end{equation} Typically $d<<L_f$. Using the same cylindrical coordinate system of Section \ref{dragregime}, projecting onto $\hat{r}$ and $\hat{\phi}$, gives \begin{equation} \frac{4d}{L_f}T_v\sin\theta_d-f_{\rm mag}L_f=F_{p,r} \quad \mbox{and} \quad \frac{4d}{L_f}T_v\cos\theta_d=F_{p,\phi}. \end{equation} Typically we are interested in $\theta_d<<1$. Noting that $F_{p,\phi}$ is limited by $F_p=R_p/r_0$, gives \begin{equation} L_f=\frac{4T_vdr_0}{E_p}. \label{Lf} \end{equation} As $E_p$ is lowered, $L_f$ typically exceeds $N_p$, and the vortex must unzip from many pinning bonds to reach static equilibrium at $P_2$. Meanwhile, the same processes are occurring elsewhere along the portions of the vortex that are still pinned along the initial pinning axis $P_1$. The vortex slowly migrates, or {\em slips}, at angle $\theta_d$ with respect to $\vbf_s$, but at a speed $v_v<<v_s$. As shown in Appendix \ref{antiparallel}, competing transitions against $\fbf_{\rm mag}$ to states such as $P^\prime$ are prevented by the strong vortex tension unless the dissipation angle is unexpectedly small. In the inner crust, $\sin 2\theta_d\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 10^{-2}$ is expected. With nuclear entrainment ($f_c=0.1$) and $E_p=0.04$ MeV, the lower limit for pinning, the suppression factor is ${\rm e}^{-1.5}\simeq 0.22$, with much greater suppression at higher $E_p$. Transitions to $P^\prime$ become important only for \begin{equation} \sin 2\theta_d< 1.5\times 10^{-4}\, E_p(\mbox{MeV})^{-1/2}\left (\frac{f_c\rho_{n,14}}{0.5}\right)^{-1/2} \left(\frac{a}{\mbox{50 fm}}\right)^{-1/2} \left(\frac{kT}{\mbox{10 keV}}\right) \qquad \mbox{(inner crust)} \end{equation} In the outer core, where $\sin 2\theta_d\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 10^{-3}$ is expected, transitions to $P^\prime$ are suppressed relative to transitions to $P_2$ by a typical factor ${\rm e}^{-5\times 10^4}$, independent of $E_p$. Transitions to $P^\prime$ become important only for \begin{equation} \sin 2\theta_d<2\times 10^{-8}\, \left(\frac{\rho_{n,14}}{4}\right)^{-1} \left(\frac{kT}{\mbox{10 keV}}\right) B_{12}^{1/2} \qquad \mbox{(outer core)}. \end{equation} Transitions against $\fbf_{\rm mag}$ can thus be ignored in both the inner crust and the outer core. The vortex slippage process can be described in terms of a two-state system; 1) the vortex is pinned, with zero velocity and zero energy, or, 2) the vortex is unpinned, moving under drag at velocity $\vbf_v$, with energy $A$ upon unpinning. The energy $A$ is the activation energy for unpinning, specified below. The partition function for this two-state system is \begin{equation} Z=1+{\rm e}^{-\beta A}. \label{Z} \end{equation} Here $\beta\equiv(kT)^{-1}$ where $T$ is the temperature and $k$ is Boltzmann's constant. For slow vortex slippage, ${\rm e}^{-\beta A}<<1$ and $Z\simeq 1$. The slippage velocity of a typical vortex is given by the statistical average \begin{equation} \langle\vbf_v\rangle_\beta = Z^{-1}{\rm e}^{-\beta A}\vbf_v\simeq r\omega\left(\frac{1}{2}\sin 2\theta_d\,\hat{r}+\cos^2\theta_d\,\hat{\phi}\right) {\rm e}^{-\beta A}. \label{vslippage1} \end{equation} The terms that depend on $\theta_d$ refer to the forces exerted on the unpinned vortex segment by the environment. In the inner crust, the main dissipative force arises from the excitation of Kelvin waves as a vortex segment moves past nuclei. In the outer core, dissipation arises from electron scattering, but there will also be a contribution from the interaction of the translating segment with flux tubes; see Section \ref{dissipation_with_pinning}. This treatment of vortex motion assumes the existence of a well-defined pinned state, which requires $A>>kT$. The limit of no pinning given by Equation [\ref{vv}] is recovered by taking $A\rightarrow 0$ and $Z\rightarrow 1$, the latter replacement corresponding to the vortex having only one state of motion in this limit. We are interested in the flow dynamics of the system over length scales that exceed the inter-vortex spacing: \begin{equation} l_v=\left(\frac{\kappa}{2\Omega_s}\right)^{1/2}=3\times 10^{-3} \left(\frac{\Omega_s}{100\mbox{ rad s$^{-1}$}}\right)^{1/2}\mbox{ cm}. \end{equation} We obtain the desired limit by averaging the slippage velocity of Equation [\ref{vslippage1}] for a single vortex over a bundle of $N>>1$ vortices. Consider a bundle of vortices enclosed by a contour that spans an area $S$ that satisfies $\sqrt{S}>>l_v$, and average the vortex slippage velocity over the bundle as follows: \begin{equation} \langle\vbf\rangle=\frac{1}{S}\int dS\, \langle{\vbf}\rangle_\beta. \end{equation} $S$ can be chosen so that it contains many vortices ($\sqrt{S}>>l_v$), while being small enough that $A$, $\theta_d$, and $\omega$, which vary over macroscopic scales, are nearly constant. This choice gives the same result as the velocity for a typical vortex (Equation \ref{vslippage1}): \begin{equation} \langle\vbf\rangle= r\omega\left(\frac{1}{2}\sin 2\theta_d\,\hat{r}+\cos^2\theta_d\,\hat{\phi}\right) {\rm e}^{-\beta A}, \label{vslippage} \end{equation} Note that $\langle ...\rangle$ denotes both a statistical average and a spatial average. All vortices in the bundle move in the same way on average, so that $\theta_d$ after spatial averaging takes the same value that appears in Equation [\ref{vslippage1}] for a single vortex. The spatial averaging is similar to the macroscopic averaging procedure introduced by \cite{bc83}. The equation of motion of the superfluid follows by replacing $\vbf_v$ in Equation [\ref{sfeom1}] with $\langle\vbf_v\rangle$: \begin{equation} \dot{\Omega}_s(r,t)=-\frac{1}{r}\left(2\Omega_s(r,t)+ r\frac{\partial}{\partial r}\Omega_s(r,t)\right) \langle\vbf_v\rangle\cdot\hat{r}. \label{sfeom2} \end{equation} As an aside, I now connect the treatment so far to that of introducing a {\em mutual friction} force into the fluid equations, usually taken to have the following form: \begin{equation} \fbf/\rho_s={\cal B}^\prime\,\omegabf_s\times(\vbf_s-\vbf_b)+ {\cal B}\,\hat\omega_s\times(\omega_s\times\left\{\vbf_s-\vbf_b\right\}), \end{equation} where $\omegabf\equiv\nabla\times\vbf_s$ is the vorticity in the inertial frame, $\vbf_b$ is the velocity of the background against which vortices are dragged, and ${\cal B}^\prime$ and ${\cal B}$ are the the mutual friction coefficients. In terms of the dissipation angle, the mutual friction coefficients are \begin{equation} 1-{\cal B}^\prime=\cos^2\theta_d\,{\rm e}^{-\beta A}<<1 \qquad {\cal B}=\frac{1}{2}\sin 2\theta_d {\rm e}^{-\beta A}<<1. \end{equation} Vortex pinning is sometimes considered to be equivalent to the limit of high drag ({\it e.g.}, \citealt{swc99,link06,gaj08a,vl08,ga09}). Assuming that drag is the only force on a vortex, the vortex velocity in Equation [\ref{vv}] goes to zero in the limit $\eta\rightarrow\infty$. This description of vortex pinning is inadequate because it ignores the forces exerted on the vortex by the pinning lattice, which can immobilize the vortex even for zero dissipation. As Equation [\ref{vslippage}] shows, the vortex speed is $v_v={\rm e}^{-\beta A}\,v_s<<v_s$ for $\eta=0$ ($\theta_d=0)$. Dissipation determines the direction of vortex slippage, and gives $\vbf_v$ a component parallel to $\hat{r}$ for $\omega>0$, and anti-parallel to $\hat{r}$ for $\omega<0$. For vortex slippage at low dissipation angle, the mutual friction coefficients satisfy \begin{equation} {\cal B}<< 1-{\cal B}^\prime <<1, \end{equation} whereas many treatments of vortex pinning assume ({\it e.g.}, \citealt{swc99,link06,gaj08a,vl08,ga09}) \begin{equation} 1-{\cal B}^\prime <<{\cal B} <<1. \end{equation} This condition implicitly assumes very high drag, and is overly restrictive. \section{Mechanics and Energetics of Vortex Unpinning} \label{AE} Consider a vortex in stable equilibrium, pinned to a linear array of equally-spaced pinning sites along the $z$ axis. In the presence of a Magnus force, the vortex unpins by first assuming the unstable, minimum-energy saddle-point configuration depicted in Figure \ref{saddlepoint}. In the saddle-point configuration, a segment of vortex is free of $N_p$ pinning bonds. The activation energy $A$ is equal to the energy difference between this state and the fully pinned state, and is determined by the competition between the Magnus force, the pinning force, and the self-energy (or tension) of the vortex. Let the shape of the vortex be given by the vector $\ubf(z)$. The energy of a vortex segment of length $L$ is \citep{le91} \begin{equation} E=\int_L dz\, \left[ \frac{T_v}{2}\left|\frac{d\ubf(z)}{dz}\right|^2+ \rho_p V(u) -\rho_s (\kappabf\times\vbf_s)\cdot\ubf(z) \right]. \label{Ev} \end{equation} The first term is the energy cost of bending a vortex with a self-energy per unit length $T_v$. The second term is the pinning interaction, where $\rho_p=l_p^{-1}$ is the number of pinning sites per unit length and $l_p$ is the average pinning spacing. The third term is the potential energy associated with the constant Magnus force per unit length. The dependence of the interaction energy on the separation between a vortex segment and a pinning site is unknown, but will increase from zero to a maximum value $E_p$ over a length scale $r_0$. A physically sensible parameterization for the potential is the continuous piecewise function \begin{equation} V(u)= E_p\times \left\{\begin{array}{ll} \frac{3}{4}\left(\frac{u}{r_0}\right)^2-\frac{1}{4}\left(\frac{u}{r_0}\right)^3 & 0\le u\le 2r_0 \\ 1 & u>2r_0, \end{array} \right. \label{potential} \end{equation} shown in Figure \ref{potentialplot}. The corresponding force per pinning site has the following parabolic form: \begin{equation} F_p(u)=-\frac{dV}{du}=\left\{\begin{array}{ll} -F_{\rm max}\left[1-\left(\frac{u}{r_0}-1\right)^2\right] & 0\le u\le 2r_0 \\ 0 & u>2r_0, \end{array} \right. \end{equation} where the maximum force is $F_{\rm max}=3E_p/4r_0$. The assumed potential corresponds to the second-order series expansion of the interaction force with a maximum of $F_{\rm max}$ at $r_0$, and so is of general physical relevance. For pinning of vortices to nuclei in the inner crust, the interaction range $r_0$ is $\sim \xi_n$, and the pinning spacing is $l_p>a$. For the pinning of vortices to flux tubes in the outer core, the interaction range $r_0$ is $\sim \Lambda_*$, and the pinning spacing is $l_p\sim l_\Phi$. The large vortex self energy plays an important role in determining the mechanics and energetics of the unpinning process, quantified by the dimensionless tension parameter \footnote{In the notation of \citet{le91}, $\tau={\cal T}$, the pinning energy is $U_0=E_p$, and the pinning spacing is $l=l_p$.} \begin{equation} {\cal T}\equiv \frac{T_v r_0^2}{E_pl_p}. \end{equation} If vortex tension were negligible, so that ${\cal T}$ is effectively zero, the flexible vortex could unpin one pinning bond at a time to assume a new, stable configuration. Because $T_v$ is non-zero, though, the vortex must overcome $N_p>1$ pinning bonds in an unpinning event; $N_p$ is determined by ${\cal T}$. In the inner crust, \begin{equation} {\cal T}\simeq\frac{T_v \xi_n^2}{E_p l_p}\simeq 0.31\,\rho_{n,14}\,E_p(\mbox{MeV})^{-1/2}\,\left(\frac{f_c}{0.1}\right)^{1/2} \left(\frac{\xi_n}{10\mbox{ fm}}\right)^2 \left(\frac{a}{50\mbox{ fm}}\right)^{-3/2} \qquad \mbox{(inner crust)}, \label{Tcal1} \end{equation} while in the outer core, \begin{equation} {\cal T}\simeq \frac{T_v\Lambda_*^2}{E_pl_\Phi} \simeq 0.12\,\left(\frac{\rho_{n,14}}{4}\right) \left(\frac{E_p}{100\mbox{ MeV}}\right)^{-1} \left(\frac{\Lambda_*}{50\mbox{ fm}}\right)^2 B_{12}^{1/2} \qquad \mbox{(outer core)}. \label{Tcal} \end{equation} Minimization of Equation [\ref{Ev}] gives the stable and saddle-point vortex configurations and the associated energies; see \citet{le91} for details. For a linear array of pinning sites with the potential of Equation [\ref{potential}], the number of broken pinning bonds in the saddle-point configuration is \begin{equation} N_p\simeq 3.2{\,\cal T}^{1/2}(1-\omega/\omega_{\rm crit})^{-1/4}, \end{equation} and the activation energy is \begin{equation} A\simeq 1.6 N_p E_p (1-\omega/\omega_{\rm crit})^{3/2} \simeq 5.1 E_p {\cal T}^{1/2} (1-\omega/\omega_{\rm crit})^{5/4} \label{A} \end{equation} These expressions are valid for $3/4\le\omega/\omega_{\rm crit}< 1$ and $N_p\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 3$. The latter condition gives $1-\omega/\omega_{\rm crit}<{\cal T}^2$; for typical parameters of the inner crust and the outer core, $\omega$ will always be very close to $\omega_{\rm crit}$, so that this condition is satisfied. Hence, in both the inner crust and the outer core $N_p>>1$, and the vortex is effectively stiff for all pinning energies of interest. Vortex stiffness becomes more important as $E_p$ is reduced. The important effects of vortex stiffness in determining the activation energy were not considered by AAPS or in subsequent development of vortex creep theory. To see the basic scaling of $N_p$ and $A$ with ${\cal T}^{1/2}$, note that the energy of a vortex segment bent by amplitude $u$ is, from Equation [\ref{Ev}], \begin{equation} E\sim \frac{1}{2}T_v\frac{u^2}{L_p}+\frac{L_p}{l_p}E_p, \end{equation} for zero Magnus force. Extremizing the energy in $L_p$ gives \begin{equation} E\sim \left(\frac{2T_vu^2}{E_pl_p}\right)^{1/2}E_p. \label{Eest} \end{equation} If the vortex is to unpin, it must be excited to an amplitude comparable to the range of the pinning potential, $u\sim r_0$. Hence, \begin{equation} A\sim (2{\cal T})^{1/2}E_p\sim N_pE_p, \end{equation} where $N_p=L_p/l_p$ is the characteristic number of bonds set by vortex tension that must be broken for the vortex to unpin. \section{Relaxation Dynamics with Vortex Slippage} \label{relaxation} Suppose a glitch occurs at $t=0$, through some instability that need not be specified for this perturbative analysis, and let us consider the dynamics in the $C+S+L$ model of Equation [\ref{jcons}]. For axisymmetric rotation, the dynamics is given by \begin{equation} I_c\dot\Omega_c+\int dI_s\,\dot\Omega_s=-I\vert\dot\Omega_0\vert+J_l\,\delta(t), \label{Jcons7} \end{equation} \begin{equation} \dot\Omega_s(r,t)=\left(\Omega_s(r,t)+\frac{r}{2}\frac{\partial\Omega_s(r,t)} {\partial r}\right)\omega(r,t)\sin2\theta_d\, {\rm e}^{-\beta A(\omega(r,t))}, \label{Omsdot_full} \end{equation} where $r$ is the cylindrical radius. The response of the crust is determined by integrated dynamics of the superfluid on cylindrical shells of radius $r$. The system can be reduced to a single integral equation with some approximations. The second term in parenthesis in Equation [\ref{Omsdot_full}] is \begin{equation} \frac{r}{2}\frac{\partial\Omega_s(r,t)}{\partial r}\simeq \frac{r}{2}\frac{\partial\omega_{\rm crit}(r))}{\partial r}, \end{equation} where, from Equations [\ref{omc_crust}] and [\ref{omc_core}], $\omega_{\rm crit}\propto r^{-1}$. Hence, \begin{equation} \frac{r}{2}\frac{\partial\Omega_s(r,t))}{\partial r} \simeq -\frac{1}{2}\omega_{\rm crit}(r). \end{equation} The critical lag $\omega_{\rm crit}$ is always much smaller than $\Omega_s$, so the second term in Equation [\ref{Omsdot_full}] can be neglected, giving \begin{equation} \dot\Omega_s(r,t)=-\Omega_s(r,t)\,\omega(r,t)\,\sin 2\theta_d\, {\rm e}^{-\beta A(\omega(r,t))}. \end{equation} Expanding $A(\omega)$ about $\omega_0$ in the exponential gives \begin{equation} \dot{\Omega}_s(r,t)\simeq\dot{\Omega}_0\,{\rm e}^{b(\omega(r,t)-\omega_0(r))} \qquad b(r)\equiv \left.-\beta\,\frac{\partial A}{\partial\omega}\right|_{\omega_0(r)}. \label{eom} \end{equation} Combining this equation with Equation [\ref{Jcons7}] gives the equation of motion for the lag at each polar radius $r$, for $t>0$ \begin{equation} \dot\omega(r,t)= \vert\dot\Omega_0\vert\left(1-{\rm e}^{b(r)(\omega(r,t)-\omega_0(r))}\right) +\frac{1}{I_c}\int dI_s\, \vert\dot\Omega_0\vert\left(1-{\rm e}^{b(r)(\omega(r,t)-\omega_0(r))} \right). \label{lageom} \end{equation} Let us assume that the regions with pinning constitute a small fraction of the star's moment of inertia, so that $I_s<<I_c$. In this limit the spatial integral in the above equation can be neglected. With these approximations, the evolution of the lag is determined entirely by its value locally, and the response of the crust can be calculated by integrating $\omega(r,t)$ over cylindrical shells. These simplifications were introduced by \citet{alpar_etal84a}. Ignoring the spatial integral in Equation [\ref{lageom}], the solution is \citep{alpar_etal84a,leb93} \begin{equation} \dot{\omega}(r,t)=\vert\dot{\Omega}_0\vert \left[1-\frac{1}{1+({\rm e}^{t_d(r)/t_r(r)}-1){\rm e}^{-t/t_r(r)}}\right]. \label{solution} \end{equation} where \begin{equation} t_d(r)\equiv\frac{\omega_0(r)-\omega(r,0+)} {\vert\dot{\Omega}_0\vert} \qquad\qquad t_r(r)\equiv \frac{1}{b(r)\vert{\dot{\Omega}_0\vert}} \label{tt} \end{equation} Here $t_d(r)$ is the {\em decoupling time} of the shell at radius $r$ and $t_r$ is the intrinsic {\em recoupling time} of the shell. In the limit $t_d>>t_r$, a perturbation in $\omega$ leaves components $C$ and $S$ decoupled for a time $\sim t_d$, before the perturbation is damped over a recoupling time $t_r$. In the opposite limit $t_d<<t_r$, a perturbation damps over a time $t_r$. This behavior is a consequence of the non-linear dependence of the slippage velocity on the initial value of $\omega$. Equation [\ref{solution}] was also obtained by AAPS for the special case of the activation energy having linear dependence on $\omega$; the form given here is more general with the quantity $b$ given in Equation [\ref{eom}]. As a simplification, let us specify the pre-glitch state by assuming that all transient rotational dynamics has damped before the glitch, so that the system is in in {\em spin-down equilibrium} at $t=0$, such that $\dot{\omega}=0$, $\dot{\Omega}_s=\dot{\Omega}_c=\dot{\Omega}_0$ for lag $\omega_0$. We will find that the post-glitch recovery time of the inner crust greatly exceeds the average interval between glitches, so rotational equilibrium in the inner crust is not reached between glitches (more on this point in Section \ref{timescales}). By contrast, we will find that the core has ample time to recover rotational equilibrium between glitches. The equilibrium lag $\omega_0$ is given by the solution to \begin{equation} \dot{\Omega}_0=-\Omega_s(r)\,\omega_0(r) \sin2\theta_d\, {\rm e}^{-\beta A(\omega_0(r))}. \label{equilibrium} \end{equation} The residual spin rate of the crust is given by Equation [\ref{resid}]: \begin{equation} \delta\Omega_c(t)=-\frac{1}{I}\int dI_s\, \left[ \omega(r,t)-\omega_0(r)\right]+\frac{J_l}{I}, \end{equation} To determine the form and timescale of the response, I assume for simplicity that the glitch is driven by component $L$, so that \begin{equation} \omega(r,0+)=\omega_0(r)+\Delta\Omega_c. \end{equation} The decoupling time is independent of both $r$ and the details of pinning: \begin{equation} t_d=\frac{\Delta\Omega_c}{\vert\dot\Omega_0\vert}=2t_{sd}\frac{\Delta\Omega_c}{\Omega_c}, \label{td} \end{equation} where $t_{sd}\equiv\Omega_c/2\vert\dot\Omega_c\vert$ is the spin-down age. The spin rate residual $\delta\Omega_c$, in units of the initial spin jump $\Delta\Omega_c$, is \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}= 1-\frac{1}{I}\int dI_s\, \left[\frac{t}{t_d}-\int_{0+}^t\frac{dt^\prime}{t_d}\, \frac{1}{1+({\rm e}^{t_d/t_r(r)}-1){\rm e}^{-t^\prime/t_r(r)}}\right]. \label{fullsolution} \end{equation} At late times, the time integral approaches unity, giving \begin{equation} \lim_{t\rightarrow\infty}\frac{\delta\Omega_c(t)}{\Delta\Omega_c}= \frac{I_c}{I}=1-Q; \end{equation} the glitch recovers by a fraction $Q=I_s/I$. Now consider the pinning region to be in a spherical shell of constant density $\rho$ and thickness $\Delta$, inner radius $R$, and outer radius $R+\Delta$. The height of the shell in each hemisphere is $h(r)$. The response of the crust is given by \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}= 1-\frac{1}{I} \int_0^{R+\Delta}dr\, 2\pi\rho h(r) r^3\, \left[ \frac{t}{t_d}- \int_{0+}^t \frac{dt^\prime}{t_d}\, \frac{1}{1+({\rm e}^{t_d/t_r(r)}-1){\rm e}^{-t^\prime/t_r(r)}} \right], \label{fullsolution1} \end{equation} where \begin{eqnarray} h(r)= \left\{ \begin{array}{ll} \sqrt{(R+\Delta)^2-r^2}-\sqrt{R^2-r^2} & r<R \\ \sqrt{(R+\Delta)^2-r^2} & r\ge R \end{array} \right. \end{eqnarray} We will see that $t_r$ scales as $\omega_{\rm crit}\propto r^{-1}$. Cylindrical shells with smaller radii thus have the longest recoupling times, but they also have smaller moments of inertia $\propto r^4\Delta$. The response will be largely determined by the relaxation time in the region $R<r<R+\Delta$. Let the characteristic recoupling time there be $t_r(R)$. Equation [\ref{fullsolution1}] then becomes \[ \frac{\delta\Omega_c(t)}{\Delta\Omega_c}\simeq 1-\frac{I_s}{I} \left[ \frac{t}{t_d}- \int_{0+}^t \frac{dt^\prime}{t_d}\, \frac{1}{1+({\rm e}^{t_d/t_r(R)}-1){\rm e}^{-t^\prime/t_r(R)}} \right]= \] \begin{equation} 1-\frac{{I_s}}{t_dI}\left[t +t_r(R)\ln\left(\frac{1+c(R)}{{\rm e}^{t/t_r(R)}+c(R)}\right) \right] \qquad c(R)\equiv {\rm e}^{t_d/t_r(R)}-1. \label{fullsolution2} \end{equation} Figure \ref{shellmodel} shows that Equation [\ref{fullsolution2}] is a good approximation to Equation [\ref{fullsolution1}]. Henceforth I write $t_r(R)=t_r$. Now suppose we add concentric spherical shells, each with superfluid moment of inertia $I_{s,i}$ and characteristic relaxation time $t_r(R_i)\equiv t_{r,i}$. Because of the additive nature of the contributions to the post-glitch recovery within the approximations made so far, the total response of the star is \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}\simeq 1-\frac{1}{t_dI}\sum_i I_{s,i} \left[t +t_{r,i}\ln\left(\frac{1+c_i}{{\rm e}^{t/t_{r,i}}+c_i}\right) \right] \qquad c_i\equiv {\rm e}^{t_d/t_{r,i}}-1, \label{fullsolution3} \end{equation} and \begin{equation} Q=\frac{1}{I}\sum_i I_{s,i}. \end{equation} For simplicity I continue to consider the response of a star with a single pinning zone, since the results are easily generalized for multiple pinning zones. In general, the response of the crust has non-linear dependence on the size of the glitch through the appearance of $t_d$, and is referred to as the ``non-linear creep regime'' by \citet{alpar_etal89}. The integral of Equation [\ref{fullsolution2}] can be evaluated analytically in two limits. If the glitch is small enough that $t_d<<t_r$, expanding the integrand to first order in $t_d/t_r$ gives \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}=1-Q(1-{\rm e}^{-t/t_r}) \qquad t_d<<t_r, \label{exponential} \end{equation} as for the case for linear frictional coupling (Equation \ref{friction}), but now with a restriction on the glitch magnitude; the response to the glitch consists of {\em exponential relaxation} over a timescale $t_r$. The residual in the spin-down rate, $\delta\dot{\Omega}_c(t)$, is \begin{equation} \frac{\delta\dot{\Omega}_c(t)}{\vert\dot{\Omega}_0\vert}= -Q\frac{\Delta\Omega_c}{\vert\dot{\Omega}_0\vert t_r}{\rm e}^{-t/t_r}. \end{equation} For a single $S$ component, the glitch-induced discontinuity in the spin-down rate is \begin{equation} \frac{\delta\dot{\Omega}_c(0)}{\vert\dot{\Omega}_0\vert}= -\frac{I_s}{I}\frac{\Delta\Omega_c}{\vert\dot{\Omega}_0\vert} \frac{t_{sd}}{t_r}. \end{equation} In the opposite limit of a large glitch, so that $t_d>>t_r$, the integrand of Equation [\ref{fullsolution2}] becomes a Fermi function \begin{equation} \frac{1}{1+({\rm e}^{t_d/t_r}-1){\rm e}^{-t^\prime/t_r}}\simeq \frac{1}{1+{\rm e}^{-(t^\prime-t_d)/t_r}}, \end{equation} and the response of the crust is \begin{equation} \frac{\delta\Omega_c(t)}{\Delta\Omega_c}= 1-Q\left[\frac{t}{t_d} -\frac{t_r}{t_d}\ln\left(1+{\rm e}^{(t-t_d)/t_r}\right)\right] \qquad t_d>>t_r, \label{nonexponential} \end{equation} The star recovers through {\em delayed response}, as shown by the dashed curve in Figure \ref{response_resid}. The residual in the spin-down rate is \begin{equation} \frac{\delta\dot{\Omega}_c(t)}{\vert\dot{\Omega}_0\vert}= -Q\frac{\Delta\Omega_c} {\vert\dot{\Omega}_0\vert}\frac{1}{1+{\rm e}^{(t-t_d)/t_r}}= -\frac{I_s}{I_c}\frac{1}{1+{\rm e}^{(t-t_d)/t_r}} \qquad t_d>>t_r. \label{eq87} \end{equation} Right after the glitch, the magnitude of the spin rate has increased by $\vert\delta\dot\Omega_c\vert/\vert\dot\Omega_c\vert=I_s/I_c$. This constant spin rate excess persists for a time $t_d$, followed by relatively quick recovery over a timescale $t_r$, as shown by the dashed curve in Figure \ref{response_dresid}. Examples of the general behavior of $\delta\Omega_c(t)$ and $\delta\dot{\Omega}_c(t)$ are shown in Figure \ref{response_resid} and \ref{response_dresid}. The characteristic relaxation time of the system is $\tau=t_d+t_r$. Significant deviations from exponential response occur for $t_d\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} 3t_r$, corresponding to $\tau\hbox{${_{\displaystyle>}\atop^{\displaystyle\sim}}$} t_d$. The recoupling time $t_r$ depends on the value of the equilibrium lag $\omega_0$. The equilibrium lag is given by the solution to Equations [\ref{equilibrium}] and [\ref{A}]: \begin{equation} 1-\frac{\omega_0}{\omega_{\rm crit}}= \left[ \frac{kT}{5.1{\cal T}^{1/2}E_p}\ln\left(2t_{sd}\,\omega_0\sin 2\theta_d\,\right)\right]^{4/5}, \label{eqlag} \end{equation} where $\Omega_s\simeq\Omega_c$ was used. For typical parameters, the right-hand side of Equation [\ref{eqlag}] is much smaller than unity, so that $\omega_0\simeq\omega_{\rm crit}$; $\omega_0$ can be replaced by $\omega_{\rm crit}$ to a very good approximation. Equations [\ref{tt}], [\ref{eom}], and [\ref{A}] give the recoupling time \begin{equation} t_r=0.22\frac{I_c}{I}\omega_{\rm crit}\vert\dot{\Omega}_0\vert^{-1} ({\cal T}^{1/2}\beta E_p)^{-4/5} \left[\ln(2t_{sd}\omega_{\rm crit}\,\sin 2\theta_d)\right]^{-1/5}. \label{trfinal} \end{equation} Note the extremely weak dependence on $t_{sd}$ and $\theta_d$, and linear dependence on $\vert\dot\Omega_0\vert^{-1}$. Unlike $t_d$, the recoupling time $t_r$ does not depend on the conditions after the glitch; $t_r$ is intrinsic to the system (or part of the system), and does not depend the simple $C$+$S$+$L$ model used to calculate $t_d$. The chief mathematical results of the vortex slippage theory developed in this paper are given by Equations [\ref{td}], [\ref{fullsolution3}], and [\ref{trfinal}]. I now proceed with numerical estimates. \section{Estimates of Post-glitch Recovery Timescales} \label{timescales} The decoupling time in regions of star in which there is no vortex motion at the time of the glitch is, from Equation [\ref{td}], \begin{equation} t_d=7\,\left(\frac{t_{sd}}{10^4\mbox{ yr}}\right) \left(\frac{\Delta\nu/\nu}{10^{-6}}\right) \mbox{ days} \label{tdest} \end{equation} where $\nu$ is the spin frequency of the pulsar and $\Delta\nu$ is the change at the glitch. For large glitches in older stars, $t_d$ can be very long; $\Delta\nu/\nu=3\times 10^{-6}$ gives $t_d=210$ d for $t_{sd}=10^5$ yr. In the region or regions of the star in which vortex motion drives the glitch, $t_d$ can be much longer; see below. The recoupling time $t_r$ is determined by the pinning parameters. For vortex slippage in the inner crust, $t_r$ follows from Equations [\ref{trfinal}], [\ref{lp}], [\ref{omc_crust}], and [\ref{Tcal1}]. Taking $\sin 2\theta_d=10^{-2}$ for dissipation through Kelvin phonon production (though the exact value is unimportant), $t_r$ in terms of fiducial values is \begin{equation} t_r\simeq 30\, \left(\frac{\vert\dot{\nu}\vert}{10^{-11}\mbox{ s$^{-2}$}}\right)^{-1} \left(\frac{f_c}{0.1}\right)^{-1.7}\left(\frac{\rho_{n,14}}{0.5}\right)^{-1.7} E_p(\mbox{MeV})^{0.9}\times \left(\frac{kT}{10\mbox{ keV}}\right)^{4/5} \label{trcrust} \end{equation} \[ \hspace*{3.cm} \left(\frac{kT}{10\mbox{ keV}}\right)^{4/5} \left(\frac{a}{50\mbox{ fm}}\right)^{-0.9} \left(\frac{\xi_n}{10\mbox{ fm}}\right)^{-9/5} \left(\frac{R}{10\mbox{ km}}\right)^{-1} \mbox{ yr}. \] The population of glitching pulsars shows a range in $\dot{\nu}$ from $\sim -10^{-15}$ s$^{-2}$ to $\sim -10^{-9}$ s$^{-2}$, giving values of $t_r$ from $\sim 10^2$ d to decades. Nuclear entrainment has the important effect of making $t_r$ very long by reducing the density of free neutrons. The Magnus force (for given lag) that drives the system's recovery to rotational equilibrium is correspondingly reduced, so the recovery time is increased. Recoupling of the core through vortex slippage occurs much more quickly than in the inner crust. Using Equations [\ref{lphi}], [\ref{omc_core}], and [\ref{Tcal}] in Equation [\ref{trfinal}], and taking fiducial values for the outer core, gives: \[ t_r\simeq 3\, \left(\frac{\vert\dot\nu\vert}{10^{-11} \mbox{ s$^{-2}$}}\right)^{-1} \left(\frac{\rho_{n,14}}{4}\right)^{-7/5} \left(\frac{kT}{10\mbox{ keV}}\right)^{4/5} \left(\frac{E_p}{100\mbox{ MeV}}\right)^{3/5}\times \hspace*{3.cm} \] \begin{equation} \left(\frac{\Lambda_*}{50\mbox{ fm}}\right)^{-9/5} \left(\frac{R}{10\mbox{ km}}\right)^{-1} B_{12}^{3/10} \mbox{ days}. \label{trcore} \end{equation} While the dipole components of the field typically have $B_{12}=1$, it is the internal field that determines the recoupling time. With relatively weak sensitivity to $E_p$ and $B$, this estimate is more robust than for the inner crust. Higher multipoles could plausibly make $B$ about ten times larger than the surface field \citep{braithwaite09}. For large glitches, $t_r$ is typically less than $t_d$, and so the response will be the delayed response illustrated by the dashed curves of Figures \ref{response_resid} and \ref{response_dresid}; see the discussion following Equation [\ref{eq87}]. If a moment of inertia $\Delta I_s$ is decoupled by the glitch, post-glitch response of the core consists of a fixed increase in the magnitude of the spin-down rate by a factor $\Delta I_s/I_c$, followed by a relatively quick recovery over a timescale $t_r$, after time $t_d$. The large vortex self-energy acts to reduce the recoupling time $t_r$; see Equation [\ref{trfinal}]. Increasing the tension by a factor of 10 over that used in Equation [\ref{tension}] decreases $t_r$ by a factor of $\simeq 5$ in the inner crust and $\simeq 3$ in the outer core. Vortex tension generally increases vortex mobility, giving shorter recoupling times. The decoupling time $t_d$ is unaffected. The estimate of $t_d$ in Equation [\ref{tdest}] applies to all parts of the star in which pinned vortices do not move during the glitch. If some portion $S^\prime$ of the superfluid interior with a moment of inertia $I_s^\prime$ drives the glitch, and $I_s^\prime$ is much less than $I_c$, the decoupling time of this region could be very long. By angular momentum conservation \begin{equation} I_c\Delta\Omega_c + I_s^\prime \Delta\Omega_s=0, \end{equation} giving a change in lag in this region given by \begin{equation} I\Delta\Omega_c+I_s^\prime\Delta\omega=0, \end{equation} and a decoupling time of this region of \begin{equation} t_d^\prime=\frac{\vert\Delta\omega\vert}{\vert\dot\Omega_c\vert}= \frac{I}{I_s^\prime}\frac{\Delta\Omega_c}{\vert\dot\Omega_0\vert} =2\, \left(\frac{I/I_s^\prime}{10^{-2}}\right) \left(\frac{t_{sd}}{10^4\mbox{ yr}}\right) \left(\frac{\Delta\nu/\nu}{10^{-6}}\right) \mbox{ yr} \end{equation} Such a region could serve as the $L$ component in the simple model of this paper as follows. After the glitch, $S^\prime$ remains decoupled from $C$ for a time $t_d^\prime$, and stores angular momentum as the star spins down. After a time $t_d^\prime$, $S^\prime$ recovers to spin-down equilibrium over the recoupling time $t_r^\prime$ of that region. An instability drives vortex motion in $S^\prime$, creating a glitch, and the process repeats itself. Parts of the inner crust with typically very long coupling times, could serve as the $L$ component. For all other parts of the star in which there is pinning, $t_d$ is the same, and is set by the size of the glitch according to Equation [\ref{tdest}]. \section{Comparison with observed post-glitch response} \label{data} I now give a brief comparison for two pulsars for which post-glitch response has been measured in detail: the Vela and Crab pulsars. Vela has shown post-glitch response times that range between about 0.5 d and about 400 d. The Crab pulsar has shown post-glitch response times that range from about two days to 100 d. Most glitches show multiple timescales; see the ATNF Pulsar Glitch Catalogue \citep{manchester_etal05} and references therein. \subsection{Inner-crust slippage} Equation [\ref{trcrust}] predicts $t_r\simeq 10^4$ d for the Vela pulsar, with a relatively small addition to the total relaxation of $t_d\simeq 20$ d for a typical glitch. For the Crab pulsar the predicted recoupling time is $t_r\simeq 300$ d, with $t_d<<t_r$ because the glitches are generally small. It appears that post-glitch response due to vortex slippage in the inner crust is too slow to reconcile with observations in these two pulsars. Note that $t_r$ decreases by a factor of about 50 if entrainment is negligible for some reason. \subsection{Outer-core crust slippage} For the Vela pulsar, the total relaxation time $t_d+t_r$ from Equations [\ref{tdest}] and [\ref{trcore}] is not in significant disagreement with observed relaxation times but the predictions for the Crab pulsar are $\tau<0.1$ d, far shorter than observed; details will be given in a subsequent publication. \bigskip These preliminary comparisons suggest that most (if not all) aspects of post-glitch response are not due to thermally-activated vortex activation, but arise from a different process. The dominant process in post-glitch response might be vortex drag in regions were there is no pinning. \section{Comparison with Earlier Work} \label{comparison} The theory developed here differs significantly from the vortex creep theory of AAPS applied to the inner crust (see, also, \citealt{alpar_etal84b,alpar_etal85,alpar_etal86,alpar_etal88,alpar_etal89,alpar_etal93,alpar_etal94,alpar_etal96}) and the extensions of the AAPS theory to vortex creep in the outer core \citep{sa09}, and gives different predictions. The differences are due to different treatments of the vortex velocity and different forms for the activation energy. AAPS and \citet{alpar_etal84b} adopt the following expression for the radial component of the creep velocity: \begin{equation} \langle\vbf_v\rangle\cdot\hat{r}=v_0{\rm e}^{-\beta E_p(1-\omega/\omega_{\rm crit})} -v_0{\rm e}^{-E_p\beta (1+\omega/\omega_{\rm crit})} =2v_0 {\rm e}^{-\beta E_p}\sinh\left(\beta E_p\frac{\omega}{\omega_{\rm crit}}\right); \label{valpar} \end{equation} {\sl c.f.}, Equation [\ref{vslippage}]. The first term accounts for motion away from the rotation axis, and the second term for motion toward the rotation axis. The factor $v_0$ is a velocity that AAPS refer to as ``a typical velocity of microscopic motion of the vortex lines between pinning centers''; a calculation of $v_0$ is not given, and $v_0$ is fixed at a constant value of $10^7$ cm s$^{-1}$ for both the inner crust and outer core. The activation energies for outward (+) and inward (-) vortex motion are assumed to be \begin{equation} A_\pm=E_p(1\mp\omega/\omega_{\rm crit}). \label{Aalpar} \end{equation} This form assumes that vortices are moving in a {\em linear} bias imposed by the average Magnus force; {\sl c.f.}, Equation [\ref{A}]. Note that vortex tension does not appear in Equations [\ref{valpar}] and [\ref{Aalpar}]; each pinned segment of a vortex is treated effectively as a point particle, without coupling of adjacent segments through tension. The chief conclusions of the vortex creep theory of AAPS and its extensions are these: \begin{enumerate} \item For relatively large $E_p$, spin-down drives the system to an equilibrium lag $\omega_0\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$}\omega_{\rm crit}$. The radial component of the creep velocity becomes, in the theory of AAPS: \begin{equation} \langle\vbf_v\rangle\cdot\hat{r}=v_0{\rm e}^{-\beta E_p(1-\omega/\omega_{\rm crit})} \qquad \omega_0\hbox{${_{\displaystyle<}\atop^{\displaystyle\sim}}$}\omega_{\rm crit}. \label{valpar_nonlinear} \end{equation} In the nomenclature of \citet{alpar_etal89}, this limit is the regime of ``non-linear creep''. In this regime, AAPS obtain a decoupling time \begin{equation} t_d=2t_{sd}\frac{\Delta\Omega_c}{\Omega_c}, \end{equation} as in Equation [\ref{td}]. AAPS obtain a relaxation time, for perturbations about spin-down equilibrium, of \begin{equation} t_r=\frac{\omega_{\rm crit}}{\vert\dot{\Omega}_0\vert}\frac{kT}{E_p} =\frac{kT}{\vert\dot\Omega_0\vert\rho_s\kappa r \xi_n a} \qquad\qquad \mbox{``non-linear creep''}. \label{traaps} \end{equation} As in this paper, post-glitch response consists of exponential decay if $t_r>>t_d$ and delayed response if $t_d>>t_r$. An important difference between the results of this paper and those of AAPS is the dependence of the $t_r$ on $E_p$. In the vortex creep theory of AAPS, $\omega_{\rm crit}\propto E_p$, giving a relaxation time that is {\em independent of $E_p$}. Equation [\ref{traaps}] from the theory of AAPS gives \begin{equation} t_r\simeq 8\, (f_c \rho_{n,14})^{-1} \left(\frac{\vert\dot\nu\vert}{10^{-11}\mbox{ s$^{-2}$}}\right)^{-1} \left(\frac{kT}{\mbox{10 keV}}\right) \left(\frac{\xi_n}{\mbox{10 fm}}\right)^{-1} \left(\frac{a}{\mbox{50 fm}}\right)^{-1} \left(\frac{R}{\mbox{10 km}}\right)^{-1} \mbox{ yr}, \end{equation} far longer than observed, especially if nuclear entrainment with $f_c<0.1$ is taken into account; nuclear entrainment could increase this timescale to over a century. These timescales are incompatible with observations. \item In the opposite limit of low equilibrium lag, $\omega_0 <<\omega_{\rm crit}$, corresponding to relatively small $E_p$, Equations [\ref{valpar}] and [\ref{Aalpar}] give for the theory of AAPS: \begin{equation} \langle\vbf_v\rangle\cdot\hat{r}=2v_0\beta\, E_p{\rm e}^{-\beta E_p}\frac{\omega}{\omega_{\rm crit}} \qquad \omega_0<<\omega_{\rm crit}. \label{valpar_linear} \end{equation} In the nomenclature of \citet{alpar_etal89}, this limit is the regime of ``linear creep''. In this regime, post-glitch response consists of exponential decay over a timescale that {\em depends exponentially on $E_p$}: \begin{equation} t_r=\frac{r\omega_{\rm crit}}{4\Omega_s v_0}\left(\beta E_p\right)^{-1}{\rm e}^{\beta E_p} \qquad\qquad \mbox{``linear creep''}, \label{taulinear} \end{equation} and the decoupling time $t_d$ is effectively zero. Equation [\ref{taulinear}] gives much shorter coupling times than Equation [\ref{traaps}]; consequently, the regime of ``linear creep'' was invoked by \citet{alpar_etal93,alpar_etal94,alpar_etal96} to explain observed post-glitch response. \end{enumerate} The ``linear regime'' of \citet{alpar_etal89} depends on competition between the two terms in Equation [\ref{valpar}] that gives linear dependence on $\omega$ for $\omega<<\omega_{\rm crit}$. As shown in Appendix \ref{antiparallel}, though, thermal activation of vortex segments {\em against} the Magnus force is prevented by the large vortex tension unless $\sin 2\theta_d$ is orders of magnitude smaller then expected. Hence, the second term in Equation [\ref{valpar}] should not be present, and the ``linear regime'' introduced by \citet{alpar_etal88,alpar_etal89} is not physically realized. The analysis of this paper gives, instead, \begin{equation} \langle\vbf_v\rangle\cdot\hat{r}=\frac{1}{2}r\omega\sin 2\theta_d\, {\rm e}^{-\beta A(\omega)}, \end{equation} for any lag. The above expression does not have a linear expansion about $\omega=0$.\footnote{For $\omega<<\omega_{\rm crit}$, the activation energy diverges as $\omega^{-1}$ \citep{le91}.} The ``linear creep'' regime was invoked to fit glitch data by \citet{alpar_etal93,alpar_etal94,alpar_etal96}. In fitting to eight glitches in the Vela pulsar, \citet{alpar_etal93} included three linear response terms, that is simple exponentials with no time offsets, with relaxation times of $\sim 30$ d and $\sim 3$ d, and also 10 hours for the glitch of December 24, 1988. The three contributions to the linear response are assumed to correspond to regions through which no vortex motion occurred, and the pinning is ``superweak''. Under this assumption, the change in lag at the glitch for each region is $\omega_0-\omega(0+)=\Delta\Omega_c$. Assuming corotation of the core, as did \citet{alpar_etal93}, the decoupling time for each region follows from Equation [\ref{tt}]. The glitch magnitudes were in the range $1.14\times 10^{-6}\le \Delta\Omega_c/\Omega_c\le 3.06\times 10^{-6}$, giving a range of decoupling times $8\mbox{ d}\le t_d\le 21\mbox{ d}$, significantly larger than the relaxation times of 3 d and 10 hours. Hence, the glitches were too large to exhibit exponential recovery according to the treatment of this paper. Whether or not the second term in Equation [\ref{valpar}] is retained, the creep velocity of AAPS violates the vortex equations of motion, Equations [\ref{vv}] and [\ref{vslippage}]; Equation [\ref{vslippage}] contains a factor of $\omega$ that multiplies the exponential. AAPS effectively fixed \begin{equation} \frac{v_s}{2}\sin 2\theta_d=\frac{r\omega}{2}\sin 2\theta_d=v_0=10^7\mbox{ cm s$^{-1}$}, \label{v0} \end{equation} For a critical lag of $\omega_{\rm crit}=4$ \hbox{\rm\hskip.35em rad s}$^{-1}$\ (Equation \ref{omc_crust}), $v_0$ cannot exceed $2\times 10^6$ \hbox{\rm\hskip.35em cm s}$^{-1}$\ for any drag process; Equation [\ref{v0}] corresponds to unphysically large drag. For ``non-linear creep'' the value of $v_0$ is not very important, since the assumption of spin-down equilibrium gives a weak, logarithmic dependence on the prefactor in Equation [\ref{vslippage}]; see Equation [\ref{trfinal}]. More significant is the choice of AAPS of the activation energy of Equation [\ref{Aalpar}], which gives a recoupling time $t_r$ that is independent of of $E_p$ in this limit, so that the recoupling time depends only on the fact that there is pinning, and not on the strength of the pinning. Equation [\ref{Aalpar}] used by \citet{alpar_etal84a} does not follow from consideration of a realistic pinning potential or the large vortex self energy. These considerations give a dependence of the activation energy on $(1-\omega/\omega_{\rm crit})^{5/4}$ for a parabolic pinning force, giving $t_r\propto E_p^{9/10}$ for the inner crust and $t_r\propto E_p^{3/5}$ for the outer core. As Equation [\ref{fullsolution}] shows, the response of the crust to a glitch generally depends non-linearly on the the magnitude of glitch, through the appearance of $t_d$. Exponential decay occurs only in the limit $t_d<<t_r$, as for a small glitch. For large glitches, exponential decay occurs only if $t_r$ exceeds $t_d$. For a large glitch ($\Delta\nu/\nu=3\times10^{-6}$), the decoupling time is $21\,(t_{sd}/10^4\mbox{ yr})$ d; in regions with pinning, {\em exponential decay can occur only over longer timescales than the decoupling time.} \citet{sa09} studied response of the outer core assuming ``linear creep'', a regime that is not realized for physically reasonable parameters (see Appendix \ref{antiparallel}). With Equation [\ref{valpar_linear}], \citet{sa09} obtained a sub-day relaxation time assuming $E_p\simeq 0.1$ MeV, much smaller than the estimate of 100 MeV discussed in Section \ref{pinning}. From the analysis of Appendix \ref{transition}, pinning does not exist for such a low value of $E_p$. For larger values of $E_p$ in the pinning regime, $t_r$ can be short ({\it e.g.}, $E_p=1$ MeV gives $t_r=3$ hours for Vela), but $t_d$ will still be approximately one week for $\Delta\nu/\nu=10^{-6}$. \section{Conclusions} \label{predictions} Two chief conclusions of this paper are: \begin{enumerate} \item The ``linear creep'' regime introduced by \citet{alpar_etal88,alpar_etal89} and invoked in fits to post glitch response \citep{alpar_etal93,alpar_etal94,alpar_etal96} is not realized for physically reasonable parameters, a conclusion that strongly constrains possible post-glitch response through thermal activation. The validity of fits of vortex creep theory to data, and the conclusions drawn about pinning parameters, should be reassessed. \item A regime of ``superweak pinning'', crucial to the theory of AAPS and its extensions, is probably precluded by thermal fluctuations. \end{enumerate} The vortex slippage theory given here has clear, falsifiable predictions: \begin{enumerate} \item Post-glitch recovery associated with regions in which there is vortex pinning, whether in the inner crust or outer core, requires a time that is {\em at least} as long as the decoupling time: \begin{equation} \mbox{Recovery time } > t_d\simeq 7\,\left(\frac{t_{sd}}{10^4\mbox{ yr}}\right) \left(\frac{\Delta\nu/\nu}{10^{-6}}\right) \mbox{ days}. \end{equation} This timescale is independent of the details of vortex pinning, and so is a robust conclusion within the framework of thermally-activated, post-glitch response. \item Post-glitch response depends non-linearly on the size of the glitch. For a glitch that is sufficiently large that $t_d>>t_r$ is satisfied, the response will consist of {\em delayed response} over a time $t_d$, wherein the star {\em spins down at a constant, higher rate, before relaxing over a time $t_r$}. Detection of delayed response over a timescale proportional to the glitch magnitude would represent strong evidence for thermally-activated vortex slippage in neutron stars. \item Exponential decay can occur only if the glitch is sufficiently small that $t_d<<t_r$ is satisfied. Hence, exponential decay can occur only over timescales that exceed $t_d$, which is determined by the size of the glitch. Quicker decay is prevented by the non-linear nature of coupling through thermal activation. This conclusion is an important difference between vortex slippage theory and vortex creep theory. \item Response over a relatively quick timescale $<t_d$ as has been observed in some pulsars, most notably Vela, cannot be due to thermally-activated vortex motion but must represent a different process, such as drag on vortices in regions where there is no pinning. \end{enumerate} Comparison of these predictions with the wealth of glitch data now available will be given in a forthcoming publication, and constraints upon the theory will be obtained. \begin{acknowledgments} I am grateful to M. A. Alpar, D. Antonopoulou, N. Chamel, and C. Espinoza for very useful discussions, and to R. I. Epstein for valuable suggestions on the manuscript. I also thank the anonymous referee for very useful criticism. This work was supported by NSF Award AST-1211391 and NASA Award NNX12AF88G. \end{acknowledgments}
\section{Introduction} Interest in the study of analysis in fractals has increased since the publication of Kigami's papers \cite{Kigami89} and \cite{Kigami93}. In particular, there has been interest in the explicit construction of harmonic functions and the eigenfunctions of the Laplacian of a postcritically finite (PCF) set. In \cite{DSV99}, Dalrymple, Strichartz and Vinson described algorithms for the construction of harmonic functions and the eigenfunctions in the Sierpinski gasket. The construction of harmonic functions is achieved by the general algorithm for PCF sets described in \cite[Chapter 3]{Kigami}, while the construction of the eigenfunctions is achieved by decimation \cite{Shima91} (see also \cite[Chapter 3]{StrichartzDEF} for a detailed explanation of the decimation method). The explicit construction of harmonic functions or eigenfunctions has been done in other fractals, as the Vicset set \cite{CSW11}, where one also has decimation, and the pentakun \cite{ASST03}, where one does not have decimation. In \cite{DSV99}, the authors also described algorithms for the restriction of harmonic functions and the eigenfunctions to the edges of the Sierpinski gasket, allowing us to visualize them as functions on the interval $[0,1]$. Demir, Dzhafarov, Ko{\c{c}}ak and {\"U}reyen \cite{DDKU07} observed that these functions have zero derivatives on a dense subset of $[0,1]$. Later De Amo, D\'iaz Carrillo and Fern\'andez S\'anchez \cite{ADF13} proved that such restrictions are singular functions whenever they are monotone, {\em i.$\,$e.$\!$} that have zero derivatives almost everywhere. In this paper we construct the harmonic functions and the Dirichlet eigenfunctions for the harmonic structure of the Hata set \cite{Hata85}, and their restrictions to the interval $[0,1]$, the longest edge contained in the set. Since the Hata set does not have decimation, we have to construct the eigenfunctions by the finite element method \cite{DSV99}. Moreover, the Hata set has a natural family of harmonic structures, so we study the properties of these restrictions with respect to the parameter of the family. In section \ref{structuresection}, we describe the family of harmonic structures of the Hata set, as well as explicitly calculate its Laplacians. In section \ref{harmonicsection} we construct the harmonic functions, its restrictions, and study whether these restrictions are singular functions. In section \ref{Laplaciansection} we explicitly describe the Laplacian with respect to a self-similar homogenous measure, and in section \ref{Dirichletsection} we study the Dirichlet eigenfunctions, as well as its restrictions, and give numerical evidence to decide whether they are also singular. \section{Harmonic structure}\label{structuresection} The Hata tree set is the unique compact set $K\subset\mathbb{C}$ such that \[ K = F_1(K) \cup F_2(K), \] where the functions $F_1, F_2$ are given by \[ F_1(z) = \alpha \bar z \qquad\text{and}\qquad F_2(z) = (1 - |\alpha|^2)\bar z + |\alpha|^2, \] and $\alpha\in\mathbb{C}$ is such that $0 < |\alpha|,|1-\alpha| < 1$ \cite{Hata85,YHK97}. Observe that the points $0$ and $1$ are the fixed points of $F_1$ and $F_2$, respectively, $\alpha = F_1(1)$, and \[ |\alpha|^2 = F_1(\alpha) = F_2(0). \] Hence, the critical set is given by $\mathcal C = \{|\alpha|^2\}$ and the post-critical set, its \emph{boundary}, is \[ V_0 = \{\alpha, 0, 1\}. \] (See Figure \ref{Hata-fig}.) \begin{figure}[ht] \includegraphics[width=3in]{Hata.pdf} \caption{Hata tree set, with $\alpha=1/2 + \sqrt{3}i/6$. The critical set consists of the point $|\alpha|^2 = 1/3$.} \label{Hata-fig} \end{figure} We will denote the points $\alpha, 0$ and $1$ by $p_0, p_1$ and $p_2$, respectively. We observe that $K\setminus V_0$ is disconnected. We call $L$ the closure of the connected component of $K\setminus V_0$ that contains the interval $(0,1)$, and $M$ the closure of the component that contains the open segment from $0$ to $\alpha$. If $w\in W_m$, we define $p_{wi} = F_w(p_i)$, for $i=0,1,2$. We note that $p_0 = p_{12}$, and that \[ |\alpha|^2 = p_{10} = p_{21}. \] As points in $V_m$, $m\ge1$, we see that $p_0 = p_{12\ldots2}$, $p_1=p_{1\ldots1}$ and $p_2 = p_{2\ldots2}$. If $w\not=w'$ are in $W_m$ and $p\in K_w\cap K_{w'}$, then either \[ p = p_{w1} = p_{w'0} \qquad\text{or}\qquad p = p_{w1} = p_{w'2}. \] We have the former case if $p\in V_m\setminus V_{m-1}$, and the latter if $p\in V_{m-1}$. If $p\in K_w$, and in no other $K_{w'}$, then \[ p = p_{w0} \qquad\text{or}\qquad p = p_{w2}. \] Again, we have the former case if $p\in V_m\setminus V_{m-1}$, and the latter if $p\in V_{m-1}$ (and $p\not= p_1$). A point in $V_m$ has only one adyacent vertex if it's of the form $p_{w20}$ or $p_{w22}$; otherwise it has three adyacent vertices in $V_m$, except for $p_1$, which has only two.\footnote{We note that, since $p_0 = p_{12}$, $p_{w20} = p_{w212}$ and $p_{w10} = p_{w112}$. Thus, points of the form $p_{w12}$ may have one or three adyacent vertices, depending on the last term of $w$.} To construct a harmonic structure on the Hata set $K$, we need a Laplacian $D$ on $V_0$. Using the standard base $\{\chi_\alpha, \chi_0, \chi_1\}$, we set \[ D = \begin{pmatrix} -h & h & 0\\ h & -(h+1) & 1\\ 0 & 1 & -1 \end{pmatrix}. \] Then, if $\mathbf r = (r_1, r_2)$, $(D,\mathbf r)$ is a regular harmonic structure \cite{Kigami} for $K$ if $h > 1$, \[ r_1 = \frac{1}{h} \qquad\text{and}\qquad r_2 = 1 - \frac{1}{h^2}. \] Explicitly, the Laplacian $H_m$ at a point $p\in V_m\setminus V_0$ is given by, if $p=p_{w1}=p_{w'0}$, \begin{subequations}\label{LaplacianPoint} \begin{equation}\label{LaplacianPointNew} \begin{split} H_mf(p) = \frac{1}{r_w} & \Big(f(p_{w2}) - f(p_{w1}) + h\big(f(p_{w0}) - f(p_{w1})\big)\Big) \\ &+ \frac{h}{r_{w'}}\Big(f(p_{w'1}) - f(p_{w'0})\Big); \end{split} \end{equation} if $p = p_{w1} = p_{w'2}$, \begin{equation}\label{LaplacianPointOld} \begin{split} H_mf(p) = \frac{1}{r_w} & \Big(f(p_{w2}) - f(p_{w1}) + h\big(f(p_{w0}) - f(p_{w1})\big)\Big) \\ &+ \frac{1}{r_{w'}}\Big(f(p_{w'1}) - f(p_{w'2})\Big); \end{split} \end{equation} if $p=p_{w0}$ and in no other cell, \begin{equation}\label{LaplacianPointNewOne} H_mf(p) = \frac{h}{r_{w}}\Big(f(p_{w1}) - f(p_{w0})\Big); \end{equation} and, if $p=p_{w2}$ and in no other cell, \begin{equation}\label{LaplacianPointOldOne} H_mf(p) = \frac{1}{r_{w}}\Big(f(p_{w1}) - f(p_{w2})\Big). \end{equation} \end{subequations} The Laplacian at the points $p_0$ and $p_2$ is given by formula \eqref{LaplacianPointOldOne} for $m\ge 1$, while for $p=p_1$ is given by \[ H_mf(p)= \frac{1}{r_w}\Big( f(p_{w2} - f(p_{w1}) + h\big(f(p_{w0}) - f(p_{w1})\big) \Big), \] $w=(1,1,\ldots,1)\in W_m$ is the word with $m$ ones. Note that $r_w = r_1^m$. Observe that we have a family of harmonic structures for $K$, parameterized by $h$. For each $h>1$, the Hausdorff dimension with respect to the effective resistance metric \cite{Kigami} is the unique $d$ such that \[ \Big(\frac{1}{h}\Big)^d + \Big(1 - \frac{1}{h^2}\Big)^d = 1. \] We note that, since $F_1$ and $F_2$ are affine linear contractions with constants $|\alpha|$ and $1 - |\alpha|^2$, respectively, $d$ coincides with the Hausdorff dimension with respect to the Euclidean metric if $h = 1/|\alpha|$, by Hutchinson theorem \cite{Hutchinson}. \section{Harmonic functions}\label{harmonicsection} A function $u$ on $V_*$ is harmonic if $H_mu(p)=0$ for any $p\in V_m\setminus V_0$. In this section we describe an algorithm to construct harmonic functions on $V_*$ from any boundary values on $V_0$. We say that $T_w\subset V_m$ is a \emph{minimal cell in $V_m$}, if $T_w$ a set of three vertices of the form $p_{w0}$, $p_{w1}$ and $p_{w2}$, with $w\in W_m$. As points in $V_{m+1}$, we have that \[ T_w = \{p_{w12}, p_{w11}, p_{w22}\}, \] and the ``new points'' in $V_{m+1}$, contained in $K_w$, are then $p_{w10} = p_{w21}$ and $p_{w20}$. In other words, $K_w\cap V_{m+1}$ is the union of the minimal cells in $V_{m+1}$ \[ T_{w1} = \{p_{w10}, p_{w11}, p_{w12}\} \qquad\text{and}\qquad T_{w2} = \{p_{w20}, p_{w21}, p_{w22}\} \] Given a harmonic function on $V_m$, we want to extend to a harmonic function on $V_{m+1}$. The extension from $T_w$ to $T_{w1}\cup T_{w2}$ is given by the following algorithm. \begin{algorithm}\label{alg-harm} Let $u$ be a harmonic function on $V_m$. If, for each $w\in W_m$, $T_w$ is a minimal cell in $V_m$, and we extend $u$ to $T_{w1}\cup T_{w2}$ with \begin{subequations}\label{alg-eqn} \begin{eqnarray} u(p_{w10}) &=& \Big(1 - \frac{1}{h^2}\Big) u(p_{w1}) + \frac{1}{h^2} u(p_{w2}),\label{alg-eqn10}\\ u(p_{w20}) &=&u(p_{w10}),\label{alg-eqn20} \end{eqnarray} \end{subequations} then $u$ is harmonic in $V_{m+1}$. \end{algorithm} \begin{proof} With respect to the basis $\{\chi_{p_{12}},\chi_{p_{11}},\chi_{p_{22}},\chi_{p_{10}},\chi_{p_{20}}\}$, the matrix of $H_1$ is given by \[ H_1 = \begin{pmatrix} -\dfrac{1}{r_1} & \dfrac{1}{r_1} & 0 & 0 & 0\\ \dfrac{1}{r_1} & - \dfrac{1+h}{r_1} & 0 & \dfrac{h}{r_1} & 0\\ 0 & 0 & - \dfrac{1}{r_2} & \dfrac{1}{r_2} & 0\\ 0 & \dfrac{h}{r_1} & \dfrac{1}{r_2} & -\dfrac{h}{r_1} - \dfrac{1+h}{r_2} & \dfrac{h}{r_2}\\ 0 & 0 & 0 & \dfrac{h}{r_2} & - \dfrac{h}{r_2} \end{pmatrix}. \] Writing the matrix as \[ \begin{pmatrix} T & J^t \\ J & X \end{pmatrix}, \] where $T$ takes functions on $V_0$ to functions on $V_0$, $J$ functions on $V_0$ to functions on $V_1$, and $X$ functions on $V_1$ to functions on $V_1$, Theorem 2.1.6 in \cite{Kigami} implies that, if $u$ is harmonic, then \[ u|_{V_1\setminus V_0} = -X^{-1}J(u|_{V_0}). \] Multiplying the matrices, and by the compatibility of the sequence of $H_m$ \cite{Kigami}, we obtain the result. \end{proof} Algorithm \ref{alg-harm} allows us to construct harmonic functions on the Hata set with arbitrary precision, with any particular value of the parameter $h$. Figure \ref{graphs-harmonic} shows examples of harmonic functions, with distinct boundary values and distinct values of $h$. \begin{figure}[ht] \includegraphics[width=4in]{harmonic.pdf} \caption{Harmonic functions on the Hata set. The three on the top correspond to $\alpha=1/2 + \sqrt{3}i/6$ and $h=2$, with boundary values $\chi_c$, $\chi_0$ and $\chi_1$, respectively. On the bottom, harmonic funcions with boundary values $\chi_1$, with values of $h$ equal to $3/2$, $\sqrt 3$ and $3$, respectively.} \label{graphs-harmonic} \end{figure} We observe that, since $K\setminus V_0$ is disconnected, we have harmonic functions supported in each one of the connected components $L$ and $M$ of $K\setminus V_0$, as we see in Figure \ref{graphs-harmonic} for harmonic functions with boundary values $\chi_1$ and $\chi_\alpha$, respectively. Moreover, from equation \eqref{alg-eqn10}, the value of a harmonic function on each point $p$ in the line segment from 0 to 1 only depends on its values on the adyacent points to $p$ in the same line, from a previous iteration. Thus, one can easily construct restrictions of such harmonic functions to the line segment, following the work of \cite{DSV99}. Figure \ref{restr-harmonic} shows examples of such restrictions with boundary values $\chi_1$. \begin{figure}[ht] \includegraphics[width=4in]{harmonic-restrictions.pdf} \caption{Restriction to $[0,1]$ of harmonic functions on the Hata set, with $\alpha=1/2 + \sqrt{3}i/6$, boundary values $\chi_1$ and values of $h$ equal to $3/2$, $\sqrt 3$ and $3$, respectively. Observe that, in the case $h=1/|\alpha|$, such restriction is a line. Otherwise, is a singular function.} \label{restr-harmonic} \end{figure} We note, as readily verified from equation \eqref{alg-eqn10}, that the restriction of a harmonic function to this line segment is a line if $h = 1/|\alpha|$, because the ``middle'' point in each iteration corresponds to the convex combination of its adyacent points in the line with weights $|\alpha|^2$ and $1 - |\alpha|^2$. However, if $h \not= 1/|\alpha|$, this restriction is a singular function, {\em i.$\,$e.$\!$} monotonic and with derivative 0 almost everywhere \cite{YHK97}. \begin{theorem}\label{harm-sing} Assume $h \not= 1/|\alpha|$ and let $u$ be a harmonic function on the Hata set with boundary values $u|_{V_0} = \chi_1$, and let $f$ be its restriction to the interval $[0,1]$. Then \begin{enumerate} \item $f$ is increasing; and \item $f$ is differentiable on $[0,1]\setminus V_*$, with \[ f'(x) = 0 \] for every $x\in[0,1]\setminus V_*$. \end{enumerate} \end{theorem} \begin{proof} Since $|\alpha|^2 = p_{10} = F_1(F_1(1))$, \[ F_1(F_1(x)) = |\alpha|^2 x \qquad\text{and}\qquad F_2(x) = (1 - |\alpha|^2)x + |\alpha|^2, \] we see that equation \eqref{alg-eqn10} implies that $f$ satisfies the system of equations \begin{equation*} \begin{cases} f(x) = \dfrac{1}{h^2} f\bigg(\dfrac{1}{|\alpha|^2}x\bigg) & 0 \le x \le |\alpha|^2\vspace*{.05in}\\ f(x) = \bigg( 1 - \dfrac{1}{h^2}\bigg) f\bigg(\dfrac{1}{|\alpha|^2}x - 1\bigg) + \dfrac{1}{h^2} & |\alpha|^2 \le x \le 1. \end{cases} \end{equation*} Hence $f$ is essentially the same as Lebesgue's singular function, and the result of the theorem follows as in \cite[Section 3.4]{YHK97}. \end{proof} We note that, if $h = 1/|\alpha|$, then we have $f(x) = x$. Recall that $d$ coincides with the Euclidean Hausdorff dimension precisely when $h = 1/|\alpha|$. \section{Laplacian}\label{Laplaciansection} In this section we calculate the Laplacian $\Delta$ on the Hata set $K$ with respect to the self-similar measure $\mu$ with weights \[ \mu_1 = \mu(F_1(K)) = r_1^d = \Big(\frac{1}{h}\Big)^d \quad\text{and}\quad \mu_2 = \mu(F_2(K)) = r_2^d = \Big( 1 - \frac{1}{h^2} \Big)^d. \] This measure is comparable to the Hausdorff measure with respect to the effective resistance metric \cite{Kigami94}. Moreover, $\mu$ is homogenous with respect to this metric \cite{Saenz12}. As in \cite[Section 3.7]{Kigami}, the domain $\mathcal{D}$ of the operator $\Delta$ is given by the set of continuous functions $u$ on $K$ such that there exists a continuous function $f$ with \begin{equation}\label{Laplacian} \lim_{m\to\infty} \max_{p\in V_m\setminus V_0} \big|\frac{1}{\mu_p^m}H_mu(p) - f(p) \big| = 0, \end{equation} where $\mu_p^m=\int_K \psi_p^m d\mu$, the integral of the $m$-harmonic spline $\psi_p^m$ that satisfies \[ \psi_p^m(p) = 1 \qquad\text{and}\qquad \psi_p^m(q) = 0 \text{ for }q\in V_m, q\not=p. \] If $u\in\mathcal{D}$ and $f$ is as in \eqref{Laplacian}, then we write $\Delta u = f$. We can then approximate explicitly $\Delta u$ once we calculate the normalizing numbers $\mu_p^m$. In order to calculate these numbers we observe that, by the self-similarity of $\mu$, \begin{equation*} \begin{split} \mu_p^m &= \int_K \psi_p^m d\mu = \sum_{\substack{w\in W_m\\p\in K_w}} \int_{K_w}\psi_p^m d\mu\\ &= \sum_{\substack{w\in W_m\\p\in K_w}} \mu_w \int_K \psi_{F_w^{-1}(p)}^0 d\mu = \sum_{\substack{w\in W_m\\p\in K_w}} \mu_w \mu_{F_w^{-1}(p)}^0, \end{split} \end{equation*} where $\mu_w = \mu_{w_1}\cdots\mu_{w_m}$. Thus it is sufficient to calculate the three numbers $\mu_\alpha^0, \mu_0^0$ and $\mu_1^0$ corresponding to $\alpha$, $0$ and $1$, the points in $V_0$. Again, using the self-similarity of $\mu$, we observe that, by Algorithm \ref{alg-harm} \[ \int_K \psi_\alpha^0 d\mu = \mu_1 \int_K \psi_1^0 d\mu, \] \begin{multline*} \int_K \psi_0^0 d\mu = \mu_1 \int_K \bigg(\Big(1 - \frac{1}{h^2}\Big)\psi_\alpha^0 + \psi_0^0 \bigg) d\mu \\+ \mu_2 \int_K \bigg( \Big(1 - \frac{1}{h^2}\Big)\psi_\alpha^0 + \Big(1 - \frac{1}{h^2}\Big)\psi_0^0 \bigg) d\mu, \end{multline*} and \[ \int_K \psi_1^0 d\mu = \mu_1 \int_K \frac{1}{h^2} \psi_\alpha^0 d\mu + \mu_2 \int_K \Big( \frac{1}{h^2} \psi_\alpha^0 + \frac{1}{h^2} \psi_0^0 + \psi_1^0 \Big) d\mu, \] Thus, $\mu_c^0$, $\mu_0^0$ and $\mu_1^0$ satisfy the system of equations \[ \begin{split} \mu_\alpha^0 &= \mu_1 \mu_1^0\\ \mu_0^0 &= \Big(1 - \frac{1}{h^2}\Big)\mu_\alpha^0 + \mu_2 \Big(1 - \frac{1}{h^2}\Big)\mu_0^0\\ \mu_1^0 &= \frac{1}{h^2}\mu_\alpha^0 + \frac{\mu_2}{h^2}\mu_0^0 + \mu_2 \mu_1^0, \end{split} \] where we have already use the fact $\mu_1 + \mu_2 = 1$. Moreover, as the sum $\psi_\alpha^0 + \psi_0^0 + \psi_1^0$ is the constant function 1, we also have \[ \mu_\alpha^0 + \mu_0^0 + \mu_1^0 = 1. \] Solving this system we obtain \[ \mu_\alpha^0 = \frac{\mu_1\mu_2}{\mu_1\mu_2 + (h^2-1)\mu_1 + \mu_2}, \quad \mu_0^0 = \frac{(h^2-1)\mu_1}{\mu_1\mu_2 + (h^2-1)\mu_1 + \mu_2}, \] and \[ \mu_1^0 = \frac{\mu_2}{\mu_1\mu_2 + (h^2-1)\mu_1 + \mu_2}. \] \section{Dirichlet Spectrum}\label{Dirichletsection} We now proceed to study the Dirichlet spectrum of $\Delta$. As there is no decimation on the Hata set, we have to use the finite element method in order to approximate the eigenvalues and eigenfunctions of $\Delta$. We present a summary of the observations obtained numerically by solving the system of equations \[ \begin{cases} -\Delta_m u(x) = \lambda u(x) & x\in V_m\setminus V_0 \\ u(p) = 0 & p\in V_0, \end{cases} \] where $\Delta_m u(x) = \frac{1}{\mu_p^m}H_mu(x)$. Recall that $L$ is the closure of the connected component of $K\setminus V_0$ that contains the interval $(0,1)$, and $M$ the closure of the component that contains the open segment from $0$ to $\alpha$. Thus, linearly independent Dirichlet eigenfunctions are supported either in $L$ or $M$. We observe numerically that the Dirichlet ground state is supported in $L$, as observed in Figure \ref{ground-state} (for $h=2$), corresponding to $\lambda_1 \approx 9.888$. \begin{figure}[ht] \includegraphics[width=2in]{GroundStatem8.pdf} \includegraphics[width=2in]{DerivedGroundStatem8.pdf} \caption{Dirichlet ground state $\phi$ for $h=2$, $\lambda_1 \approx 9.888$, and its derived eigenfunction $\tilde\phi = \chi_{K_1}\cdot\phi\circ F_1^{-1}$, $\lambda_3 \approx 56.21$, approximated to the eigth iteration, with $\alpha=1/2 + \sqrt{3}i/6$. Note that $\phi$ is supported in $L$ and $\tilde\phi$ is supported in $M$.} \label{ground-state} \end{figure} For each eigenfunction $\phi$ supported in $L$, there is a corresponding eigenfunction supported in $M$. \begin{proposition}\label{derived-eigenvalues} Let $\phi$ be a Dirichlet eigenfunction of $\Delta$, supported in $L$, with respect to the eigenvalue $\lambda$. Then $\chi_{K_1}\cdot\phi\circ F_1^{-1}$ is an eigenfunction supported in $M$ with respect to the eigenvalue $\lambda/(r_1\mu_1)$. \end{proposition} \begin{proof} Let $\tilde\phi = \chi_{K_1}\cdot\phi\circ F_1^{-1}$. If $x\in L$, then $x\in K_2$ or $x\in L\cap K_1$. In the first case, clearly $\tilde\phi(x)=0$ because $\chi_{K_1}(x)=0$, unless $x=|\alpha|^2$, the critical point. But in that case $x\in L\cap K_1$, and $F_1^{-1}(x)\in M$, so $\phi\circ F_1^{-1}(x)=0$ since $\phi$ is supported in $L$. Therefore $\tilde\phi$ is supported in $M$. Now, for $u\in\mathcal{F}_0$, by the self-similarity of the Dirichlet form $\mathcal{E}$, \[ \begin{split} \mathcal{E}(\tilde\phi, u) &= \frac{1}{r_1}\mathcal{E}(\tilde\phi\circ F_1, u\circ F_1) + \frac{1}{r_2}\mathcal{E}(\tilde\phi\circ F_2, u\circ F_1)\\ &= \frac{1}{r_1}\mathcal{E}(\phi, u\circ F_1) = \frac{\lambda}{r_1}\int_K \phi \, u\circ F_1 d\mu, \end{split} \] where we have used the fact that $\tilde\phi$ is supported in $K_1$ and $\phi$ is a Dirichlet eigenfunction with respect to $\lambda$. One the other hand, by the self-similarity of the measure $\mu$, \[ \int_K \tilde\phi\, u d\mu = \mu_1 \int_K \tilde\phi\circ F_1 \, u\circ F_1 d\mu + \mu_2 \int_K \tilde\phi\circ F_2 \, u\circ F_2 d\mu = \mu_1 \int_K \phi \, u\circ F_1 d\mu, \] so we obtain \[ \mathcal{E}(\tilde\phi, u) = \frac{\lambda}{r_1\mu_1} \int_K \tilde\phi\, u d\mu, \] and thus we conclude $\Delta\tilde\phi = - \dfrac{\lambda}{r_1\mu_1}\tilde\phi$. \end{proof} Figure \ref{ground-state} (right) shows the eigenfunction corresponding to the eigenvalue $\lambda_3 = \lambda_1/r_1\mu_1 \approx 56.21$, where $\lambda_1$ is the first Dirichlet eigenvalue ($h=2$). Proposition \ref{derived-eigenvalues} lets us classify the Dirichlet eigenvalues (and eigenfunctions) in two classes, which we will call ``primary'' and ``derived''. Table \ref{eigentable} shows the approximations to the first 20 Dirichlet eigenvalues, for $h=3/2,\sqrt 3$ and $3$. We note that the derived eigenvalues appear in different positions in the sequence $\lambda_k$, depending on $h$, and they seem to be more sparse as $h$ increases. \begin{table}[ht] \caption{Approximations $\lambda_k^{10}$ to the first 20 Dirichlet eigenvalues for values of $h$ equal to $3/2$, $\sqrt 3$ and $3$, indicating whether they correspond to a primary or a derived eigenfunction.} \label{eigentable} \begin{tabular}{|c|c|c|c|c|c|} \hline \multicolumn{2}{|c|}{$h=3/2$} & \multicolumn{2}{|c|}{$h=\sqrt 3$} &\multicolumn{2}{|c|}{$h=3$} \\ \hline $\lambda_k^{10}$ & Type & $\lambda_k^{10}$ & Type &$\lambda_k^{10}$ & Type \\ \hline 2.12748 & Primary & 10.012 & Primary & 38.7802 & Primary \\ 5.80965 & Derived & 31.037 & Primary & 139.978 & Primary \\ 8.3776 & Primary & 38.7455 & Derived & 255.362 & Primary \\ 13.7502 & Primary & 83.3496 & Primary & 336.428 & Primary \\ 22.8762 & Derived & 106.366 & Primary & 435.129 & Primary \\ 33.3334 & Primary & 120.11 & Derived & 566.447 & Primary \\ 34.0196 & Primary & 193.982 & Primary & 741.34 & Primary \\ 37.5447 & Derived & 226.027 & Primary & 972.052 & Primary \\ 53.6119 & Primary & 244.389 & Primary & 1067.3 & Derived \\ 59.5265 & Primary & 322.541 & Derived & 1266.44 & Primary \\ 91.007 & Derived & 401.184 & Primary & 1623.74 & Primary \\ 92.8821 & Derived & 411.618 & Derived & 1814.55 & Primary \\ 98.8091 & Primary & 503.566 & Primary & 2248.34 & Primary \\ 109.503 & Primary & 580.894 & Primary & 2574.77 & Primary \\ 133.249 & Primary & 613.579 & Primary & 2909.76 & Primary \\ 136.474 & Primary & 654.576 & Primary & 3013.28 & Primary \\ 146.338 & Derived & 750.644 & Derived & 3299.2 & Primary \\ 162.469 & Derived & 783.032 & Primary & 3812.6 & Primary \\ 195.591 & Primary & 874.566 & Derived & 4001.01 & Derived \\ 213.997 & Primary & 945.709 & Primary & 4147.5 & Primary \\ \hline \end{tabular} \end{table} Figure \ref{eigenfunctionsm10} shows the first four Dirichlet eigenfunctions for those values of $h$, where we can observe the appearance of the derived eigenfunctions corresponding to $\lambda_2^{10}$ ($h=3/2$) and $\lambda_3^{10}$ ($h=\sqrt 3$). \begin{figure}[ht] \includegraphics[width=4.5in]{Graficasm10.pdf} \caption{The first four Dirichlet eigenfunctions for the values $h=3/2,\sqrt 3$ and $3$, respectively, approximated to the 10th iteration, with $\alpha=1/2 + \sqrt{3}i/6$.} \label{eigenfunctionsm10} \end{figure} More interestingly, we show in Figure \ref{eigenfunctionsm10-rest} the \begin{figure}[ht] \includegraphics[width=4.5in]{Restriccionesm10.pdf} \caption{The restrictions to $[0,1]$ for the first three primary Dirichlet eigenfunctions for the values $h=3/2,\sqrt 3$ and $3$, respectively, approximated to the 10th iteration, with $\alpha=1/2 + \sqrt{3}i/6$.} \label{eigenfunctionsm10-rest} \end{figure} restrictions of these eigenfunctions to the interval $[0,1]$ (only the primary ones, as the derived are zero in $[0,1]$). One can ask whether these functions have singularity properties as in the case of the harmonic functions. Recall that, in the case of the harmonic functions, if $q\in V_{m+1}\cap[0,1]$ is the middle point between the points $x,y\in V_m\cap[0,1]$, then the harmonic function $u$ satisfies \begin{equation}\label{proportions-eq} u(q) = (1-\theta) u(x) + \theta u(y), \end{equation} where $\theta = 1/h^2$. In Figure \ref{proportions}, we show the \begin{figure}[ht] \includegraphics[width=4.5in]{Propor.pdf} \caption{The values of $\theta$ that solve equation \eqref{proportions-eq} for each middle point $q\in V_9\cap[0,1]$ (top of each pair of graphs) and $q\in V_{10}\cap[0,1]$ (bottom), for the approximations at $m=9$ and $m=10$ to the restrictions to $[0,1]$ for the first three primary Dirichlet eigenfunctions for the values $h=3/2,\sqrt 3$ and $3$, respectively, with $\alpha=1/2 + \sqrt{3}i/6$. The line corresponds to $1/h^2$, and the plot range is $1/h^2 \pm .01$.} \label{proportions} \end{figure} values of $\theta$ for such middle points $q\in V_9\cap[0,1]$ and $q\in V_{10}\cap[0,1]$, with respect to their adjacent points in $V_8$ and $V_9$, respectively, for the approximations at $m=9$ and $m=10$ to the restrictions to $[0,1]$ for the first three primary Dirichlet eigenfunctions for the values $h=3/2,\sqrt 3$ and $3$. We observe that they are closely equal to $1/h^2$, with better approximations as $h$ increases. We show the two iterations $m=9$ and $m=10$ in order to find out if the same proportions are preserved through two levels. As the $\theta$ are preserved through different iterations, we are lead to conjecture that the restrictions to $[0,1]$ of the Dirichlet eigenfunctions of the Laplacian with respect to $h\not=1/|\alpha|$ are singular functions whenever they are monotone, as in the case of the harmonic functions. \section{Conclusions}\label{Conclusionssection} We have studied harmonic and Dirichlet eigenfunctions for the family of harmonic structures of the Hata set $K$ parametrized by $h>1$. The former can be constructed by means of the known algorithms for harmonic functions on PCF sets, and one observes that, when restricted to the interval $[0,1]$, the longest edge in $K$, one obtains singular functions for all but one value of the parameter $h$. In fact, we observed that the only value of $h$ for which these restrictions are not singular coincides with the value such that the Hausdorff dimensions of $K$ with respect to the Euclidean and effective resistance metrics are the same. As is known, restrictions of harmonic functions to the edges of the Sierpinski gasket are also singular. This lead us to ask whether this behaviour is typical for harmonic functions on PCF sets. Moreover, since we know that in the case of the Hata set $K$ such functions are not singular for a particular embedding $K$ in the plane (given $h$, one can choose $\alpha$ such that $|\alpha|=1/h$), one can also ask whether, for every PCF set, there exists an embedding such that these restrictions are not singular. In particular, is this true for the Sierpinski gasket? The same questions can be asked for the eigenfunctions of a PCF set. We have numerical evidence for the case of the Hata set, and one can start by studying those PCF sets with decimation. In particular, do the eigenfunctions on the Sierpinski gasket have singular restrictions to the edges?
\section{Introduction} A great deal of attention has been applied to studying new and better ways to perform learning tasks involving static finite vectors. Indeed, over the past century the fields of statistics and machine learning have amassed a vast understanding of various learning tasks like density estimation, clustering, classification, and regression using simple real valued vectors. However, we do not live in a world of simple objects. From the contact lists we keep, the sound waves we hear, and the distribution of cells we have, complex objects such as sets, functions, and distributions are all around us. Furthermore, with ever-increasing data collection capacities at our disposal, not only are we collecting more data, but richer and more bountiful complex data are becoming the norm. This paper aims to make learning on massive data-sets of distributions tractable; we study distribution to real regression (DRR) where input covariates are arbitrary distributions and output responses are real values. We provide an estimator that scales well with data-set size and is efficient at evaluation-time. Furthermore, we prove that the estimator achieves a fast rate of convergence for a broad class of functions. We consider a mapping $f: \mathcal{I} \mapsto \mathbb{R}$ that takes $P \in \mathcal{I}$, an input distribution from a family of distributions $\mathcal{I}$, and produces $Y$ a real-valued response as: \begin{align} Y=f(P)+\epsilon, \mathrm{where }\ \EE{\epsilon}=0,\ \EE{\epsilon^2}\leq\sigma^2_\epsilon. \end{align} Of course, it is infeasible to directly observe a distribution in practice. Thus, we will work on a data-set of $N$ input sample-sets/responses: \begin{gather} \mathcal{D} = \{(\mathcal{X}_i,Y_i)\}_{i=1}^N,\ \mathrm{where } \label{eq:dataset}\\ \mathcal{X}_i = \{ X_{i1},\ldots,X_{in_i} \},\ X_{ij}\overset{iid}{\sim } P_i \in \mathcal{I}, \end{gather} and $Y_i = f(P_i)+\epsilon_i$. Further, $P_i \overset{iid}{\sim } \Phi$, where $\Phi$ is some measure over $\cal I$ (see Figure \ref{fig:gmodel}). \begin{figure}[h!] \label{fig:gmodel} \centering \includegraphics[width=0.4\textwidth]{figures/graphical_model_d2y.pdf} \caption{A graphical representation of our model. We observe a data-set of input sample-set/output response pairs $\{(\mathcal{X}_i,Y_i) \}_{i=1}^N$, where $\mathcal{X}_i = \{X_{i1},\ldots, X_{in_i} \}$, $X_{ij}\sim P_i$ and $Y_i = f(P_i) + \epsilon_i$, for some noise $\epsilon_i$. From these sample sets $\{\mathcal{X}_i \}_{i=1}^N$ we build density estimates $\{\tilde{P}_i \}_{i=1}^N$ using projection series estimates \eqref{eq:coef-est}. These estimates will then be used in our response estimator \eqref{eq:OLSest}. } \vspace{-2.5mm} \end{figure} Many interesting problems across various domains fit the DRR model. For instance, one may be interested in studying the mapping that takes in the distribution of star locations in a galaxy and outputs the galaxy's age. Also, one may be consider a mapping that takes in the distribution of prices for stocks of a particular sector and outputs the future average change in stock price for that sector. In fact, many estimation tasks in statistics can be framed as a distribution to regression problem. For instance, in parameter estimation one studies a mapping that takes in a distribution (usually restricted to be in a parametric class of distributions) and outputs a corresponding parameter. We will see that our estimator can be used to leverage previously seen sample sets to outperform standard estimation procedures, to perform model selection when cross validation is expensive, or to perform parameter estimation when no analytical sample estimate is available. In effect, we shall show that this estimator, and the concept of distribution to real regression, is powerful enough to itself learn how to perform general statistical procedures. At its core, the problem of distribution to real value regression is a learning task over infinite dimensional objects (distributions) and would benefit greatly from learning on data-sets with a large number of input/output pairs. Hence, this paper focuses on the case where one has a massive data-set in terms of instances, i.e.\ $n_i = o(N)$. DRR for the case of general input distributions in a H\"{o}lder class and a smooth class of mappings has been previously studied in \cite{poczos2013distribution}. There, an estimator---the Kernel-Kernel estimator---analogous to the Nadaraya-Watson estimator \cite{tsybakov2008introduction} for functional distribution inputs was shown to have a polynomial rate of convergence. This rate is dependent on the dimensionality of the domains of the distributions, sample sizes, and a doubling dimension on the measure $\Phi$ over distributions, which, roughly speaking, controls the degrees of freedom of the input distributions. However, evaluating the estimator in \cite{poczos2013distribution} for new predictions scales as $\Omega(N)$ in the number of input/output instances in a data-set. Thus, the Kernel-Kernel estimator is not feasible for data-sets where the number of distributions, $N$, is in the high-thousands, millions, or even billions. Furthermore, the doubling dimension of $\Phi$ may be rather large, producing a slow convergence rate. In this paper we shall introduce an estimator for DRR, the Double-Basis estimator, which does not depend on $N$ for evaluating an estimate for a new input distribution. Furthermore, we shall show that this estimator achieves a better rate of convergence that does not depend on the doubling dimension over a broad class of distribution to real mappings. \section{Related Work} As previously mentioned, the problem of DRR was studied in \cite{poczos2013distribution}, where the Kernel-Kernel estimator was introduced. Since the data-set one works with is \eqref{eq:dataset}, first one uses kernel density estimation (KDE) \cite{tsybakov2008introduction} on $\{\mathcal{X}_1,\ldots,\mathcal{X}_N\}$ to make density estimates $\{\tilde{P}_1,\ldots,\tilde{P}_N\}$. Similarly for an unseen query input sample set $\mathcal{X}_0 \sim P_0$, one makes a KDE $\tilde{P}_0$. Then, the Kernel-Kernel estimator works as follows: \begin{align} &\hat{f}(\tilde{P}) = \sum_{i=1}^N W(\tilde{P}_i,\tilde{P}_0) Y_i,\ \mathrm{where } \label{eq:KKest}\\ &W(\tilde{P}_i,\tilde{P}_0) \nonumber \\ &= \begin{cases} \frac{K(D(\tilde{P}_i,\tilde{P}_0))}{\sum_j K(D(\tilde{P}_j,\tilde{P}_0))} &\mbox{if } \sum_j K(D(\tilde{P}_j,\tilde{P}_0))>0 \\ 0 & \mbox{otherwise }. \end{cases} \end{align} Here $K$ is taken to be a symmetric Kernel with bounded support, and $D$ is some metric over functions. Clearly, \eqref{eq:KKest} scales as $\Omega(N)$ in terms of the number of input distributions in ones data-set. Furthermore, if one uses a Gaussian KDE, and takes $D(\tilde{P}_i,\tilde{P}_0)=\norm{\tilde{P}_i-\tilde{P}_0}_2=\sqrt{\int(\tilde{p}_i-\tilde{p}_0)^2}$ (where $\tilde{p}_i$ is the pdf of $\tilde{P}_i$) and $n_i\asymp n$, then the computation required for evaluating \eqref{eq:KKest} is $\Omega(Nn^2)$. DRR is related to the functional analysis, where one regresses a mapping whose input domain are functions \cite{ferraty2006nonparametric}. However, the objects DRR works over--distributions and their pdfs--are inferred through sets of samples drawn from the objects, with finite sizes. In functional analysis, the functions are inferred through observations of $(X,Y)$ pairs that are often taken to be an arbitrarily dense grid in the domain of the functions. For a comprehensive survey in functional analysis see \cite{ferraty2006nonparametric,ramsay2002applied}. Also, recently \cite{olivadistribution} studied the problem of distribution to distribution regression, where both input and output covariates are distributions. A common approach to performing ML tasks with distributions is to embed the distributions in a Hilbert space, then solve the tasks using kernel machines. Perhaps the most clear-cut of these methods is to fit a parametric model to distributions for estimating kernels \cite{jebara2004probability, jaakkola1999exploiting, moreno2003kullback}. Nonparametric methods over distributions have also been developed using kernels. For example, since we only observe distributions through finite sets, set kernels may be used \cite{smola2007hilbert}. Futhermore, the representer theorem was recently generalized for the space of distributions \cite{muandet2012learning}. Also, kernels based on nonparametric estimators of divergences have been explored \cite{poczos2012nonparametric,poczos2012image}. \section{Double-Basis Estimator} We introduce the Double-Basis Estimator for DRR. First, we shall use orthonormal basis projection estimators \cite{tsybakov2008introduction} for estimating the densities of $P_i$ from $\mathcal{X}_i$. Suppose that $\Lambda^l \subseteq \mathbb{R}^l$, the domain of input densities is compact s.t. $\Lambda = [a,b]$. Let $\{\varphi_i\}_{i\in\mathbb{Z}}$ be an orthonormal basis for $L_2(\Lambda)$. Then, the tensor product of $\{\varphi_i\}_{i\in\mathbb{Z}}$ serves as an orthonormal basis for $L_2(\Lambda^l)$; that is, \begin{gather*} \{\varphi_\alpha\}_{\alpha\in\mathbb{Z}^l} \quad \mathrm{where} \quad \varphi_\alpha(x) = \prod_{i=1}^l \varphi_{\alpha_i}(x_i),\ x\in \Lambda^l \end{gather*} serves as an orthonormal basis (so we have $\forall \alpha,\rho\in \mathbb{Z}^l,\ \langle\varphi_\alpha,\varphi_\rho \rangle= I_{\{\alpha=\rho\}}$). Let $P\in \mathcal{I} \subseteq L_2(\Lambda^l)$, then \begin{gather} p(x)=\sum_{\alpha\in\mathbb{Z}^l} a_\alpha(P)\varphi_\alpha(x) \ \mathrm{where }\\ \quad a_\alpha(P) = \langle\varphi_\alpha, p\rangle = \int_{\Lambda^l} \varphi_\alpha(z)\mathrm{d} P(z)\ \in \mathbb{R} \nonumber. \end{gather} where $p(x)$ denotes the probability density function of the distribution $P$. Suppose that the projection coefficients $a(P) = \{a_\alpha(P)\}_{\alpha\in \mathbb{Z}^l}$ are as follows for $P\in\mathcal{I}$: \begin{gather} \mathcal{I} = \{ P : a(P) \in \Theta_l(\nu,\gamma,A),\ \norm{P}_2^2 \leq A \} \quad \mathrm{where } \label{eq:sob-ellp}\\ \Theta_l(\nu,\gamma,A) = \left\{\{a_\alpha\}_{\alpha\in \mathbb{Z}^l} : \sum_{\alpha\in\mathbb{Z}^l} a_\alpha^2 \kappa_\alpha^2(\nu,\gamma) < A \right\},\nonumber\\ \kappa_\alpha^2(\nu,\gamma) = \sum_{i=1}^l(\nu_i|\alpha_i|)^{2\gamma_i}\ \mathrm{for}\ \nu_i,\gamma_i,A > 0. \nonumber \end{gather} See \cite{ingster2011estimation,laurent1996efficient} for other analyses with this type of assumption. The assumption in \eqref{eq:sob-ellp} will control the tail-behavior of projection coefficients and allow us to effectively estimate $P\in \mathcal{I}$ using a finite number of projection coefficients on the empirical distribution of a sample. Given a sample $\mathcal{X}_i = \{X_{i1},\ldots,X_{in_i}\}$ where $X_{ij}\overset{iid}{\sim } P_i \in \mathcal{I}$, let $\widehat{P}_i$ be the empirical distribution of $\mathcal{X}_i$; i.e. $\widehat{P}_i(X=X_{ij})=\frac{1}{n_i}$. Our estimator for $p_i$ will be: \begin{align} \tilde{p}_i(x) = \sum_{\alpha\ :\ \kappa_\alpha(\nu,\gamma)\leq t}a_\alpha(\widehat{P}_i)\varphi_\alpha(x) \quad \mathrm{where } \label{eq:coef-est}\\ a_\alpha(\widehat{P}_i) = \int_{\Lambda^l} \varphi_\alpha(z)\mathrm{d} \widehat{P}_i(z) = \frac{1}{n_i}\sum_{j=1}^{n_i} \varphi_\alpha(X_{ij}) \label{eq:coef-hat}. \end{align} Choosing $t$ optimally\footnote{See appendix for details.} can be shown to lead to $\mathbb{E}[\norm{\tilde{p}_i-p_i}_2^2]=O(n_i^{-\frac{2}{2+\gamma^{-1}}})$, where $\gamma^{-1}=\sum_{j=1}^{l}\gamma_j^{-1}$, $n_i \rightarrow \infty$ \cite{nussbaum1983optimal}. Next, we shall use random basis functions from Random Kitchen Sinks (RKS) \cite{rahimi2007random} to compute our estimate of the response. \cite{rahimi2007random} shows that if one has a shift-invariant kernel $K$ (in particular we consider the RBF kernel $K(x)=\exp(-x^2/2)$) then for $x,y \in \mathbb{R}^d$: \begin{align} &K(\Norm{x-y}_2/\sigma) \approx z(x)^Tz(y),\ \mathrm{where }\\ &z(x) \equiv\nonumber \\ & \sqrt{\tfrac{2}{D}}\left[\cos(\omega_1^Tx+b_1) \cdots \cos(\omega_D^Tx+b_D)\right]^T \end{align} with $\omega_i \stackrel{iid}{\sim} \mathcal{N}(0,\sigma^{-2}I_d)$, $b_i \stackrel{iid}{\sim} \text{Unif}[0,2\pi]$ Let $M_t = \{\alpha\ :\ \kappa_\alpha(\nu,\gamma) \leq t \} = \{\alpha_1,\ldots,\alpha_S\}$. First note that: \begin{align*} \idot{\tilde{p}_i}{\tilde{p}_j} =& \Idot{\sum_{\alpha\in M_t} a_{\alpha}(\widehat{P}_i) \varphi_{\alpha}}{\sum_{\alpha\in M_t} a_{\alpha}(\widehat{P}_j) \varphi_{\alpha}} \\ =& \sum_{\alpha\in M_t} \sum_{\beta\in M_t}a_{\alpha}(\widehat{P}_i) a_{\beta}(\widehat{P}_j)\Idot{\varphi_{\alpha}}{\varphi_{\beta}} \\ =& \sum_{\alpha\in M_t} a_{\alpha}(\widehat{P}_i) a_{\alpha}(\widehat{P}_j) \\ =& \Idot{\vec{a}_t(\widehat{P}_i)}{\vec{a}_t(\widehat{P}_j)}, \end{align*} where $\vec{a}_t(\widehat{P}_i) = (a_{\alpha_1},\ldots,a_{\alpha_s})$, $M_t = \{\alpha_1,\ldots,\alpha_s\}$, and the last inner product is the vector dot product. Thus, \begin{align*} \Norm{\tilde{p}_i-\tilde{p}_j}_2 = \Norm{\vec{a}_t(\widehat{P}_i)-\vec{a}_t(\widehat{P}_j)}_2, \end{align*} where the norm on the LHS is the $L_2$ norm and the $\ell_2$ on the RHS. Consider a fixed $\sigma$. Let $\omega_i \stackrel{iid}{\sim} \mathcal{N}(0,\sigma^{-2}I_s)$, $b_i \stackrel{iid}{\sim} \text{Unif}[0,2\pi]$, be fixed. Let $K_\sigma(x) = K(x/\sigma)$. Then, \begin{align} \sum_{i=1}^N\theta_i K_\sigma(\norm{\tilde{p}_i-\tilde{p}_0}_2) \approx& \sum_{i=1}^N \theta_iz(\vec{a}_t(\widehat{P}_i))^Tz(\vec{a}_t(\widehat{P}_0)) \nonumber \\ =& \left(\sum_{i=1}^N \theta_iz(\vec{a}_t(\widehat{P}_i))\right)^Tz(\vec{a}_t(\widehat{P}_0)) \nonumber\\ =& \psi^T z(\vec{a}_t(\widehat{P}_0)) \label{eq:lin_est_approx} \end{align} where $\psi = \sum_{i=1}^N \theta_iz(\vec{a}_t(\widehat{P}_i)) \in \mathbb{R}^s$. Hence, we consider estimators of the form \eqref{eq:lin_est_approx}; that is, we consider linear estimators in the non-linear space induced by $z(\vec{a}_t(\cdot))$. In particular, we consider the OLS estimator using the data-set $\{(z(\vec{a}_t(\widehat{P}_i)),Y_i)\}_{i=1}^N$ : \begin{align} \hat{f}(\tilde{P}_0) \equiv& \hat{\psi}^T z(\vec{a}_t(\widehat{P}_0))\ \mathrm{where } \label{eq:OLSest} \\ \hat{\psi} \equiv& \argmin_\beta \norm{\vec{Y}-{\bf{Z}}\beta}_2^2 \\ =& ({\bf{Z}}^T{\bf{Z}})^{-1}{\bf{Z}}^T\vec{Y} \end{align} for $\vec{Y}=(Y_1,\ldots,Y_N)^T$, and with ${\bf{Z}}$ being the $N\times D$ matrix: ${\bf{Z}}=[z(\vec{a}_t(\widehat{P}_1))\cdots z(\vec{a}_t(\widehat{P}_N)) ]^T$. \subsection{Evaluation Computational Complexity} We see that after computing $\hat{\psi}$, evaluating our estimator on a new distribution $P_0$ amounts to taking an inner product with a $D \times 1$ vector. Including the time required for computing $z(\vec{a}_t(\widehat{P}_0))$, the computation required for the evaluation, $\hat{f}(\tilde{P}_0) = \hat{\psi}^Tz(\vec{a}_t(\widehat{P}_0))$, is: one, the time for evaluating the projection coefficients $\vec{a}_t(\widehat{P}_1)$, $O(sn)$; two, the time to compute the RKS features $z(\cdot)$, $O(Ds)$; three, the time to compute the inner product, $\langle \hat{\psi}, \cdot \rangle$, $O(D)$. Hence, the total time is $O(D+Ds+sn)$. We'll see that $D=O(n\log(n))$ and $s=O(n)$ hence the total run-time for evaluating $\hat{f}(\tilde{P}_0)$ is $O(n^2\log(n))$. Since we are considering data-sets where the number of instances $N$ far outnumbers the number of points per sample set $n$, $O(n^2\log(n))$ is a substantial improvement over $O(Nn^2)$. \subsection{Ridge Double-Basis Estimator} We note that a straightforward extension to the Double-Basis estimator is to use a ridge regression estimate on features $z(\vec{a}_t(\cdot))$ rather than a OLS estimate. That is, for $\lambda\geq 0$ let \begin{align} \hat{\psi}^T_\lambda \equiv& \argmin_\beta \norm{\vec{Y}-{\bf{Z}}\beta}_2^2 + \lambda \norm{\beta}_2 \label{eq:ridgeest}\\ =& ({\bf{Z}}^T{\bf{Z}}+\lambda I)^{-1}{\bf{Z}}^T\vec{Y}. \end{align} Clearly the Ridge Double-Basis estimator is still evaluated via a dot product with $\hat{\psi}^T_\lambda$, and our above complexity analysis holds. Furthermore, we note that the Double-Basis estimator is a special case of the Ridge Double-Basis estimator with $\lambda =0$. \section{Theory} \subsection{Assumptions} We shall assume the following: \begin{enumerate}[label=\textbf{A.\arabic*}] \item{ \label{asmp:sob} {\em Sobolev Input Distributions.} Suppose that \eqref{eq:sob-ellp} holds. } \item{ \label{asmp:sob} {\em RKHS Mapping.} We shall assume that $f \in \mathcal{F}(\sigma,B)$ for $f: \mathcal{I} \mapsto \mathbb{R}$, where $\sigma, B \in \mathbb{R}$ and \begin{align} \mathcal{F}(\sigma,B) = \Big\{& f : f(P) = \sum_{i=1}^\infty \theta_i K_\sigma\left(G_i,P\right) , \\ &\mathrm{where }\ G_i \in \mathcal{I},\ \norm{\theta}_1 \leq B\Big\}. \end{align} Here we take $K_\sigma(G_i,P) = K_\sigma(\norm{g_i-p}_2) = K({\norm{g_i-p}_2}/{\sigma})$ to be a shift-invariant kernel. In particular, we take $K$ to be the RBF kernel: $K(x) = \exp(-x^2/2)$. Note further that: \begin{align} |K(x)-K(x')| \leq e^{-\frac{1}{2}}|x-x'|. \label{eq:lips} \end{align} } \item{ \label{asmp:sampsize} {\em Input Sample Set Sizes. } Suppose that $\forall i$ $|\mathcal{X}_i| \asymp n$. } \end{enumerate} \subsection{Convergence Rate} Since by \ref{asmp:sob} we have that $|f(P)|\leq B$, we consider an upperbound for the risk of a truncated version of our estimator \eqref{eq:OLSest}. Let $T_B(x) \equiv \sign(x)\min(|x|,B)$. For readability, let $Z(P)=z(\vec{a}_t(P))$. Let a small real $\delta>0$ be fixed. We look to show that: \begin{theorem} \begin{align*} &\EE{\left(T_B\left(\hat{\psi}^TZ(\widehat{P}_0)\right)-f(P_0)\right)^2} \\ &= O\left(n^{-1/(2+\gamma^{-1})}\right) + O\left(\frac{n\log(n)\log(N)}{N}\right) \end{align*} with probability at least $1-\delta$. \end{theorem} Roughly speaking, our proof will work as follows: first, we show that a population optimal linear model in the non-linear features $Z(\cdot)$ is close to the function $f$; then we will show that a population optimal linear model is close to the OLS (sample optimal) linear model. Thus, we proceed to show that predictions from the optimal linear model using $Z(P_0)$ is close to $f(P_0)$, that is: \begin{align*} \frac{1}{2}\mathbb{E}_{P_0}\left[ \left(f(P_0) -\beta^TZ(\widehat{P}_0) \right)^2\right] \end{align*} is small, where $\beta$ is an optimal weight vector. Note that $\beta$ minimizes: \begin{align} &\mathbb{E}\left[ \left(Y_0 -\beta^TZ(\widehat{P}_0) \right)^2\right] = \label{eq:opti_linear}\\ & \mathbb{E}\left[ Y_0^2\right] - 2\mathbb{E}\left[Y_0Z(\widehat{P}_0)\right]^T\beta +\beta^T\mathbb{E}\left[Z(\widehat{P}_0)Z(\widehat{P}_0)^T\right]\beta \nonumber. \end{align} Let \begin{align} \varsigma_i \equiv \sum_{j=1}^\infty \theta_j \left( K_\sigma\left(g_j,p_i\right) - Z(G_j)^T Z(\widehat{P}_i)\right). \end{align} Furthermore, note that: \begin{align} Y_i = f(p_i) + \epsilon_i = \sum_{j=1}^\infty \theta_j K_\sigma\left(g_j,p_i\right) + \epsilon_i. \label{eq:realmodel} \end{align} Let \begin{align*} \bar{g}_i = \sum_{\alpha\in M_t }a_\alpha(G_i)\varphi_\alpha(x). \end{align*} Also, let $\vec{a}_t(G_j) = (a_{\alpha_1}(G_j),\ldots,a_{\alpha_S}(G_j))^T$. When using kitchen sinks, we will see that $Y$ is approximately a linear model. Precisely, \begin{align*} Y_i =& \sum_{j=1}^\infty \theta_j Z(G_j)^T Z(\widehat{P}_i) + \varsigma_i + \epsilon_i \\ =& \psi^T Z(\widehat{P}_i) + \varsigma_i + \epsilon_i, \end{align*} where $\psi = \sum_{i=1}^\infty \theta_i Z(G_i)$. First we prove the following bound for the error using the optimal linear model $\beta$: \begin{lemma} \begin{align*} \mathbb{E}_{P_0}\left[ \left(f(P_0) -\beta^TZ(\widehat{P}_0) \right)^2\right] \leq \mathbb{E}_{P_0}\left[ \varsigma_0^2\right]+4B\sqrt{\mathbb{E}_{P_0}\left[ \varsigma_0^2\right]} \end{align*} \end{lemma} \begin{proof} Since \eqref{eq:opti_linear} is a quadratic function bounded below, an optimal $\beta$ may be found by satisfying stationarity (taking the gradient \eqref{eq:opti_linear} and setting to zero). We take $\beta = \Sigma^{+}\Sigma_Y$ where $\Sigma = \mathbb{E}[Z(\widehat{P}_0)Z(\widehat{P}_0)^T]$ is the uncentered covariance matrix, $\Sigma^{+}$ is its Moore-Penrose inverse, and $\Sigma_Y=\mathbb{E}[Y_0Z(\widehat{P}_0)]$ is the vector of uncentered covariances to the response\footnote{Note that if $\Sigma$ is nonsingular, $\Sigma^{+}=\Sigma^{-1}$ and $\beta$ is unique.}. Hence, \begin{align*} &\frac{1}{2}\mathbb{E}_{P_0}\left[ \left(f(P_0) -\beta^TZ(\widehat{P}_0) \right)^2\right] \\ &= \frac{1}{2}\mathbb{E}_{P_0}\left[ \left(f(P_0) -\Sigma_Y^T\Sigma^{+}Z(\widehat{P}_0) \right)^2\right]\\ &= \frac{1}{2}\mathbb{E}_{P_0}\left[ \left(f(P_0) \right)^2\right] - \Sigma_Y^T\Sigma^{+}\mathbb{E}_{P_0}\left[ f(P_0) Z(\widehat{P}_0) \right] \\ &\quad + \frac{1}{2}\Sigma_Y^T\Sigma^{+}\mathbb{E}_{P_0}\left[ Z(\widehat{P}_0)Z(\widehat{P}_0)^T\right]\Sigma^{+}\Sigma_Y \\ &= \frac{1}{2}\mathbb{E}_{P_0}\left[ \left(\psi^T Z(\widehat{P}_0)+\varsigma_0 \right)^2\right] \\ &\quad - \Sigma_Y^T\Sigma^{+}\mathbb{E}_{P_0,\epsilon_0}\left[ (f(P_0)+\epsilon_0) Z(\widehat{P}_0) \right] \\ &\quad + \frac{1}{2}\Sigma_Y^T\Sigma^{+}\Sigma\Sigma^{+}\Sigma_Y\\ &= \frac{1}{2}\psi^T\Sigma\psi + \mathbb{E}_{P_0}\left[ \varsigma_0z(\widehat{P}_0)^T \right]\psi+ \frac{1}{2}\mathbb{E}_{P_0}\left[ \varsigma_0^2\right] \\ &\quad - \frac{1}{2}\Sigma_Y^T\Sigma^{+}\Sigma_Y. \end{align*} Also, \begin{align*} \Sigma_Y &= \mathbb{E}_{P_0,\epsilon_0}\left[(\psi^T Z(\widehat{P}_0))Z(\widehat{P}_0)+\varsigma_0Z(\widehat{P}_0)+\epsilon_0Z(\widehat{P}_0) \right] \\ &= \Sigma\psi + \mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)\right]. \end{align*} Thus, \begin{align*} &\Sigma_Y^T\Sigma^{+}\Sigma_Y \\ = & (\psi^T\Sigma + \mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right])\Sigma^{+}(\Sigma\psi + \mathbb{E}_{P_0}\left[\varsigma_0Z(P_0)\right])\\ = & \psi^T\Sigma\Sigma^{+}\Sigma\psi + \mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\Sigma\psi\\ &+\psi^T\Sigma\Sigma^{+}\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)\right]\\ &+\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)\right]\\ = & \psi^T\Sigma\psi + 2\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\Sigma\psi\\ &+\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)\right]. \end{align*} Hence, \begin{align} &\frac{1}{2}\mathbb{E}_{P_0}\left[ \left(f(P_0) -\beta^TZ(\widehat{P}_0) \right)^2\right] \nonumber\\ &= \frac{1}{2}\psi^T\Sigma\psi + \mathbb{E}_{P_0}\left[ \varsigma_0Z(\widehat{P}_0)^T \right]\psi+ \frac{1}{2}\mathbb{E}_{P_0}\left[ \varsigma_0^2\right]\nonumber\\ &\quad -\frac{1}{2}\psi^T\Sigma\psi - \mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\Sigma\psi\nonumber\\ &\quad-\frac{1}{2}\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)^T\right]\Sigma^{+}\mathbb{E}_{P_0}\left[\varsigma_0Z(\widehat{P}_0)\right]\nonumber\\ &\leq \frac{1}{2}\mathbb{E}_{P_0}\left[ \varsigma_0^2\right]+4B\sqrt{\mathbb{E}_{P_0}\left[ \varsigma_0^2\right]}\label{eq:blin-bnd}, \end{align} see Appendix for details on the last bound \end{proof} \begin{lemma} \begin{align*} \mathbb{E}_{P_0}\left[ \varsigma_0^2\right] = O\left(n^{\frac{-2}{2+\gamma^{-1}}}\right) \end{align*} with probability at least $1-\delta$. \end{lemma} \begin{proof} $|\varsigma_i| \leq \sum_{j=1}^\infty |\theta_j| \left| K_\sigma\left(g_j,p_i\right) - Z(G_j)^T z(\widehat{P}_i)\right|$ and \begin{align*} &\left| K_\sigma\left(g_j,p_i\right) - Z(G_j)^T Z(\widehat{P}_i)\right| \\ &\leq \left| K_\sigma\left(g_j,p_i\right) - K_\sigma\left(\bar{g}_j,\tilde{p}_i\right) \right| \\ &+ \left| K_\sigma\left(\bar{g}_j,\tilde{p}_i\right) - Z(G_j)^T Z(\widehat{P}_i)\right|. \end{align*} Also, using \eqref{eq:lips}: \begin{align*} &\left| K_\sigma\left(g_j,p_i\right) - K_\sigma\left(\bar{g}_j,\tilde{p}_i\right) \right| \\ &\leq \frac{e^{-\frac{1}{2}}}{\sigma} \left|\norm{g_j - p_i}_2 - \norm{\bar{g}_j - \tilde{p}_i}_2 \right|. \end{align*} Moreover, using the triangle inequality: \begin{align*} \left| \norm{g_j - p_i}_2 - \norm{\bar{g}_j - \tilde{p}_i}_2 \right| \leq & \norm{g_j - \bar{g}_j}_2 + \norm{\tilde{p}_i - p_i}_2 . \end{align*} Thus, \begin{align*} &\EE{\left| K_\sigma\left(g_j,p_i\right) - K_\sigma\left(\bar{g}_j,\tilde{p}_i\right) \right|}\\ &\leq \EE{\frac{e^{-\frac{1}{2}}}{\sigma} \left(\norm{g_j - \bar{g}_j}_2 + \norm{\tilde{p}_i - p_i}_2 \right)} \\ &= O\left(n^{\frac{1}{2+\gamma^{-1}}}\right), \end{align*} where the last line follows\footnote{\label{note1}See Appendix for details.} by choosing $t\asymp n^{\frac{1}{2+\gamma^{-1}}}$, and the expectation is w.r.t. $\mathcal{X}_i\sim P_i$, $P_i\sim\Phi$. Also, note that the dimensionality of $\vec{a}_t(G_i)$ and $\vec{a}_t(\widehat{P}_i)$ is\footnotemark[\value{footnote}] $S=|M(t)|=O(n^{{\gamma^{-1}}/{(2+\gamma^{-1}})})$. Let $\mathcal{M} = \{ v\in \mathbb{R}^S \ : \ \norm{v}_2^2\leqA\}$. Then, $\vec{a}_t(G_j),\ \vec{a}_t(\widehat{P}_i)\in \mathcal{M}$. Hence, by {\em Claim 1} in \cite{rahimi2007random}: \begin{align*} &\mathbb{P}\left[ \sup_{u,v\in\mathcal{M}}|K(u,v)-z(u)^Tz(v)|\geq \xi \right] \\ &\leq 2^8\left(\frac{\sqrt{S}\diam(\mathcal{M})}{\sigma \xi} \right)^2 \exp\left(- \frac{D \xi^2}{4(S+2)} \right). \end{align*} Thus, with probability at least $1-\delta$: \begin{align*} \sup_{u,v\in\mathcal{M}}|K(u,v)-z(u)^Tz(v)|< n^{-\frac{1}{2+\gamma^{-1}}}, \end{align*} if we choose $D$ such that: \begin{align*} D &= \Omega\left( 4(S+4)n^{\frac{2}{2+\gamma^{-1}}}\log\left( \delta^{-1} 2^{10} A S n^{\frac{2}{2+\gamma^{-1}}}/\sigma^2 \right) \right), \end{align*} which is satisfied setting $D \asymp n\log(n)$. Hence, probability at least $1-\delta$: \begin{align*} &\frac{1}{2}\mathbb{E}_{P_0}\left[ \varsigma_0^2\right]\\ &\leq \mathbb{E}_{P_0}\Bigg[ \Bigg(\sum_{j=1}^\infty |\theta_j| \Big(\tfrac{e^{-\frac{1}{2}}}{\sigma} \norm{g_j - \bar{g}_j}_2 + \tfrac{e^{-\frac{1}{2}}}{\sigma}\norm{\tilde{p}_0 - p_0}_2 \\ &\quad\quad\quad\quad +\left| K_\sigma\left(\bar{g}_j,\tilde{p}_0\Big)- Z(G_j)^T Z(\widehat{P}_0)\right| \right) \Bigg)^2 \Bigg]\\ &= \mathbb{E}_{P_0}\left[ \left(\sum_{j=1}^\infty |\theta_j| \left(\norm{\tilde{p}_0 - p_0}_2 + O\left(n^{\frac{1}{2+\gamma^{-1}}}\right) \right) \right)^2 \right] \\ &= \left(\sum_{j=1}^\infty |\theta_j| \right)^2\mathbb{E}_{P_0}\left[ \left(\norm{\tilde{p}_0 - p_0}_2 + O\left(n^{\frac{1}{2+\gamma^{-1}}}\right) \right)^2 \right] \\ &= O\left(n^{\frac{-2}{2+\gamma^{-1}}}\right). \end{align*} \end{proof} Thus, we see that $f(P)$ is close to the linear model in the non-linear spaced induced by the $O(n\log(n))$ features $Z(\cdot)$: Then, {\em Theorem 11.3 of} \cite{gyorfi2002distribution} states that the estimated linear predictor $\hat{\beta}\in\mathbb{R}^d$ has an error to the mean conditional response (when truncated) relative an optimal linear predictor $\beta$ as follows: \begin{align} &\EE{\left(T_B\left(\hat{\beta}^Tx\right)-\EE{y|x}\right)^2} \leq \nonumber\\ &8 \EE{\left(\beta^Tx-\EE{y|x}\right)^2}+ O(\max\{\sigma_\epsilon^2,B^2\}d \log(N)/N) \label{eq:claim-rate}. \end{align} Using our notation we have that: \begin{align} &\EE{\left(T_B\left(\hat{\psi}^TZ(\widehat{P}_0)\right)-f(P_0)\right)^2} \nonumber\\ &= O\left(n^{-1/(2+\gamma^{-1})}\right) + O\left(\frac{n\log(n)\log(N)}{N}\right) \label{eq:rate}, \end{align} where we have bounded $\EE{\left(T_B\left(\hat{\beta}^Tx\right)-\EE{y|x}\right)^2}$ using Lemmas 4.2 and 4.3, giving us our desired rate. \section{Experiments} We perform experiments that demonstrate the ability of the Double-Basis estimator to learn distribution-to-real mappings from large training datasets, which can be applied to yield fast, accurate, and useful predictions. We illustrate this on a few statistical estimation tasks, which aim to take a set of samples from a distribution as input and yield some estimated quantity as output. For many such tasks, we can generate large amounts of relevant output quantities and associated input samples synthetically, and can train the Double-Basis estimator on these big datasets, giving us an automated procedure to learn a mapping for these statistical estimation tasks. We will show that, in some cases, this mapping can be more accurate, faster, and more robust than existing statistical procedures. In all of the following experiments, we train on data of the form $\mathcal{D} = \{(\mathcal{X}_i,Y_i)\}_{i=1}^N$. \subsection{Synthetic Mapping} First, we look to emphasize the computational improvement in evaluation time of the Double-Basis estimator over the Kernel-Kernel estimator using experiments with synthetic data. Our experiments are as follows. We first set $N\in\{1\Ep4,1\Ep5,1\Ep6\}$. Then, we generate a random mapping $f$ such that $f(P) = \sum_{i=1}^{10}\theta_iK_\sigma(G_i,P)$. We took $\sigma=1$, $\theta_i \sim \text{Unif}[-5,5]$, and $G_i$ to be the pdf of a mixture of two truncated Gaussians (each with weight $.5$) on the interval $[0,1]$, whose mean locations are chosen uniformly at random in $[0,1]$, and whose variance parameters are taken uniformly at random in $[.05,.1]$. For $j=\{1,\ldots,N\}$ we also set $P_j$ to be a randomly generated mixture of two truncated Gaussians as previously described. We then generate $Y_i$ under the the noiseless case, i.e $Y_i = f(P_i)$ (kernel values were computed numerically). Then, we generated $\mathcal{X}_i=\{X_{i1},\ldots,X_{in}\}$ where $n\propto N^{3/5}$ and $X_{i1}\overset{iid}{\sim } P_i$. $\tilde{P}_i$ was then estimated using the samples $\mathcal{X}_i$. We compared the performance of both the Double-Basis (BB), and the Kernel-Kernel (KK) estimator on a separate test set of $\mathcal{D}_t = \{(\mathcal{X}_j,Y_j)\}_{j=1}^{N_t}$ where $N_t = 1\Ep5$, that was generated as $\mathcal{D}$ was. We measured performance in terms of mean squared error (MSE) and mean evaluation time per new query $\mathcal{X}_0$ (Figures \ref{fig:big_mse} and \ref{fig:big_time} respectively). One can see that in this case both estimators have similar MSEs, with the BB estimator doing somewhat better in each configuration of the data-set size. However, one can observe a striking difference in the average time to evaluate a new estimate $\hat{f}(\tilde{P})$. Figure \ref{fig:big_time} is presented in a log scale, and illustrates the Kernel-Kernel estimator's lack of scaling on data-set size, $N$. On the other hand, the Double-Basis estimator is considerbly efficient even at large $N$ and has a speed-up of about $\times12$, $\times67$, and $\times139$ over the Kernel-Kernel estimate for $N=1\Ep4,1\Ep5,1\Ep6$ respectively. \begin{figure} \centering \subfigure[Estimation Error]{\raisebox{0mm}{\label{fig:big_mse}\includegraphics[width=.2325\textwidth]{figures/big_mse.pdf}}} \subfigure[Estimation Time]{\raisebox{0mm}{\label{fig:big_time}\includegraphics[width=.2325\textwidth]{figures/big_time_log.pdf}}} \vspace{-0.3cm} \caption{Results on predicting synthetic mapping $f$. } \vspace{-0.4cm} \end{figure} \vspace{-0.2cm} \subsection{Choosing $k$: model selection for Gaussian mixtures} \vspace{-0.2cm} Many common statistical tasks involve producing a mapping from a distribution to a real value, and may be tackled using DRR. One such task is that of model selection, where one is given a set $\mathcal{X}_0 = \{X_{01},\ldots,X_{0n_0}\}$ drawn from an unknown distribution $P$ and wants to find some parameter that is indicative of the complexity of the true distribution. In other words, the mapping of interest takes in a distribution and outputs a hyperparameter of the distribution that is often illustrative of the distribution's complexity. In particular, we shall consider the model selection problem of selecting $k$, the number of components in a Gaussian mixture model (GMM). GMMs are often used in modeling data, however the selection of how many components to use is often a difficult choice. Attempting an MLE fit to training data will lead to choosing $k=n_0$ with one mixture component corresponding to each data-point. Hence, in order to effectively select $k$, one must fit a GMM for each potential choice of $k$ using an algorithm such as the expectation maximization algorithm (EM)~\cite{moon1996expectation}, then select the choice of $k$ that optimizes some score. In practice this often becomes computationally expensive. Typically scores used include Akaike information criterion (AIC), Bayesian information criterion (BIC), or a cross-validated data-fitting score on a holdout set (CV). We note that often GMMs are used to cluster data, where each data-point $X_{0i}$ is a assigned to a cluster based on which mixture component most likely generated it. Hence, the problem of selecting the number of mixture components in a GMM is closely related to the problem of selecting the number of clusters to use, which is itself a difficult problem. Since selecting $k$ in GMMs is a DRR problem, and it is a relatively smooth mapping (that is, similar distributions should have a similar number of components), we hypothesize that one may learn to perform model selection in GMMs using the Double-Basis estimator. Particularly, by using a supervised dataset of $\{$sample-set, $k \}$ pairs, the Double-Basis estimator will be able to leverage previously seen data to perform model selection for a new unseen input sample set. Our experiment proceeds as follows. We can generate our own training data for this task by randomly drawing a value for $k$ (over some bounded range), then drawing 2-dimensional Gaussian mixture parameters for each of the $k$ components\footnote{See Appendix for figures of typical GMMs.}, and finally drawing samples from each Gaussian. That is, we generate $N=28,000$ input sample set/$k$ response pairs: $\mathcal{D} = \{(\mathcal{X}_i,k_i)\}_{i=1}^N$, where $\mathcal{X}_i = \{X_{i1},\ldots,X_{in}\}$, $X_{ij}\in\mathbb{R}^2$, $X_{ij}\overset{iid}{\sim }\GMM(k_i)$, $k_i\sim\text{Unif}\{1,\ldots,10\}$, and $\GMM(k_i)$ is a random GMM generated as follows, for $j=1,\ldots,k_i$: the prior weights for each component is taken to be $\pi_j=1/k_i$; the means are $\mu_j\sim\text{Unif}[-5,5]^2$; and covariances are $\Sigma_j = a^2AA^T+B$, where $a\sim\text{Unif}[1,2]$ $A_{uv}\sim\text{Unif}[-1,1]$, and $B$ is a diagonal $2\times2$ matrix with $B_{uu}\sim\text{Unif}[0,1]$. We train and get results using $n$ in the following range: $n \in {10,25,50,200,500,1000}$. We perform model selection using the mapping learned by the Ridge Double-Basis estimator \eqref{eq:ridgeest} (denoted BB in experiments), and compare it with model selection via AIC, BIC, and CV. We also compare agasint the Kernel-Kernel (KK) smoother. For all methods we computed the mean squared error between the true and predicted value for $k$ over 2000 test sample sets (Figure~\ref{fig:gmm_results}). We see that the Double-Basis estimator has both the lowest MSE and the lowest average evaluation time for computing a new prediction. In fact, the Double-Basis estimator can carry out the model selection prediction orders of magnitude faster than the CV, AIC, or BIC procedures. \begin{figure} \centering \subfigure[Estimation Error]{\raisebox{0mm}{\includegraphics[width=.235\textwidth]{figures/gmmExpResults_b.pdf}}} \subfigure[Estimation Time]{\raisebox{1mm}{\includegraphics[width=.23\textwidth]{figures/gmmExpResults_a.pdf}}} \vspace{-0.3cm} \caption{Results on predicting the number of GMM components. } \label{fig:gmm_results} \vspace{-0.5cm} \end{figure} \vspace{-0.2cm} \subsection{Low Sample Dirichlet Parameter Estimation} \vspace{-0.2cm} Similar to model selection, general parameter point estimation is a statistical task that may be posed as a DRR problem. That is, in parameter estimation one considers a set $\mathcal{X}_0 = \{X_{01},\ldots,X_{0n_0}\}$ where points are drawn from some distribution $P(\eta_0)$ that is parameterized by $\eta_0$, and attempts to estimate $\eta_0$. In particular, we use DRR and the Double-Basis estimator to perform parameter estimation for Dirichlet distributions. The Dirichlet distribution is a family of continuous, multivariate distributions parameterized by a vector $\alpha \in \mathbb{R}_+^d$, with support over the $d$-simplex. Since every element of the support sums to one, the Dirichlet is often used to model distributions over proportion data. As before, we hypothesize that the Double-Basis estimator will serve as a way to leverage previously seen sample sets to help perform parameter estimation for new unseen sets. Effectively, our estimator will be able to ``boost'' the sample-size of a new input sample set by making use of what was learned on previously seen labeled sample sets. Maximum likelihood parameter estimation for $\alpha$, given a set of Dirichlet samples, is often performed via iterative optimization algorithms, such as gradient ascent or Newton's method \cite{minka2000estimating}, as a closed form solution for the MLE does not appear to exist in the literature. In this experiment, we aim to use DDR as a new method for Dirichlet parameter estimation. In particular, we generate samples from Dirichlet distributions with parameter values in a prespecified range, and use these as training data to learn a mapping from data samples to Dirichlet $\alpha$ parameter values. In our experiments, we first fix the range of $\alpha$ values to be constrained such that the $i^{th}$ component $\alpha_i \in [0.1,10]$. For each $28,000$ training instances, we uniformly sample a new $\alpha$ parameter vector within this range, and then generate $n$ points from the associated Dirichlet$(\alpha)$ distribution, where $n \in \{10,25,50,200,500,1000\}$. We compare the Ridge Double-Basis estimator \eqref{eq:ridgeest} against a Newtons-method procedure for maximum likelihood estimation (MLE) from the fastfit toolbox \cite{minka2006fastfit}, and again against the Kernel-Kernel smoother. For all methods, for each $n$, we compute the mean squared error between the true and the estimated $\alpha$ parameter. We also record the time taken to perform the parameter estimation in each case. Results are shown in Figure~\ref{fig:dir_results}. We see that the Double-Basis estimator achieves the lowest MSE in all cases, and has the lowest average compute time. It is worth noting that the Double-Basis estimator performs particularly well relative to the MLE in cases where the sample size is low. We envision that Double-Basis estimator is particularly well suited for cases where one hopes to quickly, and in an automatic fashion, construct an estimator that can achieve highly accurate results for a statistical estimation problem for which an optimal estimator might be hard to derive analytically. \begin{figure} \centering \subfigure[Estimation Error]{\raisebox{0mm}{\includegraphics[width=.23\textwidth]{figures/dirExpResults_b.pdf}}} \subfigure[Estimation Time]{\includegraphics[width=.235\textwidth]{figures/dirExpResults_a.pdf}} \vspace{-0.3cm} \caption{Results predicting Dirichlet parameters. } \label{fig:dir_results} \vspace{-0.5cm} \end{figure} \vspace{-0.2cm} \section{Conclusion} \vspace{-.2cm} In conclusion, this paper presents a new estimator, the Double-Basis (BB) estimator, for performing distribution to real regression. In particular, this estimator scales independently of $N$ (the number input sample-set/response pairs) in a large dataset for performing evaluations for response predictions. This is a great improvement over the linear scaling with $N$ that the Kernel-Kernel (KK) estimator has and allows one to explore DRR in new domains with large collections of distributions, such as astronomy and finance. Furthermore, we prove an efficient upper bound on the risk for the BB estimator. Also, we empirically showed the improved scaling of the Double-Basis estimator, as well improvements in risk over the KK estimator. It is worth noting that while the BB estimator regresses a mapping in a nonlinear space (induced by RKS features), the KK estimator is outputs only a weighted average of training set responses \vspace{-.2cm} \subsubsection*{Acknowledgements} \vspace{-.2cm} This work is supported in part by NSF grants IIS1247658 and IIS1250350. \clearpage {\small \bibliographystyle{amsplain} \section{GENERAL FORMATTING INSTRUCTIONS} Both submitted and camera-ready versions of the paper are 8 pages, plus any additional pages needed for references. Papers are in 2 columns with the overall line width of 6.75~inches (41~picas). Each column is 3.25~inches wide (19.5~picas). The space between the columns is .25~inches wide (1.5~picas). The left margin is 1~inch (6~picas). Use 10~point type with a vertical spacing of 11~points. Times Roman is the preferred typeface throughout. Paper title is 16~point, caps/lc, bold, centered between 2~horizontal rules. Top rule is 4~points thick and bottom rule is 1~point thick. Allow 1/4~inch space above and below title to rules. Author descriptions are center-justified, initial caps. The lead author is to be listed first (left-most), and the Co-authors are set to follow. If up to three authors, use a single row of author descriptions, each one center-justified, and all set side by side; with more authors or unusually long names or institutions, use more rows. (But, do not include author names in the initial double-blind submission! Instead leave a row of ``Anonymous Author'' descriptions as above.) One-half line space between paragraphs, with no indent. \section{FIRST LEVEL HEADINGS} First level headings are all caps, flush left, bold, and in point size 12. One line space before the first level heading and 1/2~line space after the first level heading. \subsection{Second Level Heading} Second level headings are initial caps, flush left, bold, and in point size 10. One line space before the second level heading and 1/2~line space after the second level heading. \subsubsection{Third Level Heading} Third level headings are flush left, initial caps, bold, and in point size 10. One line space before the third level heading and 1/2~line space after the third level heading. \paragraph{Fourth Level Heading} Fourth level headings must be flush left, initial caps, bold, and Roman type. One line space before the fourth level heading, and place the section text immediately after the heading with, no line break, but an 11 point horizontal space. \subsection{CITATIONS, FIGURES, REFERENCES} \subsubsection{Citations in Text} Citations within the text should include the author's last name and year, e.g., (Cheesman, 1985). References should follow any style that you are used to using, as long as their style is consistent throughout the paper. Be sure that the sentence reads correctly if the citation is deleted: e.g., instead of ``As described by (Cheesman, 1985), we first frobulate the widgets,'' write ``As described by Cheesman (1985), we first frobulate the widgets.'' Be sure to avoid accidentally disclosing author identities through citations. \subsubsection{Footnotes} Indicate footnotes with a number\footnote{Sample of the first footnote.} in the text. Use 8 point type for footnotes. Place the footnotes at the bottom of the column in which their markers appear, continuing to the next column if required. Precede the footnote section of a column with a 0.5 point horizontal rule 1~inch (6~picas) long.\footnote{Sample of the second footnote.} \subsubsection{Figures} All artwork must be centered, neat, clean, and legible. All lines should be very dark for purposes of reproduction, and art work should not be hand-drawn. Figures may appear at the top of a column, at the top of a page spanning multiple columns, inline within a column, or with text wrapped around them, but the figure number and caption always appear immediately below the figure. Leave 2 line spaces between the figure and the caption. The figure caption is initial caps and each figure should be numbered consecutively. Make sure that the figure caption does not get separated from the figure. Leave extra white space at the bottom of the page rather than splitting the figure and figure caption. \begin{figure}[h] \vspace{.3in} \centerline{\fbox{This figure intentionally left non-blank}} \vspace{.3in} \caption{Sample Figure Caption} \end{figure} \subsubsection{Tables} All tables must be centered, neat, clean, and legible. Do not use hand-drawn tables. Table number and title always appear above the table. See Table~\ref{sample-table}. One line space before the table title, one line space after the table title, and one line space after the table. The table title must be initial caps and each table numbered consecutively. \begin{table}[h] \caption{Sample Table Title} \label{sample-table} \begin{center} \begin{tabular}{ll} {\bf PART} &{\bf DESCRIPTION} \\ \hline \\ Dendrite &Input terminal \\ Axon &Output terminal \\ Soma &Cell body (contains cell nucleus) \\ \end{tabular} \end{center} \end{table} \section{SUPPLEMENTARY MATERIAL} If you need to include additional appendices during submission, you can include them in the supplementary material file. \newpage \section{INSTRUCTIONS FOR CAMERA-READY PAPERS} For the camera-ready paper, if you are using \LaTeX, please make sure that you follow these instructions. (If you are not using \LaTeX, please make sure to achieve the same effect using your chosen typesetting package.) \begin{enumerate} \item Install the package \texttt{fancyhdr.sty}. The \texttt{aistats2014.sty} file will make use of it. \item Begin your document with \begin{flushleft} \texttt{\textbackslash documentclass[twoside]\{article\}}\\ \texttt{\textbackslash usepackage[accepted]\{aistats2014\}} \end{flushleft} The \texttt{twoside} option for the class article allows the package \texttt{fancyhdr.sty} to include headings for even and odd numbered pages. The option \texttt{accepted} for the package \texttt{aistats2014.sty} will write a copyright notice at the end of the first column of the first page. This option will also print headings for the paper. For the \emph{even} pages, the title of the paper will be used as heading and for \emph{odd} pages the author names will be used as heading. If the title of the paper is too long or the number of authors is too large, the style will print a warning message as heading. If this happens additional commands can be used to place as headings shorter versions of the title and the author names. This is explained in the next point. \item If you get warning messages as described above, then immediately after $\texttt{\textbackslash begin\{document\}}$, write \begin{flushleft} \texttt{\textbackslash runningtitle\{Provide here an alternative shorter version of the title of your paper\}}\\ \texttt{\textbackslash runningauthor\{Provide here the surnames of the authors of your paper, all separated by commas\}} \end{flushleft} The text that appears as argument in \texttt{\textbackslash runningtitle} will be printed as a heading in the \emph{even} pages. The text that appears as argument in \texttt{\textbackslash runningauthor} will be printed as a heading in the \emph{odd} pages. If even the author surnames do not fit, it is acceptable to give a subset of author names followed by ``et al.'' \item Use the file sample\_paper.tex as an example. \item Both submitted and camera-ready versions of the paper are 8 pages, plus any additional pages needed for references. \item If you need to include additional appendices, you can include them in the supplementary material file. \item Please, don't change the layout given by the above instructions and by the style file. \end{enumerate} \subsubsection*{Acknowledgements} Use unnumbered third level headings for the acknowledgements. All acknowledgements go at the end of the paper. Be sure to omit any identifying information in the initial double-blind submission! \subsubsection*{References} References follow the acknowledgements. Use an unnumbered third level heading for the references section. Any choice of citation style is acceptable as long as you are consistent. Please use the same font size for references as for the body of the paper---remember that references do not count against your page length total. J.~Alspector, B.~Gupta, and R.~B.~Allen (1989). Performance of a stochastic learning microchip. In D. S. Touretzky (ed.), {\it Advances in Neural Information Processing Systems 1}, 748-760. San Mateo, Calif.: Morgan Kaufmann. F.~Rosenblatt (1962). {\it Principles of Neurodynamics.} Washington, D.C.: Spartan Books. G.~Tesauro (1989). Neurogammon wins computer Olympiad. {\it Neural Computation} {\bf 1}(3):321-323. \end{document}
\section{Introduction}\label{sec:introduction} Suffix sorting---the process of ordering all the suffixes of a string into lexicographical order---is the key step in construction of suffix arrays and the Burrows-Wheeler transform, two of the most important structures in text indexing and biological sequence analysis~\cite{OhlebuschBook,MBCT15,AKO04}. As such, algorithms for efficient suffix sorting have been the focus of intense research since the early 1990s~\cite{MM93,PST07}. With the rise of pangenomics, there is an increased demand for indexes that support fast pattern matching over collections of genomes of individuals of the same species (see, e.g.,~\cite{GNP20,PZ20,ROBLGB22}). With pangenomic collections constantly growing and changing, construction of these indexes---and in particular suffix sorting---is a computational bottleneck in many bioinformatics pipelines. While traditional and well-established suffix sorting tools such as {\tt divsufsort}~\cite{divsufsort,0001K17} and {\tt sais}~\cite{sais-lite,NZC11} can be applied to these collections, specialised algorithms for collections of similar sequences, perhaps most notably the so-called {\tt BigBWT} program~\cite{BoucherGKLMM19}, are beginning to emerge. In this paper we describe a suffix sorting algorithm specifically targeted to collections of highly similar genomes that makes use of the {\em matching statistics}, a data structure due to Chang and Lawler, originally used in the context of approximate pattern matching~\cite{CL94}. The core device in our suffix sorting algorithm is a novel compressed representation of the matching statistics of every genome in the collection with respect to a designated reference genome, that allows determining the relative order of two arbitrary suffixes, from any of the genomes, efficiently. We use this data structure to drive a suffix sorting algorithm that has a small working set relative to the size of the whole collection, with the aim of increasing locality of memory reference. Experimental results with a prototype implementation show the new approach to be faster or competitive with state-of-the-art methods for suffix array construction, including those targeted at highly repetitive data. We also provide a fast, practical algorithm for matching statistics computation, which is of independent interest. The remainder of this paper is structured as follows. The next section sets notation and defines basic concepts. In Section~\ref{sec:ms} we describe a compressed representation of the matching statistics and a fast algorithm for constructing it. Section~\ref{sec:comparing} then describes how to use the compressed matching statistics to determine the relative lexicographic order of two arbitrary suffixes of the collection. Section~\ref{sec:putting} describes a complete suffix sorting algorithm. We touch on several implementation details in Section~\ref{sec:implementation}, before describing experimental results in Section~\ref{sec:experiments}. Reflections and avenues for future work are then offered. \section{Basics}\label{sec:basics} A string $T$ over an ordered alphabet $\Sigma$, of size $|\Sigma| = \sigma$, is a finite sequence $T=T[1..n]$ of characters from $\Sigma$. We use the notation $T[i]$ for the $i$th character of $T$, $|T|$ for its length $n$, and $T[i..j]$ for the substring $T[i]\cdots T[j]$; if $i>j$ then $T[i..j]=\epsilon$, where $\epsilon$ is the empty string. The substring (or factor) $T[i..]=T[i..n]$ is called the $i$th suffix, and $T[..i]=T[1..i]$ the $i$th prefix of $T$. We assume throughout that the last character of each string is a special character $\$$, not occurring elsewhere in $T$, which is set to be smaller than every character in $\Sigma.$ Given a string $T$, the {\em suffix array} $\textit{SA}$ is a permutation of the index set $\{1,\ldots,n\}$ defined by: $\textit{SA}[i]=j$ if the $j$th suffix of $T$ is the $i$th in lexicographic order among all suffixes of $T$. The {\em inverse suffix array} $\textit{ISA}$ is the inverse permutation of $\textit{SA}$. The {\em \textit{LCP}-array} is given by: $\textit{LCP}[1]=0$, and for $i\geq 2$, $\textit{LCP}[i]$ is the length of the longest common prefix (lcp) of the two suffixes $T[\textit{SA}[i-1]..]$ and $T[\textit{SA}[i]..]$ (which are consecutive in lexicographic order). A variant of the $\textit{LCP}$ array is the {\em permuted $\textit{LCP}$-array}, $\textit{PLCP}$, defined as $\textit{PLCP}[i] = \textit{LCP}[\textit{ISA}[i]]$, i.e.\ the lcp values are stored in text order, rather than in $\textit{SA}$ order. We further define $\textit{LCPsum}(T)=\sum_{i=1}^{|T|} \textit{LCP}[i]$. $\textit{LCPsum}$ can be used as a measure of repetitiveness of strings, since the number of distinct substrings of $T$ equals ${(|T|^2+|T|)}/{2} - \textit{LCPsum}(T)$. All these arrays can be computed in linear time in $|T|$, see e.g.~\cite{NZC11,KMP09}. Given the suffix array $\textit{SA}$ of $T$ and a substring $U$ of $T$, the indices of all suffixes which have $U$ as prefix appear consecutively in $\textit{SA}$. We refer to this interval as {\em $U$-interval}: the $U$-interval is $\textit{SA}[s..e]$, where $\{ \textit{SA}[s],\textit{SA}[s+1],\ldots,\textit{SA}[e-1],\textit{SA}[e]\}$ are the starting positions of the occurrences of $U$ in $T$. Let ${\cal C} = \{S_1,\ldots,S_m\}$ be a collection of strings (a set or multiset). The {\em generalized suffix array} $\textit{GSA}$ of ${\cal C}$ is defined as $\textit{GSA}[i]=(d,j)$ if $S_d[j..]$ is the $i$th suffix in lexicographic order among all suffixes of the strings from ${\cal C}$, where ties are broken by the document index $d$. The \textit{GSA}\ can be computed in time ${\cal O}(N)$, where $N$ is the total length of strings in ${\cal C}$~\cite{OhlebuschBook}. Let $R$ and $S$ be two strings. The {\em matching statistics of $S$ with respect to $R$} is an array $\textit{MS}$ of length $|S|$, defined as follows. Let $U$ be the longest prefix of suffix $S[i..]$ which occurs in $R$ as a substring, where the end-of-string character $\#$ of $R$ is assumed to be different from, and smaller than that of $S$. Then $\textit{MS}[i] = (p_i,\ell_i)$, where $p_i=-1$ if $U=\epsilon$, and $p_i$ is an occurrence of $U$ in $R$ otherwise, and $\ell_i = |U|$. (Note that $p_i$ is not unique in general.) We refer to $U$ as the {\em matching factor}, and to the character $c$ immediately following $U$ in $S$ as the {\em mismatch character}, of position $i$. For a collection ${\cal C} = \{S_1,\ldots,S_m\}$ and a string $R$, the matching statistics of ${\cal C}$ w.r.t.\ $R$ is simply the concatenation of $\textit{MS}_i$'s, where $\textit{MS}_i$ is the matching statistics of $S_i$ w.r.t.\ $R$. We will discuss matching statistics in more detail in Section~\ref{sec:ms}. For an integer array $A$ of length $n$ and an index $i$, the previous and next smaller values, $\textit{PSV}$ resp.\ $\textit{NSV}$, are defined as $\textit{PSV}(A, i) = \max\{i' < i : A[i'] < A[i]\}$ resp.\ $\textit{NSV}(A, i) = \min\{i' > i : A[i'] < A[i]\}$. Note that $\textit{PSV}$ resp.\ $\textit{NSV}$ is not defined for $i = 1$ resp. $i = n$. In $O(n)$ preprocessing of $A$, a data structure of size $n \log_2(3 + 2\sqrt{2}) + o(n)$ bits can be built that supports answering arbitrary \textit{PSV}\ and \textit{NSV}\ queries in constant time per query~\cite{F11}. Let $X$ be a finite set of integers. Given an integer $x$, the predecessor of $x$, $\textit{pred}(x)$ is defined as the largest element smaller than $x$, i.e. $\textit{pred}_X(x) = \max\{y\in X \mid y \leq x\}$. Using the y-fast trie data structure of Willard~\cite{W83} allows answering predecessor queries in $O(\log\log |X|)$ time using $O(|X|)$ space. We are now ready to state our problem: \begin{quote} {\bf Problem Statement:} Given a string collection ${\cal C} = \{ S_1,\ldots, S_m\}$ and a reference string $R$, compute the generalized suffix array $\textit{GSA}$ of ${\cal C}$. \end{quote} We will denote the length of $R$ by $n$ and the total length of strings in the collection by $N=\sum_{d=1}^m |S_d|$. As before, we assume that the end-of-string character $\#$ of $R$ is strictly smaller than those of the strings in the collection ${\cal C}$. We are interested in those cases where $\textit{LCPsum}_R$ is small and the strings in ${\cal C}$ are very similar to $R$. If no reference string is given in input, we will take $S_1$ to be the reference string by default. \subsection{Efficient suffix array construction}\label{sec:SACA} Currently, the best known and conceptually simplest linear-time suffix array construction algorithm is the SAIS algorithm by Nong et al.~\cite{NZC11}. It cleverly combines, and further develops, several ideas used by previous suffix array construction algorithms, among these {\em induced sorting}, and use of a so-called {\em type array}, already used in~\cite{IT99,KoA05} (see also~\cite{PST07}). Nong et al.'s approach can be summarized as follows: assign a type to each suffix, sort a specific subset of suffixes, and compute the complete suffix array by inducing the order of the remaining suffixes from the sorted subset. There are three types of suffixes, one of which constitutes the subset to be sorted first. The definition of types is as follows (originally from~\cite{KoA05}, extended in~\cite{NZC11}): Suffix $i$ is {\em $S$-type} (smaller) if $T[i..] < T[i+1..]$, and {\em $L$-type} (larger) if $T[i..] > T[i+1..]$. An $S$-type suffix is {\em $S^*$-type} if $T[i..]$ is $S$-type and $T[i-1..]$ is $L$-type. It is well known that assigning a type to each suffix can be done with a back-to-front scan of the text in linear time. Now, if the relative order of the $S^*$-suffixes is known, then that of the remaining suffixes can be induced with two linear scans over the partially filled-in suffix array: the first scan to induce $L$-type suffixes, and the second to induce $S$-type suffixes. For details, see~\cite{NZC11} or~\cite{OhlebuschBook}. Another ingredient of SAIS, and of several other suffix array construction algorithms, is what we term the {\em metacharacter method}. Subdivide the string $T$ into overlapping substrings, show that if two suffixes start with the same substring, then their relative order depends only on the remaining part; assign metacharacters to these substrings according to their rank (w.r.t.\ the lexicographic order, or some other order, depending on the algorithm), and define a new string on these metacharacters. Then the relative order of the suffixes of the new string and the corresponding suffixes starting with these specific substrings will coincide. In SAIS~\cite{NZC11}, so-called LMS-substrings are used, while a similar method is applied in prefix-free-parsing (PFP)~\cite{BoucherGKLMM19}. Here we will apply this method using substrings starting in special positions which we term insert-heads, see Sections~\ref{sec:comparing} and~\ref{sec:putting} for details. \section{Compressed matching statistics}\label{sec:ms} Let $R,S$ be two strings over $\Sigma$ and $\textit{MS}$ be the matching statistics of $S$ w.r.t.\ $R$. Let $\textit{MS}[i] = (p_i,\ell_i)$. It is a well known fact that if $\ell_i>0$, then $\ell_{i+1} \geq \ell_i-1$. This can be seen as follows. Let $U$ be the matching factor of position $i$, and $p_i$ an occurrence of $U$ in $R$. Then $U' = U[2..\ell_i]$ is a prefix of $S[i+1..]$ of length $\ell_i-1$, which occurs in position $p_i+1$ of $R$. Let us call a position $j$ a {\em head} if $\ell_j > \ell_{j-1}-1$, and a sequence of the form $(x,x-1,x-2,\ldots)$, of length at most $x-1$, a {\em decrement run}, i.e.\ each element is one less than the previous one. Using this terminology, we thus have that the sequence $L=(\ell_1,\ell_2,\ldots, \ell_n)$ is a concatenation of decrement runs, i.e.\ $L$ has the form $(x_1,x_1-1,x_1-2,\ldots, x_2,x_2-1,x_2-2,\ldots,\ldots,x_k,x_k-1,x_k-2,\ldots)$, with each $x_j=\ell_j$ for some head $j$. We can therefore store the matching statistics in compressed form as follows: \begin{definition}[Compressed matching statistics] Let $R,S$ be two strings over $\Sigma$, and $\textit{MS}$ be the matching statistics of $S$ w.r.t.\ $R$. The {\em compressed matching statistics (\textit{CMS}) of $S$ w.r.t.\ $R$} is a data structure storing $(j,\textit{MS}[j])$ for each head $j$, and a predecessor data structure on the set of heads $H$. \end{definition} We can use $\textit{CMS}$ to recover all values of $\textit{MS}$: \begin{lemma}\label{lemma:CMS} Let $1\leq i\leq |S|$. Then $\textit{MS}[i] = (p_j+k,\ell_j-k)$, where $j = \textit{pred}_H(i)$ and $k = i-j$. \end{lemma} \begin{proof} Let $\ell_i$ be the length of the matching factor of $i$. Since there is a matching factor of length $\ell_j$ starting in position $j$ in $S$, this implies that $\ell_i \geq \max(0,\ell_j-k)$. If $\ell_i$ was strictly greater than $\ell_j-k$, this would imply the presence of another head between $j$ and $i$, in contradiction to $j = \textit{pred}_H(i)$. Since an occurrence of the matching factor $U_j$ of $j$ starts in position $p_j$ of $R$, therefore the matching factor $U'=U[k+1..\ell_j]$ of $i$ has an occurrence at position $p_j+k$. \end{proof} \begin{figure}[h!] \centering \begin{tabular}{|r|r|r|r|r|r|r|r|r|r|r|r|r|r|r|r|r|r|r} \hline $i$ & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10 & 11 & 12 & 13 & 14 & 15 & 16 & 17 \\ \hline $R$ & {\tt T} & {\tt G} & {\tt A} & {\tt T} & {\tt G} & {\tt G} & {\tt C} & {\tt A} & {\tt C} & {\tt A} & {\tt G} & {\tt A} & {\tt T} & {\tt A} & {\tt C} & {\tt T} & {\tt \#} \\ \hline \hline $S$ & {\tt G} & {\tt A} & {\tt T} & {\tt G} & {\tt G} & {\tt C} & {\tt A} & {\tt C} & {\tt A} & {\tt T} & {\tt T} & {\tt G} & {\tt A} & {\tt T} & {\tt G} & {\tt G} & {\tt \$} \\ $p_i$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 12 & 13 & 1 & 2 & 3 & 4 & 5 & 6 & -1\\ $\ell_i$ & 9 & 8 & 7 & 6 & 5 & 4 & 3 & 2 & 2 & 1 & 6 & 5 & 4 & 3 & 2 & 1 & 0 \\ head & \checkmark & & & & & & & & \checkmark & & \checkmark & & & & & & \\ \hline $q_i$ & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 3 & 4 & 1 & 2 & 3 & 4 & 5 & 11 & 17\\ i-head & \checkmark & & & & & & & & \checkmark & & \checkmark & & & & & \checkmark & \checkmark \\ \hline \end{tabular} \caption{Example for the matching statistics and the data for the \textit{CMS}\ and the \textit{eCMS}. In the first two rows, we give $\textit{MS}$ of $S$ w.r.t.\ $R$, where $\textit{MS}[i]=(p_i,\ell_i)$. In row 3, we mark the heads (for the \textit{CMS}). In rows 4, we give the position $q_i$, defined by $ip(i)$, i.e.\ $q_i=\textit{SA}_R[ip(i)]$, where $ip(i)$ is the insert-point of suffix $S[i..]$ in the suffix array of $R$. In row $5$, we mark the insert-heads (for the \textit{eCMS}).} \label{fig:example_eCMS} \end{figure} \begin{figure}[ht] \centering \begin{tabular}{|r|r|r|l|} \hline & $i$ & $\textit{SA}_R$ & $R[{\textit{SA}_R[i]}..]$\\ \hline & 1 & 17 & {\tt \#} \\ & 2 & 8 & {\tt ACAGATACT\#} \\ & 3 & 14 & {\tt ACT\#} \\ & 4 & 10 & {\tt AGATACT\#} \\ & 5 & 12 & {\tt \textcolor{blue}{AT}ACT\#} \\ $\rightarrow$ & 6 & 3 & {\tt \textcolor{blue}{AT}GGCACAGATACT\#} \\ & 7 & 7 & {\tt CACAGATACT\#} \\ & 8 & 9 & {\tt CAGATACT\#} \\ & 9 & 15 & {\tt CT\#} \\ & 10 & 11 & {\tt GATACT\#} \\ & 11 & 2 & {\tt GATGGCACAGATACT\#} \\ & 12 & 6 & {\tt GCACAGATACT\#} \\ & 13 & 5 & {\tt GGCACAGATACT\#} \\ & 14 & 16 & {\tt T\#} \\ & 15 & 13 & {\tt TACT\#} \\ $\rightarrow$ & 16 & 1 & {\tt \textcolor{blue}{TGATGG}CACAGATACT\#} \\ & 17 & 4 & {\tt TGGCACAGATACT\#} \\ \hline \end{tabular} \caption{Details of computation of the matching statistics from Figure \ref{fig:example_eCMS}. We highlight in blue the matching factors for the indices $i=9$ (matching factor ${\tt AT}$, mismatch character ${\tt T}$) and $11$ (matching factor ${\tt TGATGG}$, mismatch character ${\tt \$}$). The arrows represent the insert-points.} \label{fig:insertion-point} \end{figure} \begin{example}\label{ex:1} Consider the reference $R={\tt TGATGGCACAGATACT}$ and $S=$ ${\tt GATGGCACATTGATGG}$. The \textit{CMS}\ of $S$ w.r.t.\ $R$ is: $(1,2,9), (9,12,2), (11,1,6)$, see Figure~\ref{fig:example_eCMS}. \end{example} From Lemma~\ref{lemma:CMS} and the properties of the predecessor data structure on the set of heads we get: \begin{proposition}\label{prop:CMS} Let $R,S$ be two strings over $\Sigma$. We can store the matching statistics of $S$ w.r.t.\ $R$ in ${\cal O}(\chi)$ space such that any entry $\textit{MS}[i]$, for $1\leq i \leq |S|$, can be accessed in ${\cal O}(\log\log \chi)$ time, where $\chi=|H|$ is the number of heads. \end{proposition} For some statistics on the number $\chi$ of heads, see the end of Sec.~\ref{sec:enhanced_cms}. \subsection{Enhancing the CMS}\label{sec:enhanced_cms} Let $R,S$ be two strings over $\Sigma$, and $\textit{MS}$ the matching statistics of $S$ w.r.t.\ $R$. We now assume that all characters that occur in $S$ also occur in $R$ (see Sec.~\ref{sec:implementation}). Let $\textit{SA}_R$ be the suffix array of $R$. For position $i$ of $S$, let $U\neq \epsilon$ be the matching factor and $c$ the mismatch character of $i$. We want to compute the position that the suffix $S[i..]$ would have in $\textit{SA}_R$ if it was present. To this end, we define the {\em insert point} of $i$, $ip(i)$, as follows: \begin{align*} ip(i) = \begin{cases} 1 & \text{ if $U=\epsilon$},\\ \max\{ j \mid U \text{ occurs in } \textit{SA}_R[j] \text{ and } R[\textit{SA}_R[j]..] < Uc\} & \text{ if this set is non-empty,}\\ \min\{j \mid U \text{ occurs in } \textit{SA}_R[j]\} & \text{ otherwise.} \end{cases} \end{align*} In other words, the insert point is the lexicographic rank, among all suffixes of $R$, of the next smaller occurrence of $U$ in $R$ if such an occurrence exists, and of the smallest occurrence of $U$ in $R$ otherwise. Note that case 1 (where $U=\epsilon$) only happens for end-of-string characters. The insert point is well-defined for every $i$ because $\#$ is smaller than all other characters, including other end-of-string characters. Observe that the insert point of $i$ always lies within the $U$-interval of $\textit{SA}_R$. For an example, see Fig.~\ref{fig:insertion-point}. We will later use the insert points to bucket suffixes. First we need to slightly change the definition of our compressed matching statistics. We will add more information to the heads: we add the mismatch character and replace the position entry $p_i$, which gives just some occurrence of the matching factor, by the specific occurrence $q_i$ given by the insert point. This will imply adding more heads, so our data structure may increase in size. To this end, we define $j$ to be an {\em insert-head} if $\textit{SA}_R[ip(j)] \neq \textit{SA}_R[ip(j-1)]+1$. Note that, in particular, all heads are also insert-heads, but it is possible to have insert-heads $j$ which are not heads, namely where $\ell_j = \ell_{j-1}-1$. \begin{definition}[Enhanced compressed matching statistics] Let $R,S$ be two strings over $\Sigma$. Define the {\em enhanced matching statistics} as follows: for each $1\leq i \leq |S|$, let $\textit{ems}(i) = (q_i,\ell_i,x_i,c_i)$, where $q_i = \textit{SA}_R[ip(i)]$, $\ell_i$ is the length of the matching factor $U$ of $i$, $c_i$ is the mismatch character, and $x_i\in \{S,L\}$ indicates whether $Uc_i$ is smaller (S) or greater (L) than $R[q_i..]$. The {\em enhanced compressed matching statistics (\textit{eCMS}) of $S$ w.r.t.\ $R$} is a data structure storing $(j,\textit{ems}(j))$ for each insert-head $j$, and a predecessor data structure on the set of insert-heads $H'$. \end{definition} \begin{example}\label{ex:2} Continuing with Example~\ref{ex:1}, the enhanced \textit{CMS}\ of $S$ w.r.t.\ $R$ is: $(1,2,9,L,{\tt T})$, $(9,3,2,L,{\tt T})$, $(11,1,6,S,{\tt \$}),$ $(16,11,1,S,{\tt \$})$, $(17,17,0,L,{\tt \$})$, see Figure~\ref{fig:example_eCMS}. \end{example} We will need some properties of the insert point in the following: \begin{observation}\label{obs:ip} Let $ip(i)$ be the insert point of $i$, and $\textit{ems}(i) = (q_i,\ell_i,x_i,c_i)$. \begin{enumerate} \item $ip(i) = ip(i')$ if and only if $q_i = q_{i'}$, \item if $x_i=S$ then $R[\textit{SA}_R[ip(i)-1]..] < S[i..] < R[\textit{SA}_R[ip(i)]..] = R[q_i..]$, \item if $x_i=L$ then $R[q_i..] = R[\textit{SA}_R[ip(i)]..] < S[i..] < R[\textit{SA}_R[ip(i)+1]..]$. \end{enumerate} \end{observation} The enhanced CMS can be used in a similar way as the CMS to recover the enhanced matching statistics (including the matching statistics) of each $i$. Denote by $\textit{i-head}(i)$ the next insert-head to the left of $i$, i.e.\ $\textit{i-head}(i) = \max\{j\leq i \mid j \text{ is an insert-head}\}$. Note that $\textit{i-head}(i) = \textit{pred}_{H'}(i)$. \begin{lemma}\label{lemma:eCMS} Let $1\leq i \leq |S|$, let $\textit{eCMS}$ be the enhanced CMS of $S$ w.r.t.\ $R$. Let $j = \textit{i-head}(i)$, $k=i-j$, and $\textit{ems}(j) = (q_j,\ell_j,x_j,c_j)$. Then $\textit{ems}(i) = (q_j+k,\ell_j-k,x_j,c_j)$, and $ip(i) = \textit{ISA}_R[q_j+k]$. In particular, $q_j+k$ is an occurrence and $\ell_j-k$ is the length of the matching factor of $i$ (in other words, the matching statistics entry $\textit{MS}[i]$). \end{lemma} \begin{proof} Analogous to Lemma~\ref{lemma:CMS}, resp.\ straightforward from the definitions. \end{proof} Similarly to the \textit{CMS}\ (cp. Prop.~\ref{prop:CMS}), the enhanced \textit{CMS}\ allows access to all values for every index $i$, using space ${\cal O}(\chi')$ and time ${\cal O}(\log \log \chi')$, where $\chi'=|H'|$ is the number of insert-heads. Again, this is due to the fact that the predecessor data structure on the set $H'$ of insert-heads allows retrieving $\textit{pred}_{H'}(i) =\textit{i-head}(i)$ in ${\cal O}(\log \log |H'|)$ time, and the values of $\textit{ems}(i)$ can then be computed in ${\cal O}(1)$ time (Lemma~\ref{lemma:eCMS}). \medskip We close this subsection by remarking that for a collection of similar genomes, one can expect the number of heads to be small. Indeed, on a 500MB viral genome data set (see Section~\ref{sec:experiments}) containing approximately 10,000 SARS-cov2 genomes, we observed the number of heads to be 5,326,226 (100x less than the input size) and the number of insert heads to be 6,537,294. \subsection{Computing the CMS}\label{sec:computing_cms} It is well known that the matching statistics of $S$ w.r.t.\ $R$ can be computed in time $O(|R| + |S|\log\sigma)$ and $O(|R|)$ space by using, for example, the suffix tree of $R$, as described in Chang and Lawler's original paper~\cite{CL94}. Since then, several authors have described similar algorithms for computing matching statistics, all focussed on reducing space requirements via the use of compressed indexes instead of the suffix tree~\cite{AKO04,OGK10,BCD18}. These algorithms all incur the slowdowns typical of compressed data structures. In our setting, where end-to-end runtime is the priority, it is the speed at which the matching statistics can be computed (rather than working space) that is paramount. Moreover, because the size of the reference is generally small relative to the total length of all the strings $S_i \in {\cal C}$, we have some freedom to use large index data structures on $R$ to compute the matching statistics, without overall memory usage getting out of hand. With these factors in mind, we take the following approach to computing CMS. The algorithm is similar to that of Chang and Lawler, but makes use of array-based data structures rather than the suffix tree. Recall that, given the suffix array $\textit{SA}_R$ of string $R$ and a substring $\textit{Y}$ of $R$, the $\textit{Y}$-interval is the interval $\textit{SA}_R[s..e]$ that contains all suffixes having $\textit{Y}$ as a prefix. \begin{definition}[Right extension and left contraction] For a character $c$ and a string $\textit{Y}$, the computation of the $\textit{Yc}$-interval from the $\textit{Y}$-interval is called a {\em right extension} and the computation of the $\textit{Y}$-interval from $\textit{cY}$-interval is called a {\em left contraction}. \end{definition} We remark that a left contraction is equivalent to following a (possibly implicit) suffix link in the suffix tree of $R$ and a right extension is a downward movement (either to a child or along an edge) in the suffix tree of $R$. Given a $Y$-interval, because of the lexicographical ordering on the $\textit{SA}_R$, we can implement a right extension to a $\textit{Yc}$-interval in $O(\log |R|)$ time by using a pair of binary searches (with $c$ as the search key), one to find the lefthand end of the $\textit{Yc}$-interval and another to find the righthand end. If a right extension is empty then there are no occurrences of $\textit{Yc}$ in $R$, but we can have the binary search return to us the insert point where it would have been in $\textit{SA}_R$. On the other hand, given a $cY$-interval, $\textit{SA}_R[s..e]$, we can compute the $Y$-interval (i.e. perform a left contraction) in the following way. Let the target $Y$-interval be $\textit{SA}_R[x..y]$. Observe that both $\textit{SA}_R[s]+1$ and $\textit{SA}_R[e]+1$ must be inside the $Y$-interval, $SA_R[x..y]$---that is, $s' = \textit{ISA}_R[\textit{SA}_R[s]+1] \in [x..y]$ and $e' = \textit{ISA}_R[\textit{SA}_R[e]+1] \in [x..y]$. To finish computing $\textit{SA}_R[x..y]$ from $\textit{SA}_R[s'..e']$ there are two cases to consider. Firstly, if $s' = e'$ and $|Y| > \textit{LCP}_R[s']$, then $\textit{SA}_R[s']$ is the only occurrence of $Y$ and we are done (the $Y$-interval is a singleton). Alternatively, $s' \ne e'$ and we compute $\textit{SA}_R[x..y]$ using $\textit{NSV}/\textit{PSV}$ queries on $LCP_R$, in particular $\textit{SA}_R[x..y] = \textit{SA}_R[\textit{PSV}(\textit{LCP}_R,s')..\textit{NSV}(\textit{LCP}_R,e')]$. With these ideas in place, we are ready to describe the matching statistics algorithm. We first compute $\textit{SA}_R$, $\textit{ISA}_R$, and $\textit{LCP}_R$ for $R$ and preprocess $\textit{LCP}_R$ for $\textit{NSV}/\textit{PSV}$ queries. The elements of the $\textit{MS}$ will be computed in left-to-right order, $\textit{MS}[1], \textit{MS}[2], \ldots, \textit{MS}[|S|]$. Note that this makes it trivial to save only the heads (or iheads) and so compute the CMS (or eCMS) instead. To find $\textit{MS}[1]$ use successive right extensions starting with the interval $SA_R[1..|R|]$, searching with successive characters of $S[1..]$ until the right extension is empty, at which point we know $\ell_1$ and $p_1$. At a generic step in the algorithm, immediately after computing $\textit{MS}[i]$, we know the interval $\textit{SA}_R[s_i..e_i]$ containing all the occurrences of $R[p_i..p_i+\ell_i-1]$. To compute $\textit{MS}[i+1]$ we first compute the left contraction of $\textit{SA}_R[s_i..e_i]$, followed by as many right contractions as possible until $\ell_{i+1}$ and $p_{i+1}$ are known. \medskip When profiling an implementation of the above algorithm, we noticed that very often the sequence of right extensions ended with a singleton interval (i.e., an interval of size one) and so was the interval reached by the left contraction that followed. In terms of the suffix tree, this corresponds to the match between $R$ and the current suffix of $S_i$ being inside a leaf branch. This frequently happens on genome collections because each sequence is likely to have much longer matches with other sequences (in this case with $R$) than it does with itself (a single genome tends to look fairly random, at least by string complexity measures). A simple heuristic to exploit this phenomenon is to compare $\ell_i$ to the maximum value in the entire $\textit{LCP}_R$ array of $R$ immediately after $\textit{MS}[i]$ has been computed. If $\ell_i-1 > \max(\textit{LCP}_R)$ then $\textit{ISA}_R[p_i+1]$ will also be inside a leaf branch (i.e., the left contraction will also be a singleton interval), and so the left contraction can be computed trivially as $\textit{ISA}_R[p_i+1]$---with no subsequent $\textit{NSV}/\textit{PSV}$ queries or access to $\textit{LCP}_R$ required to expand the interval. Although this gives no asymptotic improvement, there is potential gain from the probable cache miss(es) avoided by not making random accesses to those large data structures. On a viral genome data set (see Section~\ref{sec:experiments}), $\max(\textit{LCP}_R)$ was 14, compared to an average $\ell_i$ value of over $1,100$, and this heuristic saved lots of computation. On a human chromosome data set, however, $\max(\textit{LCP}_R)$ was in the hundreds of thousands, and so we generalized the trick in the following way. We divide the LCP array up into blocks of size $b$ and compute the minimum of each block. These minima are stored in an array $M$ of size $|R|/b$, and $b$ is chosen so that $M$ is small enough to comfortably fit in cache. Now, when transitioning from $\textit{MS}[i]$ to $\textit{MS}[i+1]$, if $\ell_i > M[\textit{ISA}_R[p_i+1]/b]$ then there is a single match corresponding to $\textit{MS}[i+1]$, which we compute with right extensions. This generalized form of the heuristic has a consistent and noticeable effect in practice. For a 500MB viral genome data set its use reduced CMS computation from 12.23 seconds to 2.34 seconds. On the human chromosome data set the effect is even more dramatic: from 76.50 seconds down to 7.14 seconds. \section{Comparing two suffixes via the enhanced CMS}\label{sec:comparing} We will now show how to use the enhanced CMS of the collection ${\cal C}$ w.r.t.\ $R$ to define a partial order on the set of suffixes of strings in ${\cal C}$ (Prop.~\ref{prop:PO}), and how to break ties when the entries are identical (Lemma~\ref{lemma:sufcomp_3}). These results can then be used either directly to determine the relative order of any two of the suffixes (Prop.~\ref{prop:suffix_comparison}), or as a way of inducing the complete order once that of the subset of the insert-heads has been determined (Prop.~\ref{prop:suffix_sort}). We will prove Prop.~\ref{prop:PO} via two lemmas. Recall that in the \textit{eCMS}\ we only have the entries referring to the insert-heads; however, Lemma~\ref{lemma:eCMS} tells us how to compute them for any position. \begin{lemma}\label{lemma:sufcomp_1} Let $1\leq d,d' \leq m$ and $1\leq i\leq |S_d|$, $1\leq i'\leq |S_{d'}|$. If $ip(d,i) < ip(d',i')$, then $S_d[i..] < S_{d'}[i'..]$. \end{lemma} \begin{proof} If $ip(d',i')-ip(d,i)>1$, then there exists an index $j$ s.t.\ $ip(d,i)<j<ip(d',i')$, and therefore $S_d[i..] < R[\textit{SA}_R[ip(d,i)+1]..] \leq R[\textit{SA}_R[j]..] \leq R[\textit{SA}_R[ip(d',i')-1]..] < S_{d'}[i'..]$. Now let $ip(d',i') = ip(d,i)+1$. If $x_{d,i}=S$, then $S_d[i..]<R[\textit{SA}_R[ip(d,i)]..] = R[\textit{SA}_R[ip(d',i')-1]..] < S_{d'}[i'..]$, by Obs.~\ref{obs:ip}. Similarly, if $x_{d',i'}=L$, then $S_d[i..]<R[\textit{SA}_R[ip(d,i)+1]..] = R[\textit{SA}_R[ip(d',i')]..] < S_{d'}[i'..]$. Finally, let $x_{d,i}=L$ and $x_{d',i'}=S$. Then $R[\textit{SA}_R[ip(d,i)]..] < S_d[i..],S_{d'}[i'..] < R[\textit{SA}_R[ip(d,i)+1]..] = R[\textit{SA}_R[ip(d',i')]..]$. Let $U$ be the matching factor of $(d,i)$, $U'$ that of $(d',i')$, and $V = \textit{lcp}(U,U')$, the longest common prefix of the two. $V$ cannot be equal to $U'$ because then $U'$ would be a proper prefix of $U$, but $ip(d',i')$ is the smallest occurrence in $R$ of $U'$. If $V=U$, then $U$ is a proper prefix of $U'$, and by definition of $ip(d',i')$, the character following $U$ in $U'$ is strictly greater than the mismatch character $c_i$ of $(d,i)$. Finally, if $V$ is a proper prefix both of $U$ and of $U'$, then the character following $V$ in $U$ is smaller than the one following $V$ in $U'$, therefore $U<U'$. Since $U$ is a prefix of $S_d[i..]$ and $U'$ is a prefix of $S_{d'}[i'..]$, and neither is prefix of the other, this implies $S_d[i..] < S_{d'}[i'..]$. \end{proof} \begin{lemma}\label{lemma:sufcomp_2} Let $1\leq d,d' \leq m$ and $1\leq i\leq |S_d|$, $1\leq i'\leq |S_{d'}|$, and $ip(d,i) = ip(d',i')$. \begin{enumerate} \item If $\ell_{d,i} < \ell_{d',i'}$ and $x_{d,i}=S$, then $S_d[i..] < S_{d'}[i'..]$. \item If $\ell_{d,i} < \ell_{d',i'}$ and $x_{d,i}=L$, then $S_{d'}[i'..] < S_d[i..]$. \item If $\ell_{d,i} = \ell_{d',i'}$ and $x_{d,i}=S$ and $x_{d',i'}=L$, then $S_d[i..] < S_{d'}[i'..]$. \item If $\ell_{d,i} = \ell_{d',i'}$ and $x_{d,i}=x_{d',i'}$ and $c_{d,i}<c_{d',i'}$, then $S_d[i..] < S_{d'}[i'..]$. \end{enumerate} \end{lemma} \begin{proof} {\em 1.,2.:} Let $U$ be the matching factor of $i$, and $U'$ that of $i'$. Since $\ell_{d,i}<\ell_{d',i'}$, this implies that $U$ is a proper prefix of $U'$. If $x_{d,i}=S$, then the mismatch character $c_{d,i}$ is smaller than the character following $U$ in $U'$, therefore $S_d[i..] < S_{d'}[i'..]$. If $x_{d,i}=L$, then it is greater, and thus $S_{d'}[i'..] < S_d[i..]$. {\em 3.} follows directly from Observation~\ref{obs:ip}, since now $S[i..] < R[\textit{SA}_R[ip(i)]..] < S[i'..]$. {\em 4.: } Now both suffixes start with the same matching factor $U$, followed by different mismatch characters, which define their relative order. \end{proof} These two lemmas in fact imply the following: \begin{proposition}\label{prop:PO} The conditions of Lemmas~\ref{lemma:sufcomp_1} and~\ref{lemma:sufcomp_2} result in a partial order of the suffixes of strings in ${\cal C}$, of which the lexicographic order is a refinement. \end{proposition} What happens if two suffixes $S_d[i..]$ and $S_{d'}[i'..]$ have the same values of the enhanced matching statistics, i.e.\ $\textit{ems}(d,i) = \textit{ems}(d',i')$? The next lemma says that in this case, the relative order of the two suffixes is decided by the relative order of the heads preceding their respective mismatch characters. \begin{lemma}\label{lemma:sufcomp_3} Let $1\leq d,d' \leq m$ and $1\leq i\leq |S_d|$, $1\leq i'\leq |S_{d'}|$. If $ip(d,i) = ip(d',i')$, $\ell_{d,i} = \ell_{d',i'}$, $x_{d,i}=x_{d',i'}$, and $c_{d,i}=c_{d',i'}$, then $S_d[i..] < S_{d'}[i'..]$ if and only if $S_d[j..] < S_{d'}[j'..]$, where $(d,j) = \textit{i-head}(d,i+\ell_i)$ and $(d',j') = \textit{i-head}(d',i'+\ell_{i'})$. \end{lemma} \begin{proof} We will prove that the relative position of the insert-head of $i$'s and $i'$'s mismatch character is the same, i.e.\ that $j-i = j'-i'$. The claim then follows. First note that $j>i$. This is because the matching factor of position $i$ ends in position $i+\ell_{d,i}-1$, so there must be a new insert-head after $i$ and at most at $i+\ell_{d,i}$, the position of the mismatch character. Similarly, $j'>i'$. The fact that $j=\textit{i-head}(i+\ell_{d,i})$ implies that there is a matching factor starting in position $j$ which spans the mismatch character $c=c_{d,i}=c_{d',i'}$. Let's write $Vc$ for the prefix of length $i+\ell_{d,i}-j$ of this matching factor. $V$ is a suffix of the matching factor $U$ of position $i$, but $Vc$ is not. However, $Vc$ is also a prefix of $S_{d'}[i'..]$. Therefore, $j'=i'+(j-i)$ is also an insert-head in $S_{d'}$. An analogous argument shows that any insert-head between $i'$ and $i'+\ell_{d',i'}$ in $S_{d'}$ is also an insert-head in $S_d$, in the same relative position. \end{proof} \begin{proposition}\label{prop:suffix_comparison} Let $R,S_1,\ldots,S_m$ be strings over $\Sigma$. Using the enhanced \textit{CMS}\ of ${\cal C}=\{S_1,\ldots,S_m\}$ w.r.t.\ $R$, we can decide, for any $1\leq d,d' \leq m$ and $1\leq i\leq |S_d|$, $1\leq i'\leq |S_{d'}|$, the relative order of $S_d[i..]$ and $S_{d'}[i'..]$ in ${\cal O}(\log\log\chi'\cdot \max_d \{\text{no.\ of insert-heads of } S_d\})$ time. \end{proposition} \begin{proof} Let $(d,j) = \textit{i-head}(d,i+\ell_i)$ and $(d',j') = \textit{i-head}(d',i'+\ell_{i'})$. From Lemma~\ref{lemma:eCMS} we get the four \textit{eCMS}-entries of $(d,i)$ and $(d',i')$, namely the insert positions $q_i$ resp.\ $q_{i'}$, the length of the matching factor, whether the mismatch characters is smaller or larger, and the mismatch character itself. If any of these differ for the two suffixes, then Lemmas~\ref{lemma:sufcomp_1} and~\ref{lemma:sufcomp_2} tell us their relative order. This check takes ${\cal O}(1)$ time. Otherwise, Lemma~\ref{lemma:sufcomp_3} shows that the relative order is determined by the next relevant heads. Iteratively applying the three lemmas, in the worst case, takes us through all heads for the strings $S_d$ and $S_{d'}$. \end{proof} Instead of using Prop.~\ref{prop:suffix_comparison}, we will use these lemmas in the following way. We will first sort only the insert-heads. The following proposition states that this suffices to determine the order of any two suffixes in constant time. \begin{proposition}\label{prop:suffix_sort} Given the insert-heads in sorted order, the relative order of any two suffixes can be determined in ${\cal O}(\log\log\chi')$ time, where $\chi'$ is the number of insert-heads. \end{proposition} \begin{proof} Follows from Lemmas~\ref{lemma:sufcomp_1},~\ref{lemma:sufcomp_2}, and~\ref{lemma:sufcomp_3}, since all checks take constant time, and each of the two predecessor queries take ${\cal O}(\log\log\chi')$ time. \end{proof} \section{Putting it all together}\label{sec:putting} A high-level view of our algorithm is as follows. We first partially sort the insert-heads, then use this partial sort to generate a new string, whose suffixes we sort with an existing suffix sorting algorithm. This gives us a full sort of the insert heads. We then use this to sort the $S^*$-suffixes of the collection. Finally, we induce the remaining suffixes of the collection using the $S^*$-suffixes. We next give a schematic description of the algorithm. \medskip \begin{quote} \hrule {\tt Algorithm 1}\\ {\bf input:} string collection ${\cal C}$, reference string $R$\\ {\bf output:} the $\textit{GSA}$ of ${\cal C}$ \hrule \begin{itemize} \item {\bf Phase 1 - Augmenting and constructing data structures on $R$:} Preprocess $R$ (``augmenting'', see Sec.~\ref{sec:implementation}). Compute the data structures $\textit{SA}_R, \textit{ISA}_R, \textit{PLCP}_R, \textit{LCP}_R$ and the RMQ-data structure for \textit{PSV}- and \textit{NSV}-queries on $\textit{LCP}_R$. \item {\bf Phase 2 - Computing the \textit{eCMS}:} Compute the \textit{eCMS}\ of ${\cal C}$, as described in Sec.~\ref{sec:computing_cms}. \item {\bf Phase 3 - Bucketing:} Identify the $S^*$-suffixes in ${\cal C}$ via a backward linear scan of ${\cal C}$. Bucket $S^*$-suffixes $i$ according to $ip(i)$, computed using the \textit{eCMS}\ (Lemma~\ref{lemma:eCMS}). \item {\bf Phase 4 - Sorting the insert-heads:} \begin{itemize} \item bucket the insert-heads according to their insert point; \item for each bucket $B$, partially sort $B$, according to Lemmas~\ref{lemma:sufcomp_1} and~\ref{lemma:sufcomp_2}; \item rename insert-heads according to lexicographic rank of substring stretching up to the mismatch character (metacharacters are $S_d[j..j+\ell_{d,j}]$); \item generate new string $C$ as concatenation of these metacharacters; \item compute the suffix array of $C$, map back to corresponding suffixes of ${\cal C}$. \end{itemize} \item {\bf Phase 5 - Fully sorting the $S^*$-suffixes:} for each bucket $B$ from Phase 3, sort $B$, according to Lemmas~\ref{lemma:sufcomp_2} and~\ref{lemma:sufcomp_3} \item {\bf Phase 6 - Inducing the \textit{GSA}:} With two scans, induce $L$-suffixes, induce $S$-suffixes. \end{itemize} \hrule \end{quote} \medskip We next give a worst-case asymptotic analysis of the algorithm. \begin{proposition} Algorithm 1 computes the \textit{GSA}\ of a string collection ${\cal C}$ of total length $N$ in worst-case time ${\cal O}(N\log N)$. \end{proposition} \begin{proof} Let $|R|=n$. Phase 1 takes ${\cal O}(n+N)$ time, since constructing all data structures on $R$ can be done in linear time in $n$ and scanning the collection ${\cal C}$ takes time ${\cal O}(N)$. Phase 2 takes time ${\cal O}(N\log n)$ using the algorithm from Sec.~\ref{sec:computing_cms}. In Phase 3, identifying the $S^*$ suffixes, takes time ${\cal O}(N)$. Since at this point, the $\textit{eCMS}$ is in text-order, identifying $\textit{i-head}(i)$ takes constant time, also computing the insert-point takes constant time, so altogether ${\cal O}(N)$ time. In Phase 4, all steps are linear in $\chi'$, the number of insert-heads, including the partial sort of the buckets, since this can be done with radix-sort (three passes over each bucket), so this phase takes time ${\cal O}(\chi')$. Phase 5 takes time ${\cal O}(|B|\log|B|)$ for each bucket $B$, thus ${\cal O}(N\log |B_{\max}|)$ for the entire collection, where $B_{\max}$ is a largest bucket. Since all strings in the collection are assumed to be highly similar to the reference, the size of the buckets can be expected to vary around the number of strings in the collection $m$; however, in the worst case the largest bucket can be $\Theta(N)$. Finally, Phase 6 takes linear time ${\cal O}(N)$. Altogether, the running time is dominated by Phase 5, ${\cal O}(N\log N)$. \end{proof} \section{Implementation details}\label{sec:implementation} In Phase 1, the augmentation step involves, for every character $c$ not occurring in $R$ but occurring in ${\cal C}$, appending $c^{n_c}$ to $R$, where $n_c$ is the length of the longest run of $c$ in ${\cal C}$. This avoids having $0$-length entries in the matching statistics and is necessary in order to have a well defined $ip$. To compute $\textit{SA}_R$ in Phase 1, we use {\tt sais}~\cite{sais-lite} as implemented by Yuta Mori, a well engineered version of SAIS \cite{NZC11}, which was chosen due to its consistent speed on many different inputs. For the computation of $\textit{PLCP}_R$ and $\textit{LCP}_R$ we use the $\Phi$ method~\cite{KMP09}. This is the fastest method to compute the LCP array we know of. We constructed the data structure of C{\'a}novas and Navarro~\cite{CN10} for NSV/PSV queries on the LCP array, as it has low space overheads and was fast to query and initialize. For the predecessor data structure, we use the following two-layered approach in practice (rather than~\cite{W83}). We sample every $b$th head starting position and store these in an array. In a separate array we store a differential encoding of all head positions. The array of differentially encoded starting positions takes 32 bits per entry. Predecessor search for a position $x$ proceeds by first binary searching in the sampled array to find the predecessor sample at index $i$ of that array. We then access the differentially encoded array starting at index $ib$ and scan, summing values until the cumulative sum is greater than $x$, at which point we know the predecessor. This takes $O(\chi'/b + b)$ time, where $\chi'$ is the number of insert-heads. For Phase 4, when we have to sort $C$ (the concatenation of metacharacters representing partially sorted heads), we use a SACA-K implementation that handles integer alphabets~\cite{LouzaGT17}. This choice was made because of this algorithm's low space requirement, in particular, ${\cal O}(K)$, where $K$ is the number of distinct $\textit{ems}$-entries of insert-heads in $H'$ (note $K = O(\chi')$). \section{Experiments}\label{sec:experiments} We implemented our algorithm for computing the generalized suffix array in C++. Our prototype implementation, {\tt sacamats}, is available at \url{https://github.com/fmasillo/sacamats}. The experiments were conducted on a laptop equipped with 16GB of RAM DDR4-2400MHz and an Intel(R) Core(R) i5-8250U@3.4GHz with 6MB of cache. The operating system is Ubuntu 20.04 LTS, the compiler used is {\tt g++} version 9.4.0 with options {\tt -std=c++17 -O3 -funroll-loops} enabled. In the following experiments, we compare ${\tt sacamats}$ to two well known suffix array construction tools, both implementations by Yuta Mori~\cite{sais-lite, divsufsort}. The first, {\tt sais}, is an implemenation of the well-known SAIS algorithm by Nong et al.~\cite{NZC11}; the second, ${\tt divsufsort}$~\cite{0001K17}, is perhaps the most widely used tool for suffix array construction. We also compare against ${\tt gsufsort}$~\cite{LouzaTGPR20}, which is an extension of the SACA-K algorithm~\cite{Nong13} to a collection of strings, and to ${\tt bigBWT}$ \cite{BoucherGKLMM19}, a tool computing the BWT and the suffix array, designed specifically for highly repetitive data. \subsection{Datasets} For our tests, we used two publicly available datasets, one consisting of copies of human chromosome 19 from the 1000 Genomes Project~\cite{1000genomes}, and another of copies of SARS-CoV2 genomes taken from NCBI Datasets\footnote{\url{https://www.ncbi.nlm.nih.gov/datasets/coronavirus/genomes/}}. For both datasets we selected subsets of different sizes in order to study the scalability of our algorithm. The sizes are 250MB, 500MB, 800MB and 1GB. More information can be found in Table~\ref{tab:datasets}. We observe that on both datasets the number of i-heads is around 100x less than the input size, and on {\tt chr19} it is 8x less than the number of BWT runs. \begin{table}[th] \centering \begin{tabular}{|l|l|r|r|r|r|} \hline Name & Description & $\sigma$ & $r$ & no.\ of $S^*$-suffixes & no.\ of i-heads \\ \hline {\tt chr19} & Human Chromosome 19 & 5 & 32\,018\,267 & 129\,129\,636 & 4\,220\,033 \\ {\tt sars-cov2} & SARS-CoV2 genome & 14 & 351\,596 & 143\,588\,463 & 6\,537\,294 \\ \hline \end{tabular} \caption{Datasets used in experiments. In column 3, we specify the alphabet size $\sigma$, in column 4 the number $r$ of runs of the BWT, in column 5 the number of $S^*$-suffixes, and column 6 the number of insert-heads. In our experiments we use prefixes of each dataset up to 1GB. The last three columns refer to the 500MB prefix. } \label{tab:datasets} \end{table} \subsection{Results} In Figures \ref{fig:chr19} and \ref{fig:covid}, information about running time for both datasets is displayed. The line plot represents a direct comparison of different algorithms, whereas the stacked bar plot is to visualize how much each phase of ${\tt sacamats}$ takes w.r.t. the total running time (cp.~Sec.~\ref{sec:putting}). These tools all produce slightly different outputs: ${\tt sais}$ and ${\tt divsufsort}$ output the $\textit{SA}$, ${\tt gsufsort}$ and ${\tt sacamats}$ the $\textit{GSA}$, and ${\tt bigBWT}$ both the $\textit{BWT}$ and the $\textit{SA}$. Because of these differences, if one were to write to disk each result, the running time would be affected accordingly by the size of the output. Therefore, we only compare the building time, i.e.\ the time spent constructing the SA and storing it in a single array in memory, without the time spent writing it to disk. For this reason, we made slight changes to the ${\tt bigBWT}$ code to enable storing the $\textit{SA}$ in main memory. By looking at the line plots, one can see that ${\tt sacamats}$ is competitive in both scenarios, i.e., it is faster than all tools on {\tt sars-cov2}, except ${\tt bigBWT}$. The same is true for {\tt chr19}, where it is the fastest method, especially on larger inputs, but here the main competitor becomes ${\tt divsufsort}$. More precisely, for the first dataset ({\tt chr19}) and considering 1GB of data, ${\tt sacamats}$ takes less than a third of the time of ${\tt gsufsort}$, is 20\% faster than ${\tt sais}$, 12\% faster than ${\tt bigBWT}$, and 5\% faster than ${\tt divsufsort}$. For the second dataset ({\tt covid}), ${\tt sacamats}$ takes again less than a third of the time of ${\tt gsufsort}$, is 37\% faster than ${\tt divsufsort}$, 16\% faster than ${\tt sais}$, and 30\% slower than ${\tt bigBWT}$. Shifting our attention to the stacked bar plots, Figure~\ref{fig:chr19-phases} indicates that a lot of time is spent in the first phase, consisting in the augmentation of $R$ and the construction of various data structures for the augmented version of $R$. In the setting of DNA strings it is not too hard to think that the augmentation process will not elongate $R$, due to the very restricted alphabet. If the application lends itself to it, one could compute beforehand all the data structures listed in Phase 1, gaining roughly 20 seconds of run time. In our experiment on {\tt chr19} we would then be clearly the best algorithm, further distancing from the others. Alternatively, the common method of replacing {\tt N} symbols with random nucleotide symbols would be another way to speed up this phase. Finally, we comment on memory usage, which is highest for ${\tt sacamats}$ and ${\tt gsufsort}$ at 8 bytes per input symbol, and 4 bytes per input symbol for ${\tt divsufsort}$ and ${\tt sais}$, and ${\tt bigBWT}$ (including the 4 bytes per input symbol of the $\textit{SA}$ when it is saved in memory, see above). We have not yet optimized for memory usage and note that a semi-external implementation of our approach, in which buckets reside on disk, presents itself as an effective way to reduce main memory usage. In all phases, the actual working set---the amount of data active in main memory---is small (for the most part, proportional to the number of i-heads), and other authors have shown that the inducing phase is amenable to external memory, too~\cite{KKPZ17}. We leave these optimizations as future work. \begin{figure}[hb] \centering \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\textwidth]{imgs/chr19_comparison_no_output.png} \caption{Running time comparison.} \end{subfigure}% \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\textwidth]{imgs/chr19_phases_no_output.png} \caption{Phases breakdown, see Sec.~\ref{sec:putting} for details.} \label{fig:chr19-phases} \end{subfigure} \caption{Experiments on different subsets of copies of Human Chromosome 19.} \label{fig:chr19} \end{figure} \begin{figure}[ht] \centering \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\textwidth]{imgs/covid_comparison_no_output.png} \caption{Running time comparison.} \end{subfigure}% \begin{subfigure}{.5\textwidth} \centering \includegraphics[width=\textwidth]{imgs/covid_phases_no_output.png} \caption{Phases breakdown.} \end{subfigure} \caption{Experiments on different subsets of SARS-CoV2 genomes.} \label{fig:covid} \end{figure} \section{Conclusion}\label{sec:conclusion} We have presented a new algorithm for computing the generalized suffix array of a collection of highly similar strings. It is based on a compressed representation of the matching statistics, and on efficient handling of string comparisons. Our experiments show that a relatively straightforward implementation of the new algorithm is competitive with the fastest existing suffix array construction algorithms on datasets of highly similar strings, as are common in computational biology applications. A byproduct of our suffix sorting algorithm is a heuristic for fast computation of the matching statistics of a collection of highly similar genomes w.r.t. a reference sequence, which is of independent interest. We also envisage uses for our compressed matching statistics (CMS) data structure beyond the present paper, for example as a tool for sparse suffix sorting, or for distributed suffix sorting in which the CMS is distributed to all sorting nodes together with a lexicographic range of the suffixes that each particular node is responsible for sorting. From the CMS alone, each node can extract the positions of its suffixes and then sort them with the aid of the CMS. We believe there to be a great deal of room for further practical improvements, both through algorithm engineering and parallelism. Interestingly, in an initial attempt along the second line, simply assigning $S^*$-suffix buckets to one of four different sorting threads reduces runtime significantly, for example, from 122 to 89 seconds on the 1GB {\tt sars-cov2} dataset. Further studies will be conducted on how the size of \textit{eCMS}\ impacts on the competitiveness of our tool. \newpage
\section{Introduction}\label{sec:Intro} In the course of an undergraduate to first year graduate student's physcis education, the student will learn quantum mechanics (QM) encompassing the solution of bound states of the hydrogen atom, and the application and manipulation of spin. In specialized physics courses the student may have the opportunity to take an introduction to general relativity (GR), typically encompassing particle and photon orbits in the Schwarzschild spacetime (SST). Finally, in an advanced undergraduate (or first year graduate) QM course, the student will be exposed to the special relativistic (SR) Dirac equation (DE), and its solutions in special cases. Given the above archetypal physics education, a curious student may make the following observations, leading to the the subsequently posed question to a professor, or better yet, to themselves. ``In my introductory GR class, we saw that in the Schwarzschild metric, the radial equation for particles and photons orbits has a form reminiscent of a classical energy equation (see Hartle\cite{Hartle:2009}, p195), which for particles ($m>0$) takes the form \bea{Hartle:9.32} E_{Newt} &=& \frac{m}{2}\,\left( \frac{dr}{d\tau}\right)^2 +V_{eff}(r), \\ \textrm{where}\quad V_{eff}(r) &=& -\frac{G M m}{r}+\frac{L^2}{2 m r^2} - \frac{G M L^2}{c^2 m r^3}, \end{eqnarray} where $M$ is the mass of the gravitating body, $m$ is the test mass, $e = (E_{Newt} + m c^2)/ (m c^2)$ and $\ell = L/m$ are the conserved energy and angular momentum per unit rest mass, respectively (constants of the motion along the trajectory). Now if we ignore the $1/r^3$ term, this looks just like the central potential with the angular momentum barrier term included for an attractive $1/r$ potential, like the Coulomb problem. In my (non-relativistic) quantum mechanics (NRQM) class, we solved for the wave functions and eigenvalues of the bound quantum states of a hydrogen-like atom in a central potential $V_{C} = -(Ze)e/r$. Thus, can one expect to find bound states of $V_{Schw} \equiv V_{eff}$ if one were to simply replace $M\to Z e$ and $m\to e$? Would this make any sense? The scale of $M$, say for a solar mass black hole (BH), and $Ze$ are so vastly different? How would one go about answering this (intuitive) question?" The purpose of this article is to assemble the disparate, requisite components for the students in order to answer this question in the affirmative, at a level within the students undergraduate, to first year graduate physics background, aided by detailed, readable textbook material that would enable the student to answer this question for themselves. Let us first ask, what component physics, in order of difficulty, is required by the student to answer this question, and point out (from the author's personal perspective) those reference books containing enough explicit worked out material/derivations that would enable the student solve their question. First, once has to be able to solve for the bound states of hydrogen atom. This is standard fare in almost all introductory QM textbooks, but is lucidly well described and worked out in detail in one of the classic undergraduate QM texts, ``Quantum Mechanics: Volume 1," by Cohen-Tannoudgi, Diu and Laloe (CTQMv1) \cite{Cohen-Tannoudji:1977}. The eigenenergies of the hydrogen atom ($Z=1$) are known to be given by $E_n = -E_I/n^2$ where the ionizations energy is given by (CTQMv1, Chapter VII) $E_I = \frac{\alpha^2}{2} \mu c^2 = 13.6$ eV, where $\mu = (m_e\,m_p)/(m_e+m_p)\sim m_e$ is the reduced mass of the electron-proton hydrogen atom system, $\mu c^2\sim m_e c^2$ is the rest mass of the electron, $e>0$ is the charge of the proton, and $\alpha = e^2/(\hbar c) \sim 1/137$ is the fine structure constant. However, since we are inquiring about possible quantum bound states in a classical background GR spacetime, it is not enough to have a NRQM solution. One could attempt to invoke the special relativity (SR) energy equation and posit that by including the rest mass, the SRQM energy $\mathcal{E}$ should be given by $\mathcal{E} \sim m c^2 - E_I$ (where from now on we just denote $m_e\to m$). Working backwards, one might surmise that this might the small energy approximation to either \bea{SRQM:eval:guess} \mathcal{E}_s \overset{?}{=} \sqrt{1 - \frac{E_I}{n^2}}, \label{E_s}\\ \mathcal{E}_v \overset{?}{=} \frac{1}{\sqrt{1 + \frac{E_I}{n^2}}}. \label{E_v} \end{eqnarray} (The meaning of the $s$ and $v$ subscripts on $\mathcal{E}$ above, denoting scalar and vector coupling, respectively, will be explain in due time). But, which formula is more appropriate, and how would one physically decide between the two? It appears that a much more complete description of the Coulomb problem is required - i.e. the solution of the Coulomb problem in a SR setting using the Dirac equation (DE). The DE and its solution in a central potential are treated in the extremely informative worked out example problems in one of the extraordinary set of textbooks by Walther Greiner, namely ``Relativistic Quantum Mechanics: Wave Equations (RQMWE)."\cite{Greiner:1990} This textbook series, including volumes on QM, SRQM, Electrodynamics, Nuclear Physics, Thermodynamics, Quantum Field Theory and Statistical and Thermal Physics, is essentially (in the author's opinion) a ``reworking" of the classic Landau and Lipfshitz (informative, but terse) set of graduate level physics, now geared for advanced undergraduates and/or first year graduate students (the prefaces just state that it is for ``students"). In addition to its exceptional lucidity, each book involves typically over 100 solved problems, worked out in full detail, and showing all intermediate steps, which greatly aids the students proficiency in solving problems. Chapter 2 introduces the DE and Chapter 9 solves the Coulomb problem in a central potential for both (electromagnetic) vector and scalar (mass) coupling. Here, vector coupling means that the potential $V_v(r)=e\phi(r) /c$ arises from the timelike component of a SR 4-vector potential $A^\mu=(\phi, \bs{A})$ such that the relativistic energy $E$ is modified to $E\to E-V_v$ in the DE. Scalar coupling means that the potential $V_s(r)$ couples to the mass, so that $m c^2 \to mc^2 + V_s(r)$ in the DE. So intuition would lead one to suspect favoring the scalar coupling formula \Eq{E_s}. We shall explore if this intuition has any validity. If one is able to master the solution of the DE wave equation in SR in a central potential (aided by RQMWE), the student still has to tackle the thorny question of how to incorporate QM into GR. Of course, this is an outstanding unsolved problem in physics, but such a full theory is not required here. What one requires is the solution of a quantum mechanical wave function in a classical background GR spacetime. In effect, one needs to solve a wave equation in a medium with a spatially (here radially) dependent index of refraction. Thus, what is required is the DE in classical GR (curved spacetime) background. But how does one obtain this equation, since in RQM we solve the DE in a flat (Minkowski) background spacetime, incorporating only SR? The solution is well-known (to the QM/GR Illuminati), but the key concept is easy to grasp - Einstein's Equivalence Principle. Namely, the global inertial frames of SR no longer hold in GR, and are only locally valid at each point (actually within the observer's local laboratory\cite{Hartle:2009}, if small on the length scale of changes of the spacetime curvature). Thus, one needs to ``project" the SR DE into the observer's instantaneous local laboratory, and solve the resulting GR DE in this local (moving) frame. This involves the use of \tit{tetrads}, describing the four axes of the observer's frame $\mathbf{e}_a$, with $a\in\{0,1,2,3\}$, moving in the surrounding spacetime, with world vector components $\{e_a^{\;\;\mu}(x)\}=(\mathbf{e}_a)^\mu(x)$, for $\mu=\{0,1,2,3\}$. In the local observer's frame the GR DE takes on a first order form analogous to the SR DE, with all the complication going into the \tit{connection} describing how the spinor (2-vector with complex entries) components change (parallel transport and all that) as one moves from one point to another in a curved spacetime. There are now numerous introductory GR textbooks aimed at the upper level undergraduate student. The only one that tackles the GR DE, and in lucid, informative detail (but just setting up the equation, not solving it) is Lewis Ryder's (a particle physicist) wonderful ``Introduction to General Relativity, Chapter 11."\cite{Ryder:2009}. This is the formulation of the GR DE that we will use in this work. As one can see, there is much the student must master in order to answer their own intuitive question. And it may seem too daunting a task to tackle; a mountain too high to climb even for the motivated student. In NRQM, the solution of the Schr\"{o}dinger equation (SE) in a central potential brings in the additional complication of orbitial angular momentum operators, spherical harmonics and power series solutions of the SE, which are all standard fare in an undergraduate QM class (and clearly elucidated in many textbook, including CTQMv1). Mastering just the SR DE involves understanding an appreciation for covariant formalism, spinor transformations under Lorentz transformations (LT), and clever manipulation of the Dirac matrices to bring out the orbital and spin angular momentum terms (as explained in detail in RQMWE). Finally, GR itself presents conceptual and mathematical manipulation hurdles itself to the average student, even before one attempts tackling the GR DE. The latter, then involves additional concepts and mathematical manipulations of ``gauging" Lorentz symmetry, and constructing the appropriate tetrads for the local observer. In spite of this seeming mountain of required background, the claim here in this work is that the above mentioned textbooks provide more than sufficient background to step by step educate the motivated student, and allow them to learn the necessary prerequisites required to answer their self-posed question. The purpose of this article is to summarize the highlights of the relevant, requisite material (rather than reproduce it in full) and guide the student to the those portions of the above mentioned textbooks that give vastly more comprehensive and worked out details. The end result is a solution to the bound states in the background Schwarzschild spacetime (SST) and a comparison to the SR DE formulas and their simplified Schrodinger plus special relativity (SE+SR) forms \Eq{E_s} and \Eq{E_v}, answering the student's intuitively posed question. To keep the problem as simple as possible, we focus on spherically symmetric $s$-orbital states with $L=0$, so that $V_{eff}$ is directly analogous to $V_C$ (i.e. we drop the angular momentum barrier terms in the former). We include Appendices with enough background material to make this as work self-contained as possible when supplemented by the aforementioned textbooks. We will also write down equations in dimensionless form using the length, energy, etc. scales appropriate to the Coulomb and gravitational problems (see \App{app:Units}), so that both equations have the same form for purposes of comparison. So let us now embark on our investigative journey of the student's intuitive question. \section{Wave functions and eigenenergies of the NRQM Coulomb problem: a refresher}\label{NRQM:Coulomb:solns} As any student exposed to a NRQM course is well award, the Schrodinger equation (SE) $i \hbar\,\partial \psi(\bs{r},t)/\partial t = \left( \bs{p}^{\,2}/2m + V_C(r) \right)\, \psi(\bs{r},t)$ for stationary states $\psi(\bs{r},t) = e^{-i E t/\hbar}\,\phi(\mathbf{r})$ for $E<0$ constant, with $E\to i \hbar \frac{\partial}{\partial t}$ and $\mathbf{p}\to -i \hbar\,\mathbf{\nabla}$ yields \be{SchrEqn} \mathcal{H}\, \phi(\mathbf{r})\equiv \left[-\frac{\hbar^2}{2 m} \nabla^2 - \frac{e^2}{r} \right]\,\phi(\mathbf{r}) = E\,\phi(\bs{r}). \end{equation} Using the differential operator identities (see the inside cover of CTQMv1 and Arfken\cite{Arfken:2012}, Chapter 3) \bea{Del:Opr} \boldsymbol{\nabla} &=& \mathbf{e}_r \frac{\partial}{\partial r} - \frac{i}{\hbar r^2}\,\mathbf{r}\times\mathbf{L}, \\ \mathbf{L} &=& \frac{\hbar}{i}\,\mathbf{r}\times\boldsymbol{\nabla} \end{eqnarray} where $\mathbf{e}_r = \bs{r}/r$ is the orthonormal unit vector in the radial direction. Following CTQMv1 p662 and p778, we can write the well known formula for the Laplacian $\boldsymbol{\nabla}^2$ in spherical coordinates as \bea{Laplacian} \boldsymbol{\nabla}^2 &=& -\frac{\hbar^2}{2 m}\,\frac{1}{r}\, \frac{\partial^2}{\partial r^2}\, r + \frac{\mathbf{L}^2}{2 m r^2}, \\ \ell^2 = \frac{\mathbf{L}^2}{\hbar^2} &=& -\left( \frac{\partial^2}{\partial\theta^2} + \frac{1}{\tan\theta}\,\frac{\partial}{\partial \theta} + \frac{1}{\sin^2\theta}\,\frac{\partial^2}{\partial\varphi^2 } \right) \end{eqnarray} Since the Coulomb force is radially directed, there is no torque, so that the angular momentum $\mathbf{L}$ is constant of the motion. As such, we can consider the classical motion of particles as taking place in a plane, typically taken to be the equator $\theta=\pi/2$. Thus, we seek solutions $\phi(\mathbf{r})$ involving a complete set of commuting observables $\{ \mathcal{H}, \mathbf{L}^2, L_z\}$ such that \be{NRQM:sep:vars} \mathcal{H}\,\phi(\mathbf{r}) = E\,\phi(\mathbf{r}),\quad \mathbf{L}^2\,\phi(\mathbf{r}) = \hbar^2\,\ell\,(\ell+1)\,\phi(\mathbf{r}),\quad L_z\,\phi(\mathbf{r}) = m\hbar\,\phi(\mathbf{r}). \end{equation} Using the well known properties of the spherical harmonics\cite{Cohen-Tannoudji:1977,Arfken:2012} $Y_{\ell, m}(\theta,\varphi)$ as the eigenfunctions of the angular momentum operator, \be{Ylm} \ell^2\, Y_{\ell, m}(\theta,\varphi) = \ell\,(\ell + 1)Y_{\ell, m}(\theta,\varphi),\quad -i \frac{\partial}{\partial \varphi}Y_{\ell, m}(\theta,\varphi) = m\,Y_{\ell, m}(\theta,\varphi), \end{equation} for $m\in\{-\ell, -\ell+1,\ldots, \ell-1, \ell\}$, we choose the separation of variables as $\phi(\mathbf{r}) = R(r)\,Y_{\ell, m}(\theta,\varphi)$ to obtain \be{radial:SE:R} \left[ -\frac{\hbar^2}{2 m}\,\frac{1}{r}\, \frac{\partial^2}{\partial r^2}\, r + \frac{\ell\,(\ell + 1)\,\hbar^2}{2 m r^2} + V_C(r) \right]\,R(r) = E\,R(r), \end{equation} where the spherical harmonics have been canceled from both sides of the equation, leading to a purely radial Schrodinger equation. In addition to \Eq{radial:SE:R} we impose, on physical grounds, the boundary condition that the radial wave function $R(r)$ decay to zero as $r\to\infty$, It is convenient to make one last substitution $R(r) = u_{k,\ell}(r)/r$, and based on physical ground argue that one needs to impose the boundary condition $u_{k,\ell}(0)=0$, at the singularity that exists at the origin for the point particle electron. This leads to the final, simplified final form of the radial equation \be{radial:SE:u} \left[ -\frac{\hbar^2}{2 m}\, \frac{\partial^2}{\partial r^2}\, + \frac{\ell\,(\ell + 1)\,\hbar^2}{2 m r^2} + V_C(r) \right]\,u_{k,\ell}(r) = E\,u_{k,\ell}(r), \quad V_C(r) = -\frac{e^2}{r}. \end{equation} Here, the $\ell$ subscripts on $u_{k,\ell}(r)$ indicates the dependence on the angular momentum, and the index $k$ represents all other non-angular momentum dependencies of the eigensolutions. We can now write \Eq{radial:SE:u} in two dimensionless forms: \bea{radial:SE:u:dimless} \left[ \frac{\partial^2}{\partial \rho^2}\, - \frac{\ell\,(\ell + 1)}{\rho^2} + \frac{2}{\rho} - \lambda^2_{k,\ell} \right]\,u_{k,\ell}(\rho) &=& 0, \quad \rho = \frac{r}{a_0}, \quad \lambda_{k,\ell} = \sqrt{\frac{-E_{k,\ell}}{E_I}}\ge 0, \\ \label{radial:SE:u:dimless:atomic:units} &{}& \nonumber \\ \left[ \frac{\partial^2}{\partial \rho^2}\, - \frac{\ell\,(\ell + 1)}{\rho^2} + \frac{2 \alpha_C}{\rho} - \epsilon_{k,\ell} \right]\,u_{k,\ell}(\rho) &=& 0, \quad \rho = \frac{r}{\lambda_c}, \quad \epsilon_{k,\ell} = \frac{E_{k,\ell}}{m c^2}. \label{radial:SE:u:dimless:fundamental:units} \end{eqnarray} \Eq{radial:SE:u:dimless:atomic:units} is expressed in atomic units: lengths and energy in terms of the Bohr radius $a_0$ and ionization energy $E_I$ of the hydrogen atom (see \App{app:Units}), and is the form that is typically employed (see CTQMv1, p794). The second form \label{radial:SE:u:dimless:fundamental:units} is expressed in terms of fundamental units of the electron: lengths and energy in terms of the Compton wavelength $\lambda_c$ and rest mass $m c^2$ of the electron (see \App{app:Units}). While this latter form is not typically employed, we list it here for later reference when we want to consistently compare to the DE, especially in the case of SST. This latter form explicitly brings out the role of fine structure constant $\alpha_C$. Forgoing the use of (non-intuitive) special (confluent hypergeometric) functions, we will follow CTQMv1 and outline the solution to \Eq{radial:SE:u:dimless:atomic:units} using a power series expansion, the typical approach taught in an undergraduate QM course. Following CTQMv1, pp794-797, we outline the essentials of the solution method. First one looks at the asymptotic form of the equation for $\rho\to\infty$, keeping only the dominant terms (i.e. dropping terms involving $\rho^{-1}$ and $\rho^{-2}$), which yields $[\partial^2_\rho - \lambda^2_{k,\ell}]\, u_{k,\ell}(\rho)=0$ with solutions $e^{\pm\lambda_{k,\ell}\,\rho}$. Since we want the wavefunction to decay to zero at spatial infinity (so that they are square integrable for finite total probability) we reject the positive exponent, and look for solutions of the form $u_{k,\ell}(\rho) =e^{-\lambda_{k,\ell}\,\rho} \,y_{k,\ell}(\rho)$, where we take $y_{k,\ell}(\rho)$ of the power series form $y_{k,\ell}(\rho) = \rho^s\,\sum_{q=0}^\infty c_q\,\rho^q$ with $c_0\ne 0$. The equation satisfied by $y_{k,\ell}(\rho)$ is \be{y:eqn} \left[ \frac{d^2}{d\rho^2} - 2\,\lambda_{k,\ell}\,\frac{d}{d\rho} + \left( \frac{2}{\rho}-\frac{\ell(\ell+1)}{\rho^2} \right)\, \right]\,y_{k,\ell}(\rho)=0, \end{equation} which must be satisfied for each coefficient $c_q$ term by term. In particular, the $c_0$ term yields the simple equation $[ s(s+1) - \ell(\ell+1) ]\,c_0=0$ yielding the solutions $s=\{\ell+1, -\ell\}$. Since we have the boundary condition $y_{k,\ell}(0)=0$ we must reject $s=-\ell$ (since $\ell\ge0$) and retain $s=\ell+1$. It is then straight forward to develop the following one term recursion $q (q+ 2\,\ell +1)\,c_q = 2\,[(q+\ell)\,\lambda_{k,\ell} -1]\,c_{q-1}$, for $q\ge 1$. The key quantization concept is that if series is infinite then $\frac{c_q}{c_{q-1}}\underset{q\to\infty}{\sim}\frac{2\,\lambda_{k,\ell}}{q}$ which is the same ratio of coefficients in the power series expansion of the function $e^{2\,\lambda_{k,\ell}\,\rho}$, which then violates our requirement for solutions decaying as we approach infinity. The only way to avoid this, is for the above one term recursion relation to terminate. By examining the coefficient of $c_{q-1}$ above, this is easily satisfied if $\lambda_{k,\ell} = \frac{1}{k+\ell}$. Thus, one obtains that the bound energies are given by $E_{k,\ell}= -E_I/(k+\ell)^2$ for a given $\ell$ for $k\in\{1, 2, 3, \ldots\}$ and $\ell=\{0, 1, 2, \ldots\}$. We therefore rename $k+\ell\to n$ and recover the Bohr energies in terms of the principle quantum number $n$. The radial wavefunctions are then given by (CTQMv1, p797) \bea{R:wavefunctions} n=1, \ell=0:\quad R_{k=1,\ell=0}(r) &=& 2\, (a_0)^{-3/2}\, e^{-r/a_0}, \\ n=2, \ell=0:\quad R_{k=2,\ell=0}(r) &=& 2\, (2\,a_0)^{-3/2}\,\left(1- \frac{r}{2\,a_0} \right) \,e^{-r/2\,a_0}, \\ n=2, \ell=1:\quad R_{k=2,\ell=0}(r) &=& (2\,a_0)^{-3/2}\,\frac{1}{\sqrt{3}}\,\frac{r}{a_0} \,e^{-r/2\,a_0} \end{eqnarray} Finally, since $E_I\ll m\,c^2$, one is justified in using the NRQM SE. \section{The SR Dirac Equation for spin $\mathbf{1/2}$ particles and its solutions for vector (Coulomb) and scalar (mass) coupling}\label{sec:SRDE:vector:scalar} Having traversed familiar ground, we now steadfastly venture in territory that may be less familiar to some students. \subsection{Free DE: no potentials}\label{sec:DE:no:V} As is historically well-known, Dirac was bothered by the inconsistency of the SE being first order in time, yet second order in space. Since SR treats time and space equivalently he was led to seek an equation that was first order in both time and space. For the following we will closely follow the exposition of Chapter 2 of Greiner\cite{Greiner:1990} which we denote as RQMWE. Dirac postulated a wave equation given by $i\,\hbar\,\frac{\partial \psi}{\partial t} = \hat{H}\,\psi$ of the form (RQMWE, p75, p82) \be{DE:form} i\,\hbar\,\frac{\partial \psi}{\partial t} =\left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi \equiv \hat{H}_f\,\psi. \end{equation} Here $\hat{\mathbf{p}} = -i\,\hbar\boldsymbol{\nabla}$ is the usual QM momentum operator and again we have taken $\hat{E}\to i\,\hbar\,\partial/\partial t$ as the energy operator. The quantities $\hat{\boldsymbol{\alpha}}= \{\hat{\alpha}_1, \hat{\alpha}_2,\hat{\alpha}_3\}$ and $\hat{\beta}$ cannot be numbers if one wishes to preserve the form invariance of the equation under simple spatial rotations, and these are considered as $N\times N$ matrices. Therefore, $\psi$ itself must be a column vector of $N$ complex functions of space and time \be{psi:DE} \psi = \left(\begin{array}{c} \psi_1(\mathbf{x},t) \\ \psi_2(\mathbf{x},t) \\ \vdots \\ \psi_N(\mathbf{x},t) \end{array}\right) \quad\Rightarrow\quad \rho(x) = \psi^\dagger\,\psi(x) = \sum_{i=1}^N\, \psi^*(x)\,\psi(x)\ge 0, \end{equation} where $\rho(x)$ is the positive probability density of the particle. One also wishes to preserve the SR energy equation $E^2 = \mathbf{p}^2\,c^2 + (m\,c^2)^2$ where $\mathbf{p}$ without the operator circumflex $\verb+^+$ denotes the ordinary classical momentum vector. Under the usual NR substitution (quantization) $E\to i\,\hbar\,\partial/\partial t$ and $\mathbf{p}\to -i\,\boldsymbol{\nabla}$ the SR energy equations becomes the Klein-Gordon wave equation $-\hbar^2\,\frac{\partial^2}{\partial t^2}\psi_\sigma = \left(-\hbar^2\,c^2\nabla^2 + m^2\,c^4 \right)\psi_\sigma$ for each component of the wave function $\psi_\sigma$ (where one typically denotes the components of the wave function with Greek indices). Carrying out this straightforward algebra (RQMWE, p76-77) the net result is that the matrices $\hat{\alpha}_i$ and $\hat{\beta}$ must satisfy \bea{alpha:beta:DE} \{ \hat{\alpha}_i, \hat{\alpha}_j\} &\equiv& \hat{\alpha}_i\,\hat{\alpha}_j + \hat{\alpha}_j\, \hat{\alpha}_i = 2\,\delta_{ij}\,\mathbb{I} , \label{alpha:beta:DE:line1} \\ \{ \hat{\alpha}_i, \hat{\beta}\} &=& 0, \label{alpha:beta:DE:line2} \\ \hat{\alpha}^2_i &=& \hat{\beta}^2= \mathbb{I}, \label{alpha:beta:DE:line3} \end{eqnarray} where $\mathbb{I}$ is the $N\times N$ identity matrix. It is not hard to show that no $N=2$ solution exists, and the first non-trivial solution that exists is for $N=4$. This is called the Dirac 4-spinor, denoted by \be{Dirac:4spinor} \psi(x) = \left(\begin{array}{c} \varphi(\mathbf{x},t) \\ \chi(\mathbf{x},t) \end{array}\right) \end{equation} where $\varphi$ and $\chi$ are each themselves 2-spinors (a 2-vector with complex entries), familiar from NRQM. The physical interpretation is that upper two components $\varphi$ of the 4-spinor $\psi$ represents particles of spin-$1/2$, while the lower two components $\chi$ represents anti-particles of spin-$1/2$ (same mass, but opposite charge). A conventional set of $4\times 4$ matrices satisfying \Eq{alpha:beta:DE:line1} -\Eq{alpha:beta:DE:line3} are given by \be{DE:alpha:beta} \hat{\alpha}_i = \left(\begin{array}{cc} 0 & \hat{\sigma}_i \\ \hat{\sigma}_i & 0 \end{array}\right), \qquad % \hat{\beta} = \left(\begin{array}{cc} \mathbb{I} & 0 \\ 0& -\mathbb{I} \end{array}\right), \end{equation} where $\sigma_i$ are the standard $2\times 2$ Pauli matrices, and $\mathbb{I}$ is the $2\times 2$ identity matrix, \be{Pauli:matrices} \sigma_1 = \left(\begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array}\right),\quad % \sigma_2 = \left(\begin{array}{cc} 0 & -i \\ i & 0 \end{array}\right),\quad % \sigma_3 = \left(\begin{array}{cc} 1 & 0 \\ 0& -1 \end{array}\right), \quad % \mathbb{I} = \left(\begin{array}{cc} 1 & 0 \\ 0& 1 \end{array}\right). \end{equation} Another commonly employed set of matrices $\gamma^\mu$ satisfying $\{\gamma^\mu, \gamma^\nu\} = 2\,\eta^{\mu,\nu}\mathbb{I}$ when using the particle physicist convention for the SR metric (RQMWE\cite{Greiner:1990}) $\eta_{\mu\nu}=\eta^{\mu\nu} = \trm{diagonal}\{1,-1,-1,-1\}$, (note: SR and GR books, e.g. see Chapter 11.3, Ryder\cite{Ryder:2009}, most commonly use the alternative convention, $\eta_{\mu\nu}=\eta^{\mu\nu} = \trm{diagonal}\{-1,1,1,1\}$) called the Dirac gamma matrices, are given by \be{DE:gamma:matrices} \gamma^0= \left(\begin{array}{cc} \mathbb{I} & 0 \\ 0& -\mathbb{I} \end{array}\right), \quad \gamma^i = \left(\begin{array}{cc} 0 & \hat{\sigma}_i \\ -\hat{\sigma}_i & 0 \end{array}\right). \end{equation} These are employed so that the DE can be written in covariant form \be{DE:covariant:form} i\,\hbar\,\gamma^\mu\,\partial_\mu\,\psi = m\,c\,\psi, \qquad \eta_{\mu\nu} =\eta^{\mu\nu} = \trm{diagonal}\{1,-1,-1,-1\}, \end{equation} where the $4\times 4$ identity matrix $\mathbb{I}_{4\times 4}$ is implied (and often omitted) on the righthand side. In the particle physicist convention $\hat{p}^\mu = i\,\hbar\, \{\frac{1}{c}\frac{\partial}{\partial t}, -\boldsymbol{\nabla}\}$ so the covariant DE is given by $\gamma^\mu\,\hat{p}_\mu\,\psi~=~m\,c\,\psi$, where $\hat{p}_\mu = \eta_{\mu\nu}\,\hat{p}^\nu = i\,\hbar\, \{\frac{1}{c}\frac{\partial}{\partial t}, \boldsymbol{\nabla}\}$. Lastly, the operator $\hat{\mathbf{S}}$ \be{DE:spin:opr} \hat{\mathbf{S}} = \frac{\hbar}{2}\,\hat{\mathbf{\Sigma}}\equiv \frac{\hbar}{2}\, \left(\begin{array}{cc} \hat{\boldsymbol{\sigma}} & 0\\ 0 & \hat{\boldsymbol{\sigma}} \end{array}\right), \qquad \hat{\mathbf{J}} = \hat{\mathbf{L}}+\hat{\mathbf{S}}. \end{equation} which commutes with the Dirac Hamiltonian $\hat{H}_f$ is called the spin vector operator, and represents the intrinsic spin of the electron, in addition to its orbital angular momentum. The total angular momentum $\hat{\mathbf{J}}$ is the sum of the orbital and intrinsic angular momentum, and is a constant (of the motion). For stationary states (which is our primary interest) we let $\psi(\mathbf{x},t) = e^{-i\,E/\hbar}\psi(\mathbf{x})$ for $E$ constant, such that $ i\,\hbar\,\partial/\partial t \,\psi(\mathbf{x},t) = E\,\psi(\mathbf{x},t)$, and the DE simply becomes of the form (RQMWE, p82) \be{DE:form:E} E\,\psi =\left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi \equiv \hat{H}_f\,\psi. \end{equation} \subsection{DE with potentials}\label{sec:DE:with:V} If the electron interacts with an electromagnetic potential $A^\mu = \{A^0, \mathbf{A}\}$ one obtains the usual NRQM substitutions (RQMWE, p94) $E\to i\,\hbar\,\partial/\partial t - e\,A_0$ and $\hat{\mathbf{p}} \to \hat{\mathbf{p}} - \frac{e}{c}\, \mathbf{A}$ so that the DE for stationary states becomes \be{DE:form:EM:general} i\,\hbar\,\frac{\partial \psi}{\partial t} = \left[c\,\hat{\boldsymbol{\alpha}}\cdot\left(\hat{\mathbf{p}} - \frac{e}{c}\, \mathbf{A}\right) + e\,A_0 + \hat{\beta}\,m\,c^2\right]\,\psi, \\ \end{equation} If we define say the Coulomb potential in the Coulomb gauge via $V_v \overset{\trm{def}}{=} e\,A_0$ and $\mathbf{A}=0$. the DE for stationary states becomes \be{DE:form:EM:vector:coupling} \left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi =(E-V_v)\,\psi. \end{equation} One can also introduce a scalar potential $V_s$ which couples to the mass so that $m\,c^2~\to~m\,c^2+ V_s$. Thus, the DE for stationary states, with both vector and scalar coupling, is given by (RQMWE, section 9.8, p184-187) \be{DE:form:EM:scalar:coupling} \left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,(m\,c^2 + V_s)\right]\,\psi =(E-V_v)\,\psi. \end{equation} \subsection{DE in a central potential}\label{sec:DE:central:potential} The DE in a central potential (vector coupling), in this section denoted by $V_v \to V(r)$, presents some involved matrix and vector identity manipulations which are explicitly worked out in RQMWE (see section 9.3, p169-172). The goal of this exercise is to (i) explicitly reveal the terms involving the angular momentum $\hat{\mathbf{L}}$ (as in the NRQM SE), as well as the intrinsic spin $\hat{\mathbf{S}}$, and (ii) ultimately remove the spinor dependency of the equations to produce a purely radial equation (actually, a pair of first order radial equations, vs a single second order radial equation as in the NRQM SE). Rather than reproducing all these algebraic machinations in detail here, we simply highlight the essential points which lead to our objectives (i) and (ii). Using the representation of $\hat{\boldsymbol{\alpha}}= \tiny{\left(\begin{array}{cc} 0 & \hat{\boldsymbol{\sigma}} \\ \hat{\boldsymbol{\sigma}} & 0 \end{array}\right)}$ and $\hat{\beta}= \tiny{\left(\begin{array}{cc} \mathbb{I} &0 \\0& -\mathbb{I} \end{array}\right)}$ given previously, and the representation of the Dirac 4-spinor in terms of a pair of 2-spinors $\psi = \tiny{\left(\begin{array}{c}\varphi \\ \chi \end{array}\right)}$ the DE becomes the pair of coupled ordinary differential equations \bea{DE:RQMWE:p170:5} c(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}})\chi &=& [(E-V) - m\,c^2 ]\,\varphi, \label{DE:RQMWE:p170:5:line:1} \\ c(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}})\varphi &=& [(E-V) + m\,c^2 ]\,\chi. \label{DE:RQMWE:p170:5:line:2} \end{eqnarray} One must now make an ansatz about the form of the 2-spinors $\varphi$ and $\chi$. From the NRQM SE, we suspect that they must contain the spherical harmonics $Y_{l,m}(\theta,\phi)$ in some form. This is the trickiest part of the separation of variables problem of the SR DE in a central potential, and involves much detailed algebra (explicitly worked out in RQMWE, section 9.3 pp169-172. For an intricate symmetry proof also see (10.54) pp212-214). These results are required in order that the angular portion of the 2-spinors drop out the of DE (analogous to the NRQM SE), leaving only two coupled, scalar radial equations - which is the end goal of all these algebraic manipulations. As mentioned above, we will state some of these important results, without proof, but with explanation, primarily directing the reader to the detailed derivations in RQMWE.\cite{Greiner:1990} The ansatz for the pair of 2-spinors is \be{2:spinor:ansatx} \varphi\to \varphi_{j\ell m} = i\,g(r)\,\Omega_{j\ell m}(\mathbf{r}/r), \qquad \chi\to \chi_{j\ell' m} = -f(r)\,\Omega_{j\ell' m}(\mathbf{r}/r), \end{equation} where $\Omega_{j\ell m}$ is called a spherical spinor and $\mathbf{r}/r = \mathbf{e}_r=(\sin\theta\,\cos\phi, \sin\theta\,\sin\phi, \cos\theta)$ so that the former only depend on the angles $(\theta,\phi)$ and not on $r$. Here the indices $\{j\ell m\}$ indicate the total angular moment $j$ (the eigenvalues of $\mathbf{J}^2$ are $\hbar^2\,j(j+1)$), orbital angular momentum $\ell$ (the eigenvalues of $\mathbf{L}^2$ are $\hbar^2\,\ell(\ell+1)$), and the magnetic quantum number $m$ takes values from $\{-j, -j+1,\ldots,j\}$. Here, the value of the intrinsic spin is $s=\tfrac{1}{2}$ (the eigenvalues of $\mathbf{S}^2$ are $\hbar^2\,s(s+1)$). Note that the expression for $\chi$ carries an index $\ell'$ given by \be{ell:prime} \ell' = \begin{cases} \ell+1, \quad \trm{if}\; j=\ell+\frac{1}{2},\\ \ell-1, \quad \trm{if}\; j=\ell-\frac{1}{2}. \end{cases} \end{equation} This is the subtlety that warrants much algebra to prove. The physics of it lies with the fact that solutions of the SR DE (as well as the NRQM SE) must also be eigenstate of the parity operator, which simply changes $\mathbf{r}\to-\mathbf{r}$, and is equivalent to $\{\theta, \phi\}\to\{\pi-\theta, \pi+\phi\}$. This is manifested in the choice of spherical harmonic functions $Y_{\ell m}(\theta,\phi)$ that appear in $\Omega_{j\ell m}$ and $\Omega_{j\ell' m}$. Under the operation of parity we have have $Y_{\ell m}(\theta,\phi)\to Y_{\ell m}(\pi-\theta, \pi+\phi) = (-1)^\ell\,Y_{\ell m}(\theta,\phi)$ so that the parity of the spherical harmonics depends solely on $\ell$. The implication of \Eq{ell:prime}, and the primary physical point, is that the parity of $\varphi_{j\ell m}$ is the negative (i.e. opposite) of that of $\chi_{j\ell' m}$. Thus, under parity, the orbital orbital angular momentum value $\ell'$ of $\chi_{j\ell' m}$ can only change by one unit from that $\ell$ of $\varphi_{j\ell m}$. Formally, this statement is expressed as \be{Greiner:p171:12} \left( \hat{\boldsymbol{\sigma}}\cdot \frac{\mathbf{r}}{r} \right)\,\Omega_{j\ell m} = -\Omega_{j\ell' m}. \end{equation} Note the use of $\ell$ on the lefthand side of \Eq{Greiner:p171:12}, but $\ell'$ on the righthand side. Here $\hat{\boldsymbol{\sigma}}\cdot \mathbf{r}/r$ is a scalar operator that changes sign under parity ($\mathbf{r}\to-\mathbf{r}$). \Eq{Greiner:p171:12} is a non-trivial statement which should be proved (explicitly, see RQMWE, (10.54) pp212-214), but for which here, we will simply utilize. Ultimately, the above additional complexity arises from the addition of the orbital and spin angular momentum to yield a fixed total angular momentum. Addition of angular momentum, and the manipulation of Clebsch-Gordon (CG) coefficients are standard topics covered in most undergraduate QM textbooks, but also quite explicitly and lucidly (with many worked examples) in the second volume (CTQMv2) of the two volume classic QM text by Cohen-Tannoudji, Diu and Laloe.\cite{Cohen-Tannoudji:1977:2} The key concept is that for a given orbital angular momentum $\ell$, the addition of an intrinsic spin-1/2 ($s=\frac{1}{2}$) leads to the two possible values for the angular momentum $j\in\{\ell-\frac{1}{2}, \ell+\frac{1}{2}\}$. This is often expressed as $\ell\otimes\frac{1}{2} = (\ell-\frac{1}{2})\oplus(\ell+\frac{1}{2})$ to indicate how the composition (addition, tensor product $\otimes$) of two angular momentum leads to the (direct sum, $\oplus$) of subspaces of total angular momentum $j=\ell\pm\frac{1}{2}$. Thus, if $\varphi_{j\ell m}$ has orbital angular momentum $\ell$ and total angular momentum, say, $j=\ell + \frac{1}{2}$, then since $\chi_{j\ell' m}$ has opposite parity with orbital angular momentum differing by one unit, then it can only have $\ell'=\ell+1$. This is the case, since by addition of its angular momenta, $\chi_{j\ell m}$ can only possibly have total the angular momentum values $\ell'=(\ell+1)\otimes\frac{1}{2} = (\ell+\frac{1}{2}\equiv j) \oplus (\ell+\frac{3}{2})$. The other potential choice $\ell'=\ell-1$ must be rejected since it can only result in total angular momentum values of $\ell'\overset{?}{=}(\ell-1)\otimes\frac{1}{2} = (\ell-\frac{3}{2}) \oplus (\ell-\frac{1}{2})$, neither which are $j=\ell+\frac{1}{2}$. (Similar arguments hold if we take $j=\ell-\frac{1}{2}$ implying that $\ell'=\ell-1$). The explicit expressions for the spherical 2-spinors are then given by (RQMWE, (11a) and (11b), p170 - which are not really needed - just their symmetry property \Eq{Greiner:p171:12}) \be{Greiner:p170:11a:11b} \Omega_{j=(\ell+\frac{1}{2}),\ell,m} = \left(\begin{array}{c} \sqrt{\frac{j+m}{2j}}\,Y_{\ell,m-\frac{1}{2}} \\ \sqrt{\frac{j-m}{2j}}\,Y_{\ell,m+\frac{1}{2}}\end{array}\right), \quad \Omega_{j=(\ell-\frac{1}{2}),\ell,m} = \left(\begin{array}{c} -\sqrt{\frac{j-m+1}{2(j+1)}}\,Y_{\ell,m-\frac{1}{2}} \\ \sqrt{\frac{j+m+1}{2(j+1)}}\,Y_{\ell,m+\frac{1}{2}}\end{array}\right). \end{equation} With this (important but involved) preliminary out of the way, one can now turn to the desired goal of extracting the pair of radial equations from our coupled \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2}. The solution method is straightforward ``in theory," namely we want to apply $\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}$ to \Eq{DE:RQMWE:p170:5:line:1} so that it can act on $\varphi$ on the righthand side, for which we can then substitute in \Eq{DE:RQMWE:p170:5:line:2}. The ``rub" comes in that $\hat{\mathbf{p}} = -i\,\hbar\,\boldsymbol{\nabla}$, and so it acts on both the radial and angular parts of $\varphi$ and $\chi$. We provide the relevants steps here (see the more complete details in RQMWE, p171) so that the student can see where the orbital angular momentum operator $\hat{\mathbf{L}}$ arises. Let us first consider the following calculation: \bea{Greiner:p171:10} (\hbs{\sigma}\cdot\hbs{p})\,\varphi_{j\ell m} &=& (\hbs{\sigma}\cdot\hbs{p})\Big(i\,g(r)\,\Omega(\bs{r}/r)_{j\ell m}\Big), \nonumber \\ &=& (\hbs{\sigma}\cdot\hbs{p})\big(i\,g(r)\big)\,\Omega_{j\ell m} + i\,g(r)\,(\hbs{\sigma}\cdot\hbs{p})\,\Omega_{j\ell m}, \nonumber \\ &=& \hbar\,\frac{d g(r)}{dr} \left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right) \,\Omega_{j\ell m} + i\,g(r)\,(\hbs{\sigma}\cdot\hbs{p})\,\Omega_{j\ell m}, \end{eqnarray} where the parity term of \Eq{Greiner:p171:12} $\left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right) \Omega_{j\ell m}$ is clearly exhibited. Let us now apply $\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}$ to $\Omega_{j\ell m}$ and use \Eq{Greiner:p171:12} (with $\left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right)^2=\mathbb{I}$) to obtain \be{Greiner:p171:13} -\left(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}\right)\, \Omega_{j\ell m} = \left( \hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}} \right)\, \left( \hat{\boldsymbol{\sigma}}\cdot\frac{\mathbf{r}}{r} \right)\, \Omega_{j\ell' m}. \end{equation} We can now use the well known product formula for Pauli matrices $(\hat{\boldsymbol{\sigma}}\cdot\mathbf{A}) (\hat{\boldsymbol{\sigma}}\cdot\mathbf{B}) = \mathbf{A}\cdot\mathbf{B} + i\,\hat{\boldsymbol{\sigma}}\cdot\mathbf{A}\times\mathbf{B}$, a relationship which is straightforwardly proved by expanding the vector operations in terms of components and using the well known Pauli matrix relations $\sigma_i\,\sigma_j = \delta_{ij} + i\,\epsilon_{ijk}\,\sigma_j\,\sigma_k$. Applying this to the righthand side of \Eq{Greiner:p171:13} one obtains \be{Greiner:p171:15} -\left(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}\right)\, \Omega_{j\ell m} = \left( \hat{\mathbf{p}}\cdot\frac{\mathbf{r}}{r} + i\, \hat{\boldsymbol{\sigma}}\cdot \left( \hat{\mathbf{p}}\times\frac{\hat{\mathbf{r}}}{r} \right) \right) \,\Omega_{j\ell' m}, \end{equation} from which one can identify the angular momentum operator $\hat{\mathbf{L}}= \mathbf{r}\times\hat{\mathbf{p}}$. However, one must be careful of operator ordering since $\hat{\mathbf{p}} = -i\hbar\,\boldsymbol{\nabla}$ is a differential operator that also acts on $\mathbf{r}$. Thus, we write the above term in the parentheses on the righthand side as \bea{Greiner:p171:16} \big( \hat{\mathbf{p}}\cdot\mathbf{r} &+& i\, \hat{\boldsymbol{\sigma}}\cdot \left( \hat{\mathbf{p}}\times\mathbf{r} \right) \big) \,\frac{1}{r}\,\Omega_{j\ell' m}, \nonumber \\ &=& \big( -i\hbar\,(\boldsymbol{\nabla}\cdot\mathbf{r}) -i\,\hbar\,\mathbf{r}\cdot\boldsymbol{\nabla} -i\, \hat{\boldsymbol{\sigma}}\cdot(\mathbf{r}\times\hat{\mathbf{p}}) \big) \,\frac{1}{r}\,\Omega_{j\ell' m}, \nonumber \\ &=& \left( -i\,\hbar\,\frac{3}{r} -i\,\hbar\,r\,\left(-\frac{1}{r^2}\right) -i\,\frac{\hbs{\sigma}\cdot\hbs{L}}{r} \right) \,\Omega_{j\ell' m}, \nonumber \\ &=& -i\,\frac{1}{r}\, \left( 2\,\hbar + \hbs{L}\cdot\hbs{\sigma} \right) \,\Omega_{j\ell' m}. \end{eqnarray} One then performs the standard calculation $\hbs{J}^2 = \left(\hbs{L}+\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2 = \hbs{L}^2+\left(\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2+\hbar\,\hbs{\sigma}\cdot\hbs{L}$ in order to write \bea{Greiner:p171:18} \hbar\,\hbs{\sigma}\cdot\hbs{L}\,\Omega_{j\ell' m} &=& \left( \hbs{J}^2 -\hbs{L}^2-\left(\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2 \right) \,\Omega_{j\ell' m}, \nonumber \\ &=& \Big( j(j+1) - \ell(\ell+1) - s(s+1) \Big) \,\hbar^2\,\Omega_{j\ell' m}, \quad \trm{for}\; s=\frac{1}{2}. \end{eqnarray} Further, one can now cleverly define $\kappa$ via \bea{kappa:defn} \kappa = \mp\,(j+\frac{1}{2})= \begin{cases} -(\ell+1), \quad\trm{for}\quad j=\ell+\frac{1}{2}, \\ \quad\ell, \hspace{0.55in}\trm{for}\quad j=\ell-\frac{1}{2}, \end{cases} \end{eqnarray} with $|\kappa| = j+\frac{1}{2}$ or $j=|\kappa|-\frac{1}{2}$. Then using $\ell' \equiv 2\,j-1$ and \Eq{Greiner:p171:18} one has (after some clever, but simple algebra, RQMWE, p171) that the righthand side of \Eq{Greiner:p171:16} can be written as $(2\hbar + \hbs{L}\cdot\hbs{\sigma})\,\Omega_{j\ell' m} = (1+\kappa)\,\hbar\Omega_{j\ell' m}$. If one had instead begun this whole calculation with $\Omega_{j\ell m}$ one would similarly arrive at $(2\hbar + \hbs{L}\cdot\hbs{\sigma})\,\Omega_{j\ell m} = (1-\kappa)\,\hbar\,\Omega_{j\ell m}$. These can be rearranged (by simply subtracting $\hbar\,\Omega_{j\ell' m}$ from both sides of the equality, and similarly with $\hbar\,\Omega_{j\ell m}$) in the form of and eigenvalue equation for the operator $\hat{\kappa}\overset{\trm{def}}{=} \hbar + \hbs{L}\cdot\hbs{\sigma}$ yielding $\hat{\kappa}\,\Omega_{j\ell m}= -\hbar\,\kappa\,\Omega_{j\ell m}$ and $\hat{\kappa}\,\Omega_{j\ell' m}= \hbar\,\kappa\,\Omega_{j\ell' m}$. Thus, one often defines the spinors $\chi_{{\kappa},\mu}\equiv\Omega_{j\ell m}$ and $\chi_{-{\kappa},\mu}\equiv\Omega_{j\ell' m}$ where $\mu=\pm\frac{1}{2}$ is the magnetic quantum number indexing spin up and spin down, respectively, so that these eigenvalue equations read equivalently as $\hat{{\kappa}}\,\chi_{{\kappa},\mu}= -\hbar\,{\kappa}\,\chi_{{\kappa},\mu}$ and $\hat{{\kappa}}\,\chi_{-{\kappa},\mu}= \hbar\,{\kappa}\,\chi_{-{\kappa},\mu}$. We can therefore write the (spatial portion) of the 4-spinor in a central potential for stationary states as \bea{Greiner:172:4:spinor} \psi_{j\ell m}(\bs{r}) &=& \left(\begin{array}{c}\varphi_{j\ell m}(\bs{r}) \\ \chi_{j\ell' m}(\bs{r},t)\end{array}\right) = \left(\begin{array}{c}i\,g(r)\,\Omega_{j\ell m}(\bs{r}/r) \\ -f(r)\,\Omega_{j\ell' m}(\bs{r}/r)\end{array}\right), \\ &=& \left(\begin{array}{c}i\,g(r)\,\chi_{{\kappa}, m} \\ -f(r)\,\chi_{-{\kappa}, m}\end{array}\right) = i\,\left(\begin{array}{c}\,g(r)\,\chi_{{\kappa}, m} \\ i\,f(r)\,\chi_{-{\kappa}, m}\end{array}\right). \end{eqnarray} Finally, using \Eq{Greiner:p171:16} and \Eq{Greiner:p171:12}, \Eq{Greiner:p171:10} takes the sought after form \be{Greiner:p172:22} (\hbs{\sigma}\cdot\hbs{p})\,\varphi_{j\ell m} = -\Omega_{j\ell' m} \left( \hbar\,\frac{d g(r)}{dr} + \hbar\,\frac{{\kappa}+1}{r}\,g(r) \right). \end{equation} Notice the parity flip from $\ell$ to $\ell'$ in going from the righthand to the lefthand side. Similarly, one derives \be{Greiner:p172:23} (\hbs{\sigma}\cdot\hbs{p})\,\chi_{j\ell' m} = -\Omega_{j\ell m} \left( \hbar\,\frac{d f(r)}{dr} - \hbar\,\frac{{\kappa}-1}{r}\,f(r) \right). \end{equation} We can now insert these expression into our coupled SR DE \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2} with the spherical spinors cancelling from both sides of both equations (the fruition of all the above involved labor) to obtain one of our final forms \bea{Greiner:p172:24} \hbar c\, \frac{dg(r)}{dr} + (1+{\kappa})\,\hbar c\,\frac{g(r)}{r} - [(E-V(r)) + m\,c^2]\,f(r) &=& 0, \\ \hbar c\, \frac{df(r)}{dr} + (1-{\kappa})\,\hbar c\,\frac{f(r)}{r} + [(E-V(r)) - m\,c^2]\,g(r) &=& 0. \end{eqnarray} With the standard substitution $G= r\,g$ and $F= r\,f$ ensuring $G(0)=F(0)=0$ at the origin singularity, the above equations further simplify to our final form \bea{Greiner:p172:24} \hbar c\, \frac{dG(r)}{dr} + \hbar c\,\frac{{\kappa}}{r}\,G(r) - [(E-V(r)) + m\,c^2]\,F(r) &=& 0, \\ \hbar c\, \frac{dF(r)}{dr} - \hbar c\,\frac{{\kappa}}{r}\,F(r)+ [(E-V(r)) - m\,c^2]\,G(r) &=& 0. \end{eqnarray} Again, if we wish to include a scalar coupling $V_s(r)$ to the mass, in addition to the vector coupling $V\to V_v(r)$ (as computed above), these equations become \bea{Greiner:with:vector:scalar:coupling} \hbar c\, \frac{dG(r)}{dr} + \hbar c\,\frac{{\kappa}}{r}\,G(r) - [(E-V_v(r)) + \big(m\,c^2+V_s(r)\big)]\,F(r) &=& 0, \label{Greiner:with:vector:scalar:coupling:line:1} \\ \hbar c\, \frac{dF(r)}{dr} - \hbar c\,\frac{{\kappa}}{r}\,F(r)+ [(E-V_v(r)) - \big(m\,c^2+V_s(r)\big)]\,G(r) &=& 0. \label{Greiner:with:vector:scalar:coupling:line:2} \end{eqnarray} The coupled first order equations \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2} are the SR extension of the second order radial equations in the NRQM SE. \subsection{Eigenenergies of the DE in a central potential with vector and scalar coupling}\label{DE:Solns:vector:scalar:coupling} We now wish to outline the solution for the eigenenergies of the stationary states of the DE containing both vector (Coulomb-like) and scalar (gravitational-like) coupling. Typically, for scalar coupling, the mass acts like a position dependent mass term, and is often associated with very massive particles such as the $\sigma$ meson so that the range is very short (inversely proportional to the mass). In quantum field theory (QFT) one describes this interaction in terms of the exchange of massless scalar mesons mediating the force, in analogy with the exchange of massless photons mediating the Coulomb force. Since we are dealing with SR particles, bound states mean that the magnitude of the energy of the stationary state is less than the rest mass, i.e. $-m\,c^2<E< m\,c^2$, where positive energies are associated with bound particles and negative energies with bound anti particles. Energies outside this range are associated with continuum states for both particles and anti-particles. The strategy to solve \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2} is analogous to that of the solution of the RNQM SE, except that now one must contend with two coupled first order radial equations for the DE, vs a single second order equation for the SE. This means there will ultimately be two coupled power series expansions, but once again the quantization condition will involve the truncations of these infinite series to polynomials to ensure that the boundary condition at spatial infinite (i.e. the wave function approaches zero) is satisfied, which in turn implies the functions are square integrable, thus ensuring the finiteness of the associate probabilities over all space. Again, one begins by examining the solution near the origin to develop a leading non-negative exponent to $r$ so that the wave function is also well behaved (i.e. zero) at the origin (the location of the point source). The details of such a calculation for the case of the DE with both vector and scalar coupling is explicitly worked out in RQMWE (section 9.8, pp184-187, see also section 9.6 p178-182 for the Coulomb solution only). Here, we just outline the highlights, but strongly encourage the reader to go through the detailed worked problems in RQMWE.\cite{Greiner:1990}. Using the convention of RQMWE we write $V_v(r) = -\alpha/r$ for the vector coupling (e.g. $\alpha = e^2/\hbar c = \alpha_C$) and $V_s(r) = -\hbar c \alpha'/r$ for the scalar coupling (e.g. $\alpha' = GMm/\hbar c = \alpha_G$). From \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2}, the radial equations take the form \bea{Greiner:p184:2} \frac{dG(r)}{dr} &=& -\frac{{\kappa}}{r}\,G(r) + \left[ \frac{E + m\,c^2}{\hbar c} + \frac{\alpha-\alpha'}{r}\right]\,F(r), \label{Greiner:p184:2:line1} \\ \frac{dF(r)}{dr} &=& \frac{{\kappa}}{r}\,F(r) - \left[ \frac{E - m\,c^2}{\hbar c} + \frac{\alpha+\alpha'}{r}\right]\,G(r). \label{Greiner:p184:2:line2} \end{eqnarray} We first consider the region $r\sim 0$, where we can then keep only the $1/r$ terms and drop the constant terms. With the ansatz $G=a\,r^{\gamma}$ and $G=b\,r^{\gamma}$ were are lead to a homogeneous set of linear equations $a({\gamma}+{\kappa}) - b(\alpha-\alpha')=0$ and $a(\alpha-\alpha')+b({\gamma}-{\kappa}) =0$, which upon setting the determinant of the coefficients equal to zero yields ${\gamma} = \pm \sqrt{{\kappa}^2 -\alpha^2+\alpha^{'2}}$. To allow for normalization of the wave functions, one must select the positive sign for ${\gamma}$. One can now define $\lambda \overset{\textrm{def}}{=} \frac{m^2 c^4-E^2}{\hbar c}$ and $\rho = 2\lambda r$ so that the radial equations become \bea{Greiner:p185:11} \frac{dG}{d\rho} &=& -\frac{{\kappa}}{\rho}\,G + \left[ \frac{E + m\,c^2}{2\hbar c\,\lambda} + \frac{\alpha-\alpha'}{\rho}\right]\,F, \\ \frac{dF}{d\rho} &=& \frac{{\kappa}}{\rho}\,F -\left[ \frac{E - m\,c^2}{2\hbar c\,\lambda} + \frac{\alpha+\alpha'}{\rho}\right]\,G. \end{eqnarray} It turns out to be convenient to define the functions $\phi_1(\rho)$ and $\phi_2(\rho)$ via $G = \sqrt{mc^2+E}\, e^{-\rho/2}\,$ $(\phi_1+\phi_2)$ and $F = \sqrt{mc^2+E}\, e^{-\rho/2}\,(\phi_1-\phi_2)$ and expand each in a power series given by $\phi_1=\rho^{\gamma}\,\sum_{m=0}^\infty \alpha_m\,\rho^m$ and $\phi_2=\rho^{\gamma}\,\sum_{m=0}^\infty \beta_m\,\rho^m$. The justification for introducing $\phi_1$ and $\phi_2$ is that resulting equation for $\phi_2$ allows one to write the coefficients $\beta_m$ in terms of the $\alpha_m$, via (see RQMWE, p186) \bea{Greiner:p186:17} \frac{\beta_m}{\alpha_m} &=& \frac { -{\kappa} + \alpha m c^2/\hbar c\lambda + \alpha' E/\hbar c\lambda } { m+{\gamma} - \alpha m c^2/\hbar c\lambda - \alpha'E/\hbar c\lambda }, \nonumber \\ &=& \frac { {\kappa} - \alpha m c^2/\hbar c\lambda - \alpha' E/\hbar c\lambda } {n'-m}, \nonumber \\ \trm{with}\;\; n' &\overset{\textrm{def}}{=}& \frac{\alpha E}{\hbar c \lambda} + \frac{\alpha' m c^2}{\hbar c \lambda}-{\gamma}. \end{eqnarray} Inserting this into the series relationship from $\phi_1$, which involves $\alpha_m$ expressed in terms of $\alpha_{m-1}$ and $\beta_m$, one can develop an expression for $\alpha_m/\alpha_{0}$ whose numerator involves the product $(n'-1)(n'-2)\ldots(n'-m)$. Once again, if the series do not terminate, they lead to a sum that scales as $e^{\rho}$ so that the wave function does not converge at spatial infinity. Thus, as in the SE quantization case, requiring the series to terminate, demands that $n'=\{0,1,2,\ldots\}$. Inserting this result back into \Eq{Greiner:p186:17} this equation produces a quadratic equation for the energy $E$ (since $\lambda\propto E$), which can be easily solved to finally give \bea{Greiner:p186:20} \frac{E}{mc^2} &=& \left\{ \frac{-\alpha\alpha'}{\alpha^2 + (n-(j+\frac{1}{2})+\gamma)^2} \pm \left[ \left( \frac{\alpha\alpha'}{\alpha^2 + (n-(j+\frac{1}{2})+\gamma)^2} \right)^2 \right. \right. \nonumber \\ &-& \left. \left. \frac { \alpha^{'2} - (n-(j+\frac{1}{2})+\gamma)^2 } { \alpha^{2} + (n-(j+\frac{1}{2})+\gamma)^2 } \right]^{1/2} \right\}. \end{eqnarray} In the above, the principal quantum number $n$ is defined as $n=n'+|{\kappa}| = n' + j + \frac{1}{2}$ with $n\in\{1, 2, \ldots\}$. Special cases of the above are both interesting and relevant. \subsubsection{Vector Coupling only} Here $\alpha'=0$ and ${\gamma} = \sqrt{{\kappa}^2-\alpha^2}$ apropos for the Coulomb field and one obtains \be{Greiner:p186:case:2} \frac{E}{mc^2} = \left[ 1+ \frac{\alpha^2}{(n-(j+\frac{1}{2})+{\gamma})^2} \right]^{-1/2} \approx 1 - \frac{1}{2}\frac{\alpha^2}{n^2} + \ldots, \end{equation} recovering the SE result since $\alpha=\alpha_C \approx 1/137\ll 1$. (Note: the negative energies do not fulfill the original quadratic equation \Eq{Greiner:p186:17} for $E$ in the case of vector coupling since both $n'+{\gamma}$ and $\frac{\alpha\,E}{\hbar c \lambda} + \frac{\alpha' m c^2}{\hbar c \lambda}$ are positive, and so that the $E<0$ solutions must be excluded). \subsubsection{Scalar Coupling only} In the case of scalar coupling $\alpha=0$ and ${\gamma} = \sqrt{{\kappa}^2+\alpha^{'2}}$ (apropos for a gravitational-like field), one obtains instead the expression \be{Greiner:p186:case:1} \frac{E}{mc^2} = \pm\left[ 1- \frac{\alpha^{'2}}{(n-(j+\frac{1}{2})+{\gamma})^2} \right]^{1/2} \approx \pm\left(1 - \frac{1}{2}\frac{\alpha'^2}{n^2} + \ldots\right), \quad\trm{if}\quad \alpha'\ll 1. \end{equation} Thus, the above two formulas \Eq{Greiner:p186:case:2} and \Eq{Greiner:p186:case:1} almost answer the student's original question when we note the similarity of the placement of the square roots when comparing \Eq{E_v} with \Eq{Greiner:p186:case:2}, and \Eq{E_s} with \Eq{Greiner:p186:case:1}. We see that in the small coupling limit both formulas have the form of $1-\frac{\alpha^2}{ 2 n^2}$, but this requires that $\alpha'=GMm/\hbar c = \frac{M}{M_p} \frac{m}{M_p}\ll 1$ (see \Eq{alpha:G}). While this is true say for $\sigma$ mesons and other elementary particles, for the originally posed question of a solar mass BH, $\frac{M}{M_p}$ is astronomically huge, so that $\alpha'\to\alpha_G\approx 4\times 10^{15}\gg 1$ (to say the least!). Still, one's intuition is somewhat borne out by this analysis, but not fully yet answered to the student's satisfaction. Thus, in order to fully answer the student's question, we need to turn now to the DE in curved spacetime where the role of the metric encoding the gravitational field enters in a fundamentally new way. Further, the strange new concept of a horizon, dividing spacetime into two distinct regions: outside and inside, with its conceptually unusual (and confounding) ``one-way membrane" property (i.e. particles can cross the horizon from outside to in, but not the reverse!) rears it strange and complicated head. \section{The DE in curved spacetime}\label{sec:DE:CST} General Relativity (GR) presents its own unique challenges, both conceptually and mathematically, even to the motivated student. However, most students are familiar at least with some basic concepts, including the curved spacetime metric, black holes (BH) and particle orbits about them, and the bending of light rays around a massive object, even if they might not be quite capable of deriving them. Therefore, while there is some degree of heavy lifting involved when learning GR (the fundamentally new concept of spacetime vs space and time, tensors, covariant derivatives, parallel transport, etc...) there currently exists enough excellent GR textbooks now geared to the upper level undergraduate that attempts to ease these burdens. One excellent text that we will use here is that by James Hartle\cite{Hartle:2009}, which emphasizes ``physical concepts first" before mathematical manipulation. For most students (of all ages!) it's the concepts that both SR and GR introduce that are at first glance difficult to wrap one's head around. But, 100 years on now, there has been enough well explained exposition (and textbooks) that physics such as BHs and the bending of light are now part of general (and popular) knowledge. Therefore, the goal here is not review all of GR, but just those salient points that will get us to our goal most expeditiously. On top of this, we also want to include the non-standard (for the novice) topic of the DE in curved spacetime (CST). To that end, there does exist an excellent undergraduate GR text by Lewis Ryder\cite{Ryder:2009} that clearly and lucidly deals with this topic, leading to the DE in CST (see Ryder, Chapter 11.3, pp409-416). Again, it is not our goal to reproduce all of Ryder's illuminating discussion. In the spirit of wanting to drive our car first before knowing how the engine under the hood was built (but allowing ourselves to read the service manual now and again), we again will point out the key concepts involved and refer the reader to Ryder's excellent GR textbook. In the following we will first draw from Hartle\cite{Hartle:2009} (see section 9.3, p191-204) in order to get to a form of the classical GR radial equation (no quantum here!) that looks like an ordinary NR energy equation. This involves being given the Schwarzschild metric (sans derivation), and manipulating it (the 4-velocity) in order to produce the desired equation. The goal is to identify the GR radial potential for comparison to the DE in CST. We then next switch gears, and focus on the derivation of the DE in CST (Ryder, Chapter 11.3, pp409-416). The key concept from GR is the covariant derivative, required to explain how to take derivatives in the surrounding CST as one moves from point to point. As mentioned earlier, the global inertial frames of SR are now ``collapsed" to local regions around each point which (following Hartle) we refer to as the observer's local laboratory. As long as the extent of the observer's spatial axes, and the duration of time measured, are in a sense ``small" (with respect to changes in the curvature of the CST), the Equivalence Principle holds, and hence physics appears special relativistic. Thus, one must be able to describe the observer's local (frame) laboratory, and hence we are led down the road of introducing a tetrad of 4-vectors describing the local laboratory. Measurements and description of particles (e.g. their 4-momentum) passing through the observer's local laboratory are then made by ``projecting" then onto the observer's local axes (3 space and 1 time axes). The introduction of tetrads is the key concept required then to similarly ``project" the DE in the surrounding CST ``into" the observer's local frame, which is locally (Minkowski) flat (i.e where SR holds). The relevant concept could be stated as such: ``Since we already know how to write down and solve the DE in SR (see the previous section), then by use of the Equivalence Principle, the strategy is to ``simply" project the CST DE into the observer's locally flat frame at each spacetime point $x$, where we already know how to solve it." This is what Ryder does in Chapter 11.3. Again, our objective is to get to the DE in CST as quickly as possible, so again we point out the major highlights of this derivation (with an emphasis more on the why and how we get there, which can be daunting to the un-initiated, rather than on the detailed proofs - which are shown in Ryder). So let us begin. \subsection{The effective radial potential for particles in the Schwarzschild metric}\label{sec:Veff:SST} As is well known, the Schwarzschild metric for a central symmetric gravitational source of mass M is (in full units, see Hartle, p186) \bea{Schw:metric} \hspace{-0.5in} ds^2 &=&-{\mathcal{F}}(r)\,c^2\,dt^2 + \frac{1}{{\mathcal{F}}(r)}\,dr^2 + r^2\,d\Omega^2, \;\; {\mathcal{F}}(r) \overset{\textrm{def}}{=} 1-\frac{2GM}{c^2 r}, \;\; d\Omega^2 = \left(d\theta^2 + \sin^2\theta\,d\phi^2 \right), \label{Schw:metric:line:1} \\ \hspace{-0.5in} &=& g_{\mu\nu}\,dx^\mu\,dx^\nu, \qquad \mu,\nu\in\{0,1,2,3\}, \label{Schw:metric:line:2}\\ \hspace{-0.5in} &\approx& -\left(1+\frac{2 V_G(r)}{c^2}\right)\,c^2\,dt^2 + \left(1-\frac{2 V_G(r)}{c^2}\right)\,dr^2 + d\Omega^2, \qquad V_G(r) = -\frac{G M}{r},\label{Schw:metric:line:3} \end{eqnarray} where we have defined the Schwarzschild factor ${\mathcal{F}}(r) \overset{\textrm{def}}{=} 1-\frac{2GM}{c^2 r}$ that will figure prominently throughout the discussions of Schwarzschild spacetime (SST). \Eq{Schw:metric:line:3} is the Schwarzschild metric in the weak field approximation (apropos for, say about the Sun or the Earth), in which the Newtonian potential $V_G(r) = -\frac{G M}{r}$ explicitly appears. The relevant scale length (see \App{app:Units}) is the well known Schwarzschild radius $r_s\overset{\textrm{def}}{=} 2GM/c^2$ (so that $2 V_G/c^2 = -\left(\frac{r_s}{r}\right)$ is unitless). For a stationary observer (i.e. one that sits at a fixed coordinate position) at fixed $r$, the metric yields $ds^2 \overset{\textrm{def}}{=} -d\tau^2 = -{\mathcal{F}}\,c^2\,dt^2 = g_{00} (dx^0)^2$ (taking $x^0= c t$) where $\tau$ is then seen as the \tit{proper time}, i.e. the time as measured on a clock carried with the observer. Writing the metric as $-d\tau^2 =ds^2 = g_{\mu\nu}\,\left(\frac{dx^\mu}{d\tau^2}\right)\, \left(\frac{dx^\nu}{d\tau^2}\right)\, d\tau^2 \overset{\textrm{def}}{=} \textrm{\bf u}^2_{obs}\, d\tau^2 $ we have $\textrm{\bf u}^2_{obs}~=~-~1$, where we have defined the observer's 4-velocity $\textrm{\bf u}^\mu = dx^\mu/d\tau$, i.e. the velocity of the observer (rate of change of their coordinates) with respect their proper time $\tau$ (vs their coordinate time $t=t(\tau)$). The observer's 4-velocity will be in fact the ``time axis" of the local local laboratory, which we denote as $\textrm{\bf e}_0(x) \overset{\textrm{def}}{=} \textrm{\bf u}_{obs}(x)$ a timelike unit vector (i.e. has magnitude $-1$) and is tangent to the (freely falling geodesic) trajectory of the observer's motion (worldline) through the surrounding CST. The other three spatial axes of the observer's local laboratory, denoted as $\{\textrm{\bf e}_1(x), \textrm{\bf e}_2(x), \textrm{\bf e}_3(x)\}$ are chosen orthogonal to $\textrm{\bf e}_0(x)$ (see \Fig{fig:observers:local:laboratory}), and will be discussed later. For now, in order to extract radial orbits of the Schwarzschild metric, we only require $\textrm{\bf u}_{obs}(x)$. Note that if a particle of 4-momentum $\textrm{\bf p}$ passes through the observer's local laboratory, then the observer will measure its energy as $E/c= -\textrm{\bf p}\cdot\textrm{\bf u} = -\textrm{\bf p}\cdot\textrm{\bf e}_0$, and its 3-momentum components as $p_a=\textrm{\bf p}\cdot\textrm{\bf e}_a$ for $a\in\{1,2,3\}$ yielding a local description of the particle with a SR-like local laboratory 4-vector $(E/c, \bs{p})$ (as an instantiation of the Equivalence Principle). \begin{figure}[h] \includegraphics[width=4.5in,height=3.0in]{fig_1_DE_SST_AJP_26Jun2022} \caption{The observer's \tit{local laboratory} (small room with physicist) at the curved spacetime (CST) point $x$, defined by the orthonormal tetrad $\textrm{\bf e}_a(x)$, $a=(0,1,2,3)$. The three spatial axes $\textrm{\bf e}_i(x)$, $i=(1,2,3)$ are located at the origin of the observer's laboratory (the universal coordinates system ``in the corner of the room"), while $\textrm{\bf e}_0(x)=\textrm{\bf u}_{obs}(x)$ is the temporal axis, defined as the observer's 4-velocity, which is the tangent to the physicist's geodesic trajectory. A particle of 4-momentum $\textrm{\bf p}(x) = m \textrm{\bf u}(x)$ and world components $p^\alpha(x)$ passes through the observer's local laboratory. The observer measures the \tit{local} components $p^a(x) = e^a_{\;\;\alpha}(x)\,p^\alpha(x)$. At a small proper time later $d\tau$, the particle has moved from $x^\alpha \to x^{'\alpha} = x^\alpha + u^\alpha(x)\,d\tau$, which is measured by the observer in their local laboratory at the spacetime point $x^{'\alpha}$. } \label{fig:observers:local:laboratory} \end{figure} Particles (here both massive and massless until otherwise specified) under no other external forces (i.e. accelerations) travel on \tit{geodesics}, the analogue of straight line motion in Euclidean space, extended to CST. Such observers are called freely falling (FF). Now without delving into the geodesic equation per say (since we will not need to solve this equation directly in the subsequent discussion), an important concept is that of conserved quantities along the geodesic motion. If the metric is independent of a particular coordinate, then a conserved quantity exist along this motion (called an isometry; this is essentially Noether's theorem applied to geodesics, see Hartle, section 8.2, pp175-178). For example, for the Schwarzschild metric, we see that it is independent of the coordinates $t$ (stationarity) and azimuthal angle $\phi$ (rotationally invariant about the $z$-axis). Let $\bs{\xi}$ be a coordinate vector along the the \tit{isometry} (i.e. motion along which the metric does not change). % $\bs{\xi}$ is called a \tit{Killiing vector} (after Wilhelm Killing (1847-1923), see Hartle, Chapter 8.2, pp175-178) and there exists an equation named after him, that allows one to derive the isometries systematically for any metric. However, quite often one can deduce the Killing vectors by the symmetry of the metric by inspection. % The key result is that $\bs{\xi}\cdot\textrm{\bf u}_{obs}$ is a conserved quantity all along the geodesic. For the Schwarzschild metric, the metric does not change for translations in time $t$, so that $\bs{\xi}_t = (1,0,0,0)$, and therefore we define the quantity $e\overset{\textrm{def}}{=} -\bs{\xi}_t\cdot\textrm{\bf u}_{obs}$, which can be physically interpreted as the particles rest energy per unit mass (at large $r$). % In the following we will follow the convention of Hartle (and most GR textbooks) and work in units of $G=c=1$. Physical units can be restored by resorting to dimensional analysis. A second conserved quantity for particles on Schwarzschild geodesics is the $\ell = \bs{\xi}_\phi\cdot\textrm{\bf u}_{obs}$ the orbital angular momentum per unit mass (at large $r$), with $\bs{\xi}_\phi = (0,0,0,1)$ (in spherical polar coordinates $(t, r, \theta, \phi)$) indicating the independence of the Schwarzschild metric in the azimuthal coordinate $\phi$. Using these two conserved quantities and the timelike unit magnitude of the observer's 4-velocity, there exists enough symmetry to deduce the particle orbits directly. Note that since the orbital angular momentum is conserved, the geodesic motion occurs in a plane, which is conventionally (for convenience) take to be the equator, $\theta=\pi/2$. In the following we will work in a \tit{coordinate basis} which means that $\textrm{\bf e}_a$ are just coordinate derivatives in the direction indicated and their inner products gives the metric components. This means that for two arbitrary 4-vectors $\bs{a}$ and $\bs{b}$ with coordinate components $a^\mu$ and $b^\nu$, respectively, we have $\bs{a}\cdot\bs{b}= g_{\mu\nu}a^\mu\,b^\nu$. Thus, for the Schwarzschild metric we define the conserved quantities as (following Hartle and using units of $G=c=1$) \bea{Hartle:p193:9.21:9.22} e &=& -\bs{\xi}_t\cdot\textrm{\bf u} = \left(1-\frac{2 M}{r}\right)\,\frac{dt}{d\tau}={\mathcal{F}}(r)\,\frac{dt}{d\tau}, \label{Hartle:p193:9.21} \\ \ell &=& \bs{\xi}_\phi\cdot\textrm{\bf u} = r^2\sin^2\theta\,\frac{d\phi}{d\tau}. \label{Hartle:p193:9.22} \end{eqnarray} With the above conserved quantities in hand, and $\textrm{\bf u} = (u^t, u^r, u^\theta, u^\phi)$ we have from its normalization $\textrm{\bf u}_{obs}\cdot\textrm{\bf u}_{obs}=-1$, \bea{Hartle:p194:9.25} -1 &=&-\left(1-\frac{2M}{r} \right)\,\left(u^t\right)^2 + \left(1-\frac{2M}{r} \right)^{-1}\,\left(u^r\right)^2 + r^2\,\left(u^\phi\right)^2, \\ \trm{or}\quad -1 &=& -\left(1-\frac{2M}{r} \right)^{-1}\,e^2 + \left(1-\frac{2M}{r} \right)^{-1}\,\left(\frac{dr}{d\tau}\right)^2 + \frac{\ell^2}{r^2}, \\ \Rightarrow\quad \mathcal{E} \overset{\textrm{def}}{=} \frac{e^2-1}{2} &=& \frac{1}{2}\,\left(\frac{dr}{d\tau}\right)^2 + \left[ \left(1-\frac{2M}{r} \right)\,\left(1+\frac{\ell^2}{r^2} \right)-1 \right], \\ \trm{or}\quad \mathcal{E}&=& \frac{1}{2}\,\left(\frac{dr}{d\tau}\right)^2 + V_{eff}(r), \label{Netwon:E:eqn} \end{eqnarray} where we have defined the effective potential (see Hartle, p194) \bea{V:eff} V_{eff}(r)\equiv \left[ \left(1-\frac{2M}{r} \right)\,\left(1+\frac{\ell^2}{r^2} \right)-1 \right] &=& -\frac{M}{r} + \frac{\ell^2}{2\,r^2} - \frac{M\,\ell^2}{r^3}, \label{V:eff:line1} \\ &=& \frac{1}{c^2}\, \left( -\frac{GM}{r} + \frac{\ell^2}{2\,r^2} - \frac{G M\,\ell^2}{c^2\,r^3} \label{V:eff:line2} \right), \end{eqnarray} where in the last line we have restored all the physical constants. \Eq{Netwon:E:eqn} has the form of a NR energy equation with a Newtonian-like potential with orbital angular momentum barrier (first two terms of \Eq{V:eff:line2}), but with an additional attractive GR correction (last term of \Eq{V:eff:line2}) that scales as $1/r^3$. This last term dominates for small $r$ and is responsible for the characteristic GR effects such as the precession of the perihelion, bending of light, etc. We can make the Newtonian analogy even stronger by defining $E_{Newt}$ via $e^2 = (mc^2 + E_{Newt})/mc^2$, i.e. as the small correction to the particle's rest mass, in strong analogy to SR. With this substitution, the radial equation becomes (see Hartle p195, restoring full units) \be{Hartle:p195:9.32} E_{Newt} \approx \frac{m}{2}\,\left(\frac{dr}{d\tau}\right)^2 + \frac{L^2}{2 m r^2} -\frac{G M m}{r} - \frac{G M L^2}{c^2 m r^3}, \qquad L\overset{\textrm{def}}{=} m\,\ell, \end{equation} where we have approximated $\mathcal{E}=(e+1)(e-1)/2\approx (e-1)$. \Eq{Hartle:p195:9.32} now has the exact same form as the energy integral in Newtonian gravity with an additional relativistic correction to the potential proportional to $1/r^3$. Note that \be{Hartle:p195:9.33} V_{eff}(r) \underset{r\to\infty}{\longrightarrow} -\frac{G M}{c^2\,r} = -\frac{r_s/2}{r}, \qquad V_{eff}(r_s) = -\frac{1}{2}. \end{equation} Further, at the Schwarzschild radius $r=r_s$ the first term of $V_{eff}(r_s)=-\frac{1}{2}$, and the second and thrid terms exactly cancel each other. A detailed investigation for both radial plunge orbits ($\ell=0$), circular and hyperbolic orbits ($\ell\ne 0$) are well explored in Hartle Chapter 9 (and many other GR textbooks), and will not be covered here. For our goals, we are primarily interested in radial plunge orbits, since in this case $V_{eff}(r) = -\frac{G M}{c^2\,r}=-\frac{r_s/2}{r}$ has the exact form of the scalar coupling potential, discussed previously for the SR DE. The question now, is how to incorporate the GR effects into the DE. We turn to this next, leveraging the discussion in Ryder\cite{Ryder:2009} (Chapter 11.3). \subsection{The DE in CST: GR preliminaries}\label{sec:DE:CST:preliminaries} In order to arrive to our destination, the DE in CST, we first need to stop once more at the discussion of the description of the observer's local laboratory. We need to distinguish between the use of various ``basis vectors" and corresponding ``1-forms" used to describe the observer's frame. The reason is that since the observer's local (frame) laboratory is described in terms of four basis vectors $\textrm{\bf e}_a$, we want to distinguish between coordinate bases (cryptically called \tit{holonomic} in the literature) and an orthonormal set of basis vectors (called, \tit{non-holonomic}). Why the two descriptions? In the spirit of GR, \tit{any} set of basis vectors are allowed, but these two are the one's most conveniently employed. Coordinate basis vectors are the ``easiest" to use and formally of the form $\textrm{\bf e}_\mu = \partial_\mu \overset{\textrm{def}}{=} \partial/\partial x^\mu$ for coordinates $x^\mu$. (Here we have used $a\to\mu$ since the index $\mu$ reflects the surrounding CST). The defining property of a coordinate basis set is that their inner (dot) product defines the metric (to be shown shortly), and that the basis vectors commute, namely $[\textrm{\bf e}_\mu, \textrm{\bf e}_\nu]=[\partial_\mu, \partial_\nu]= 0$ for $\mu\ne\nu$ reflecting the independence of the order of coordinate differentiation, i.e. $\partial_\mu\,\partial_\nu= \partial_\nu\,\partial_\mu$. On the other hand, the orthonormal basis is more physical, and better suited to what the observer actually measures. It's defining property is that the inner product between the orthonormal basis vectors defines the \tit{local} metric, which in this case is the flat Minkowskian metric of SR. The orthonormal basis vectors are of the form $\textrm{\bf e}_a = e_a^{\;\;\mu}(x)\,\partial_\mu$ where the \tit{tetrad components} $e_a^{\;\;\mu}(x)$ are spatially dependent. Thus, $[\textrm{\bf e}_a, \textrm{\bf e}_b]\ne 0$ in general. However, they do have the simplifying property that $\textrm{\bf e}_a\cdot\textrm{\bf e}_b = \eta_{ab} = \trm{diagonal}(-1,1,1,1)$ the constant Minkowski SR metric - which is just a statement of the Equivalence Principle. (Note, the relativist's minus sign convention in front of the time component, vs the particle physicists convention of using $\eta_{ab} = \trm{diagonal}(1,-1,-1,-1)$. If you pick up a random book an look at the ``sign of the times", you can instantly tell if the author is a particle physicist or a relativist, without even looking at the title. Try it!). To recap this important distinction, we recap this once more below: \bea{basis:vectors} \trm{coordinate basis:}\qquad \textrm{\bf e}_\mu &=& \partial_\mu, \qquad\quad \textrm{\bf e}_\mu\cdot\textrm{\bf e}_\nu = g_{\mu\nu}(x), \\ \trm{orthonormal basis:}\qquad \textrm{\bf e}_a &=& e_a^\mu\partial_\mu, \qquad \textrm{\bf e}_a\cdot\textrm{\bf e}_a = \eta_{ab}(x). \end{eqnarray} (Note: to distinguish the two basis, some authors use a circumflex $\verb+^+$ over the index $\hat{a}$ so that $\textrm{\bf e}_0$ and $\textrm{\bf e}_{\hat{0}}$ denote the observer's 4-velocity in a coordinate and orthonormal basis, respectively. We will no have occasion to do this, since we will primarily use an orthonormal basis for the DE in CST). An example speaks a thousand words. Consider the simplest case of polar coordinates in two spatial dimensions, with coordinates $(r,\theta)$ such that $x=r\,\cos\theta$ and $y=r\,\sin\theta$ with line element $ds^2 = dr^2 + r^2 d\phi^2$. Then the coordinate basis vectors would be $\textrm{\bf e}_r = \partial_r$ and $\textrm{\bf e}_\phi = \partial_\phi$ and we have $\textrm{\bf e}_r\cdot\textrm{\bf e}_r = 1 = g_{rr}$ and $\textrm{\bf e}_\phi\cdot\textrm{\bf e}_\phi = r^2= g_{\phi\phi}$ with $\textrm{\bf e}_r\cdot\textrm{\bf e}_\phi=0$, leading to the metric $g_{\mu\nu} = \tiny{\left(\begin{array}{cc} 1 & 0 \\0 & r^2\end{array}\right)}$. Clearly $[\partial_r, \partial_\phi]=0$. In an orthonormal basis we would instead define $\textrm{\bf e}_r = \partial_r$ and $\textrm{\bf e}_\phi = \frac{1}{r}\,\partial_\phi$, with $\textrm{\bf e}_r\cdot\textrm{\bf e}_r = 1 = g_{rr}$ and $\textrm{\bf e}_\phi\cdot\textrm{\bf e}_\phi = 1= g_{\phi\phi}$ with $\textrm{\bf e}_r\cdot\textrm{\bf e}_\phi=0$, leading to the metric $g_{ab} = \tiny{\left(\begin{array}{cc} 1 & 0 \\0 & 1\end{array}\right)} = \delta_{ab}$. However, we now have $[\textrm{\bf e}_r, \textrm{\bf e}_\phi]\,f(r,\phi) = [\partial_r, \frac{1}{r}\,\partial_\phi]\,f =-\frac{1}{r^2}\partial_\phi\,f = -\frac{1}{r}\textrm{\bf e}_\phi\,f$ which allows one to conclude (since $f$ was an arbitrary function) that $[\textrm{\bf e}_r, \textrm{\bf e}_\phi]= -\frac{1}{r}\textrm{\bf e}_\phi \equiv \textrm{\bf e}_\phi\,C^{\phi}_{\;\;r \phi}(x)$. In the last step we have introduced the structure constants $C(x)$ defined by the commutators of the basis vector, $[\textrm{\bf e}_a, \textrm{\bf e}_b] =\textrm{\bf e}_c\, C^{ c}_{\;\;a b}(x)\,$ (with sum over the index $c$, see Hartle, p109), which just states that the commutator of the basis can be expanded in terms of a linear combination of the basis vectors. A general vector (an, in general, tensor) $\bs{v}$ is a geometric object, independent of the coordinates and basis vectors (user's laboratory frame) used to describe it. So we can write this as $\bs{v} = v^\mu(x)\,\textrm{\bf e}_\mu(x)$ using a coordinate basis vector description (with \tit{contravariant} coordinate components $v^\mu$), or as $\bs{v} = v^a(x)\,\textrm{\bf e}_a(x)$ using an orthonormal basis vector description (with physical orthonormal components $v^a(x)$). Note that the transition from coordinate basis to orthonormal basis (for this diagonal metric) can be performed by inspection by examining the metric and grouping terms as $ds^2 = (dr)^2 + (r d\phi)^2$ and somehow considering it's ``inverse," namely $(\partial_r, \frac{1}{r}\partial_\phi)$. Note that $r\,d\phi$ is not a total differential, so there is no coordinate associated with such an object. This intuition can be formalized by defining basis \tit{1-forms} which are \tit{dual} to the orthonormal basis vectors $\textrm{\bf e}_a$. We denote these 1-forms (generalized differentials) as $\bs{\theta}$ (Ryder's notation) which can be decomposed in terms of the coordinate (true) differentials $dx^\mu$ via $\bs{\theta}^a = e^a_{\;\;\mu}(x) dx^\mu$. Note that we have purposely used the same symbol $e$ for the components of the one form, with the important distinction that the coordinate index $\mu$ is now on the bottom and the orthonormal index $a$ is on top (vs $\textrm{\bf e}_a^{\;\;\mu}$ associated with the basis vectors $\textrm{\bf e}_a = e_a^{\;\;\mu}(x)\,\partial_\mu$). Thus, even without a metric, we can define what we mean by dual by saying that a 1-form is an object that ``eats" vectors in the following sense (in both coordinate and orthonormal bases) $dx^\mu(\partial_\nu)\overset{\textrm{def}}{=} \delta^\mu_{\;\;\nu}$, and $\bs{\theta}^a(\textrm{\bf e}_b)\overset{\textrm{def}}{=} \delta^a_{\;\;b}$. Thus, consider a general 1-form $\bs{w} = w_a(x)\bs{\theta}^a$ with (covariant) components $w_a(x)$. Then, we can have this 1-form act on a vector $\bs{v} = v^a(x)\,\textrm{\bf e}_a$ to give $\bs{w} (\bs{v}) = w_a\bs{\theta}^a (v^b\,\textrm{\bf e}_b) = w_a \,\bs{\theta}^a(\textrm{\bf e}_b)\,v^b = w_a\,\delta^a_{\;\;b}\, v^b = w_a v^a \overset{\textrm{def}}{=} \bs{w}\cdot\bs{v}$. Therefore, we can define and inner product without the need for a metric, and the metric merely serves to raise and lower components via $g_{\mu\nu} v^\nu = v_\mu$ and $\eta_{ab} v^a = v_b$. \subsection{The DE in CST: covariant derivatives in GR}\label{sec:DE:CST:covar:deriv:GR With these preliminaries out of the way, we now get to the crux of the matter (and the part that is often an initial learning bottleneck for the beginning GR student). How does one define the derivative of basis vectors? Let's first work in the simpler coordinate basis vectors. Consider the following calculation for a posited, yet unknown, derivative which we denote as $\nabla_\mu$: $\nabla_\mu \bs{v} = \nabla_\mu\big(v^\nu(x) \textrm{\bf e}_\nu(x)\big) = (\partial_\mu v^\nu)\,\textrm{\bf e}_\nu + v^\mu (\nabla_\mu\textrm{\bf e}_\nu)$. Here, we have made the (natural) assumption that $\nabla_\mu\to\partial_\mu$ on functions (of which $v^\nu$ are). Acting on basis vectors, we are yet unsure, so we just leave it as $\nabla_\mu$. This states that we must not only differentiate the components $v^\nu(x)$ of $\bs{v}$, but also its basis vectors $\textrm{\bf e}_\nu(x)$. But we are already familiar with this latter concept, since even in our simple 2D polar coordinate example above, the basis vector $\textrm{\bf e}_\phi = \partial_\phi$ is tangent to circles of constant radius $r$, and thus change direction (in the surrounding 2D Euclidean $\mathbb{R}^2$ space) as we vary the coordinate $\phi$. As in the case of the 1-forms above, we assume that $\nabla_\mu\textrm{\bf e}_\nu$ can be expanded in terms of the basis vectors, and so we write $\nabla_\mu\textrm{\bf e}_\nu(x) = {\Gamma}_{\mu\nu}^\lambda(x)\,\textrm{\bf e}_{\lambda}$ with proportionality functions ${\Gamma}_{\mu\nu}^\lambda(x)$, the famous \tit{Levi-Civita connection}, which informs us as to how the basis vectors change as we move from point $x^\mu$ to point $x^\mu+u^\mu\,d\tau$ in the CST. Inserting this into the full expression we have $\nabla_\mu \bs{v} = (\partial_\mu v^\nu)\,\textrm{\bf e}_\nu + {\Gamma}_{\mu\nu}^\lambda\textrm{\bf e}_\lambda$ which we can relabel dummy indices via $\lambda\leftrightarrow\nu$ to obtain \be{cov:deriv} \nabla_\mu \bs{v} = \left(\partial_\mu v^\nu + {\Gamma}_{\mu\lambda}^\nu v^{\lambda} \right) \textrm{\bf e}_\nu \overset{\textrm{def}}{=} v^\nu_{;\mu}\,\textrm{\bf e}_\nu, \end{equation} where the last expression defines the \tit{covariant derivative} (denoted conventionally by a semicolon vs a comma apropos for an ordinary coordinate derivative) of the components $\nabla_\mu v^\nu \equiv v^\nu_{;\mu} = \partial_\mu v^\nu + {\Gamma}_{\mu\lambda}^\nu v^{\lambda}$. The values of ${\Gamma}_{\mu\lambda}^\nu$ are tied down (see \Eq{connection:coord}) by invoking the \tit{constancy of the metric} condition $\nabla_\mu(g_{\alpha\beta})=0$, (which reduces to the identity $\partial_\mu(\eta_{\alpha\beta})=0$ in the observer's local laboratory). The covariant derivative is of fundamental importance in GR since the commutator of the covariant derivatives acting on a vector is proportional to the Riemann curvature tensor, from which Einstein's fundamental equations are derived: $[\nabla_\mu , \nabla_\nu]\,v^\alpha= v^\alpha_{;\mu;\nu} -v^\alpha_{;\nu;\mu} = -R_{\beta\mu\nu}^\alpha v^\beta$. Note: the analogy in E\&M is the potential $A^\mu$ which acts as a $U(1)$ (\tit{gauge}) potential, so that the covariant derivative is $\nabla_\mu = \partial_\mu - A_\mu(x)$ and the analogue of the curvature is the Faraday tensor (containing the electric and magnetic fields as components) such that $[\nabla_\mu,\nabla_\nu] = \partial_\mu A_\nu - \partial_\nu A_\mu = F_{\mu\nu}$ (see Ryder, section 11.1). The important point here is that the 4-potential $A^\mu(x)$ is spacetime dependent, and that changes in $A^\mu(x)$ (called \tit{gauge transformations}) do not change the physical field $F_{\mu\nu}(x)$. This is called ``gauging" the E\&M field. \subsection{The DE in CST: the spinor covariant derivative in GR}\label{sec:DE:CST:spinor:covar:deriv:GR The new question to ask is: ``How does one gauge gravity?" This is the subject of Ryder\cite{Ryder:2009}, Chapter 11.3 (This chapter is titled ``Gauging Lorentz symmetry: torsion"). The main point is that up to now we've been discussing the covariant derivative for vectors (and tensors) which are associated with integer values of spin ($j = \{0, 1, 2,\ldots\}$, with $2 j+1$ components, i.e. scalars, vectors (e.g. photons), 2-tensor, (e.g. gravitons), etc\ldots). But how does one define the covariant derivative for half-integer spin objects, specifically spin $1/2$ with 2 components? Recall that for the SR the total angular momentum is given by matrix $J_{\mu\nu} = -i\,(x_\mu\,\partial_\nu - x_\nu\,\partial_\mu)\,\mathbb{I} + \Sigma_{\mu\nu}$, where $\Sigma_{\mu\nu}=\frac{i}{4}\,[{\gamma}_\mu, {\gamma}_\nu]$, and ${\gamma}_\mu = \eta_{\mu\nu}\,{\gamma}^\nu$ are the constant Dirac matrices of \Eq{DE:gamma:matrices} (since we are operating in the observer's local SR tangent plane/laboraotry). As in SR, the underlying symmetry of GR (at least locally) is the Poincare group, which is the 10 parameter group of (3) rotations, (3) boosts, and (4) spacetime translations. These matrix operators satisfy as set of (involved) commutation relations $[J_{\mu\nu},J_{\alpha\beta}] = f_{\mu\nu,\alpha\beta}^{\rho\sigma}\,J_{\rho\sigma}$ (which reproduces the usual commutation relations for rotations if we set $\Sigma_{\mu\nu}\to 0$). The Poincare group admits both integer (vector, tensor) representation, as well as spinor (half-integer) representations. These representations are derived by consider small changes in the quantity under study. Following Ryder, Chapter 11.3, Herman Weyl proposed the following ansatz: the ($N$-dimensional) spinor $\psi$ transforms like a \tit{scalar} with respect to the ``world" transformations (i.e. with respect to the coordinate index $\mu$), but as a spinor \tit{with respect to local Lorentz transformations (LLT) in the local laboratory} (i.e. with respect to the index $a$ in the \tit{tangent space at $x$ in the CST where the observer's local laboratory instantaneously exists}). These LLT transform the observer's instantaneous state of motion (in the flat Minkowski tangent plane) at $x$ from one type of motion to another, e.g. from a stationary observer at $x$, to an instantaneous freely falling observer at $x$, or to say an observer executing circular motion instantaneously at $x$, or to any kind of instantaneous motion. Thus, under infinitesimal changes in the surrounding CST small changes in the spinor $\delta\psi$ are given as $\delta\psi = -\xi^\mu\,\partial_\mu\psi$, i.e. with respect to the ordinary coordinate derivative $\partial_\mu$ apropos for a (world) scalar field. Here, $\xi^\mu = \omega^\mu_\nu x^\mu$ where for now, we consider $\xi^\mu$ as constants (since we are acting \tit{within} a given tangent plane at $x$). As an example, for an rotation infinitesimal in the $x-y$ plane by angle $\phi$ given by $\tiny{ \left(\begin{array}{c}x'\\ y'\end{array}\right) = R(\phi) \left(\begin{array}{c}x\\ y\end{array}\right) }$ (suppressing the $t$ and $z$ components for now, i.e. this should be embedded in a $4\times 4$ matrix) where $\tiny{ R(\phi) = \left(\begin{array}{cc}\cos\phi & \sin\phi \\-\sin\phi & \cos\phi\end{array}\right) }$ then $\omega^\mu_\nu$ is the anti-symmetric matrix given by $\tiny{ \frac{d R(\phi)}{d\phi}|_{\phi=0} = \left(\begin{array}{cc}0 & 1 \\-1 & 0\end{array}\right). }$ However, for small changes in the spin, in the local observer's frame, the spinor changes are given by $\delta\psi = -i\,\frac{1}{2}\,\omega^{ab}\,\Sigma_{ab}\,\psi$. Where $\omega^{ab}$ are some constants describing the LLT (as in SR). Thus, for infinitesimal changes, the total change in the spinor is just the sum (to first order) of the two changes (variations) give by $\delta\psi = -\xi^\mu\,\partial_\mu\,\psi -i\,\frac{1}{2}\,\omega^{ab}\,\Sigma_{ab}\,\psi$. We now ``gauge" this transformation by allowing both $\xi^\mu(x)$ and $\omega^{ab}(x)$ to be spacetime dependent (i.e. LLT varying at each point $x$) in order to develop a spinor covariant derivative. The derivation is not that hard but somewhat lengthy (detailed in Ryder, Chapter 11.3) yielding a form $\psi_{|\mu} = \partial_\mu\,\psi + \frac{1}{2}\,A^{ab}_\mu\,\Sigma_{ab}\,\psi$ such that the changes in the spinor $\delta\psi$ under LLT transform the same as $\psi$ itself, namely, $\delta(\psi_{|\mu}) = \frac{1}{2}\omega^{ab}\,\Sigma_{ab}\,(\psi_{|\mu})$. (Note: the spinor covariant derivative is denoted by $\psi_{|\mu}$ to distinguish it from the covariant derivative acting on vectors and tensors, e.g. $v_{;\mu}$). Most significantly, the commutator of the spinor covariant derivatives acting on $\psi$ are once again proportional to (a spinor version of) the Riemann tensor (having both mixed tangent plane (Latin), and world (Greek) indices), namely $\psi_{|\mu|\nu} - \psi_{|\nu|\mu} = -\frac{1}{2} R^{ab}_{\mu\nu}\,\Sigma_{ab}\,\psi$. With these preliminaries under one's belt, one can then derive the DE in CST, as detailed in Ryder, Chapter 11.4, pp416-418. The derivation is quite elegant, but here we indicate only the highlights. We begin with the flat Minkowski SR DE (restoring factors of $\hbar$ and $c$) $i\,\hbar\,{\gamma}^\mu\partial_\mu\,\psi = - m\,c\, \psi$ (note the sign change on the mass, due to the local GR (e.g. Hartle, Ryder) metric, vs the particle physicist's (e.g. Greiner) Minkowski metric), noting that the Dirac matrices are the \tit{constant} ones discussed earlier for the SR DE \Eq{DE:gamma:matrices} (since we are in the observer's local laboratory, i.e. the instantaneous (locally flat, Minkowski) tangent space to the CST at the point $x$). The net result of the gauging of gravity (really, gauging Lorentz symmetry) is that the covariant derivative for spinors boils down to \be{Ryder:p416:128:cov:deriv} \partial_\mu \to D_\mu \overset{\textrm{def}}{=} \partial_\mu - \frac{i}{2}\,{\Gamma}_{\alpha\beta\mu}\,\Sigma^{\alpha\beta} = \partial_\mu + \frac{1}{8}\,{\Gamma}_{\alpha\beta\mu}\,[ {\gamma}^\alpha, {\gamma}^\beta]. \end{equation} Here, ${\Gamma}_{\alpha\beta\mu} = g_{\alpha\rho}\,{\Gamma}^{\rho}_{\beta\mu}$ where for coordinate basis vectors ${\Gamma}^{\rho}_{\beta\mu}$ are the usual Levi-Civita connenction given in terms of the metric by \be{connection:coord} \trm{coordinate basis:}\quad {\Gamma}^{\mu}_{\alpha\beta} = \frac{1}{2}\,g^{\mu\lambda}\,(\partial_\alpha\,g_{\lambda\beta} + \partial_\beta\,g_{\lambda\alpha}-\partial_\lambda\,g_{\alpha\beta}). \end{equation} The ${\gamma}^\alpha$ appearing in \Eq{Ryder:p416:128:cov:deriv} are the \tit{constant} Dirac gamma matrices given prior in \Eq{DE:gamma:matrices}. Thus, we penultimately arrive at our desired goal, the DE in CST \be{Ryder:p416:128} i\,\hbar\,{\gamma}^\mu\,D_\mu\psi = i\,\hbar\,\,{\gamma}^\mu\, \left(\partial_\mu + \frac{1}{8}\,{\Gamma}_{\alpha\beta\mu}\,[ {\gamma}^\alpha, {\gamma}^\beta]\right)\,\psi = m\,c\,\psi. \end{equation} We now introduce one more complication, namely, the translation of the above DE written in a coordinate basis, to a physical orthonormal set of basis vectors. This entails having an expression for the connection ${\Gamma}_{\alpha\beta\mu}$ in an orthonormal basis. While straightforward, yet somewhat lengthy to derive (see Ryder, Chapter 3.13, pp107-110, Eq(3.259)) the results are a pleasing generalization, denoted by ${\Gamma}_{abc}$, of the coordinate-based Levi-Civita connection ${\Gamma}_{\mu\nu\lambda}$, given by \bea{Ryder:p108:3.259} \trm{orthonormal basis:}\quad {\Gamma}_{abc}&=& -\frac{1}{2}\, ( C_{abc} + C_{bca} - C_{cab} ) \\ C_{abc} &=& \eta_{ad}\,C^{d}_{\;\;bc}, \quad\trm{where}\quad [\textrm{\bf e}_a,\textrm{\bf e}_b] = \textrm{\bf e}_c\,C^{c}_{\;\;bc}. \end{eqnarray} (Note: many GR books, including Ryder (but not Hartle), use Greek indices on all basis vectors, coordinate and orthonormal, and the metric as well. One just has to be conscious of the context of the specific calculation to discern if the indices indicate global or local basis/metric, and hence which formula to utilize for the connection, \Eq{connection:coord} or \Eq{Ryder:p108:3.259}). Finally, the DE in CST in an orthonormal basis is given by \be{Ryder:p417:11.129} i\,\hbar\,\gamma^a (e_a + {\Gamma}_a)\,\psi = m\,c\,\psi, \qquad {\Gamma}_a\overset{\textrm{def}}{=}\frac{1}{8}\,{\Gamma}_{abc}[{\gamma}^b,{\gamma}^c]. \end{equation} Here, $e_a(\psi(x)) = e_a^\mu(x)\partial_\mu\psi(x)$ is the action of the orthonormal basis vector acting on the spinor (or any object), and we write $``{\gamma}^a\equiv{\gamma}^\mu"$ by which we mean (abusing notation) that the ${\gamma}^a$ are numerically the \tit{same} constant Dirac gamma matrices as ${\gamma}^\mu$ (in SR, see \Eq{DE:gamma:matrices}). For the Schwarzschild metric, one can straightforwardly work out the commutators of the orthonormal basis vectors (see Ryder, pp418-419, and \App{app:commutators:SST}) which are defined by (restoring again the boldface vector notation) \be{Ryder:p417:11.131} \textrm{\bf e}_0 = \frac{1}{c}\,\left(1- \frac{2 M_s}{r}\right)^{-1/2}\,\frac{\partial}{\partial t},\;\; \textrm{\bf e}_1 = \left(1- \frac{2 M_s}{r}\right)^{1/2}\,\frac{\partial}{\partial r},\;\; \textrm{\bf e}_2 = \frac{1}{r}\,\frac{\partial}{\partial \theta},\;\; \textrm{\bf e}_3 = \frac{1}{r \sin\theta}\,\frac{\partial}{\partial \phi} \end{equation} where $M_s\overset{\textrm{def}}{=} G M/c^2 = \frac{1}{2}\,r_s$, to finally arrive at our desired destination (see Ryder, Eq(11.139), p418) \bea{Ryder:p418:11.139} &{}& i\,\hbar\, \left\{ \left(1- \frac{2 M_s}{r}\right)^{-1/2}\, \left[ {\gamma}^0\frac{1}{c}\frac{\partial\psi}{\partial t} - \frac{M_s}{2 r^2}{\gamma}^1\psi \right] + \left(1- \frac{2 M_s}{r}\right)^{1/2}{\gamma}^1\,\frac{\partial \psi}{\partial r} \right. \nonumber \\ &{}& \left. \hspace{0.20in} +\; {\gamma}^2\,\frac{1}{r}\frac{\partial \psi}{\partial \theta} + \frac{M_s}{r^2}\,\left(1- \frac{2 M_s}{r}\right)^{1/2}{\gamma}^1\,\psi + {\gamma}^3\,\frac{1}{r \sin\theta}\frac{\partial \psi}{\partial \phi} + \frac{\cot\theta}{2 r}\,{\gamma}^2\,\psi \right\} = m\,c\,\psi, \label{DE:CST} \end{eqnarray} of a Dirac spin-1/2 particle of mass $m$ in the Schwarzschild spacetime. One thing we are struck by right away is that nowhere in the above derivation has use been made of the classical GR conserved quantities $e$ and $\ell$ of \Eq{Hartle:p193:9.21} and \Eq{Hartle:p193:9.22}, respectively. However, upon further reflection, for a QM derivation this makes sense, since these constants of the motion involve $dt(\tau)/d\tau$ and $d\phi(\tau)/d\tau$ implying the classical notion of trajectories in spacetime, for which QM abandons, and replaces with the concept of stationary eigenstates over all space. Thus, even though tempting, it would make not make sense to replace quantities such as $\frac{\partial \psi}{\partial t}$ by $\left(\frac{d t}{d\tau}\right)\,\frac{\partial \psi}{\partial \tau}\to e\,(1-\frac{2 M_s}{r(\tau)})^{-1}\,\frac{\partial \psi(\tau)}{\partial \tau}$, etc. since the the wave function $\psi$ would then be solely a function of $\tau$, and the QM concept of spatial eigenstates would not be possible. \section{Bound states of the DE in SST}\label{sec:bound:states:DE:SST We are now finally able to attempt to answer the student's original posed question: ``Does there exist bound states of the DE in SST, analogous to the bound states of the SR DE for either vector or scalar coupling?" Before we can answer this, it is helpful to massage \Eq{Ryder:p418:11.139} into a much more amenable, dimensionless form (especially for numerical calculations). By using the SR Dirac ${\gamma}^a$ matrices \Eq{DE:gamma:matrices} and multiplying through by ${\sqrt{\mathcal{F}}} \overset{\textrm{def}}{=} \sqrt{1-2 M_s/r}$, and again letting $\psi = \tiny{\left(\begin{array}{c}\varphi \\ \chi\end{array}\right)}$, \Eq{DE:CST} can be rearranged into the form \bea{DE:CST:pma:p5.2} &{}&\left( i\,\hbar\,\frac{\partial}{\partial t} - m\,c^2\,{\sqrt{\mathcal{F}}}\right)\,\varphi = c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\chi, \label{DE:CST:pma:p5.2:line1} \\ &{}&\left( i\,\hbar\,\frac{\partial}{\partial t} + m\,c^2\,{\sqrt{\mathcal{F}}}\right)\,\chi = c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\varphi,\label{DE:CST:pma:p5.2:line2} \\ \trm{with}\quad \hat{\textrm{\textbf{p}}} &=& -i\,\hbar\, \left( {\mathcal{F}}\,\frac{\partial}{\partial r} + {\mathcal{F}}\, \frac{M_s}{r^2} - \frac{M_s}{2\,r^2},\; \frac{{\sqrt{\mathcal{F}}}}{r}\left(\frac{\partial}{\partial \theta} + \frac{1}{2}\,\cot\theta \right),\; {\sqrt{\mathcal{F}}}\,\frac{1}{r \sin\theta}\frac{\partial}{\partial \phi} \right). \label{DE:CST:pma:p5.2:line3} \end{eqnarray} The first thing we note is that in going from SR to GR the rest mass goes from $m\,c^2\to m\,c^2\,{\sqrt{\mathcal{F}}}$, acting as a variable mass, that is the ordinary rest mass $m c^2$ at $r\to\infty$ and goes to zero at the Schwarzschild horizon $r\to 2 M_s$. This is one of the new features introduced by GR, the role of the horizon. Secondly, we see that \Eq{DE:CST:pma:p5.2:line1} and \Eq{DE:CST:pma:p5.2:line2} has the form of a SR free field DE, but with the radial potential terms buried within $\hat{\textrm{\textbf{p}}}$, especially, $\hat{p}_r$ in \Eq{DE:CST:pma:p5.2:line3}. If we now postulate the existence of stationary states, with each spinor having an $e^{-i\,E\,t/\hbar}$ temporal dependence, with E constant, we then have \bea{DE:pma:p5.2:middle} c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\chi &=& (E - m\,c^2\,{\sqrt{\mathcal{F}}})\,\varphi, \\ c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\varphi &=& (E + m\,c^2\,{\sqrt{\mathcal{F}}})\,\chi, \end{eqnarray} in strong analogy with the SR DEs \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2}, but now with no explicit vector or scalar coupling potential terms, $V_v(r)$ and $V_s(r)$ on the righthand side, and now additionally with a variable mass $m c^2\,{\sqrt{\mathcal{F}}}$. Note that for $r\gg r_s=2 M_s$ we have $E + m\,c^2\,{\sqrt{\mathcal{F}}}\approx E \pm m\,c^2\,(1-G M/ (r c^2)) = E \pm (mc^2 + V_G(r))$ where $V_G(r) = -G M m/r$ the Newtonian potential. Hence, we observe scalar (mass) coupling far from the horizon, as our intuition would expect. We now write the above equations in dimensionless form using natural units. Defining the operator $\hat{\textrm{\textbf{q}}}$ via $\hat{\textrm{\textbf{p}}} = -i\,\hbar\,\hat{\textrm{\textbf{q}}}$, and then dividing through by $\hbar c$ and recalling $\lambda_C = \frac{\hbar}{m c}$, we will define lengths as $r = \lambda_C\,\rho$, and energies in terms of the rest mass via $\epsilon = \frac{E}{m c^2}$. We then obtain \bea{DE:pma:p5.3:top} \hspace{-0.25in} &{}& (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{q}}} )\,\chi = i\,(\epsilon - {\sqrt{\mathcal{F}}})\,\varphi, \label{DE:pma:p5.3:top:line1} \\ \hspace{-0.25in} &{}& (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{q}}} )\,\varphi = i\,(\epsilon + \,{\sqrt{\mathcal{F}}})\,\chi, \label{DE:pma:p5.3:top:line2} \\ \hspace{-0.25in} \hat{\textrm{\textbf{q}}} &=& \left( {\mathcal{F}}\,\frac{\partial}{\partial \rho} + {\mathcal{F}}\, \frac{m_s}{\rho^2} - \frac{m_s}{2 \rho^2},\; \frac{{\sqrt{\mathcal{F}}}}{\rho}\left(\frac{\partial}{\partial \theta} + \frac{1}{2}\,\cot\theta \right),\; {\sqrt{\mathcal{F}}}\, \frac{1}{\rho \sin\theta}\frac{\partial}{\partial \phi} \right), \;\; m_s \overset{\textrm{def}}{=} \frac{M_s}{\lambda_C} = \frac{\frac{1}{2} r_s}{\lambda_C}, \qquad \label{DE:pma:p5.3:top:line3} \\ % &\overset{\textrm{def}}{=}& \left( {\hat{\mathcal{Q}}}_1, {\hat{\mathcal{Q}}}_2, {\hat{\mathcal{Q}}}_3 \label{DE:pma:p5.3:top:line4} \right) \end{eqnarray} Since our primary goal to find the simplest possible solution, not the most general solution, we will cut to the chase and look for only radial solutions (avoiding all the complications of the orbital and spin angular momentum that arose in the SR DE) and define \bea{PMA:ansatz} \varphi &=& g(\rho)\,\tilde{\varphi}, \\ \chi &=& i\,f(\rho)\,\tilde{\chi}, \end{eqnarray} where $\tilde{\varphi}$ and $\tilde{\chi}$ are constant 2-spinors. (Recall in the SR DE, if $\ell = 0\Rightarrow m=0$ which implies $Y_{\ell=0,m=0}(\theta,\phi) = \frac{1}{\sqrt{4 \pi}}$ is a constant). Let us now, with foresight, chose $\varphi = \tiny{\left(\begin{array}{c} a \\ b\end{array}\right)}$ and $\chi = \tiny{\left(\begin{array}{c} a \\ -b\end{array}\right)}$ with $a, b$ constant such that $|a|^2 + |b|^2=1$. We orient our axes so that $r$ and hence $\rho$ are along the $\hat{z}$ direction so that $\hat{\sigma}_1 = \hat{\sigma}_z = \tiny{\left(\begin{array}{cc} 1 & 0 \\ 0 &-1\end{array}\right)}$. Then, substituting these definitions into \Eq{DE:pma:p5.3:top:line1} and \Eq{DE:pma:p5.3:top:line2} we have \bea{DE:pma:p5.3:bottom} \hspace{-0.65in} \left(\begin{array}{cc} {\hat{\mathcal{Q}}}_1 & 0 \\ 0 &-{\hat{\mathcal{Q}}}_1\end{array}\right)\,(i\,f)\, \left(\begin{array}{c} a \\ -b\end{array}\right) = i\,(\epsilon -{\sqrt{\mathcal{F}}})\,g\, \left(\begin{array}{c} a \\ b\end{array}\right) &\Rightarrow& \begin{cases} a\,{\hat{\mathcal{Q}}}_1\,f = a (\epsilon -{\sqrt{\mathcal{F}}})\,g, \\ b\,{\hat{\mathcal{Q}}}_1\,f = b (\epsilon -{\sqrt{\mathcal{F}}})\,g, \end{cases} \Rightarrow {\hat{\mathcal{Q}}}_1\,f = (\epsilon -{\sqrt{\mathcal{F}}})\,g, \;\;\qquad \\ \hspace{-0.65in} \left(\begin{array}{cc} {\hat{\mathcal{Q}}}_1 & 0 \\ 0 &-{\hat{\mathcal{Q}}}_1\end{array}\right)\,g\, \left(\begin{array}{c} a \\ b\end{array}\right) = i\,(\epsilon +{\sqrt{\mathcal{F}}})\,(i\,f)\, \left(\begin{array}{c} a \\ -b\end{array}\right) &\Rightarrow& \begin{cases} a\,{\hat{\mathcal{Q}}}_1\,g = -a (\epsilon +{\sqrt{\mathcal{F}}})\,f, \\ -b\,{\hat{\mathcal{Q}}}_1\,g = b (\epsilon +{\sqrt{\mathcal{F}}})\,f, \end{cases} \hspace{-0.5em} \Rightarrow {\hat{\mathcal{Q}}}_1\,g = -(\epsilon+{\sqrt{\mathcal{F}}})\,f. \;\;\qquad \end{eqnarray} Thus, for this choice of constant $\tilde{\varphi}$ and $\tilde{\chi}$, each spinor produces two equations, which are self-consistently the same. Therefore, our final dimensionless radial DE in CST equations are \bea{DE:CST:p5.3:final} {\hat{\mathcal{Q}}}_1(\rho)\,f(\rho) &=& \left(\epsilon -\sqrt{\mathcal{F}(\rho)}\right)\,g(\rho), \label{DE:CST:p5.3:final:line1} \\ {\hat{\mathcal{Q}}}_1(\rho)\,g(\rho) &=& -\left(\epsilon+\sqrt{\mathcal{F}(\rho)}\right)\,f(\rho), \label{DE:CST:p5.3:final:line2} \\ \trm{with}\quad \mathcal{F}(\rho) = 1 - \frac{2 m_s}{\rho},\quad {\hat{\mathcal{Q}}}_1(\rho)&=&{\mathcal{F}}\,\frac{\partial}{\partial \rho} + {\mathcal{F}}\, \frac{m_s}{\rho^2} - \frac{m_s}{2\,\rho^2}, \;\; \rho = r/\lambda_C, \;\; m_s = \frac{\frac{1}{2} r_s}{\lambda_C}. \quad \label{DE:CST:p5.3:final:line3} \end{eqnarray} For a 2-spinor in the standard form $\tiny{\left(\begin{array}{c} \cos\theta'/2 \\ \sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ we can associate the point $\hat{n} = (\sin\theta'\,\cos\phi',$ $\sin\theta'\,\sin\phi',\cos\theta')$ with polar and azimuthal angles $(\theta', \phi'$) on an ordinary 2-sphere $S^2\in\mathbb{R}^3$, called the \tit{Bloch} sphere. (While we have oriented the spin $\hat{z}$-axis of the Bloch sphere with the world radial coordinate $r$ direction, the internal spinor space angles $(\theta', \phi')$ should not be conflated with the Schwarzschild spacetime coordinates $(\theta, \phi)$). Spin up corresponds to $\theta'=0$ oriented along $\textrm{\bf e}_r$, and spin down with $\theta'=\pi$ oriented along $-\textrm{\bf e}_r$ (with $\phi'$ indeterminate at the poles, so it can be taken to be zero there). The spinor solutions we have used above are then of form, $\tilde{\varphi} = \tiny{\left(\begin{array}{c} \cos\theta'/2 \\ \sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ and $\tilde{\chi} = \tiny{\left(\begin{array}{c} \cos\theta'/2 \\ -\sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ with overlap $\IP{\tilde{\varphi}}{\tilde{\chi}} = \cos\theta'$, and thus are orthogonal at the Bloch sphere equator $\theta'=\pi/2$. For other possible choices of the spinors leading to a modified form of the radial equations \Eq{DE:CST:p5.3:final:line1} and \Eq{DE:CST:p5.3:final:line2}, see \App{app:other:spinor:solns}. \subsection{Numerical solutions of the DE in SST}\label{subsec:Num:Solns:DE:SST} The square root factor ${\sqrt{\mathcal{F}}}$ in \Eq{DE:CST:p5.3:final:line1} and \Eq{DE:CST:p5.3:final:line2} presents difficulties for the standard power series in $\rho$ solutions of these equations, as was employed in the NRQM SE and SR DE. Hence, here we make a \tit{conformal transformation} in order to map $\rho=\infty$ to a finite value, via the definition $\sin\Theta\overset{\textrm{def}}{=} \sqrt{1 - 2 m_s/\rho}$ such that $0\le\Theta\le\pi/2\leftrightarrow 2 m_s\le\rho\le\infty$. Therefore, $\rho = \frac{2 m_s}{\cos^2\Theta}$ and $\frac{\partial }{\partial \rho} = \frac{1}{\partial \rho/\partial \Theta}\,\frac{\partial }{\partial \Theta} = \frac{\cos^3\Theta}{4 m_s \sin\Theta}\frac{\partial}{\partial \Theta}$ hence our DE in SST in a more numerically amenable form is given by \bea{DE:CST:Qhat:Theta} {\hat{\mathcal{Q}}}_1(\Theta)\,f(\Theta) &=& \left(\epsilon -\sin\Theta\right)\,g(\Theta), \label{DE:CST:Qhat:Theta:line1} \\ {\hat{\mathcal{Q}}}_1(\Theta)\,g(\Theta) &=& -\left(\epsilon+\sin\Theta\right)\,f(\Theta), \label{DE:CST:Qhat:Theta:line2} \\ {\hat{\mathcal{Q}}}_1(\rho)\to {\hat{\mathcal{Q}}}_1(\Theta)&=& \frac{\cos^4\Theta}{4 m_s }\, \left( \tan\Theta\,\frac{\partial}{\partial \Theta} + \sin^2\Theta -\frac{1}{2} \right), \\ \label{DE:CST:Qhat:Theta:line3} \trm{where}\quad 0\le\Theta\le\pi/2 &\leftrightarrow& 2\,m_s\le\rho\le\infty \leftrightarrow 2\,M_s\le r\le\infty, \label{DE:CST:Qhat:Theta:line4} \end{eqnarray} for the coordinate region \tit{outside} the Schwarzschild horizon. Note that the dimensionless constant $m_s=\frac{\frac{1}{2} r_s}{\lambda_C}$ \tit{cannot} be absorbed into either $f$ or $g$ nor into the the definition of $\Theta$ itself (even if we were to have defined $\tilde{\rho} = \rho/2m_s$ with ${\mathcal{F}}\to 1-\frac{1}{\tilde{\rho}}$), and thus sets the scale for the problem. As shown in \App{app:Units}, for a solar mass BH, $r_s=1.48$ km $= 1.48\times 10^3$m and $\lambda_C = 2.246\times 10^{-12}$m, so that $m_s\sim 10^{15}$, an astronomically huge number. Nonetheless, with the courage of our intuition, but with warranted trepidation, we now seek numerical solutions of \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} for various values of $1\le m_s\le 10^{15}$. A few other forms of the above equations are worthwhile to obtain a feel for part of the structure of the radial wave functions $f(\Theta)$ and $g(\Theta)$. By using an integrating factor, we could simplify the above equations with the substitutions \bea{DE:CST:Qhat:F:G} f(\Theta)= \sqrt{\sin\Theta}\,e^{-\frac{1}{2}\,\sin^2\Theta}\,F(\Theta) &\;\;\Rightarrow\;\;& \frac{\sin\Theta\cos^3\Theta}{4 m_s }\,\frac{\partial F(\Theta)}{\partial \Theta} = \left(\epsilon -\sin\Theta\right)\,G(\Theta), \\\label{DE:CST:Qhat:F:G:line1} g(\Theta)= \sqrt{\sin\Theta}\,e^{-\frac{1}{2}\,\sin^2\Theta}\,G(\Theta) &\;\;\Rightarrow\;\;& \frac{\sin\Theta\cos^3\Theta}{4 m_s }\,\frac{\partial G(\Theta)}{\partial \Theta} = -\left(\epsilon +\sin\Theta\right)\,F(\Theta). \label{DE:CST:Qhat:F:G:line2} \end{eqnarray} Lastly one could define a coordinate $x$ to the remove the prefactors in front of the above derivatives to yield \bea{DE:CST:Qhat:F:G:x} \frac{1}{4 m_s }\,\frac{\partial F(x)}{\partial x} &=& \big(\epsilon -\sin\Theta(x)\big)\,G(x), \label{DE:CST:Qhat:F:G:x:line1} \\ \frac{1}{4 m_s }\,\frac{\partial G(x)}{\partial x} &=& -\big(\epsilon +\sin\Theta(x)\big)\,G(x), \label{DE:CST:Qhat:F:G:x:line2}\\ x(\Theta) = \int\frac{d \Theta}{\sin\Theta\cos^3\Theta} &=& \ln(\tan\theta) + \frac{1}{2\,\cos^2\theta}\equiv h(\Theta) \;\Rightarrow\; \Theta(x) \overset{\textrm{def}}{=} h^{-1}(x), \label{DE:CST:Qhat:F:G:x:line3} \end{eqnarray} where $\Theta=\Theta(x)$ is now the inverse function of $x(\Theta)$ in \Eq{DE:CST:Qhat:F:G:x:line3}. For numerical solutions we have found that all three of these formulations produce the same eigenvalues (as one would expect), and thus, in the following we will work directly with \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2}. The numerical procedure to find both the eigenvalues $\epsilon$ and eigenfunctions $f(\Theta)$ and $g(\Theta)$ \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} is essentially a \tit{shooting method}. As in the SR DE we require the boundary conditions based on physical arguments, that $f(0) = f(\pi/2)=0$ and $g(0) = g(\pi/2)=0$. Since we are looking for bound states, we search for eigenvalues in the range $-1\le\epsilon\le 1$. Thus, by choosing a value of $\epsilon$ near zero and integrating inwards from $\Theta=\pi/2$ $\Theta=0$ to one obtains solutions for $f(0)$ and $g(0)$, which are in general non-zero. One then adjusts the value of $\epsilon$ (essentially performing a line search in $\epsilon$) and re-integrates again from $\Theta=\pi/2$ to $\Theta=0$ until the desired boundary conditions on $f$ and $g$ are met at $\Theta=0$. In this fashion one can find the lowest eigenvalue of $\epsilon_1$ that yields both $f(0)=0$ and $g(0)=0$. One then repeats this procedure by searching for $\epsilon_2$ in the range $|\epsilon_1|\le\epsilon_2\le 1$ and $-1\le\epsilon_2\le-|\epsilon_1|$, and similarly for larger eigenvalues. \tit{Mathematica}\cite{Mathematica} has such a routine called \ttt{NDEigensystem} which performs this numerical procedure for coupled differential operators $\mathcal{L}_1\big(u_i(x,y,\ldots), v_i(x,y,\ldots),\ldots\big) = \lambda_i\,u_i(x,y,\ldots)$, $\mathcal{L}_2\big(u_i(x,y,\ldots), v_i(x,y,\ldots),\ldots\big) = \lambda_i\,v_i(x,y,\ldots)$, \ldots, returning the eigenvalue and eigenfunction pair $\{\lambda_i, \{u_i, v_i\}\}$ for $i=\{1,2,\ldots,78\}$ (the maximumn number of eigenvalues that \ttt{NDEigensystem} can return). We therefore write our coupled pair of radial wave equations \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} as the coupled eigenvalue equations \bea{FG:eval:oprs} \mathcal{L}_1(f,g) &=& \hspace{1.5em}{\hat{\mathcal{Q}}}(\Theta) f(\Theta) + \sin\Theta\,g(\Theta) \quad\Rightarrow\quad \mathcal{L}_1(f,g) = \epsilon\,g(\Theta), \label{FG:eval:oprs:line1} \\ \mathcal{L}_2(f,g) &=& -\left( {\hat{\mathcal{Q}}}(\Theta) g(\Theta) + \sin\Theta\,f(\Theta) \right) \;\Rightarrow\quad \mathcal{L}_2(f,g) = \epsilon\,f(\Theta). \label{FG:eval:oprs:line2} \end{eqnarray} For now we simply treat $m_s$ as a variable parameter and plot the eigenvalues from \ttt{NDEigensystem} for $m_s\in\{5, 10, 10^2, 10^3, 10^{15}\}$ as shown in \Fig{fig:DE:SST:evals:ms:5:10:100:1000:1e15}. \begin{figure}[h] \includegraphics[width=7.0in,height=2.5in]{fig_2_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s\in\{5, 10, 10^2, 10^3, 10^{15}\}$, from outside to inside, with the dashed black lines at $\pm 1$ as a guide lines for the boundary of the bound state energy region. }\label{fig:DE:SST:evals:ms:5:10:100:1000:1e15} \end{figure} The dashed black lines at $\pm 1$ denote the upper bounds for the bound state energy region for $|\epsilon_k = E_k/m\,c^2|<1$. The gray-dashed curve (working outside to inside) for $m_s=5$ shows missing values, since the these eigenvalues had imaginary components. But as the curves for $m_s=10$ (gray-dotted), $m_s=100$ (black-dashed), $m_s=1000$ (overlapping black-solid) show, the eigenvalues quickly settle down to the black-solid curve for $m_s>10^2-10^3$. In fact, this latter curve is also valid when we used $m_s=10^{15}$ appropriate for a solar mass BH. The value of the first and last five eigenvalues are found to be $\epsilon_k=\pm\,\{0.0379246, 0.0781011, 0.117483, 0.156412, 0.195074,\ldots,$ $0.980712, 0.987621, 0.992999, 0.996871, 0.999206\}$.\cite{SR:DE:Numerical:Note} \subsection{Adding numerical diffusion}\label{subsec:numerical:diffusion} \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} are \tit{convection dominated} equations, which \ttt{NDEigensystem} warns, and suggests that one add (artificial) numerical diffusion. In fact, an examination of the numerical wave function solutions reveal cusp-like behaviors in places, typical of such convection dominated problems. There is a standard numerical remedy to this problem (resulting from numerical integration instability) described in many computational physics books, but particularly lucidly (as are other numerous other topics treated similarly) in \tit{Numerical Recipes in C} \cite{NumRecinC:1992} (or in Fortran, or Fortran90, or C++, if you prefer), written by a group of numerical relativists, see Chapter 19.1, pp834-839), by Press, Teukolsky, Vetterling and Flannery. It entails adding to the righthand side of the following 1D example problem $\partial_t f(x) = v(x) \partial_x f(x)$, a second order term of the form $\alpha_d\,\partial^2_{xx}$. As discussed in Press \tit{et al}., this added diffusive term stabilizes the equation, when one performs von Neumann stability analysis, for the growth of small fluctuations (in this example, with $v(x)$ and $\alpha_d$ treated as constants). In our problem, the stability analysis is bit more complicated. As such, we have found that adding a term $\frac{\cos\Theta}{m_s}\, \partial^2_{\Theta\Theta} \,g$ to $\mathcal{L}_1$ (since it contains $\partial_{\Theta}\,g$), and similarly, a term $\frac{\cos\Theta}{m_s}\, \partial^2_{\Theta\Theta} \,f$ to $\mathcal{L}_2$ (since it contains $\partial_{\Theta} \,f$), smooths the wave functions $f(\Theta)$ and $g(\Theta)$ somewhat, without severely changing the eigenvalue spectrum. This is shown in \Fig{fig:DE:SST:evals:with:without:diffusion}. \begin{figure}[h] \includegraphics[width=5.5in,height=2.5in]{fig_3_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (black curve) without numerical diffusion, and (gray curve) with numerical diffusion. }\label{fig:DE:SST:evals:with:without:diffusion} \end{figure} The reason for the use of $\alpha_d = \cos\Theta/m_s$ is that the added diffusion term of the same order of magnitude $1/m_s$ as the coefficients in $\mathcal{L}_1$ and $\mathcal{L}_2$, and the function $\cos\Theta$ ``turns on" the diffusion as we proceed from $\Theta=\pi/2$ inwards to $\Theta=0$ where the spikiness of the wave functions without numerical diffusion are observed. The role of the diffusion term on the wave functions can be seen in \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} and \Fig{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4}, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. \begin{figure}[h] \begin{tabular}{cc} \includegraphics[width=3.0in,height=1.5in]{fig_4_top_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_4_top_right_DE_SST_AJP_26Jun2022} \\ \includegraphics[width=3.0in,height=1.5in]{fig_4_bottom_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_4_bottom_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Wave functions of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (left column) without numerical diffusion, and (right column) with numerical diffusion, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. }\label{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} \end{figure} \begin{figure}[h] \begin{tabular}{cc} \includegraphics[width=3.0in,height=1.5in]{fig_5_top_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_5_top_right_DE_SST_AJP_26Jun2022} \\ \includegraphics[width=3.0in,height=1.5in]{fig_5_bottom_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_5_bottom_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Wave functions of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (left column) without numerical diffusion, and (right column) with numerical diffusion, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. }\label{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4} \end{figure} \clearpage \newpage The left column are plots without numerical diffusion, while the right column are with numerical diffusion. While in \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2}(left) the plot of $f_1(\Theta)$ appears to be suppressed, it is actually non-zero, but so small that is does not register on the plot. In \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2}(right), with diffusion, $f_1(\Theta)$ is brought out (black curve) and the sharp cusps of $g_1(\Theta)$ on the left (numerical non-smoothness due to lack of diffusion) is smoothed out (as is $f_1(\Theta)$) on the right. The smoothing for wave functions for larger (magnitude) eigenvalues is still uneven in places as can be seen \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} and \Fig{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4}, but this resulted from our ``best guess" diffusion coefficient taken to be $\alpha_d = \cos\Theta/m_s$, with $m_s=25,000$. These plots indicate that a more sophisticated $\Theta$-dependent diffusion coefficient is most likely warranted. But for our purpose of exploring the student's intuition, our chose form is sufficient for illustration. There is a balancing act involved here since the choice of a $\Theta$-dependent diffusion coefficient can alter the values of the eigenvalues. For the choice of diffusion we employed, the altered eigenvalues were $\epsilon^{(\trm{diffusion})}_k=\pm\,\{0.0642606, 0.120916, 0.171706, 0.212431, 0.25394,\ldots,$ $0.981319, 0.988035, 0.993213, 0.996967, 0.999226\}$, for $m_s=25,000$, which are not too different from the eigenvalues obtained without diffusion, given by $\epsilon_k=\pm\,\{0.0379246, 0.0781011,$ $0.117483, 0.156412, 0.195074,\ldots,$ $0.980712, 0.987621, 0.992999, 0.996871, 0.999206\}$. We consider the latter eigenvalues without diffusion (with $m_s = 10^{15}$ apropos for a solar mass BH) as the true numerical eigenvalues. For a reasonable, desired degree of smoothness in the corresponding wave functions, we allow for the numerical diffusion we have chosen (after several ``numerical experiments") which does not significantly alter the eigenspectrum, as shown in \Fig{fig:DE:SST:evals:with:without:diffusion}. \subsection{Comparison of eigenvalues of DE in SST with SR DE and SE+SR} Lastly, in this section we compare the eigenvalues $\epsilon_k$ of the DE in SST (gray-curve: with diffusion; black curve, without diffusion) in \Fig{fig:DE:SST:evals:with:without:diffusion} against previous eigenvalue formulas. In \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(left) the outer cyan curve is the SR-DE for vector coupling \Eq{Greiner:p186:case:2} with $\alpha\to~\alpha_v=0.999281$ fitted to $E_1^{(vector)}/mc^2=\epsilon_1$ using $n=1$ and $j=\frac{1}{2}$. The inner purple curve is the the SR-DE for scalar coupling \Eq{Greiner:p186:case:1} with $\alpha'\to\alpha_s=26.25$ fitted to $E_1^{(scalar)}/mc^2=\epsilon_1$ using $n=1$ and $j=\frac{1}{2}$. As our intuition would suspect, the (purple) scalar coupling formula for the SR DE more closely approximates the DE in SST (black curve). \begin{figure}[h] \begin{tabular}{cc} \hspace{-0.75in} \includegraphics[width=4.0in,height=2.25in]{fig_6_left_DE_SST_AJP_26Jun2022} & \hspace{-0.1in} \includegraphics[width=4.0in,height=2.25in]{fig_6_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Eigenvalues of (Left:, top to bottom) (cyan curve) SR DE formula with vector coupling coefficient $\alpha_v=0.999281$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (purple curve) SR DE formula with scalar coupling coefficient $\alpha_s=26.25$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (gray curve) GR DE with diffusion $m_s=25,000$, (black curve) GR DE without diffusion. (Right: top to bottom) (purple curve) SE+SR scalar coupling formula with $E_I=0.998562$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (gray curve) GR DE with diffusion $m_s=25,000$, (black curve) GR DE without diffusion. (cyan curve) SE+SR vector coupling formula with $E_I=694.278$ fitted to $\epsilon_1$ of DE in SST (no diffusion), }\label{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling} \end{figure} In \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(right) we compare the DE in SST with the SE+SR formulas \Eq{E_s} and \Eq{E_v} which we had term SE ``scalar coupling" and SE ``vector coupling" respectively, since the position of the square root being either in the numerator or in the denominator, respectively, mimics that of the SR DE scalar and vector coupling formulas \Eq{Greiner:p186:case:1} and \Eq{Greiner:p186:case:2}, respectively. Here ``SE+SR" means that we ``add" $m c^2$ to the NRQM resulting SE energies $E_k$, and interpret the result as the firs order approximation to some unknown (from a NR POV) SR DE formula, involving square roots (either in the numerator or denominator). For the ``scalar coupling" case we fit $\epsilon_1 =\sqrt{1-\frac{E_I}{n^2}}$ for $n=1$ to find $E_I=0.998562$, while for the ``vector coupling" case we fit $\epsilon_1 =\frac{1}{\sqrt{1+\frac{E_I}{n^2}}}$ for $n=1$ to find $E_I=694.278$. The are plotted in \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(right) using the same colors as the (left) plot, namely, cyan for scalar coupling and purple for vector coupling. Counter to our intuition, now it is the SE ``vector coupling" formula that most closely mimics the eigenvalues of the DE in SST (black curve). This flipping of our intuitive interpretation of the SE+SR formulas \Eq{E_s} and \Eq{E_v} is most likely due to the fact that at the values of $E^{(scalar)}_I=0.998562$ and $E^{(vector)}_I=694.278$ a first order expansion of the square root (as we assumed by adding $m\,c^2$ to $E_I$) is not warranted at these large values of $E_I$. Nonetheless, it is curious that the roles of the SE ``scalar" and ``vector" coupling flip, and nearly approximate their SR DE opposites, vector and scalar coupling, respectively, as shown in \Fig{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All}. \begin{figure}[h] \includegraphics[width=5.0in,height=2.5in]{fig_7_DE_SST_AJP_26Jun2022} \caption{Composite of the left and right plots \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}. }\label{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All} \end{figure} \section{Solutions of the DE in SST inside the horizon?}\label{sec:DE:SST:inside:horizon} In the previous section, for solutions of the DE in SST we defined the angle $\Theta$ by $\sin\Theta~=~\sqrt{1-\frac{2 M_s}{r}}$, for $0\le\Theta\le\pi/2\leftrightarrow 2 M_s\le r \le \infty$, the region outside the horizon. Following our intuition, suppose we \tit{anlaytically continue} $\Theta\to i\,x$ so that $\sin\Theta\to i\,\sinh x$ and $\sqrt{1-\frac{2 M_s}{r}}\to i\,\sqrt{\frac{2 M_s}{r}-1}$, i.e. defining $\sinh x \overset{\textrm{def}}{=} \sqrt{\frac{2 M_s}{r}-1}$, for $0\le r\le 2 M_s \leftrightarrow \infty\ge x\ge 0$. To keep the resulting radial equations real (since the above procedure introduces explicit factor of $i$), we are led to consider redefining the ``energies" as $\epsilon\to i\,\eta$. Holding off on interpretation for now, and bolding (but not blindly) pushing forward, this leads to the following set of eigenvalue equations that we can once again use \ttt{NDEigensystems} to solve \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2}, now with $f(\Theta)\to F(x)$ and $g(\Theta)\to i\,G(x)$ % \bea{L1:L2:oprs:inside:horizon} \hspace{-0.25in} \mathcal{L}_1 &=& \frac{\cosh^4 x}{4 m_s} \left( -\tanh x \,\frac{\partial}{\partial x} + \sinh^2 x + \frac{1}{2} \right)\,G(x) - \sinh x\,F(x), \;\; \Rightarrow\; \mathcal{L}_1(F,G) = \eta\,F(x), \qquad \label{L1:opr:inside:horizon} \\ \hspace{-0.25in} \mathcal{L}_2 &=& -\frac{\cosh^4 x}{4 m_s} \left( \tanh x \,\frac{\partial}{\partial x} + \sinh^2 x + \frac{1}{2} \right)\,F(x) - \sinh x\,G(x), \;\; \Rightarrow\; \mathcal{L}_2(F,G) = \eta\,G(x). \qquad \label{L2:opr:inside:horizon} \end{eqnarray} These equations produce the ``eigenvalues" $\eta_k$ (black curve) in \Fig{fig:DE:SST:evals:inside:outside:horizon} with only negative values inside the horizon. Again, we show the eigenvalues (gray curve) $\epsilon_k$ outside the horizon, for comparison. \begin{figure}[h] \includegraphics[width=6.5in,height=2.5in]{fig_8_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of (black curve) DE in SST inside the horizon, \Eq{L1:opr:inside:horizon} and \Eq{L2:opr:inside:horizon}, \\ (gray curve) DE in SST outside the horizon, \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2}, both with $m_s=10^{15}$. } \label{fig:DE:SST:evals:inside:outside:horizon} \end{figure} We must now reconcile with imposing an interpretation on the results shown in \Fig{fig:DE:SST:evals:inside:outside:horizon}, and for the meaning of the values of $\eta_k<0$. At first glance setting $\epsilon = E/m c^2 \to i\,\eta$ seems to imply either imaginary mass, or imaginary energies, and therefore appears non-sensical. On further thought, note that the temporal and radial portion of the Schwarzschild metric $ds^2_{outside} = -\left(1-\frac{2 M_s}{r}\right)\,c^2\,dt^2 + \left(1-\frac{2 M_s}{r}\right)^{-1}\,dr^2$ for $r>2 M_s$ switches signs to $ds^2_{inside} = \left(\frac{2 M_s}{r}-1\right)\,c^2\,dt^2 - \left(\frac{2 M_s}{r}-1\right)^{-1}\,dr^2$ inside for $r<2 M_s$. The interpretation is that the timelike variable is associated with the term with the $-1$ (times a positive factor), and thus becomes $r$ inside the horizon. Thus, the variables $t$ and $r$ switch roles and become spacelike and timelike respectively, inside the horizon. Consequently, moving to the the singularity $r=0$ is inevitable, since ``time" $r$ always moves forward for inward falling particles inside the horizon. In fact, $\epsilon$ now has the interpretation of a momentum (vs an energy). Further, the stationary states, which outside had a temporal dependence of $e^{-i \epsilon_k\,t}$ now become upon replacement $\epsilon_k\to i\,\eta_k$, factors of $e^{\eta_k\,t}$, which for $\eta_k<0$ become evanescent like wave functions decaying to the singularity as $t$ increases. Therefore, the fact the numerics actually produces solutions with only $\eta_k<0$ seems to fit with our intuition. Recall that outside the horizon $e^{\pm i k r}$ are waves moving radially outward ($+$) and inward ($-$), with $\hbar k$ being the magnitude of the momentum $p_k$, with energy $p_k c = \hbar k c$. Thus it is conceivable (or at least plausibly interpretable) that $|\eta| \to \frac{(E=\hbar k c)}{m c^2}$ measures a quantized momentum inside the BH. \section{Summary and Conclusion}\label{sec:Summary:Conclusion} It has been a somewhat lengthy journey to attempt to answer a seeming simple question posed by an astute student (in 15 words or less): ``In analogy with the Coulomb problem, can one talk about bounds states in Schwarzschild spacetime?" To answer this question, we had to assemble theoretical background concepts, and analytic and numerical solution techniques for (i) the Schrodinger Equation (SE) in a central potential (the Coulomb problem), (ii) the Special Relativity (SR) generalization to the Dirac Equation (DE) in flat (Minkowski spacetime, MST), and finally (iii) the General Relativity (GR) generalization of the DE to curved spacetime (CST), specifically, the central symmetric Schwarzschild spacetime (SST). Driven by intuition, and bolstered by the courage of our convictions, we attempted to put some theoretical, analytic, and numerical terra firma beneath feet, though knowing at times we might be pushing an analogy (or two) beyond the limits of credulity. The results in \Sec{sec:bound:states:DE:SST}, and in particular, the bound state eigenvalues $|E_k|< m c^2$ numerically found in \Fig{fig:DE:SST:evals:with:without:diffusion} seem to indicate that the answer to the student's question appears to be affirmative. Further, the comparison of the the eigenvalues for the DE in SST with formulas from both the SR DE and the SE+SR, as shown in \Fig{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All}, shows that the student's intuition, and question, had merit. The intent of this investigation was to highlight a research strategy that could be employed by the student, drawing upon knowledge obtained in upper undergraduate to first year graduate physics education, supplemented by a host of well written and highly useful expository QM and GR textbooks currently available to the student, with enough worked exercises that this whole endeavor could be the seed of an independent research project. Much more remains that could be investigated, as we have only scratched the surface here, and peaked under a few lids, on a host container of fascinating NR and SR QM, and GR topics worthy of deeper investigation (e.g. angular momentum and spin details in the SR and GR Dirac equation, solutions of massless particles in Schwarzschild and other spacetimes often used to model photons, and the role of spinor transformations under the Poincare group of rotations, boost and translations, to name just a few). It is hoped that this investigation has served as a vehicle that might peak the research interest of the inquisitive and curious student, and spur them to further guided or self study, complemented by a host available instructive textbooks, and hopefully beyond, to the research literature itself. \flushleft{\underline{\bf Extra credit (research) project:}} \newline For the ambitious student (or professor!), you can verify your understanding of this material by redoing the DE in CST calculation presented in the main text for the \tit{Reissner-N{\o}rdstrom} metric (see Ryder\cite{Ryder:2009}, p265) for a star with mass $M$ and electric charge $Q$ given by spherically symmetric metric \be{RN:metric} \hspace{-0.25in} ds^2 = -c^2 \left(1- \frac{2 G M }{c^2 r} + \frac{G Q^2}{c^4 r^2}\right) dt^2 + \left(1- \frac{2 G M }{c^2 r} + \frac{G Q^2}{c^4 r^2}\right)^{-1}\,dr^2 + r^2\,\left(d\theta^2 + \sin^2\theta\,d\phi^2\right). \end{equation} As before, $\frac{2 G M }{c^2 r} = \frac{r_s}{r}$, while $\frac{G Q^2}{c^4 r^2} = \left(\frac{e^2}{\hbar c}\right) \left(\frac{Q^2}{e^2}\right) \left(\frac{G \hbar}{c^3 r^2}\right) = \alpha_c\, \left( \frac{Q}{e} \right)^2\, \left( \frac{\ell_p}{r} \right)^2, $ where $\ell_p$ is the Planck length and $\alpha_C$ is the fine structure constant (see \App{app:Units}). In the spirit of John Wheeler (who coined the term black hole), what might one anticipate about any possible solutions even before embarking on the lengthy calculation (to back up one's intuition)? Hint: compare the magnitude and length scales of the two $r$-dependent terms in \Eq{RN:metric}. Note: Astronomically speaking, the Reissner-N{\o}rdstrom is not considered particularly physical, since any charge on a BH will most likely be rapidly neutralized by surrounding in-falling matter. However, the metric is an intuitively satisfying extension of the Schwarzschild metric, extending the sole extensive properties of the BH from mass, to mass plus charge - and it is amenable to the analysis presented in this work. \section{Introduction}\label{sec:Intro} In the course of an undergraduate to first year graduate student's physcis education, the student will learn quantum mechanics (QM) encompassing the solution of bound states of the hydrogen atom, and the application and manipulation of spin. In specialized physics courses the student may have the opportunity to take an introduction to general relativity (GR), typically encompassing particle and photon orbits in the Schwarzschild spacetime (SST). Finally, in an advanced undergraduate (or first year graduate) QM course, the student will be exposed to the special relativistic (SR) Dirac equation (DE), and its solutions in special cases. Given the above archetypal physics education, a curious student may make the following observations, leading to the the subsequently posed question to a professor, or better yet, to themselves. ``In my introductory GR class, we saw that in the Schwarzschild metric, the radial equation for particles and photons orbits has a form reminiscent of a classical energy equation (see Hartle\cite{Hartle:2009}, p195), which for particles ($m>0$) takes the form \bea{Hartle:9.32} E_{Newt} &=& \frac{m}{2}\,\left( \frac{dr}{d\tau}\right)^2 +V_{eff}(r), \\ \textrm{where}\quad V_{eff}(r) &=& -\frac{G M m}{r}+\frac{L^2}{2 m r^2} - \frac{G M L^2}{c^2 m r^3}, \end{eqnarray} where $M$ is the mass of the gravitating body, $m$ is the test mass, $e = (E_{Newt} + m c^2)/ (m c^2)$ and $\ell = L/m$ are the conserved energy and angular momentum per unit rest mass, respectively (constants of the motion along the trajectory). Now if we ignore the $1/r^3$ term, this looks just like the central potential with the angular momentum barrier term included for an attractive $1/r$ potential, like the Coulomb problem. In my (non-relativistic) quantum mechanics (NRQM) class, we solved for the wave functions and eigenvalues of the bound quantum states of a hydrogen-like atom in a central potential $V_{C} = -(Ze)e/r$. Thus, can one expect to find bound states of $V_{Schw} \equiv V_{eff}$ if one were to simply replace $M\to Z e$ and $m\to e$? Would this make any sense? The scale of $M$, say for a solar mass black hole (BH), and $Ze$ are so vastly different? How would one go about answering this (intuitive) question?" The purpose of this article is to assemble the disparate, requisite components for the students in order to answer this question in the affirmative, at a level within the students undergraduate, to first year graduate physics background, aided by detailed, readable textbook material that would enable the student to answer this question for themselves. Let us first ask, what component physics, in order of difficulty, is required by the student to answer this question, and point out (from the author's personal perspective) those reference books containing enough explicit worked out material/derivations that would enable the student solve their question. First, once has to be able to solve for the bound states of hydrogen atom. This is standard fare in almost all introductory QM textbooks, but is lucidly well described and worked out in detail in one of the classic undergraduate QM texts, ``Quantum Mechanics: Volume 1," by Cohen-Tannoudgi, Diu and Laloe (CTQMv1) \cite{Cohen-Tannoudji:1977}. The eigenenergies of the hydrogen atom ($Z=1$) are known to be given by $E_n = -E_I/n^2$ where the ionizations energy is given by (CTQMv1, Chapter VII) $E_I = \frac{\alpha^2}{2} \mu c^2 = 13.6$ eV, where $\mu = (m_e\,m_p)/(m_e+m_p)\sim m_e$ is the reduced mass of the electron-proton hydrogen atom system, $\mu c^2\sim m_e c^2$ is the rest mass of the electron, $e>0$ is the charge of the proton, and $\alpha = e^2/(\hbar c) \sim 1/137$ is the fine structure constant. However, since we are inquiring about possible quantum bound states in a classical background GR spacetime, it is not enough to have a NRQM solution. One could attempt to invoke the special relativity (SR) energy equation and posit that by including the rest mass, the SRQM energy $\mathcal{E}$ should be given by $\mathcal{E} \sim m c^2 - E_I$ (where from now on we just denote $m_e\to m$). Working backwards, one might surmise that this might the small energy approximation to either \bea{SRQM:eval:guess} \mathcal{E}_s \overset{?}{=} \sqrt{1 - \frac{E_I}{n^2}}, \label{E_s}\\ \mathcal{E}_v \overset{?}{=} \frac{1}{\sqrt{1 + \frac{E_I}{n^2}}}. \label{E_v} \end{eqnarray} (The meaning of the $s$ and $v$ subscripts on $\mathcal{E}$ above, denoting scalar and vector coupling, respectively, will be explain in due time). But, which formula is more appropriate, and how would one physically decide between the two? It appears that a much more complete description of the Coulomb problem is required - i.e. the solution of the Coulomb problem in a SR setting using the Dirac equation (DE). The DE and its solution in a central potential are treated in the extremely informative worked out example problems in one of the extraordinary set of textbooks by Walther Greiner, namely ``Relativistic Quantum Mechanics: Wave Equations (RQMWE)."\cite{Greiner:1990} This textbook series, including volumes on QM, SRQM, Electrodynamics, Nuclear Physics, Thermodynamics, Quantum Field Theory and Statistical and Thermal Physics, is essentially (in the author's opinion) a ``reworking" of the classic Landau and Lipfshitz (informative, but terse) set of graduate level physics, now geared for advanced undergraduates and/or first year graduate students (the prefaces just state that it is for ``students"). In addition to its exceptional lucidity, each book involves typically over 100 solved problems, worked out in full detail, and showing all intermediate steps, which greatly aids the students proficiency in solving problems. Chapter 2 introduces the DE and Chapter 9 solves the Coulomb problem in a central potential for both (electromagnetic) vector and scalar (mass) coupling. Here, vector coupling means that the potential $V_v(r)=e\phi(r) /c$ arises from the timelike component of a SR 4-vector potential $A^\mu=(\phi, \bs{A})$ such that the relativistic energy $E$ is modified to $E\to E-V_v$ in the DE. Scalar coupling means that the potential $V_s(r)$ couples to the mass, so that $m c^2 \to mc^2 + V_s(r)$ in the DE. So intuition would lead one to suspect favoring the scalar coupling formula \Eq{E_s}. We shall explore if this intuition has any validity. If one is able to master the solution of the DE wave equation in SR in a central potential (aided by RQMWE), the student still has to tackle the thorny question of how to incorporate QM into GR. Of course, this is an outstanding unsolved problem in physics, but such a full theory is not required here. What one requires is the solution of a quantum mechanical wave function in a classical background GR spacetime. In effect, one needs to solve a wave equation in a medium with a spatially (here radially) dependent index of refraction. Thus, what is required is the DE in classical GR (curved spacetime) background. But how does one obtain this equation, since in RQM we solve the DE in a flat (Minkowski) background spacetime, incorporating only SR? The solution is well-known (to the QM/GR Illuminati), but the key concept is easy to grasp - Einstein's Equivalence Principle. Namely, the global inertial frames of SR no longer hold in GR, and are only locally valid at each point (actually within the observer's local laboratory\cite{Hartle:2009}, if small on the length scale of changes of the spacetime curvature). Thus, one needs to ``project" the SR DE into the observer's instantaneous local laboratory, and solve the resulting GR DE in this local (moving) frame. This involves the use of \tit{tetrads}, describing the four axes of the observer's frame $\mathbf{e}_a$, with $a\in\{0,1,2,3\}$, moving in the surrounding spacetime, with world vector components $\{e_a^{\;\;\mu}(x)\}=(\mathbf{e}_a)^\mu(x)$, for $\mu=\{0,1,2,3\}$. In the local observer's frame the GR DE takes on a first order form analogous to the SR DE, with all the complication going into the \tit{connection} describing how the spinor (2-vector with complex entries) components change (parallel transport and all that) as one moves from one point to another in a curved spacetime. There are now numerous introductory GR textbooks aimed at the upper level undergraduate student. The only one that tackles the GR DE, and in lucid, informative detail (but just setting up the equation, not solving it) is Lewis Ryder's (a particle physicist) wonderful ``Introduction to General Relativity, Chapter 11."\cite{Ryder:2009}. This is the formulation of the GR DE that we will use in this work. As one can see, there is much the student must master in order to answer their own intuitive question. And it may seem too daunting a task to tackle; a mountain too high to climb even for the motivated student. In NRQM, the solution of the Schr\"{o}dinger equation (SE) in a central potential brings in the additional complication of orbitial angular momentum operators, spherical harmonics and power series solutions of the SE, which are all standard fare in an undergraduate QM class (and clearly elucidated in many textbook, including CTQMv1). Mastering just the SR DE involves understanding an appreciation for covariant formalism, spinor transformations under Lorentz transformations (LT), and clever manipulation of the Dirac matrices to bring out the orbital and spin angular momentum terms (as explained in detail in RQMWE). Finally, GR itself presents conceptual and mathematical manipulation hurdles itself to the average student, even before one attempts tackling the GR DE. The latter, then involves additional concepts and mathematical manipulations of ``gauging" Lorentz symmetry, and constructing the appropriate tetrads for the local observer. In spite of this seeming mountain of required background, the claim here in this work is that the above mentioned textbooks provide more than sufficient background to step by step educate the motivated student, and allow them to learn the necessary prerequisites required to answer their self-posed question. The purpose of this article is to summarize the highlights of the relevant, requisite material (rather than reproduce it in full) and guide the student to the those portions of the above mentioned textbooks that give vastly more comprehensive and worked out details. The end result is a solution to the bound states in the background Schwarzschild spacetime (SST) and a comparison to the SR DE formulas and their simplified Schrodinger plus special relativity (SE+SR) forms \Eq{E_s} and \Eq{E_v}, answering the student's intuitively posed question. To keep the problem as simple as possible, we focus on spherically symmetric $s$-orbital states with $L=0$, so that $V_{eff}$ is directly analogous to $V_C$ (i.e. we drop the angular momentum barrier terms in the former). We include Appendices with enough background material to make this as work self-contained as possible when supplemented by the aforementioned textbooks. We will also write down equations in dimensionless form using the length, energy, etc. scales appropriate to the Coulomb and gravitational problems (see \App{app:Units}), so that both equations have the same form for purposes of comparison. So let us now embark on our investigative journey of the student's intuitive question. \section{Wave functions and eigenenergies of the NRQM Coulomb problem: a refresher}\label{NRQM:Coulomb:solns} As any student exposed to a NRQM course is well award, the Schrodinger equation (SE) $i \hbar\,\partial \psi(\bs{r},t)/\partial t = \left( \bs{p}^{\,2}/2m + V_C(r) \right)\, \psi(\bs{r},t)$ for stationary states $\psi(\bs{r},t) = e^{-i E t/\hbar}\,\phi(\mathbf{r})$ for $E<0$ constant, with $E\to i \hbar \frac{\partial}{\partial t}$ and $\mathbf{p}\to -i \hbar\,\mathbf{\nabla}$ yields \be{SchrEqn} \mathcal{H}\, \phi(\mathbf{r})\equiv \left[-\frac{\hbar^2}{2 m} \nabla^2 - \frac{e^2}{r} \right]\,\phi(\mathbf{r}) = E\,\phi(\bs{r}). \end{equation} Using the differential operator identities (see the inside cover of CTQMv1 and Arfken\cite{Arfken:2012}, Chapter 3) \bea{Del:Opr} \boldsymbol{\nabla} &=& \mathbf{e}_r \frac{\partial}{\partial r} - \frac{i}{\hbar r^2}\,\mathbf{r}\times\mathbf{L}, \\ \mathbf{L} &=& \frac{\hbar}{i}\,\mathbf{r}\times\boldsymbol{\nabla} \end{eqnarray} where $\mathbf{e}_r = \bs{r}/r$ is the orthonormal unit vector in the radial direction. Following CTQMv1 p662 and p778, we can write the well known formula for the Laplacian $\boldsymbol{\nabla}^2$ in spherical coordinates as \bea{Laplacian} \boldsymbol{\nabla}^2 &=& -\frac{\hbar^2}{2 m}\,\frac{1}{r}\, \frac{\partial^2}{\partial r^2}\, r + \frac{\mathbf{L}^2}{2 m r^2}, \\ \ell^2 = \frac{\mathbf{L}^2}{\hbar^2} &=& -\left( \frac{\partial^2}{\partial\theta^2} + \frac{1}{\tan\theta}\,\frac{\partial}{\partial \theta} + \frac{1}{\sin^2\theta}\,\frac{\partial^2}{\partial\varphi^2 } \right) \end{eqnarray} Since the Coulomb force is radially directed, there is no torque, so that the angular momentum $\mathbf{L}$ is constant of the motion. As such, we can consider the classical motion of particles as taking place in a plane, typically taken to be the equator $\theta=\pi/2$. Thus, we seek solutions $\phi(\mathbf{r})$ involving a complete set of commuting observables $\{ \mathcal{H}, \mathbf{L}^2, L_z\}$ such that \be{NRQM:sep:vars} \mathcal{H}\,\phi(\mathbf{r}) = E\,\phi(\mathbf{r}),\quad \mathbf{L}^2\,\phi(\mathbf{r}) = \hbar^2\,\ell\,(\ell+1)\,\phi(\mathbf{r}),\quad L_z\,\phi(\mathbf{r}) = m\hbar\,\phi(\mathbf{r}). \end{equation} Using the well known properties of the spherical harmonics\cite{Cohen-Tannoudji:1977,Arfken:2012} $Y_{\ell, m}(\theta,\varphi)$ as the eigenfunctions of the angular momentum operator, \be{Ylm} \ell^2\, Y_{\ell, m}(\theta,\varphi) = \ell\,(\ell + 1)Y_{\ell, m}(\theta,\varphi),\quad -i \frac{\partial}{\partial \varphi}Y_{\ell, m}(\theta,\varphi) = m\,Y_{\ell, m}(\theta,\varphi), \end{equation} for $m\in\{-\ell, -\ell+1,\ldots, \ell-1, \ell\}$, we choose the separation of variables as $\phi(\mathbf{r}) = R(r)\,Y_{\ell, m}(\theta,\varphi)$ to obtain \be{radial:SE:R} \left[ -\frac{\hbar^2}{2 m}\,\frac{1}{r}\, \frac{\partial^2}{\partial r^2}\, r + \frac{\ell\,(\ell + 1)\,\hbar^2}{2 m r^2} + V_C(r) \right]\,R(r) = E\,R(r), \end{equation} where the spherical harmonics have been canceled from both sides of the equation, leading to a purely radial Schrodinger equation. In addition to \Eq{radial:SE:R} we impose, on physical grounds, the boundary condition that the radial wave function $R(r)$ decay to zero as $r\to\infty$, It is convenient to make one last substitution $R(r) = u_{k,\ell}(r)/r$, and based on physical ground argue that one needs to impose the boundary condition $u_{k,\ell}(0)=0$, at the singularity that exists at the origin for the point particle electron. This leads to the final, simplified final form of the radial equation \be{radial:SE:u} \left[ -\frac{\hbar^2}{2 m}\, \frac{\partial^2}{\partial r^2}\, + \frac{\ell\,(\ell + 1)\,\hbar^2}{2 m r^2} + V_C(r) \right]\,u_{k,\ell}(r) = E\,u_{k,\ell}(r), \quad V_C(r) = -\frac{e^2}{r}. \end{equation} Here, the $\ell$ subscripts on $u_{k,\ell}(r)$ indicates the dependence on the angular momentum, and the index $k$ represents all other non-angular momentum dependencies of the eigensolutions. We can now write \Eq{radial:SE:u} in two dimensionless forms: \bea{radial:SE:u:dimless} \left[ \frac{\partial^2}{\partial \rho^2}\, - \frac{\ell\,(\ell + 1)}{\rho^2} + \frac{2}{\rho} - \lambda^2_{k,\ell} \right]\,u_{k,\ell}(\rho) &=& 0, \quad \rho = \frac{r}{a_0}, \quad \lambda_{k,\ell} = \sqrt{\frac{-E_{k,\ell}}{E_I}}\ge 0, \\ \label{radial:SE:u:dimless:atomic:units} &{}& \nonumber \\ \left[ \frac{\partial^2}{\partial \rho^2}\, - \frac{\ell\,(\ell + 1)}{\rho^2} + \frac{2 \alpha_C}{\rho} - \epsilon_{k,\ell} \right]\,u_{k,\ell}(\rho) &=& 0, \quad \rho = \frac{r}{\lambda_c}, \quad \epsilon_{k,\ell} = \frac{E_{k,\ell}}{m c^2}. \label{radial:SE:u:dimless:fundamental:units} \end{eqnarray} \Eq{radial:SE:u:dimless:atomic:units} is expressed in atomic units: lengths and energy in terms of the Bohr radius $a_0$ and ionization energy $E_I$ of the hydrogen atom (see \App{app:Units}), and is the form that is typically employed (see CTQMv1, p794). The second form \label{radial:SE:u:dimless:fundamental:units} is expressed in terms of fundamental units of the electron: lengths and energy in terms of the Compton wavelength $\lambda_c$ and rest mass $m c^2$ of the electron (see \App{app:Units}). While this latter form is not typically employed, we list it here for later reference when we want to consistently compare to the DE, especially in the case of SST. This latter form explicitly brings out the role of fine structure constant $\alpha_C$. Forgoing the use of (non-intuitive) special (confluent hypergeometric) functions, we will follow CTQMv1 and outline the solution to \Eq{radial:SE:u:dimless:atomic:units} using a power series expansion, the typical approach taught in an undergraduate QM course. Following CTQMv1, pp794-797, we outline the essentials of the solution method. First one looks at the asymptotic form of the equation for $\rho\to\infty$, keeping only the dominant terms (i.e. dropping terms involving $\rho^{-1}$ and $\rho^{-2}$), which yields $[\partial^2_\rho - \lambda^2_{k,\ell}]\, u_{k,\ell}(\rho)=0$ with solutions $e^{\pm\lambda_{k,\ell}\,\rho}$. Since we want the wavefunction to decay to zero at spatial infinity (so that they are square integrable for finite total probability) we reject the positive exponent, and look for solutions of the form $u_{k,\ell}(\rho) =e^{-\lambda_{k,\ell}\,\rho} \,y_{k,\ell}(\rho)$, where we take $y_{k,\ell}(\rho)$ of the power series form $y_{k,\ell}(\rho) = \rho^s\,\sum_{q=0}^\infty c_q\,\rho^q$ with $c_0\ne 0$. The equation satisfied by $y_{k,\ell}(\rho)$ is \be{y:eqn} \left[ \frac{d^2}{d\rho^2} - 2\,\lambda_{k,\ell}\,\frac{d}{d\rho} + \left( \frac{2}{\rho}-\frac{\ell(\ell+1)}{\rho^2} \right)\, \right]\,y_{k,\ell}(\rho)=0, \end{equation} which must be satisfied for each coefficient $c_q$ term by term. In particular, the $c_0$ term yields the simple equation $[ s(s+1) - \ell(\ell+1) ]\,c_0=0$ yielding the solutions $s=\{\ell+1, -\ell\}$. Since we have the boundary condition $y_{k,\ell}(0)=0$ we must reject $s=-\ell$ (since $\ell\ge0$) and retain $s=\ell+1$. It is then straight forward to develop the following one term recursion $q (q+ 2\,\ell +1)\,c_q = 2\,[(q+\ell)\,\lambda_{k,\ell} -1]\,c_{q-1}$, for $q\ge 1$. The key quantization concept is that if series is infinite then $\frac{c_q}{c_{q-1}}\underset{q\to\infty}{\sim}\frac{2\,\lambda_{k,\ell}}{q}$ which is the same ratio of coefficients in the power series expansion of the function $e^{2\,\lambda_{k,\ell}\,\rho}$, which then violates our requirement for solutions decaying as we approach infinity. The only way to avoid this, is for the above one term recursion relation to terminate. By examining the coefficient of $c_{q-1}$ above, this is easily satisfied if $\lambda_{k,\ell} = \frac{1}{k+\ell}$. Thus, one obtains that the bound energies are given by $E_{k,\ell}= -E_I/(k+\ell)^2$ for a given $\ell$ for $k\in\{1, 2, 3, \ldots\}$ and $\ell=\{0, 1, 2, \ldots\}$. We therefore rename $k+\ell\to n$ and recover the Bohr energies in terms of the principle quantum number $n$. The radial wavefunctions are then given by (CTQMv1, p797) \bea{R:wavefunctions} n=1, \ell=0:\quad R_{k=1,\ell=0}(r) &=& 2\, (a_0)^{-3/2}\, e^{-r/a_0}, \\ n=2, \ell=0:\quad R_{k=2,\ell=0}(r) &=& 2\, (2\,a_0)^{-3/2}\,\left(1- \frac{r}{2\,a_0} \right) \,e^{-r/2\,a_0}, \\ n=2, \ell=1:\quad R_{k=2,\ell=0}(r) &=& (2\,a_0)^{-3/2}\,\frac{1}{\sqrt{3}}\,\frac{r}{a_0} \,e^{-r/2\,a_0} \end{eqnarray} Finally, since $E_I\ll m\,c^2$, one is justified in using the NRQM SE. \section{The SR Dirac Equation for spin $\mathbf{1/2}$ particles and its solutions for vector (Coulomb) and scalar (mass) coupling}\label{sec:SRDE:vector:scalar} Having traversed familiar ground, we now steadfastly venture in territory that may be less familiar to some students. \subsection{Free DE: no potentials}\label{sec:DE:no:V} As is historically well-known, Dirac was bothered by the inconsistency of the SE being first order in time, yet second order in space. Since SR treats time and space equivalently he was led to seek an equation that was first order in both time and space. For the following we will closely follow the exposition of Chapter 2 of Greiner\cite{Greiner:1990} which we denote as RQMWE. Dirac postulated a wave equation given by $i\,\hbar\,\frac{\partial \psi}{\partial t} = \hat{H}\,\psi$ of the form (RQMWE, p75, p82) \be{DE:form} i\,\hbar\,\frac{\partial \psi}{\partial t} =\left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi \equiv \hat{H}_f\,\psi. \end{equation} Here $\hat{\mathbf{p}} = -i\,\hbar\boldsymbol{\nabla}$ is the usual QM momentum operator and again we have taken $\hat{E}\to i\,\hbar\,\partial/\partial t$ as the energy operator. The quantities $\hat{\boldsymbol{\alpha}}= \{\hat{\alpha}_1, \hat{\alpha}_2,\hat{\alpha}_3\}$ and $\hat{\beta}$ cannot be numbers if one wishes to preserve the form invariance of the equation under simple spatial rotations, and these are considered as $N\times N$ matrices. Therefore, $\psi$ itself must be a column vector of $N$ complex functions of space and time \be{psi:DE} \psi = \left(\begin{array}{c} \psi_1(\mathbf{x},t) \\ \psi_2(\mathbf{x},t) \\ \vdots \\ \psi_N(\mathbf{x},t) \end{array}\right) \quad\Rightarrow\quad \rho(x) = \psi^\dagger\,\psi(x) = \sum_{i=1}^N\, \psi^*(x)\,\psi(x)\ge 0, \end{equation} where $\rho(x)$ is the positive probability density of the particle. One also wishes to preserve the SR energy equation $E^2 = \mathbf{p}^2\,c^2 + (m\,c^2)^2$ where $\mathbf{p}$ without the operator circumflex $\verb+^+$ denotes the ordinary classical momentum vector. Under the usual NR substitution (quantization) $E\to i\,\hbar\,\partial/\partial t$ and $\mathbf{p}\to -i\,\boldsymbol{\nabla}$ the SR energy equations becomes the Klein-Gordon wave equation $-\hbar^2\,\frac{\partial^2}{\partial t^2}\psi_\sigma = \left(-\hbar^2\,c^2\nabla^2 + m^2\,c^4 \right)\psi_\sigma$ for each component of the wave function $\psi_\sigma$ (where one typically denotes the components of the wave function with Greek indices). Carrying out this straightforward algebra (RQMWE, p76-77) the net result is that the matrices $\hat{\alpha}_i$ and $\hat{\beta}$ must satisfy \bea{alpha:beta:DE} \{ \hat{\alpha}_i, \hat{\alpha}_j\} &\equiv& \hat{\alpha}_i\,\hat{\alpha}_j + \hat{\alpha}_j\, \hat{\alpha}_i = 2\,\delta_{ij}\,\mathbb{I} , \label{alpha:beta:DE:line1} \\ \{ \hat{\alpha}_i, \hat{\beta}\} &=& 0, \label{alpha:beta:DE:line2} \\ \hat{\alpha}^2_i &=& \hat{\beta}^2= \mathbb{I}, \label{alpha:beta:DE:line3} \end{eqnarray} where $\mathbb{I}$ is the $N\times N$ identity matrix. It is not hard to show that no $N=2$ solution exists, and the first non-trivial solution that exists is for $N=4$. This is called the Dirac 4-spinor, denoted by \be{Dirac:4spinor} \psi(x) = \left(\begin{array}{c} \varphi(\mathbf{x},t) \\ \chi(\mathbf{x},t) \end{array}\right) \end{equation} where $\varphi$ and $\chi$ are each themselves 2-spinors (a 2-vector with complex entries), familiar from NRQM. The physical interpretation is that upper two components $\varphi$ of the 4-spinor $\psi$ represents particles of spin-$1/2$, while the lower two components $\chi$ represents anti-particles of spin-$1/2$ (same mass, but opposite charge). A conventional set of $4\times 4$ matrices satisfying \Eq{alpha:beta:DE:line1} -\Eq{alpha:beta:DE:line3} are given by \be{DE:alpha:beta} \hat{\alpha}_i = \left(\begin{array}{cc} 0 & \hat{\sigma}_i \\ \hat{\sigma}_i & 0 \end{array}\right), \qquad % \hat{\beta} = \left(\begin{array}{cc} \mathbb{I} & 0 \\ 0& -\mathbb{I} \end{array}\right), \end{equation} where $\sigma_i$ are the standard $2\times 2$ Pauli matrices, and $\mathbb{I}$ is the $2\times 2$ identity matrix, \be{Pauli:matrices} \sigma_1 = \left(\begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array}\right),\quad % \sigma_2 = \left(\begin{array}{cc} 0 & -i \\ i & 0 \end{array}\right),\quad % \sigma_3 = \left(\begin{array}{cc} 1 & 0 \\ 0& -1 \end{array}\right), \quad % \mathbb{I} = \left(\begin{array}{cc} 1 & 0 \\ 0& 1 \end{array}\right). \end{equation} Another commonly employed set of matrices $\gamma^\mu$ satisfying $\{\gamma^\mu, \gamma^\nu\} = 2\,\eta^{\mu,\nu}\mathbb{I}$ when using the particle physicist convention for the SR metric (RQMWE\cite{Greiner:1990}) $\eta_{\mu\nu}=\eta^{\mu\nu} = \trm{diagonal}\{1,-1,-1,-1\}$, (note: SR and GR books, e.g. see Chapter 11.3, Ryder\cite{Ryder:2009}, most commonly use the alternative convention, $\eta_{\mu\nu}=\eta^{\mu\nu} = \trm{diagonal}\{-1,1,1,1\}$) called the Dirac gamma matrices, are given by \be{DE:gamma:matrices} \gamma^0= \left(\begin{array}{cc} \mathbb{I} & 0 \\ 0& -\mathbb{I} \end{array}\right), \quad \gamma^i = \left(\begin{array}{cc} 0 & \hat{\sigma}_i \\ -\hat{\sigma}_i & 0 \end{array}\right). \end{equation} These are employed so that the DE can be written in covariant form \be{DE:covariant:form} i\,\hbar\,\gamma^\mu\,\partial_\mu\,\psi = m\,c\,\psi, \qquad \eta_{\mu\nu} =\eta^{\mu\nu} = \trm{diagonal}\{1,-1,-1,-1\}, \end{equation} where the $4\times 4$ identity matrix $\mathbb{I}_{4\times 4}$ is implied (and often omitted) on the righthand side. In the particle physicist convention $\hat{p}^\mu = i\,\hbar\, \{\frac{1}{c}\frac{\partial}{\partial t}, -\boldsymbol{\nabla}\}$ so the covariant DE is given by $\gamma^\mu\,\hat{p}_\mu\,\psi~=~m\,c\,\psi$, where $\hat{p}_\mu = \eta_{\mu\nu}\,\hat{p}^\nu = i\,\hbar\, \{\frac{1}{c}\frac{\partial}{\partial t}, \boldsymbol{\nabla}\}$. Lastly, the operator $\hat{\mathbf{S}}$ \be{DE:spin:opr} \hat{\mathbf{S}} = \frac{\hbar}{2}\,\hat{\mathbf{\Sigma}}\equiv \frac{\hbar}{2}\, \left(\begin{array}{cc} \hat{\boldsymbol{\sigma}} & 0\\ 0 & \hat{\boldsymbol{\sigma}} \end{array}\right), \qquad \hat{\mathbf{J}} = \hat{\mathbf{L}}+\hat{\mathbf{S}}. \end{equation} which commutes with the Dirac Hamiltonian $\hat{H}_f$ is called the spin vector operator, and represents the intrinsic spin of the electron, in addition to its orbital angular momentum. The total angular momentum $\hat{\mathbf{J}}$ is the sum of the orbital and intrinsic angular momentum, and is a constant (of the motion). For stationary states (which is our primary interest) we let $\psi(\mathbf{x},t) = e^{-i\,E/\hbar}\psi(\mathbf{x})$ for $E$ constant, such that $ i\,\hbar\,\partial/\partial t \,\psi(\mathbf{x},t) = E\,\psi(\mathbf{x},t)$, and the DE simply becomes of the form (RQMWE, p82) \be{DE:form:E} E\,\psi =\left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi \equiv \hat{H}_f\,\psi. \end{equation} \subsection{DE with potentials}\label{sec:DE:with:V} If the electron interacts with an electromagnetic potential $A^\mu = \{A^0, \mathbf{A}\}$ one obtains the usual NRQM substitutions (RQMWE, p94) $E\to i\,\hbar\,\partial/\partial t - e\,A_0$ and $\hat{\mathbf{p}} \to \hat{\mathbf{p}} - \frac{e}{c}\, \mathbf{A}$ so that the DE for stationary states becomes \be{DE:form:EM:general} i\,\hbar\,\frac{\partial \psi}{\partial t} = \left[c\,\hat{\boldsymbol{\alpha}}\cdot\left(\hat{\mathbf{p}} - \frac{e}{c}\, \mathbf{A}\right) + e\,A_0 + \hat{\beta}\,m\,c^2\right]\,\psi, \\ \end{equation} If we define say the Coulomb potential in the Coulomb gauge via $V_v \overset{\trm{def}}{=} e\,A_0$ and $\mathbf{A}=0$. the DE for stationary states becomes \be{DE:form:EM:vector:coupling} \left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,m\,c^2\right]\,\psi =(E-V_v)\,\psi. \end{equation} One can also introduce a scalar potential $V_s$ which couples to the mass so that $m\,c^2~\to~m\,c^2+ V_s$. Thus, the DE for stationary states, with both vector and scalar coupling, is given by (RQMWE, section 9.8, p184-187) \be{DE:form:EM:scalar:coupling} \left[c\,\hat{\boldsymbol{\alpha}}\cdot\hat{\mathbf{p}} + \hat{\beta}\,(m\,c^2 + V_s)\right]\,\psi =(E-V_v)\,\psi. \end{equation} \subsection{DE in a central potential}\label{sec:DE:central:potential} The DE in a central potential (vector coupling), in this section denoted by $V_v \to V(r)$, presents some involved matrix and vector identity manipulations which are explicitly worked out in RQMWE (see section 9.3, p169-172). The goal of this exercise is to (i) explicitly reveal the terms involving the angular momentum $\hat{\mathbf{L}}$ (as in the NRQM SE), as well as the intrinsic spin $\hat{\mathbf{S}}$, and (ii) ultimately remove the spinor dependency of the equations to produce a purely radial equation (actually, a pair of first order radial equations, vs a single second order radial equation as in the NRQM SE). Rather than reproducing all these algebraic machinations in detail here, we simply highlight the essential points which lead to our objectives (i) and (ii). Using the representation of $\hat{\boldsymbol{\alpha}}= \tiny{\left(\begin{array}{cc} 0 & \hat{\boldsymbol{\sigma}} \\ \hat{\boldsymbol{\sigma}} & 0 \end{array}\right)}$ and $\hat{\beta}= \tiny{\left(\begin{array}{cc} \mathbb{I} &0 \\0& -\mathbb{I} \end{array}\right)}$ given previously, and the representation of the Dirac 4-spinor in terms of a pair of 2-spinors $\psi = \tiny{\left(\begin{array}{c}\varphi \\ \chi \end{array}\right)}$ the DE becomes the pair of coupled ordinary differential equations \bea{DE:RQMWE:p170:5} c(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}})\chi &=& [(E-V) - m\,c^2 ]\,\varphi, \label{DE:RQMWE:p170:5:line:1} \\ c(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}})\varphi &=& [(E-V) + m\,c^2 ]\,\chi. \label{DE:RQMWE:p170:5:line:2} \end{eqnarray} One must now make an ansatz about the form of the 2-spinors $\varphi$ and $\chi$. From the NRQM SE, we suspect that they must contain the spherical harmonics $Y_{l,m}(\theta,\phi)$ in some form. This is the trickiest part of the separation of variables problem of the SR DE in a central potential, and involves much detailed algebra (explicitly worked out in RQMWE, section 9.3 pp169-172. For an intricate symmetry proof also see (10.54) pp212-214). These results are required in order that the angular portion of the 2-spinors drop out the of DE (analogous to the NRQM SE), leaving only two coupled, scalar radial equations - which is the end goal of all these algebraic manipulations. As mentioned above, we will state some of these important results, without proof, but with explanation, primarily directing the reader to the detailed derivations in RQMWE.\cite{Greiner:1990} The ansatz for the pair of 2-spinors is \be{2:spinor:ansatx} \varphi\to \varphi_{j\ell m} = i\,g(r)\,\Omega_{j\ell m}(\mathbf{r}/r), \qquad \chi\to \chi_{j\ell' m} = -f(r)\,\Omega_{j\ell' m}(\mathbf{r}/r), \end{equation} where $\Omega_{j\ell m}$ is called a spherical spinor and $\mathbf{r}/r = \mathbf{e}_r=(\sin\theta\,\cos\phi, \sin\theta\,\sin\phi, \cos\theta)$ so that the former only depend on the angles $(\theta,\phi)$ and not on $r$. Here the indices $\{j\ell m\}$ indicate the total angular moment $j$ (the eigenvalues of $\mathbf{J}^2$ are $\hbar^2\,j(j+1)$), orbital angular momentum $\ell$ (the eigenvalues of $\mathbf{L}^2$ are $\hbar^2\,\ell(\ell+1)$), and the magnetic quantum number $m$ takes values from $\{-j, -j+1,\ldots,j\}$. Here, the value of the intrinsic spin is $s=\tfrac{1}{2}$ (the eigenvalues of $\mathbf{S}^2$ are $\hbar^2\,s(s+1)$). Note that the expression for $\chi$ carries an index $\ell'$ given by \be{ell:prime} \ell' = \begin{cases} \ell+1, \quad \trm{if}\; j=\ell+\frac{1}{2},\\ \ell-1, \quad \trm{if}\; j=\ell-\frac{1}{2}. \end{cases} \end{equation} This is the subtlety that warrants much algebra to prove. The physics of it lies with the fact that solutions of the SR DE (as well as the NRQM SE) must also be eigenstate of the parity operator, which simply changes $\mathbf{r}\to-\mathbf{r}$, and is equivalent to $\{\theta, \phi\}\to\{\pi-\theta, \pi+\phi\}$. This is manifested in the choice of spherical harmonic functions $Y_{\ell m}(\theta,\phi)$ that appear in $\Omega_{j\ell m}$ and $\Omega_{j\ell' m}$. Under the operation of parity we have have $Y_{\ell m}(\theta,\phi)\to Y_{\ell m}(\pi-\theta, \pi+\phi) = (-1)^\ell\,Y_{\ell m}(\theta,\phi)$ so that the parity of the spherical harmonics depends solely on $\ell$. The implication of \Eq{ell:prime}, and the primary physical point, is that the parity of $\varphi_{j\ell m}$ is the negative (i.e. opposite) of that of $\chi_{j\ell' m}$. Thus, under parity, the orbital orbital angular momentum value $\ell'$ of $\chi_{j\ell' m}$ can only change by one unit from that $\ell$ of $\varphi_{j\ell m}$. Formally, this statement is expressed as \be{Greiner:p171:12} \left( \hat{\boldsymbol{\sigma}}\cdot \frac{\mathbf{r}}{r} \right)\,\Omega_{j\ell m} = -\Omega_{j\ell' m}. \end{equation} Note the use of $\ell$ on the lefthand side of \Eq{Greiner:p171:12}, but $\ell'$ on the righthand side. Here $\hat{\boldsymbol{\sigma}}\cdot \mathbf{r}/r$ is a scalar operator that changes sign under parity ($\mathbf{r}\to-\mathbf{r}$). \Eq{Greiner:p171:12} is a non-trivial statement which should be proved (explicitly, see RQMWE, (10.54) pp212-214), but for which here, we will simply utilize. Ultimately, the above additional complexity arises from the addition of the orbital and spin angular momentum to yield a fixed total angular momentum. Addition of angular momentum, and the manipulation of Clebsch-Gordon (CG) coefficients are standard topics covered in most undergraduate QM textbooks, but also quite explicitly and lucidly (with many worked examples) in the second volume (CTQMv2) of the two volume classic QM text by Cohen-Tannoudji, Diu and Laloe.\cite{Cohen-Tannoudji:1977:2} The key concept is that for a given orbital angular momentum $\ell$, the addition of an intrinsic spin-1/2 ($s=\frac{1}{2}$) leads to the two possible values for the angular momentum $j\in\{\ell-\frac{1}{2}, \ell+\frac{1}{2}\}$. This is often expressed as $\ell\otimes\frac{1}{2} = (\ell-\frac{1}{2})\oplus(\ell+\frac{1}{2})$ to indicate how the composition (addition, tensor product $\otimes$) of two angular momentum leads to the (direct sum, $\oplus$) of subspaces of total angular momentum $j=\ell\pm\frac{1}{2}$. Thus, if $\varphi_{j\ell m}$ has orbital angular momentum $\ell$ and total angular momentum, say, $j=\ell + \frac{1}{2}$, then since $\chi_{j\ell' m}$ has opposite parity with orbital angular momentum differing by one unit, then it can only have $\ell'=\ell+1$. This is the case, since by addition of its angular momenta, $\chi_{j\ell m}$ can only possibly have total the angular momentum values $\ell'=(\ell+1)\otimes\frac{1}{2} = (\ell+\frac{1}{2}\equiv j) \oplus (\ell+\frac{3}{2})$. The other potential choice $\ell'=\ell-1$ must be rejected since it can only result in total angular momentum values of $\ell'\overset{?}{=}(\ell-1)\otimes\frac{1}{2} = (\ell-\frac{3}{2}) \oplus (\ell-\frac{1}{2})$, neither which are $j=\ell+\frac{1}{2}$. (Similar arguments hold if we take $j=\ell-\frac{1}{2}$ implying that $\ell'=\ell-1$). The explicit expressions for the spherical 2-spinors are then given by (RQMWE, (11a) and (11b), p170 - which are not really needed - just their symmetry property \Eq{Greiner:p171:12}) \be{Greiner:p170:11a:11b} \Omega_{j=(\ell+\frac{1}{2}),\ell,m} = \left(\begin{array}{c} \sqrt{\frac{j+m}{2j}}\,Y_{\ell,m-\frac{1}{2}} \\ \sqrt{\frac{j-m}{2j}}\,Y_{\ell,m+\frac{1}{2}}\end{array}\right), \quad \Omega_{j=(\ell-\frac{1}{2}),\ell,m} = \left(\begin{array}{c} -\sqrt{\frac{j-m+1}{2(j+1)}}\,Y_{\ell,m-\frac{1}{2}} \\ \sqrt{\frac{j+m+1}{2(j+1)}}\,Y_{\ell,m+\frac{1}{2}}\end{array}\right). \end{equation} With this (important but involved) preliminary out of the way, one can now turn to the desired goal of extracting the pair of radial equations from our coupled \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2}. The solution method is straightforward ``in theory," namely we want to apply $\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}$ to \Eq{DE:RQMWE:p170:5:line:1} so that it can act on $\varphi$ on the righthand side, for which we can then substitute in \Eq{DE:RQMWE:p170:5:line:2}. The ``rub" comes in that $\hat{\mathbf{p}} = -i\,\hbar\,\boldsymbol{\nabla}$, and so it acts on both the radial and angular parts of $\varphi$ and $\chi$. We provide the relevants steps here (see the more complete details in RQMWE, p171) so that the student can see where the orbital angular momentum operator $\hat{\mathbf{L}}$ arises. Let us first consider the following calculation: \bea{Greiner:p171:10} (\hbs{\sigma}\cdot\hbs{p})\,\varphi_{j\ell m} &=& (\hbs{\sigma}\cdot\hbs{p})\Big(i\,g(r)\,\Omega(\bs{r}/r)_{j\ell m}\Big), \nonumber \\ &=& (\hbs{\sigma}\cdot\hbs{p})\big(i\,g(r)\big)\,\Omega_{j\ell m} + i\,g(r)\,(\hbs{\sigma}\cdot\hbs{p})\,\Omega_{j\ell m}, \nonumber \\ &=& \hbar\,\frac{d g(r)}{dr} \left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right) \,\Omega_{j\ell m} + i\,g(r)\,(\hbs{\sigma}\cdot\hbs{p})\,\Omega_{j\ell m}, \end{eqnarray} where the parity term of \Eq{Greiner:p171:12} $\left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right) \Omega_{j\ell m}$ is clearly exhibited. Let us now apply $\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}$ to $\Omega_{j\ell m}$ and use \Eq{Greiner:p171:12} (with $\left(\hbs{\sigma}\cdot\frac{\bs{r}}{r} \right)^2=\mathbb{I}$) to obtain \be{Greiner:p171:13} -\left(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}\right)\, \Omega_{j\ell m} = \left( \hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}} \right)\, \left( \hat{\boldsymbol{\sigma}}\cdot\frac{\mathbf{r}}{r} \right)\, \Omega_{j\ell' m}. \end{equation} We can now use the well known product formula for Pauli matrices $(\hat{\boldsymbol{\sigma}}\cdot\mathbf{A}) (\hat{\boldsymbol{\sigma}}\cdot\mathbf{B}) = \mathbf{A}\cdot\mathbf{B} + i\,\hat{\boldsymbol{\sigma}}\cdot\mathbf{A}\times\mathbf{B}$, a relationship which is straightforwardly proved by expanding the vector operations in terms of components and using the well known Pauli matrix relations $\sigma_i\,\sigma_j = \delta_{ij} + i\,\epsilon_{ijk}\,\sigma_j\,\sigma_k$. Applying this to the righthand side of \Eq{Greiner:p171:13} one obtains \be{Greiner:p171:15} -\left(\hat{\boldsymbol{\sigma}}\cdot\hat{\mathbf{p}}\right)\, \Omega_{j\ell m} = \left( \hat{\mathbf{p}}\cdot\frac{\mathbf{r}}{r} + i\, \hat{\boldsymbol{\sigma}}\cdot \left( \hat{\mathbf{p}}\times\frac{\hat{\mathbf{r}}}{r} \right) \right) \,\Omega_{j\ell' m}, \end{equation} from which one can identify the angular momentum operator $\hat{\mathbf{L}}= \mathbf{r}\times\hat{\mathbf{p}}$. However, one must be careful of operator ordering since $\hat{\mathbf{p}} = -i\hbar\,\boldsymbol{\nabla}$ is a differential operator that also acts on $\mathbf{r}$. Thus, we write the above term in the parentheses on the righthand side as \bea{Greiner:p171:16} \big( \hat{\mathbf{p}}\cdot\mathbf{r} &+& i\, \hat{\boldsymbol{\sigma}}\cdot \left( \hat{\mathbf{p}}\times\mathbf{r} \right) \big) \,\frac{1}{r}\,\Omega_{j\ell' m}, \nonumber \\ &=& \big( -i\hbar\,(\boldsymbol{\nabla}\cdot\mathbf{r}) -i\,\hbar\,\mathbf{r}\cdot\boldsymbol{\nabla} -i\, \hat{\boldsymbol{\sigma}}\cdot(\mathbf{r}\times\hat{\mathbf{p}}) \big) \,\frac{1}{r}\,\Omega_{j\ell' m}, \nonumber \\ &=& \left( -i\,\hbar\,\frac{3}{r} -i\,\hbar\,r\,\left(-\frac{1}{r^2}\right) -i\,\frac{\hbs{\sigma}\cdot\hbs{L}}{r} \right) \,\Omega_{j\ell' m}, \nonumber \\ &=& -i\,\frac{1}{r}\, \left( 2\,\hbar + \hbs{L}\cdot\hbs{\sigma} \right) \,\Omega_{j\ell' m}. \end{eqnarray} One then performs the standard calculation $\hbs{J}^2 = \left(\hbs{L}+\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2 = \hbs{L}^2+\left(\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2+\hbar\,\hbs{\sigma}\cdot\hbs{L}$ in order to write \bea{Greiner:p171:18} \hbar\,\hbs{\sigma}\cdot\hbs{L}\,\Omega_{j\ell' m} &=& \left( \hbs{J}^2 -\hbs{L}^2-\left(\frac{1}{2}\,\hbar\,\hbs{\sigma}\right)^2 \right) \,\Omega_{j\ell' m}, \nonumber \\ &=& \Big( j(j+1) - \ell(\ell+1) - s(s+1) \Big) \,\hbar^2\,\Omega_{j\ell' m}, \quad \trm{for}\; s=\frac{1}{2}. \end{eqnarray} Further, one can now cleverly define $\kappa$ via \bea{kappa:defn} \kappa = \mp\,(j+\frac{1}{2})= \begin{cases} -(\ell+1), \quad\trm{for}\quad j=\ell+\frac{1}{2}, \\ \quad\ell, \hspace{0.55in}\trm{for}\quad j=\ell-\frac{1}{2}, \end{cases} \end{eqnarray} with $|\kappa| = j+\frac{1}{2}$ or $j=|\kappa|-\frac{1}{2}$. Then using $\ell' \equiv 2\,j-1$ and \Eq{Greiner:p171:18} one has (after some clever, but simple algebra, RQMWE, p171) that the righthand side of \Eq{Greiner:p171:16} can be written as $(2\hbar + \hbs{L}\cdot\hbs{\sigma})\,\Omega_{j\ell' m} = (1+\kappa)\,\hbar\Omega_{j\ell' m}$. If one had instead begun this whole calculation with $\Omega_{j\ell m}$ one would similarly arrive at $(2\hbar + \hbs{L}\cdot\hbs{\sigma})\,\Omega_{j\ell m} = (1-\kappa)\,\hbar\,\Omega_{j\ell m}$. These can be rearranged (by simply subtracting $\hbar\,\Omega_{j\ell' m}$ from both sides of the equality, and similarly with $\hbar\,\Omega_{j\ell m}$) in the form of and eigenvalue equation for the operator $\hat{\kappa}\overset{\trm{def}}{=} \hbar + \hbs{L}\cdot\hbs{\sigma}$ yielding $\hat{\kappa}\,\Omega_{j\ell m}= -\hbar\,\kappa\,\Omega_{j\ell m}$ and $\hat{\kappa}\,\Omega_{j\ell' m}= \hbar\,\kappa\,\Omega_{j\ell' m}$. Thus, one often defines the spinors $\chi_{{\kappa},\mu}\equiv\Omega_{j\ell m}$ and $\chi_{-{\kappa},\mu}\equiv\Omega_{j\ell' m}$ where $\mu=\pm\frac{1}{2}$ is the magnetic quantum number indexing spin up and spin down, respectively, so that these eigenvalue equations read equivalently as $\hat{{\kappa}}\,\chi_{{\kappa},\mu}= -\hbar\,{\kappa}\,\chi_{{\kappa},\mu}$ and $\hat{{\kappa}}\,\chi_{-{\kappa},\mu}= \hbar\,{\kappa}\,\chi_{-{\kappa},\mu}$. We can therefore write the (spatial portion) of the 4-spinor in a central potential for stationary states as \bea{Greiner:172:4:spinor} \psi_{j\ell m}(\bs{r}) &=& \left(\begin{array}{c}\varphi_{j\ell m}(\bs{r}) \\ \chi_{j\ell' m}(\bs{r},t)\end{array}\right) = \left(\begin{array}{c}i\,g(r)\,\Omega_{j\ell m}(\bs{r}/r) \\ -f(r)\,\Omega_{j\ell' m}(\bs{r}/r)\end{array}\right), \\ &=& \left(\begin{array}{c}i\,g(r)\,\chi_{{\kappa}, m} \\ -f(r)\,\chi_{-{\kappa}, m}\end{array}\right) = i\,\left(\begin{array}{c}\,g(r)\,\chi_{{\kappa}, m} \\ i\,f(r)\,\chi_{-{\kappa}, m}\end{array}\right). \end{eqnarray} Finally, using \Eq{Greiner:p171:16} and \Eq{Greiner:p171:12}, \Eq{Greiner:p171:10} takes the sought after form \be{Greiner:p172:22} (\hbs{\sigma}\cdot\hbs{p})\,\varphi_{j\ell m} = -\Omega_{j\ell' m} \left( \hbar\,\frac{d g(r)}{dr} + \hbar\,\frac{{\kappa}+1}{r}\,g(r) \right). \end{equation} Notice the parity flip from $\ell$ to $\ell'$ in going from the righthand to the lefthand side. Similarly, one derives \be{Greiner:p172:23} (\hbs{\sigma}\cdot\hbs{p})\,\chi_{j\ell' m} = -\Omega_{j\ell m} \left( \hbar\,\frac{d f(r)}{dr} - \hbar\,\frac{{\kappa}-1}{r}\,f(r) \right). \end{equation} We can now insert these expression into our coupled SR DE \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2} with the spherical spinors cancelling from both sides of both equations (the fruition of all the above involved labor) to obtain one of our final forms \bea{Greiner:p172:24} \hbar c\, \frac{dg(r)}{dr} + (1+{\kappa})\,\hbar c\,\frac{g(r)}{r} - [(E-V(r)) + m\,c^2]\,f(r) &=& 0, \\ \hbar c\, \frac{df(r)}{dr} + (1-{\kappa})\,\hbar c\,\frac{f(r)}{r} + [(E-V(r)) - m\,c^2]\,g(r) &=& 0. \end{eqnarray} With the standard substitution $G= r\,g$ and $F= r\,f$ ensuring $G(0)=F(0)=0$ at the origin singularity, the above equations further simplify to our final form \bea{Greiner:p172:24} \hbar c\, \frac{dG(r)}{dr} + \hbar c\,\frac{{\kappa}}{r}\,G(r) - [(E-V(r)) + m\,c^2]\,F(r) &=& 0, \\ \hbar c\, \frac{dF(r)}{dr} - \hbar c\,\frac{{\kappa}}{r}\,F(r)+ [(E-V(r)) - m\,c^2]\,G(r) &=& 0. \end{eqnarray} Again, if we wish to include a scalar coupling $V_s(r)$ to the mass, in addition to the vector coupling $V\to V_v(r)$ (as computed above), these equations become \bea{Greiner:with:vector:scalar:coupling} \hbar c\, \frac{dG(r)}{dr} + \hbar c\,\frac{{\kappa}}{r}\,G(r) - [(E-V_v(r)) + \big(m\,c^2+V_s(r)\big)]\,F(r) &=& 0, \label{Greiner:with:vector:scalar:coupling:line:1} \\ \hbar c\, \frac{dF(r)}{dr} - \hbar c\,\frac{{\kappa}}{r}\,F(r)+ [(E-V_v(r)) - \big(m\,c^2+V_s(r)\big)]\,G(r) &=& 0. \label{Greiner:with:vector:scalar:coupling:line:2} \end{eqnarray} The coupled first order equations \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2} are the SR extension of the second order radial equations in the NRQM SE. \subsection{Eigenenergies of the DE in a central potential with vector and scalar coupling}\label{DE:Solns:vector:scalar:coupling} We now wish to outline the solution for the eigenenergies of the stationary states of the DE containing both vector (Coulomb-like) and scalar (gravitational-like) coupling. Typically, for scalar coupling, the mass acts like a position dependent mass term, and is often associated with very massive particles such as the $\sigma$ meson so that the range is very short (inversely proportional to the mass). In quantum field theory (QFT) one describes this interaction in terms of the exchange of massless scalar mesons mediating the force, in analogy with the exchange of massless photons mediating the Coulomb force. Since we are dealing with SR particles, bound states mean that the magnitude of the energy of the stationary state is less than the rest mass, i.e. $-m\,c^2<E< m\,c^2$, where positive energies are associated with bound particles and negative energies with bound anti particles. Energies outside this range are associated with continuum states for both particles and anti-particles. The strategy to solve \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2} is analogous to that of the solution of the RNQM SE, except that now one must contend with two coupled first order radial equations for the DE, vs a single second order equation for the SE. This means there will ultimately be two coupled power series expansions, but once again the quantization condition will involve the truncations of these infinite series to polynomials to ensure that the boundary condition at spatial infinite (i.e. the wave function approaches zero) is satisfied, which in turn implies the functions are square integrable, thus ensuring the finiteness of the associate probabilities over all space. Again, one begins by examining the solution near the origin to develop a leading non-negative exponent to $r$ so that the wave function is also well behaved (i.e. zero) at the origin (the location of the point source). The details of such a calculation for the case of the DE with both vector and scalar coupling is explicitly worked out in RQMWE (section 9.8, pp184-187, see also section 9.6 p178-182 for the Coulomb solution only). Here, we just outline the highlights, but strongly encourage the reader to go through the detailed worked problems in RQMWE.\cite{Greiner:1990}. Using the convention of RQMWE we write $V_v(r) = -\alpha/r$ for the vector coupling (e.g. $\alpha = e^2/\hbar c = \alpha_C$) and $V_s(r) = -\hbar c \alpha'/r$ for the scalar coupling (e.g. $\alpha' = GMm/\hbar c = \alpha_G$). From \Eq{Greiner:with:vector:scalar:coupling:line:1} and \Eq{Greiner:with:vector:scalar:coupling:line:2}, the radial equations take the form \bea{Greiner:p184:2} \frac{dG(r)}{dr} &=& -\frac{{\kappa}}{r}\,G(r) + \left[ \frac{E + m\,c^2}{\hbar c} + \frac{\alpha-\alpha'}{r}\right]\,F(r), \label{Greiner:p184:2:line1} \\ \frac{dF(r)}{dr} &=& \frac{{\kappa}}{r}\,F(r) - \left[ \frac{E - m\,c^2}{\hbar c} + \frac{\alpha+\alpha'}{r}\right]\,G(r). \label{Greiner:p184:2:line2} \end{eqnarray} We first consider the region $r\sim 0$, where we can then keep only the $1/r$ terms and drop the constant terms. With the ansatz $G=a\,r^{\gamma}$ and $G=b\,r^{\gamma}$ were are lead to a homogeneous set of linear equations $a({\gamma}+{\kappa}) - b(\alpha-\alpha')=0$ and $a(\alpha-\alpha')+b({\gamma}-{\kappa}) =0$, which upon setting the determinant of the coefficients equal to zero yields ${\gamma} = \pm \sqrt{{\kappa}^2 -\alpha^2+\alpha^{'2}}$. To allow for normalization of the wave functions, one must select the positive sign for ${\gamma}$. One can now define $\lambda \overset{\textrm{def}}{=} \frac{m^2 c^4-E^2}{\hbar c}$ and $\rho = 2\lambda r$ so that the radial equations become \bea{Greiner:p185:11} \frac{dG}{d\rho} &=& -\frac{{\kappa}}{\rho}\,G + \left[ \frac{E + m\,c^2}{2\hbar c\,\lambda} + \frac{\alpha-\alpha'}{\rho}\right]\,F, \\ \frac{dF}{d\rho} &=& \frac{{\kappa}}{\rho}\,F -\left[ \frac{E - m\,c^2}{2\hbar c\,\lambda} + \frac{\alpha+\alpha'}{\rho}\right]\,G. \end{eqnarray} It turns out to be convenient to define the functions $\phi_1(\rho)$ and $\phi_2(\rho)$ via $G = \sqrt{mc^2+E}\, e^{-\rho/2}\,$ $(\phi_1+\phi_2)$ and $F = \sqrt{mc^2+E}\, e^{-\rho/2}\,(\phi_1-\phi_2)$ and expand each in a power series given by $\phi_1=\rho^{\gamma}\,\sum_{m=0}^\infty \alpha_m\,\rho^m$ and $\phi_2=\rho^{\gamma}\,\sum_{m=0}^\infty \beta_m\,\rho^m$. The justification for introducing $\phi_1$ and $\phi_2$ is that resulting equation for $\phi_2$ allows one to write the coefficients $\beta_m$ in terms of the $\alpha_m$, via (see RQMWE, p186) \bea{Greiner:p186:17} \frac{\beta_m}{\alpha_m} &=& \frac { -{\kappa} + \alpha m c^2/\hbar c\lambda + \alpha' E/\hbar c\lambda } { m+{\gamma} - \alpha m c^2/\hbar c\lambda - \alpha'E/\hbar c\lambda }, \nonumber \\ &=& \frac { {\kappa} - \alpha m c^2/\hbar c\lambda - \alpha' E/\hbar c\lambda } {n'-m}, \nonumber \\ \trm{with}\;\; n' &\overset{\textrm{def}}{=}& \frac{\alpha E}{\hbar c \lambda} + \frac{\alpha' m c^2}{\hbar c \lambda}-{\gamma}. \end{eqnarray} Inserting this into the series relationship from $\phi_1$, which involves $\alpha_m$ expressed in terms of $\alpha_{m-1}$ and $\beta_m$, one can develop an expression for $\alpha_m/\alpha_{0}$ whose numerator involves the product $(n'-1)(n'-2)\ldots(n'-m)$. Once again, if the series do not terminate, they lead to a sum that scales as $e^{\rho}$ so that the wave function does not converge at spatial infinity. Thus, as in the SE quantization case, requiring the series to terminate, demands that $n'=\{0,1,2,\ldots\}$. Inserting this result back into \Eq{Greiner:p186:17} this equation produces a quadratic equation for the energy $E$ (since $\lambda\propto E$), which can be easily solved to finally give \bea{Greiner:p186:20} \frac{E}{mc^2} &=& \left\{ \frac{-\alpha\alpha'}{\alpha^2 + (n-(j+\frac{1}{2})+\gamma)^2} \pm \left[ \left( \frac{\alpha\alpha'}{\alpha^2 + (n-(j+\frac{1}{2})+\gamma)^2} \right)^2 \right. \right. \nonumber \\ &-& \left. \left. \frac { \alpha^{'2} - (n-(j+\frac{1}{2})+\gamma)^2 } { \alpha^{2} + (n-(j+\frac{1}{2})+\gamma)^2 } \right]^{1/2} \right\}. \end{eqnarray} In the above, the principal quantum number $n$ is defined as $n=n'+|{\kappa}| = n' + j + \frac{1}{2}$ with $n\in\{1, 2, \ldots\}$. Special cases of the above are both interesting and relevant. \subsubsection{Vector Coupling only} Here $\alpha'=0$ and ${\gamma} = \sqrt{{\kappa}^2-\alpha^2}$ apropos for the Coulomb field and one obtains \be{Greiner:p186:case:2} \frac{E}{mc^2} = \left[ 1+ \frac{\alpha^2}{(n-(j+\frac{1}{2})+{\gamma})^2} \right]^{-1/2} \approx 1 - \frac{1}{2}\frac{\alpha^2}{n^2} + \ldots, \end{equation} recovering the SE result since $\alpha=\alpha_C \approx 1/137\ll 1$. (Note: the negative energies do not fulfill the original quadratic equation \Eq{Greiner:p186:17} for $E$ in the case of vector coupling since both $n'+{\gamma}$ and $\frac{\alpha\,E}{\hbar c \lambda} + \frac{\alpha' m c^2}{\hbar c \lambda}$ are positive, and so that the $E<0$ solutions must be excluded). \subsubsection{Scalar Coupling only} In the case of scalar coupling $\alpha=0$ and ${\gamma} = \sqrt{{\kappa}^2+\alpha^{'2}}$ (apropos for a gravitational-like field), one obtains instead the expression \be{Greiner:p186:case:1} \frac{E}{mc^2} = \pm\left[ 1- \frac{\alpha^{'2}}{(n-(j+\frac{1}{2})+{\gamma})^2} \right]^{1/2} \approx \pm\left(1 - \frac{1}{2}\frac{\alpha'^2}{n^2} + \ldots\right), \quad\trm{if}\quad \alpha'\ll 1. \end{equation} Thus, the above two formulas \Eq{Greiner:p186:case:2} and \Eq{Greiner:p186:case:1} almost answer the student's original question when we note the similarity of the placement of the square roots when comparing \Eq{E_v} with \Eq{Greiner:p186:case:2}, and \Eq{E_s} with \Eq{Greiner:p186:case:1}. We see that in the small coupling limit both formulas have the form of $1-\frac{\alpha^2}{ 2 n^2}$, but this requires that $\alpha'=GMm/\hbar c = \frac{M}{M_p} \frac{m}{M_p}\ll 1$ (see \Eq{alpha:G}). While this is true say for $\sigma$ mesons and other elementary particles, for the originally posed question of a solar mass BH, $\frac{M}{M_p}$ is astronomically huge, so that $\alpha'\to\alpha_G\approx 4\times 10^{15}\gg 1$ (to say the least!). Still, one's intuition is somewhat borne out by this analysis, but not fully yet answered to the student's satisfaction. Thus, in order to fully answer the student's question, we need to turn now to the DE in curved spacetime where the role of the metric encoding the gravitational field enters in a fundamentally new way. Further, the strange new concept of a horizon, dividing spacetime into two distinct regions: outside and inside, with its conceptually unusual (and confounding) ``one-way membrane" property (i.e. particles can cross the horizon from outside to in, but not the reverse!) rears it strange and complicated head. \section{The DE in curved spacetime}\label{sec:DE:CST} General Relativity (GR) presents its own unique challenges, both conceptually and mathematically, even to the motivated student. However, most students are familiar at least with some basic concepts, including the curved spacetime metric, black holes (BH) and particle orbits about them, and the bending of light rays around a massive object, even if they might not be quite capable of deriving them. Therefore, while there is some degree of heavy lifting involved when learning GR (the fundamentally new concept of spacetime vs space and time, tensors, covariant derivatives, parallel transport, etc...) there currently exists enough excellent GR textbooks now geared to the upper level undergraduate that attempts to ease these burdens. One excellent text that we will use here is that by James Hartle\cite{Hartle:2009}, which emphasizes ``physical concepts first" before mathematical manipulation. For most students (of all ages!) it's the concepts that both SR and GR introduce that are at first glance difficult to wrap one's head around. But, 100 years on now, there has been enough well explained exposition (and textbooks) that physics such as BHs and the bending of light are now part of general (and popular) knowledge. Therefore, the goal here is not review all of GR, but just those salient points that will get us to our goal most expeditiously. On top of this, we also want to include the non-standard (for the novice) topic of the DE in curved spacetime (CST). To that end, there does exist an excellent undergraduate GR text by Lewis Ryder\cite{Ryder:2009} that clearly and lucidly deals with this topic, leading to the DE in CST (see Ryder, Chapter 11.3, pp409-416). Again, it is not our goal to reproduce all of Ryder's illuminating discussion. In the spirit of wanting to drive our car first before knowing how the engine under the hood was built (but allowing ourselves to read the service manual now and again), we again will point out the key concepts involved and refer the reader to Ryder's excellent GR textbook. In the following we will first draw from Hartle\cite{Hartle:2009} (see section 9.3, p191-204) in order to get to a form of the classical GR radial equation (no quantum here!) that looks like an ordinary NR energy equation. This involves being given the Schwarzschild metric (sans derivation), and manipulating it (the 4-velocity) in order to produce the desired equation. The goal is to identify the GR radial potential for comparison to the DE in CST. We then next switch gears, and focus on the derivation of the DE in CST (Ryder, Chapter 11.3, pp409-416). The key concept from GR is the covariant derivative, required to explain how to take derivatives in the surrounding CST as one moves from point to point. As mentioned earlier, the global inertial frames of SR are now ``collapsed" to local regions around each point which (following Hartle) we refer to as the observer's local laboratory. As long as the extent of the observer's spatial axes, and the duration of time measured, are in a sense ``small" (with respect to changes in the curvature of the CST), the Equivalence Principle holds, and hence physics appears special relativistic. Thus, one must be able to describe the observer's local (frame) laboratory, and hence we are led down the road of introducing a tetrad of 4-vectors describing the local laboratory. Measurements and description of particles (e.g. their 4-momentum) passing through the observer's local laboratory are then made by ``projecting" then onto the observer's local axes (3 space and 1 time axes). The introduction of tetrads is the key concept required then to similarly ``project" the DE in the surrounding CST ``into" the observer's local frame, which is locally (Minkowski) flat (i.e where SR holds). The relevant concept could be stated as such: ``Since we already know how to write down and solve the DE in SR (see the previous section), then by use of the Equivalence Principle, the strategy is to ``simply" project the CST DE into the observer's locally flat frame at each spacetime point $x$, where we already know how to solve it." This is what Ryder does in Chapter 11.3. Again, our objective is to get to the DE in CST as quickly as possible, so again we point out the major highlights of this derivation (with an emphasis more on the why and how we get there, which can be daunting to the un-initiated, rather than on the detailed proofs - which are shown in Ryder). So let us begin. \subsection{The effective radial potential for particles in the Schwarzschild metric}\label{sec:Veff:SST} As is well known, the Schwarzschild metric for a central symmetric gravitational source of mass M is (in full units, see Hartle, p186) \bea{Schw:metric} \hspace{-0.5in} ds^2 &=&-{\mathcal{F}}(r)\,c^2\,dt^2 + \frac{1}{{\mathcal{F}}(r)}\,dr^2 + r^2\,d\Omega^2, \;\; {\mathcal{F}}(r) \overset{\textrm{def}}{=} 1-\frac{2GM}{c^2 r}, \;\; d\Omega^2 = \left(d\theta^2 + \sin^2\theta\,d\phi^2 \right), \label{Schw:metric:line:1} \\ \hspace{-0.5in} &=& g_{\mu\nu}\,dx^\mu\,dx^\nu, \qquad \mu,\nu\in\{0,1,2,3\}, \label{Schw:metric:line:2}\\ \hspace{-0.5in} &\approx& -\left(1+\frac{2 V_G(r)}{c^2}\right)\,c^2\,dt^2 + \left(1-\frac{2 V_G(r)}{c^2}\right)\,dr^2 + d\Omega^2, \qquad V_G(r) = -\frac{G M}{r},\label{Schw:metric:line:3} \end{eqnarray} where we have defined the Schwarzschild factor ${\mathcal{F}}(r) \overset{\textrm{def}}{=} 1-\frac{2GM}{c^2 r}$ that will figure prominently throughout the discussions of Schwarzschild spacetime (SST). \Eq{Schw:metric:line:3} is the Schwarzschild metric in the weak field approximation (apropos for, say about the Sun or the Earth), in which the Newtonian potential $V_G(r) = -\frac{G M}{r}$ explicitly appears. The relevant scale length (see \App{app:Units}) is the well known Schwarzschild radius $r_s\overset{\textrm{def}}{=} 2GM/c^2$ (so that $2 V_G/c^2 = -\left(\frac{r_s}{r}\right)$ is unitless). For a stationary observer (i.e. one that sits at a fixed coordinate position) at fixed $r$, the metric yields $ds^2 \overset{\textrm{def}}{=} -d\tau^2 = -{\mathcal{F}}\,c^2\,dt^2 = g_{00} (dx^0)^2$ (taking $x^0= c t$) where $\tau$ is then seen as the \tit{proper time}, i.e. the time as measured on a clock carried with the observer. Writing the metric as $-d\tau^2 =ds^2 = g_{\mu\nu}\,\left(\frac{dx^\mu}{d\tau^2}\right)\, \left(\frac{dx^\nu}{d\tau^2}\right)\, d\tau^2 \overset{\textrm{def}}{=} \textrm{\bf u}^2_{obs}\, d\tau^2 $ we have $\textrm{\bf u}^2_{obs}~=~-~1$, where we have defined the observer's 4-velocity $\textrm{\bf u}^\mu = dx^\mu/d\tau$, i.e. the velocity of the observer (rate of change of their coordinates) with respect their proper time $\tau$ (vs their coordinate time $t=t(\tau)$). The observer's 4-velocity will be in fact the ``time axis" of the local local laboratory, which we denote as $\textrm{\bf e}_0(x) \overset{\textrm{def}}{=} \textrm{\bf u}_{obs}(x)$ a timelike unit vector (i.e. has magnitude $-1$) and is tangent to the (freely falling geodesic) trajectory of the observer's motion (worldline) through the surrounding CST. The other three spatial axes of the observer's local laboratory, denoted as $\{\textrm{\bf e}_1(x), \textrm{\bf e}_2(x), \textrm{\bf e}_3(x)\}$ are chosen orthogonal to $\textrm{\bf e}_0(x)$ (see \Fig{fig:observers:local:laboratory}), and will be discussed later. For now, in order to extract radial orbits of the Schwarzschild metric, we only require $\textrm{\bf u}_{obs}(x)$. Note that if a particle of 4-momentum $\textrm{\bf p}$ passes through the observer's local laboratory, then the observer will measure its energy as $E/c= -\textrm{\bf p}\cdot\textrm{\bf u} = -\textrm{\bf p}\cdot\textrm{\bf e}_0$, and its 3-momentum components as $p_a=\textrm{\bf p}\cdot\textrm{\bf e}_a$ for $a\in\{1,2,3\}$ yielding a local description of the particle with a SR-like local laboratory 4-vector $(E/c, \bs{p})$ (as an instantiation of the Equivalence Principle). \begin{figure}[h] \includegraphics[width=4.5in,height=3.0in]{fig_1_DE_SST_AJP_26Jun2022} \caption{The observer's \tit{local laboratory} (small room with physicist) at the curved spacetime (CST) point $x$, defined by the orthonormal tetrad $\textrm{\bf e}_a(x)$, $a=(0,1,2,3)$. The three spatial axes $\textrm{\bf e}_i(x)$, $i=(1,2,3)$ are located at the origin of the observer's laboratory (the universal coordinates system ``in the corner of the room"), while $\textrm{\bf e}_0(x)=\textrm{\bf u}_{obs}(x)$ is the temporal axis, defined as the observer's 4-velocity, which is the tangent to the physicist's geodesic trajectory. A particle of 4-momentum $\textrm{\bf p}(x) = m \textrm{\bf u}(x)$ and world components $p^\alpha(x)$ passes through the observer's local laboratory. The observer measures the \tit{local} components $p^a(x) = e^a_{\;\;\alpha}(x)\,p^\alpha(x)$. At a small proper time later $d\tau$, the particle has moved from $x^\alpha \to x^{'\alpha} = x^\alpha + u^\alpha(x)\,d\tau$, which is measured by the observer in their local laboratory at the spacetime point $x^{'\alpha}$. } \label{fig:observers:local:laboratory} \end{figure} Particles (here both massive and massless until otherwise specified) under no other external forces (i.e. accelerations) travel on \tit{geodesics}, the analogue of straight line motion in Euclidean space, extended to CST. Such observers are called freely falling (FF). Now without delving into the geodesic equation per say (since we will not need to solve this equation directly in the subsequent discussion), an important concept is that of conserved quantities along the geodesic motion. If the metric is independent of a particular coordinate, then a conserved quantity exist along this motion (called an isometry; this is essentially Noether's theorem applied to geodesics, see Hartle, section 8.2, pp175-178). For example, for the Schwarzschild metric, we see that it is independent of the coordinates $t$ (stationarity) and azimuthal angle $\phi$ (rotationally invariant about the $z$-axis). Let $\bs{\xi}$ be a coordinate vector along the the \tit{isometry} (i.e. motion along which the metric does not change). % $\bs{\xi}$ is called a \tit{Killiing vector} (after Wilhelm Killing (1847-1923), see Hartle, Chapter 8.2, pp175-178) and there exists an equation named after him, that allows one to derive the isometries systematically for any metric. However, quite often one can deduce the Killing vectors by the symmetry of the metric by inspection. % The key result is that $\bs{\xi}\cdot\textrm{\bf u}_{obs}$ is a conserved quantity all along the geodesic. For the Schwarzschild metric, the metric does not change for translations in time $t$, so that $\bs{\xi}_t = (1,0,0,0)$, and therefore we define the quantity $e\overset{\textrm{def}}{=} -\bs{\xi}_t\cdot\textrm{\bf u}_{obs}$, which can be physically interpreted as the particles rest energy per unit mass (at large $r$). % In the following we will follow the convention of Hartle (and most GR textbooks) and work in units of $G=c=1$. Physical units can be restored by resorting to dimensional analysis. A second conserved quantity for particles on Schwarzschild geodesics is the $\ell = \bs{\xi}_\phi\cdot\textrm{\bf u}_{obs}$ the orbital angular momentum per unit mass (at large $r$), with $\bs{\xi}_\phi = (0,0,0,1)$ (in spherical polar coordinates $(t, r, \theta, \phi)$) indicating the independence of the Schwarzschild metric in the azimuthal coordinate $\phi$. Using these two conserved quantities and the timelike unit magnitude of the observer's 4-velocity, there exists enough symmetry to deduce the particle orbits directly. Note that since the orbital angular momentum is conserved, the geodesic motion occurs in a plane, which is conventionally (for convenience) take to be the equator, $\theta=\pi/2$. In the following we will work in a \tit{coordinate basis} which means that $\textrm{\bf e}_a$ are just coordinate derivatives in the direction indicated and their inner products gives the metric components. This means that for two arbitrary 4-vectors $\bs{a}$ and $\bs{b}$ with coordinate components $a^\mu$ and $b^\nu$, respectively, we have $\bs{a}\cdot\bs{b}= g_{\mu\nu}a^\mu\,b^\nu$. Thus, for the Schwarzschild metric we define the conserved quantities as (following Hartle and using units of $G=c=1$) \bea{Hartle:p193:9.21:9.22} e &=& -\bs{\xi}_t\cdot\textrm{\bf u} = \left(1-\frac{2 M}{r}\right)\,\frac{dt}{d\tau}={\mathcal{F}}(r)\,\frac{dt}{d\tau}, \label{Hartle:p193:9.21} \\ \ell &=& \bs{\xi}_\phi\cdot\textrm{\bf u} = r^2\sin^2\theta\,\frac{d\phi}{d\tau}. \label{Hartle:p193:9.22} \end{eqnarray} With the above conserved quantities in hand, and $\textrm{\bf u} = (u^t, u^r, u^\theta, u^\phi)$ we have from its normalization $\textrm{\bf u}_{obs}\cdot\textrm{\bf u}_{obs}=-1$, \bea{Hartle:p194:9.25} -1 &=&-\left(1-\frac{2M}{r} \right)\,\left(u^t\right)^2 + \left(1-\frac{2M}{r} \right)^{-1}\,\left(u^r\right)^2 + r^2\,\left(u^\phi\right)^2, \\ \trm{or}\quad -1 &=& -\left(1-\frac{2M}{r} \right)^{-1}\,e^2 + \left(1-\frac{2M}{r} \right)^{-1}\,\left(\frac{dr}{d\tau}\right)^2 + \frac{\ell^2}{r^2}, \\ \Rightarrow\quad \mathcal{E} \overset{\textrm{def}}{=} \frac{e^2-1}{2} &=& \frac{1}{2}\,\left(\frac{dr}{d\tau}\right)^2 + \left[ \left(1-\frac{2M}{r} \right)\,\left(1+\frac{\ell^2}{r^2} \right)-1 \right], \\ \trm{or}\quad \mathcal{E}&=& \frac{1}{2}\,\left(\frac{dr}{d\tau}\right)^2 + V_{eff}(r), \label{Netwon:E:eqn} \end{eqnarray} where we have defined the effective potential (see Hartle, p194) \bea{V:eff} V_{eff}(r)\equiv \left[ \left(1-\frac{2M}{r} \right)\,\left(1+\frac{\ell^2}{r^2} \right)-1 \right] &=& -\frac{M}{r} + \frac{\ell^2}{2\,r^2} - \frac{M\,\ell^2}{r^3}, \label{V:eff:line1} \\ &=& \frac{1}{c^2}\, \left( -\frac{GM}{r} + \frac{\ell^2}{2\,r^2} - \frac{G M\,\ell^2}{c^2\,r^3} \label{V:eff:line2} \right), \end{eqnarray} where in the last line we have restored all the physical constants. \Eq{Netwon:E:eqn} has the form of a NR energy equation with a Newtonian-like potential with orbital angular momentum barrier (first two terms of \Eq{V:eff:line2}), but with an additional attractive GR correction (last term of \Eq{V:eff:line2}) that scales as $1/r^3$. This last term dominates for small $r$ and is responsible for the characteristic GR effects such as the precession of the perihelion, bending of light, etc. We can make the Newtonian analogy even stronger by defining $E_{Newt}$ via $e^2 = (mc^2 + E_{Newt})/mc^2$, i.e. as the small correction to the particle's rest mass, in strong analogy to SR. With this substitution, the radial equation becomes (see Hartle p195, restoring full units) \be{Hartle:p195:9.32} E_{Newt} \approx \frac{m}{2}\,\left(\frac{dr}{d\tau}\right)^2 + \frac{L^2}{2 m r^2} -\frac{G M m}{r} - \frac{G M L^2}{c^2 m r^3}, \qquad L\overset{\textrm{def}}{=} m\,\ell, \end{equation} where we have approximated $\mathcal{E}=(e+1)(e-1)/2\approx (e-1)$. \Eq{Hartle:p195:9.32} now has the exact same form as the energy integral in Newtonian gravity with an additional relativistic correction to the potential proportional to $1/r^3$. Note that \be{Hartle:p195:9.33} V_{eff}(r) \underset{r\to\infty}{\longrightarrow} -\frac{G M}{c^2\,r} = -\frac{r_s/2}{r}, \qquad V_{eff}(r_s) = -\frac{1}{2}. \end{equation} Further, at the Schwarzschild radius $r=r_s$ the first term of $V_{eff}(r_s)=-\frac{1}{2}$, and the second and thrid terms exactly cancel each other. A detailed investigation for both radial plunge orbits ($\ell=0$), circular and hyperbolic orbits ($\ell\ne 0$) are well explored in Hartle Chapter 9 (and many other GR textbooks), and will not be covered here. For our goals, we are primarily interested in radial plunge orbits, since in this case $V_{eff}(r) = -\frac{G M}{c^2\,r}=-\frac{r_s/2}{r}$ has the exact form of the scalar coupling potential, discussed previously for the SR DE. The question now, is how to incorporate the GR effects into the DE. We turn to this next, leveraging the discussion in Ryder\cite{Ryder:2009} (Chapter 11.3). \subsection{The DE in CST: GR preliminaries}\label{sec:DE:CST:preliminaries} In order to arrive to our destination, the DE in CST, we first need to stop once more at the discussion of the description of the observer's local laboratory. We need to distinguish between the use of various ``basis vectors" and corresponding ``1-forms" used to describe the observer's frame. The reason is that since the observer's local (frame) laboratory is described in terms of four basis vectors $\textrm{\bf e}_a$, we want to distinguish between coordinate bases (cryptically called \tit{holonomic} in the literature) and an orthonormal set of basis vectors (called, \tit{non-holonomic}). Why the two descriptions? In the spirit of GR, \tit{any} set of basis vectors are allowed, but these two are the one's most conveniently employed. Coordinate basis vectors are the ``easiest" to use and formally of the form $\textrm{\bf e}_\mu = \partial_\mu \overset{\textrm{def}}{=} \partial/\partial x^\mu$ for coordinates $x^\mu$. (Here we have used $a\to\mu$ since the index $\mu$ reflects the surrounding CST). The defining property of a coordinate basis set is that their inner (dot) product defines the metric (to be shown shortly), and that the basis vectors commute, namely $[\textrm{\bf e}_\mu, \textrm{\bf e}_\nu]=[\partial_\mu, \partial_\nu]= 0$ for $\mu\ne\nu$ reflecting the independence of the order of coordinate differentiation, i.e. $\partial_\mu\,\partial_\nu= \partial_\nu\,\partial_\mu$. On the other hand, the orthonormal basis is more physical, and better suited to what the observer actually measures. It's defining property is that the inner product between the orthonormal basis vectors defines the \tit{local} metric, which in this case is the flat Minkowskian metric of SR. The orthonormal basis vectors are of the form $\textrm{\bf e}_a = e_a^{\;\;\mu}(x)\,\partial_\mu$ where the \tit{tetrad components} $e_a^{\;\;\mu}(x)$ are spatially dependent. Thus, $[\textrm{\bf e}_a, \textrm{\bf e}_b]\ne 0$ in general. However, they do have the simplifying property that $\textrm{\bf e}_a\cdot\textrm{\bf e}_b = \eta_{ab} = \trm{diagonal}(-1,1,1,1)$ the constant Minkowski SR metric - which is just a statement of the Equivalence Principle. (Note, the relativist's minus sign convention in front of the time component, vs the particle physicists convention of using $\eta_{ab} = \trm{diagonal}(1,-1,-1,-1)$. If you pick up a random book an look at the ``sign of the times", you can instantly tell if the author is a particle physicist or a relativist, without even looking at the title. Try it!). To recap this important distinction, we recap this once more below: \bea{basis:vectors} \trm{coordinate basis:}\qquad \textrm{\bf e}_\mu &=& \partial_\mu, \qquad\quad \textrm{\bf e}_\mu\cdot\textrm{\bf e}_\nu = g_{\mu\nu}(x), \\ \trm{orthonormal basis:}\qquad \textrm{\bf e}_a &=& e_a^\mu\partial_\mu, \qquad \textrm{\bf e}_a\cdot\textrm{\bf e}_a = \eta_{ab}(x). \end{eqnarray} (Note: to distinguish the two basis, some authors use a circumflex $\verb+^+$ over the index $\hat{a}$ so that $\textrm{\bf e}_0$ and $\textrm{\bf e}_{\hat{0}}$ denote the observer's 4-velocity in a coordinate and orthonormal basis, respectively. We will no have occasion to do this, since we will primarily use an orthonormal basis for the DE in CST). An example speaks a thousand words. Consider the simplest case of polar coordinates in two spatial dimensions, with coordinates $(r,\theta)$ such that $x=r\,\cos\theta$ and $y=r\,\sin\theta$ with line element $ds^2 = dr^2 + r^2 d\phi^2$. Then the coordinate basis vectors would be $\textrm{\bf e}_r = \partial_r$ and $\textrm{\bf e}_\phi = \partial_\phi$ and we have $\textrm{\bf e}_r\cdot\textrm{\bf e}_r = 1 = g_{rr}$ and $\textrm{\bf e}_\phi\cdot\textrm{\bf e}_\phi = r^2= g_{\phi\phi}$ with $\textrm{\bf e}_r\cdot\textrm{\bf e}_\phi=0$, leading to the metric $g_{\mu\nu} = \tiny{\left(\begin{array}{cc} 1 & 0 \\0 & r^2\end{array}\right)}$. Clearly $[\partial_r, \partial_\phi]=0$. In an orthonormal basis we would instead define $\textrm{\bf e}_r = \partial_r$ and $\textrm{\bf e}_\phi = \frac{1}{r}\,\partial_\phi$, with $\textrm{\bf e}_r\cdot\textrm{\bf e}_r = 1 = g_{rr}$ and $\textrm{\bf e}_\phi\cdot\textrm{\bf e}_\phi = 1= g_{\phi\phi}$ with $\textrm{\bf e}_r\cdot\textrm{\bf e}_\phi=0$, leading to the metric $g_{ab} = \tiny{\left(\begin{array}{cc} 1 & 0 \\0 & 1\end{array}\right)} = \delta_{ab}$. However, we now have $[\textrm{\bf e}_r, \textrm{\bf e}_\phi]\,f(r,\phi) = [\partial_r, \frac{1}{r}\,\partial_\phi]\,f =-\frac{1}{r^2}\partial_\phi\,f = -\frac{1}{r}\textrm{\bf e}_\phi\,f$ which allows one to conclude (since $f$ was an arbitrary function) that $[\textrm{\bf e}_r, \textrm{\bf e}_\phi]= -\frac{1}{r}\textrm{\bf e}_\phi \equiv \textrm{\bf e}_\phi\,C^{\phi}_{\;\;r \phi}(x)$. In the last step we have introduced the structure constants $C(x)$ defined by the commutators of the basis vector, $[\textrm{\bf e}_a, \textrm{\bf e}_b] =\textrm{\bf e}_c\, C^{ c}_{\;\;a b}(x)\,$ (with sum over the index $c$, see Hartle, p109), which just states that the commutator of the basis can be expanded in terms of a linear combination of the basis vectors. A general vector (an, in general, tensor) $\bs{v}$ is a geometric object, independent of the coordinates and basis vectors (user's laboratory frame) used to describe it. So we can write this as $\bs{v} = v^\mu(x)\,\textrm{\bf e}_\mu(x)$ using a coordinate basis vector description (with \tit{contravariant} coordinate components $v^\mu$), or as $\bs{v} = v^a(x)\,\textrm{\bf e}_a(x)$ using an orthonormal basis vector description (with physical orthonormal components $v^a(x)$). Note that the transition from coordinate basis to orthonormal basis (for this diagonal metric) can be performed by inspection by examining the metric and grouping terms as $ds^2 = (dr)^2 + (r d\phi)^2$ and somehow considering it's ``inverse," namely $(\partial_r, \frac{1}{r}\partial_\phi)$. Note that $r\,d\phi$ is not a total differential, so there is no coordinate associated with such an object. This intuition can be formalized by defining basis \tit{1-forms} which are \tit{dual} to the orthonormal basis vectors $\textrm{\bf e}_a$. We denote these 1-forms (generalized differentials) as $\bs{\theta}$ (Ryder's notation) which can be decomposed in terms of the coordinate (true) differentials $dx^\mu$ via $\bs{\theta}^a = e^a_{\;\;\mu}(x) dx^\mu$. Note that we have purposely used the same symbol $e$ for the components of the one form, with the important distinction that the coordinate index $\mu$ is now on the bottom and the orthonormal index $a$ is on top (vs $\textrm{\bf e}_a^{\;\;\mu}$ associated with the basis vectors $\textrm{\bf e}_a = e_a^{\;\;\mu}(x)\,\partial_\mu$). Thus, even without a metric, we can define what we mean by dual by saying that a 1-form is an object that ``eats" vectors in the following sense (in both coordinate and orthonormal bases) $dx^\mu(\partial_\nu)\overset{\textrm{def}}{=} \delta^\mu_{\;\;\nu}$, and $\bs{\theta}^a(\textrm{\bf e}_b)\overset{\textrm{def}}{=} \delta^a_{\;\;b}$. Thus, consider a general 1-form $\bs{w} = w_a(x)\bs{\theta}^a$ with (covariant) components $w_a(x)$. Then, we can have this 1-form act on a vector $\bs{v} = v^a(x)\,\textrm{\bf e}_a$ to give $\bs{w} (\bs{v}) = w_a\bs{\theta}^a (v^b\,\textrm{\bf e}_b) = w_a \,\bs{\theta}^a(\textrm{\bf e}_b)\,v^b = w_a\,\delta^a_{\;\;b}\, v^b = w_a v^a \overset{\textrm{def}}{=} \bs{w}\cdot\bs{v}$. Therefore, we can define and inner product without the need for a metric, and the metric merely serves to raise and lower components via $g_{\mu\nu} v^\nu = v_\mu$ and $\eta_{ab} v^a = v_b$. \subsection{The DE in CST: covariant derivatives in GR}\label{sec:DE:CST:covar:deriv:GR With these preliminaries out of the way, we now get to the crux of the matter (and the part that is often an initial learning bottleneck for the beginning GR student). How does one define the derivative of basis vectors? Let's first work in the simpler coordinate basis vectors. Consider the following calculation for a posited, yet unknown, derivative which we denote as $\nabla_\mu$: $\nabla_\mu \bs{v} = \nabla_\mu\big(v^\nu(x) \textrm{\bf e}_\nu(x)\big) = (\partial_\mu v^\nu)\,\textrm{\bf e}_\nu + v^\mu (\nabla_\mu\textrm{\bf e}_\nu)$. Here, we have made the (natural) assumption that $\nabla_\mu\to\partial_\mu$ on functions (of which $v^\nu$ are). Acting on basis vectors, we are yet unsure, so we just leave it as $\nabla_\mu$. This states that we must not only differentiate the components $v^\nu(x)$ of $\bs{v}$, but also its basis vectors $\textrm{\bf e}_\nu(x)$. But we are already familiar with this latter concept, since even in our simple 2D polar coordinate example above, the basis vector $\textrm{\bf e}_\phi = \partial_\phi$ is tangent to circles of constant radius $r$, and thus change direction (in the surrounding 2D Euclidean $\mathbb{R}^2$ space) as we vary the coordinate $\phi$. As in the case of the 1-forms above, we assume that $\nabla_\mu\textrm{\bf e}_\nu$ can be expanded in terms of the basis vectors, and so we write $\nabla_\mu\textrm{\bf e}_\nu(x) = {\Gamma}_{\mu\nu}^\lambda(x)\,\textrm{\bf e}_{\lambda}$ with proportionality functions ${\Gamma}_{\mu\nu}^\lambda(x)$, the famous \tit{Levi-Civita connection}, which informs us as to how the basis vectors change as we move from point $x^\mu$ to point $x^\mu+u^\mu\,d\tau$ in the CST. Inserting this into the full expression we have $\nabla_\mu \bs{v} = (\partial_\mu v^\nu)\,\textrm{\bf e}_\nu + {\Gamma}_{\mu\nu}^\lambda\textrm{\bf e}_\lambda$ which we can relabel dummy indices via $\lambda\leftrightarrow\nu$ to obtain \be{cov:deriv} \nabla_\mu \bs{v} = \left(\partial_\mu v^\nu + {\Gamma}_{\mu\lambda}^\nu v^{\lambda} \right) \textrm{\bf e}_\nu \overset{\textrm{def}}{=} v^\nu_{;\mu}\,\textrm{\bf e}_\nu, \end{equation} where the last expression defines the \tit{covariant derivative} (denoted conventionally by a semicolon vs a comma apropos for an ordinary coordinate derivative) of the components $\nabla_\mu v^\nu \equiv v^\nu_{;\mu} = \partial_\mu v^\nu + {\Gamma}_{\mu\lambda}^\nu v^{\lambda}$. The values of ${\Gamma}_{\mu\lambda}^\nu$ are tied down (see \Eq{connection:coord}) by invoking the \tit{constancy of the metric} condition $\nabla_\mu(g_{\alpha\beta})=0$, (which reduces to the identity $\partial_\mu(\eta_{\alpha\beta})=0$ in the observer's local laboratory). The covariant derivative is of fundamental importance in GR since the commutator of the covariant derivatives acting on a vector is proportional to the Riemann curvature tensor, from which Einstein's fundamental equations are derived: $[\nabla_\mu , \nabla_\nu]\,v^\alpha= v^\alpha_{;\mu;\nu} -v^\alpha_{;\nu;\mu} = -R_{\beta\mu\nu}^\alpha v^\beta$. Note: the analogy in E\&M is the potential $A^\mu$ which acts as a $U(1)$ (\tit{gauge}) potential, so that the covariant derivative is $\nabla_\mu = \partial_\mu - A_\mu(x)$ and the analogue of the curvature is the Faraday tensor (containing the electric and magnetic fields as components) such that $[\nabla_\mu,\nabla_\nu] = \partial_\mu A_\nu - \partial_\nu A_\mu = F_{\mu\nu}$ (see Ryder, section 11.1). The important point here is that the 4-potential $A^\mu(x)$ is spacetime dependent, and that changes in $A^\mu(x)$ (called \tit{gauge transformations}) do not change the physical field $F_{\mu\nu}(x)$. This is called ``gauging" the E\&M field. \subsection{The DE in CST: the spinor covariant derivative in GR}\label{sec:DE:CST:spinor:covar:deriv:GR The new question to ask is: ``How does one gauge gravity?" This is the subject of Ryder\cite{Ryder:2009}, Chapter 11.3 (This chapter is titled ``Gauging Lorentz symmetry: torsion"). The main point is that up to now we've been discussing the covariant derivative for vectors (and tensors) which are associated with integer values of spin ($j = \{0, 1, 2,\ldots\}$, with $2 j+1$ components, i.e. scalars, vectors (e.g. photons), 2-tensor, (e.g. gravitons), etc\ldots). But how does one define the covariant derivative for half-integer spin objects, specifically spin $1/2$ with 2 components? Recall that for the SR the total angular momentum is given by matrix $J_{\mu\nu} = -i\,(x_\mu\,\partial_\nu - x_\nu\,\partial_\mu)\,\mathbb{I} + \Sigma_{\mu\nu}$, where $\Sigma_{\mu\nu}=\frac{i}{4}\,[{\gamma}_\mu, {\gamma}_\nu]$, and ${\gamma}_\mu = \eta_{\mu\nu}\,{\gamma}^\nu$ are the constant Dirac matrices of \Eq{DE:gamma:matrices} (since we are operating in the observer's local SR tangent plane/laboraotry). As in SR, the underlying symmetry of GR (at least locally) is the Poincare group, which is the 10 parameter group of (3) rotations, (3) boosts, and (4) spacetime translations. These matrix operators satisfy as set of (involved) commutation relations $[J_{\mu\nu},J_{\alpha\beta}] = f_{\mu\nu,\alpha\beta}^{\rho\sigma}\,J_{\rho\sigma}$ (which reproduces the usual commutation relations for rotations if we set $\Sigma_{\mu\nu}\to 0$). The Poincare group admits both integer (vector, tensor) representation, as well as spinor (half-integer) representations. These representations are derived by consider small changes in the quantity under study. Following Ryder, Chapter 11.3, Herman Weyl proposed the following ansatz: the ($N$-dimensional) spinor $\psi$ transforms like a \tit{scalar} with respect to the ``world" transformations (i.e. with respect to the coordinate index $\mu$), but as a spinor \tit{with respect to local Lorentz transformations (LLT) in the local laboratory} (i.e. with respect to the index $a$ in the \tit{tangent space at $x$ in the CST where the observer's local laboratory instantaneously exists}). These LLT transform the observer's instantaneous state of motion (in the flat Minkowski tangent plane) at $x$ from one type of motion to another, e.g. from a stationary observer at $x$, to an instantaneous freely falling observer at $x$, or to say an observer executing circular motion instantaneously at $x$, or to any kind of instantaneous motion. Thus, under infinitesimal changes in the surrounding CST small changes in the spinor $\delta\psi$ are given as $\delta\psi = -\xi^\mu\,\partial_\mu\psi$, i.e. with respect to the ordinary coordinate derivative $\partial_\mu$ apropos for a (world) scalar field. Here, $\xi^\mu = \omega^\mu_\nu x^\mu$ where for now, we consider $\xi^\mu$ as constants (since we are acting \tit{within} a given tangent plane at $x$). As an example, for an rotation infinitesimal in the $x-y$ plane by angle $\phi$ given by $\tiny{ \left(\begin{array}{c}x'\\ y'\end{array}\right) = R(\phi) \left(\begin{array}{c}x\\ y\end{array}\right) }$ (suppressing the $t$ and $z$ components for now, i.e. this should be embedded in a $4\times 4$ matrix) where $\tiny{ R(\phi) = \left(\begin{array}{cc}\cos\phi & \sin\phi \\-\sin\phi & \cos\phi\end{array}\right) }$ then $\omega^\mu_\nu$ is the anti-symmetric matrix given by $\tiny{ \frac{d R(\phi)}{d\phi}|_{\phi=0} = \left(\begin{array}{cc}0 & 1 \\-1 & 0\end{array}\right). }$ However, for small changes in the spin, in the local observer's frame, the spinor changes are given by $\delta\psi = -i\,\frac{1}{2}\,\omega^{ab}\,\Sigma_{ab}\,\psi$. Where $\omega^{ab}$ are some constants describing the LLT (as in SR). Thus, for infinitesimal changes, the total change in the spinor is just the sum (to first order) of the two changes (variations) give by $\delta\psi = -\xi^\mu\,\partial_\mu\,\psi -i\,\frac{1}{2}\,\omega^{ab}\,\Sigma_{ab}\,\psi$. We now ``gauge" this transformation by allowing both $\xi^\mu(x)$ and $\omega^{ab}(x)$ to be spacetime dependent (i.e. LLT varying at each point $x$) in order to develop a spinor covariant derivative. The derivation is not that hard but somewhat lengthy (detailed in Ryder, Chapter 11.3) yielding a form $\psi_{|\mu} = \partial_\mu\,\psi + \frac{1}{2}\,A^{ab}_\mu\,\Sigma_{ab}\,\psi$ such that the changes in the spinor $\delta\psi$ under LLT transform the same as $\psi$ itself, namely, $\delta(\psi_{|\mu}) = \frac{1}{2}\omega^{ab}\,\Sigma_{ab}\,(\psi_{|\mu})$. (Note: the spinor covariant derivative is denoted by $\psi_{|\mu}$ to distinguish it from the covariant derivative acting on vectors and tensors, e.g. $v_{;\mu}$). Most significantly, the commutator of the spinor covariant derivatives acting on $\psi$ are once again proportional to (a spinor version of) the Riemann tensor (having both mixed tangent plane (Latin), and world (Greek) indices), namely $\psi_{|\mu|\nu} - \psi_{|\nu|\mu} = -\frac{1}{2} R^{ab}_{\mu\nu}\,\Sigma_{ab}\,\psi$. With these preliminaries under one's belt, one can then derive the DE in CST, as detailed in Ryder, Chapter 11.4, pp416-418. The derivation is quite elegant, but here we indicate only the highlights. We begin with the flat Minkowski SR DE (restoring factors of $\hbar$ and $c$) $i\,\hbar\,{\gamma}^\mu\partial_\mu\,\psi = - m\,c\, \psi$ (note the sign change on the mass, due to the local GR (e.g. Hartle, Ryder) metric, vs the particle physicist's (e.g. Greiner) Minkowski metric), noting that the Dirac matrices are the \tit{constant} ones discussed earlier for the SR DE \Eq{DE:gamma:matrices} (since we are in the observer's local laboratory, i.e. the instantaneous (locally flat, Minkowski) tangent space to the CST at the point $x$). The net result of the gauging of gravity (really, gauging Lorentz symmetry) is that the covariant derivative for spinors boils down to \be{Ryder:p416:128:cov:deriv} \partial_\mu \to D_\mu \overset{\textrm{def}}{=} \partial_\mu - \frac{i}{2}\,{\Gamma}_{\alpha\beta\mu}\,\Sigma^{\alpha\beta} = \partial_\mu + \frac{1}{8}\,{\Gamma}_{\alpha\beta\mu}\,[ {\gamma}^\alpha, {\gamma}^\beta]. \end{equation} Here, ${\Gamma}_{\alpha\beta\mu} = g_{\alpha\rho}\,{\Gamma}^{\rho}_{\beta\mu}$ where for coordinate basis vectors ${\Gamma}^{\rho}_{\beta\mu}$ are the usual Levi-Civita connenction given in terms of the metric by \be{connection:coord} \trm{coordinate basis:}\quad {\Gamma}^{\mu}_{\alpha\beta} = \frac{1}{2}\,g^{\mu\lambda}\,(\partial_\alpha\,g_{\lambda\beta} + \partial_\beta\,g_{\lambda\alpha}-\partial_\lambda\,g_{\alpha\beta}). \end{equation} The ${\gamma}^\alpha$ appearing in \Eq{Ryder:p416:128:cov:deriv} are the \tit{constant} Dirac gamma matrices given prior in \Eq{DE:gamma:matrices}. Thus, we penultimately arrive at our desired goal, the DE in CST \be{Ryder:p416:128} i\,\hbar\,{\gamma}^\mu\,D_\mu\psi = i\,\hbar\,\,{\gamma}^\mu\, \left(\partial_\mu + \frac{1}{8}\,{\Gamma}_{\alpha\beta\mu}\,[ {\gamma}^\alpha, {\gamma}^\beta]\right)\,\psi = m\,c\,\psi. \end{equation} We now introduce one more complication, namely, the translation of the above DE written in a coordinate basis, to a physical orthonormal set of basis vectors. This entails having an expression for the connection ${\Gamma}_{\alpha\beta\mu}$ in an orthonormal basis. While straightforward, yet somewhat lengthy to derive (see Ryder, Chapter 3.13, pp107-110, Eq(3.259)) the results are a pleasing generalization, denoted by ${\Gamma}_{abc}$, of the coordinate-based Levi-Civita connection ${\Gamma}_{\mu\nu\lambda}$, given by \bea{Ryder:p108:3.259} \trm{orthonormal basis:}\quad {\Gamma}_{abc}&=& -\frac{1}{2}\, ( C_{abc} + C_{bca} - C_{cab} ) \\ C_{abc} &=& \eta_{ad}\,C^{d}_{\;\;bc}, \quad\trm{where}\quad [\textrm{\bf e}_a,\textrm{\bf e}_b] = \textrm{\bf e}_c\,C^{c}_{\;\;bc}. \end{eqnarray} (Note: many GR books, including Ryder (but not Hartle), use Greek indices on all basis vectors, coordinate and orthonormal, and the metric as well. One just has to be conscious of the context of the specific calculation to discern if the indices indicate global or local basis/metric, and hence which formula to utilize for the connection, \Eq{connection:coord} or \Eq{Ryder:p108:3.259}). Finally, the DE in CST in an orthonormal basis is given by \be{Ryder:p417:11.129} i\,\hbar\,\gamma^a (e_a + {\Gamma}_a)\,\psi = m\,c\,\psi, \qquad {\Gamma}_a\overset{\textrm{def}}{=}\frac{1}{8}\,{\Gamma}_{abc}[{\gamma}^b,{\gamma}^c]. \end{equation} Here, $e_a(\psi(x)) = e_a^\mu(x)\partial_\mu\psi(x)$ is the action of the orthonormal basis vector acting on the spinor (or any object), and we write $``{\gamma}^a\equiv{\gamma}^\mu"$ by which we mean (abusing notation) that the ${\gamma}^a$ are numerically the \tit{same} constant Dirac gamma matrices as ${\gamma}^\mu$ (in SR, see \Eq{DE:gamma:matrices}). For the Schwarzschild metric, one can straightforwardly work out the commutators of the orthonormal basis vectors (see Ryder, pp418-419, and \App{app:commutators:SST}) which are defined by (restoring again the boldface vector notation) \be{Ryder:p417:11.131} \textrm{\bf e}_0 = \frac{1}{c}\,\left(1- \frac{2 M_s}{r}\right)^{-1/2}\,\frac{\partial}{\partial t},\;\; \textrm{\bf e}_1 = \left(1- \frac{2 M_s}{r}\right)^{1/2}\,\frac{\partial}{\partial r},\;\; \textrm{\bf e}_2 = \frac{1}{r}\,\frac{\partial}{\partial \theta},\;\; \textrm{\bf e}_3 = \frac{1}{r \sin\theta}\,\frac{\partial}{\partial \phi} \end{equation} where $M_s\overset{\textrm{def}}{=} G M/c^2 = \frac{1}{2}\,r_s$, to finally arrive at our desired destination (see Ryder, Eq(11.139), p418) \bea{Ryder:p418:11.139} &{}& i\,\hbar\, \left\{ \left(1- \frac{2 M_s}{r}\right)^{-1/2}\, \left[ {\gamma}^0\frac{1}{c}\frac{\partial\psi}{\partial t} - \frac{M_s}{2 r^2}{\gamma}^1\psi \right] + \left(1- \frac{2 M_s}{r}\right)^{1/2}{\gamma}^1\,\frac{\partial \psi}{\partial r} \right. \nonumber \\ &{}& \left. \hspace{0.20in} +\; {\gamma}^2\,\frac{1}{r}\frac{\partial \psi}{\partial \theta} + \frac{M_s}{r^2}\,\left(1- \frac{2 M_s}{r}\right)^{1/2}{\gamma}^1\,\psi + {\gamma}^3\,\frac{1}{r \sin\theta}\frac{\partial \psi}{\partial \phi} + \frac{\cot\theta}{2 r}\,{\gamma}^2\,\psi \right\} = m\,c\,\psi, \label{DE:CST} \end{eqnarray} of a Dirac spin-1/2 particle of mass $m$ in the Schwarzschild spacetime. One thing we are struck by right away is that nowhere in the above derivation has use been made of the classical GR conserved quantities $e$ and $\ell$ of \Eq{Hartle:p193:9.21} and \Eq{Hartle:p193:9.22}, respectively. However, upon further reflection, for a QM derivation this makes sense, since these constants of the motion involve $dt(\tau)/d\tau$ and $d\phi(\tau)/d\tau$ implying the classical notion of trajectories in spacetime, for which QM abandons, and replaces with the concept of stationary eigenstates over all space. Thus, even though tempting, it would make not make sense to replace quantities such as $\frac{\partial \psi}{\partial t}$ by $\left(\frac{d t}{d\tau}\right)\,\frac{\partial \psi}{\partial \tau}\to e\,(1-\frac{2 M_s}{r(\tau)})^{-1}\,\frac{\partial \psi(\tau)}{\partial \tau}$, etc. since the the wave function $\psi$ would then be solely a function of $\tau$, and the QM concept of spatial eigenstates would not be possible. \section{Bound states of the DE in SST}\label{sec:bound:states:DE:SST We are now finally able to attempt to answer the student's original posed question: ``Does there exist bound states of the DE in SST, analogous to the bound states of the SR DE for either vector or scalar coupling?" Before we can answer this, it is helpful to massage \Eq{Ryder:p418:11.139} into a much more amenable, dimensionless form (especially for numerical calculations). By using the SR Dirac ${\gamma}^a$ matrices \Eq{DE:gamma:matrices} and multiplying through by ${\sqrt{\mathcal{F}}} \overset{\textrm{def}}{=} \sqrt{1-2 M_s/r}$, and again letting $\psi = \tiny{\left(\begin{array}{c}\varphi \\ \chi\end{array}\right)}$, \Eq{DE:CST} can be rearranged into the form \bea{DE:CST:pma:p5.2} &{}&\left( i\,\hbar\,\frac{\partial}{\partial t} - m\,c^2\,{\sqrt{\mathcal{F}}}\right)\,\varphi = c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\chi, \label{DE:CST:pma:p5.2:line1} \\ &{}&\left( i\,\hbar\,\frac{\partial}{\partial t} + m\,c^2\,{\sqrt{\mathcal{F}}}\right)\,\chi = c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\varphi,\label{DE:CST:pma:p5.2:line2} \\ \trm{with}\quad \hat{\textrm{\textbf{p}}} &=& -i\,\hbar\, \left( {\mathcal{F}}\,\frac{\partial}{\partial r} + {\mathcal{F}}\, \frac{M_s}{r^2} - \frac{M_s}{2\,r^2},\; \frac{{\sqrt{\mathcal{F}}}}{r}\left(\frac{\partial}{\partial \theta} + \frac{1}{2}\,\cot\theta \right),\; {\sqrt{\mathcal{F}}}\,\frac{1}{r \sin\theta}\frac{\partial}{\partial \phi} \right). \label{DE:CST:pma:p5.2:line3} \end{eqnarray} The first thing we note is that in going from SR to GR the rest mass goes from $m\,c^2\to m\,c^2\,{\sqrt{\mathcal{F}}}$, acting as a variable mass, that is the ordinary rest mass $m c^2$ at $r\to\infty$ and goes to zero at the Schwarzschild horizon $r\to 2 M_s$. This is one of the new features introduced by GR, the role of the horizon. Secondly, we see that \Eq{DE:CST:pma:p5.2:line1} and \Eq{DE:CST:pma:p5.2:line2} has the form of a SR free field DE, but with the radial potential terms buried within $\hat{\textrm{\textbf{p}}}$, especially, $\hat{p}_r$ in \Eq{DE:CST:pma:p5.2:line3}. If we now postulate the existence of stationary states, with each spinor having an $e^{-i\,E\,t/\hbar}$ temporal dependence, with E constant, we then have \bea{DE:pma:p5.2:middle} c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\chi &=& (E - m\,c^2\,{\sqrt{\mathcal{F}}})\,\varphi, \\ c (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{p}}} )\,\varphi &=& (E + m\,c^2\,{\sqrt{\mathcal{F}}})\,\chi, \end{eqnarray} in strong analogy with the SR DEs \Eq{DE:RQMWE:p170:5:line:1} and \Eq{DE:RQMWE:p170:5:line:2}, but now with no explicit vector or scalar coupling potential terms, $V_v(r)$ and $V_s(r)$ on the righthand side, and now additionally with a variable mass $m c^2\,{\sqrt{\mathcal{F}}}$. Note that for $r\gg r_s=2 M_s$ we have $E + m\,c^2\,{\sqrt{\mathcal{F}}}\approx E \pm m\,c^2\,(1-G M/ (r c^2)) = E \pm (mc^2 + V_G(r))$ where $V_G(r) = -G M m/r$ the Newtonian potential. Hence, we observe scalar (mass) coupling far from the horizon, as our intuition would expect. We now write the above equations in dimensionless form using natural units. Defining the operator $\hat{\textrm{\textbf{q}}}$ via $\hat{\textrm{\textbf{p}}} = -i\,\hbar\,\hat{\textrm{\textbf{q}}}$, and then dividing through by $\hbar c$ and recalling $\lambda_C = \frac{\hbar}{m c}$, we will define lengths as $r = \lambda_C\,\rho$, and energies in terms of the rest mass via $\epsilon = \frac{E}{m c^2}$. We then obtain \bea{DE:pma:p5.3:top} \hspace{-0.25in} &{}& (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{q}}} )\,\chi = i\,(\epsilon - {\sqrt{\mathcal{F}}})\,\varphi, \label{DE:pma:p5.3:top:line1} \\ \hspace{-0.25in} &{}& (\hat{\boldsymbol{\sigma}}\cdot\hat{\textrm{\textbf{q}}} )\,\varphi = i\,(\epsilon + \,{\sqrt{\mathcal{F}}})\,\chi, \label{DE:pma:p5.3:top:line2} \\ \hspace{-0.25in} \hat{\textrm{\textbf{q}}} &=& \left( {\mathcal{F}}\,\frac{\partial}{\partial \rho} + {\mathcal{F}}\, \frac{m_s}{\rho^2} - \frac{m_s}{2 \rho^2},\; \frac{{\sqrt{\mathcal{F}}}}{\rho}\left(\frac{\partial}{\partial \theta} + \frac{1}{2}\,\cot\theta \right),\; {\sqrt{\mathcal{F}}}\, \frac{1}{\rho \sin\theta}\frac{\partial}{\partial \phi} \right), \;\; m_s \overset{\textrm{def}}{=} \frac{M_s}{\lambda_C} = \frac{\frac{1}{2} r_s}{\lambda_C}, \qquad \label{DE:pma:p5.3:top:line3} \\ % &\overset{\textrm{def}}{=}& \left( {\hat{\mathcal{Q}}}_1, {\hat{\mathcal{Q}}}_2, {\hat{\mathcal{Q}}}_3 \label{DE:pma:p5.3:top:line4} \right) \end{eqnarray} Since our primary goal to find the simplest possible solution, not the most general solution, we will cut to the chase and look for only radial solutions (avoiding all the complications of the orbital and spin angular momentum that arose in the SR DE) and define \bea{PMA:ansatz} \varphi &=& g(\rho)\,\tilde{\varphi}, \\ \chi &=& i\,f(\rho)\,\tilde{\chi}, \end{eqnarray} where $\tilde{\varphi}$ and $\tilde{\chi}$ are constant 2-spinors. (Recall in the SR DE, if $\ell = 0\Rightarrow m=0$ which implies $Y_{\ell=0,m=0}(\theta,\phi) = \frac{1}{\sqrt{4 \pi}}$ is a constant). Let us now, with foresight, chose $\varphi = \tiny{\left(\begin{array}{c} a \\ b\end{array}\right)}$ and $\chi = \tiny{\left(\begin{array}{c} a \\ -b\end{array}\right)}$ with $a, b$ constant such that $|a|^2 + |b|^2=1$. We orient our axes so that $r$ and hence $\rho$ are along the $\hat{z}$ direction so that $\hat{\sigma}_1 = \hat{\sigma}_z = \tiny{\left(\begin{array}{cc} 1 & 0 \\ 0 &-1\end{array}\right)}$. Then, substituting these definitions into \Eq{DE:pma:p5.3:top:line1} and \Eq{DE:pma:p5.3:top:line2} we have \bea{DE:pma:p5.3:bottom} \hspace{-0.65in} \left(\begin{array}{cc} {\hat{\mathcal{Q}}}_1 & 0 \\ 0 &-{\hat{\mathcal{Q}}}_1\end{array}\right)\,(i\,f)\, \left(\begin{array}{c} a \\ -b\end{array}\right) = i\,(\epsilon -{\sqrt{\mathcal{F}}})\,g\, \left(\begin{array}{c} a \\ b\end{array}\right) &\Rightarrow& \begin{cases} a\,{\hat{\mathcal{Q}}}_1\,f = a (\epsilon -{\sqrt{\mathcal{F}}})\,g, \\ b\,{\hat{\mathcal{Q}}}_1\,f = b (\epsilon -{\sqrt{\mathcal{F}}})\,g, \end{cases} \Rightarrow {\hat{\mathcal{Q}}}_1\,f = (\epsilon -{\sqrt{\mathcal{F}}})\,g, \;\;\qquad \\ \hspace{-0.65in} \left(\begin{array}{cc} {\hat{\mathcal{Q}}}_1 & 0 \\ 0 &-{\hat{\mathcal{Q}}}_1\end{array}\right)\,g\, \left(\begin{array}{c} a \\ b\end{array}\right) = i\,(\epsilon +{\sqrt{\mathcal{F}}})\,(i\,f)\, \left(\begin{array}{c} a \\ -b\end{array}\right) &\Rightarrow& \begin{cases} a\,{\hat{\mathcal{Q}}}_1\,g = -a (\epsilon +{\sqrt{\mathcal{F}}})\,f, \\ -b\,{\hat{\mathcal{Q}}}_1\,g = b (\epsilon +{\sqrt{\mathcal{F}}})\,f, \end{cases} \hspace{-0.5em} \Rightarrow {\hat{\mathcal{Q}}}_1\,g = -(\epsilon+{\sqrt{\mathcal{F}}})\,f. \;\;\qquad \end{eqnarray} Thus, for this choice of constant $\tilde{\varphi}$ and $\tilde{\chi}$, each spinor produces two equations, which are self-consistently the same. Therefore, our final dimensionless radial DE in CST equations are \bea{DE:CST:p5.3:final} {\hat{\mathcal{Q}}}_1(\rho)\,f(\rho) &=& \left(\epsilon -\sqrt{\mathcal{F}(\rho)}\right)\,g(\rho), \label{DE:CST:p5.3:final:line1} \\ {\hat{\mathcal{Q}}}_1(\rho)\,g(\rho) &=& -\left(\epsilon+\sqrt{\mathcal{F}(\rho)}\right)\,f(\rho), \label{DE:CST:p5.3:final:line2} \\ \trm{with}\quad \mathcal{F}(\rho) = 1 - \frac{2 m_s}{\rho},\quad {\hat{\mathcal{Q}}}_1(\rho)&=&{\mathcal{F}}\,\frac{\partial}{\partial \rho} + {\mathcal{F}}\, \frac{m_s}{\rho^2} - \frac{m_s}{2\,\rho^2}, \;\; \rho = r/\lambda_C, \;\; m_s = \frac{\frac{1}{2} r_s}{\lambda_C}. \quad \label{DE:CST:p5.3:final:line3} \end{eqnarray} For a 2-spinor in the standard form $\tiny{\left(\begin{array}{c} \cos\theta'/2 \\ \sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ we can associate the point $\hat{n} = (\sin\theta'\,\cos\phi',$ $\sin\theta'\,\sin\phi',\cos\theta')$ with polar and azimuthal angles $(\theta', \phi'$) on an ordinary 2-sphere $S^2\in\mathbb{R}^3$, called the \tit{Bloch} sphere. (While we have oriented the spin $\hat{z}$-axis of the Bloch sphere with the world radial coordinate $r$ direction, the internal spinor space angles $(\theta', \phi')$ should not be conflated with the Schwarzschild spacetime coordinates $(\theta, \phi)$). Spin up corresponds to $\theta'=0$ oriented along $\textrm{\bf e}_r$, and spin down with $\theta'=\pi$ oriented along $-\textrm{\bf e}_r$ (with $\phi'$ indeterminate at the poles, so it can be taken to be zero there). The spinor solutions we have used above are then of form, $\tilde{\varphi} = \tiny{\left(\begin{array}{c} \cos\theta'/2 \\ \sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ and $\tilde{\chi} = \tiny{\left(\begin{array}{c} \cos\theta'/2 \\ -\sin\theta'/2\,e^{i\phi'} \end{array}\right)}$ with overlap $\IP{\tilde{\varphi}}{\tilde{\chi}} = \cos\theta'$, and thus are orthogonal at the Bloch sphere equator $\theta'=\pi/2$. For other possible choices of the spinors leading to a modified form of the radial equations \Eq{DE:CST:p5.3:final:line1} and \Eq{DE:CST:p5.3:final:line2}, see \App{app:other:spinor:solns}. \subsection{Numerical solutions of the DE in SST}\label{subsec:Num:Solns:DE:SST} The square root factor ${\sqrt{\mathcal{F}}}$ in \Eq{DE:CST:p5.3:final:line1} and \Eq{DE:CST:p5.3:final:line2} presents difficulties for the standard power series in $\rho$ solutions of these equations, as was employed in the NRQM SE and SR DE. Hence, here we make a \tit{conformal transformation} in order to map $\rho=\infty$ to a finite value, via the definition $\sin\Theta\overset{\textrm{def}}{=} \sqrt{1 - 2 m_s/\rho}$ such that $0\le\Theta\le\pi/2\leftrightarrow 2 m_s\le\rho\le\infty$. Therefore, $\rho = \frac{2 m_s}{\cos^2\Theta}$ and $\frac{\partial }{\partial \rho} = \frac{1}{\partial \rho/\partial \Theta}\,\frac{\partial }{\partial \Theta} = \frac{\cos^3\Theta}{4 m_s \sin\Theta}\frac{\partial}{\partial \Theta}$ hence our DE in SST in a more numerically amenable form is given by \bea{DE:CST:Qhat:Theta} {\hat{\mathcal{Q}}}_1(\Theta)\,f(\Theta) &=& \left(\epsilon -\sin\Theta\right)\,g(\Theta), \label{DE:CST:Qhat:Theta:line1} \\ {\hat{\mathcal{Q}}}_1(\Theta)\,g(\Theta) &=& -\left(\epsilon+\sin\Theta\right)\,f(\Theta), \label{DE:CST:Qhat:Theta:line2} \\ {\hat{\mathcal{Q}}}_1(\rho)\to {\hat{\mathcal{Q}}}_1(\Theta)&=& \frac{\cos^4\Theta}{4 m_s }\, \left( \tan\Theta\,\frac{\partial}{\partial \Theta} + \sin^2\Theta -\frac{1}{2} \right), \\ \label{DE:CST:Qhat:Theta:line3} \trm{where}\quad 0\le\Theta\le\pi/2 &\leftrightarrow& 2\,m_s\le\rho\le\infty \leftrightarrow 2\,M_s\le r\le\infty, \label{DE:CST:Qhat:Theta:line4} \end{eqnarray} for the coordinate region \tit{outside} the Schwarzschild horizon. Note that the dimensionless constant $m_s=\frac{\frac{1}{2} r_s}{\lambda_C}$ \tit{cannot} be absorbed into either $f$ or $g$ nor into the the definition of $\Theta$ itself (even if we were to have defined $\tilde{\rho} = \rho/2m_s$ with ${\mathcal{F}}\to 1-\frac{1}{\tilde{\rho}}$), and thus sets the scale for the problem. As shown in \App{app:Units}, for a solar mass BH, $r_s=1.48$ km $= 1.48\times 10^3$m and $\lambda_C = 2.246\times 10^{-12}$m, so that $m_s\sim 10^{15}$, an astronomically huge number. Nonetheless, with the courage of our intuition, but with warranted trepidation, we now seek numerical solutions of \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} for various values of $1\le m_s\le 10^{15}$. A few other forms of the above equations are worthwhile to obtain a feel for part of the structure of the radial wave functions $f(\Theta)$ and $g(\Theta)$. By using an integrating factor, we could simplify the above equations with the substitutions \bea{DE:CST:Qhat:F:G} f(\Theta)= \sqrt{\sin\Theta}\,e^{-\frac{1}{2}\,\sin^2\Theta}\,F(\Theta) &\;\;\Rightarrow\;\;& \frac{\sin\Theta\cos^3\Theta}{4 m_s }\,\frac{\partial F(\Theta)}{\partial \Theta} = \left(\epsilon -\sin\Theta\right)\,G(\Theta), \\\label{DE:CST:Qhat:F:G:line1} g(\Theta)= \sqrt{\sin\Theta}\,e^{-\frac{1}{2}\,\sin^2\Theta}\,G(\Theta) &\;\;\Rightarrow\;\;& \frac{\sin\Theta\cos^3\Theta}{4 m_s }\,\frac{\partial G(\Theta)}{\partial \Theta} = -\left(\epsilon +\sin\Theta\right)\,F(\Theta). \label{DE:CST:Qhat:F:G:line2} \end{eqnarray} Lastly one could define a coordinate $x$ to the remove the prefactors in front of the above derivatives to yield \bea{DE:CST:Qhat:F:G:x} \frac{1}{4 m_s }\,\frac{\partial F(x)}{\partial x} &=& \big(\epsilon -\sin\Theta(x)\big)\,G(x), \label{DE:CST:Qhat:F:G:x:line1} \\ \frac{1}{4 m_s }\,\frac{\partial G(x)}{\partial x} &=& -\big(\epsilon +\sin\Theta(x)\big)\,G(x), \label{DE:CST:Qhat:F:G:x:line2}\\ x(\Theta) = \int\frac{d \Theta}{\sin\Theta\cos^3\Theta} &=& \ln(\tan\theta) + \frac{1}{2\,\cos^2\theta}\equiv h(\Theta) \;\Rightarrow\; \Theta(x) \overset{\textrm{def}}{=} h^{-1}(x), \label{DE:CST:Qhat:F:G:x:line3} \end{eqnarray} where $\Theta=\Theta(x)$ is now the inverse function of $x(\Theta)$ in \Eq{DE:CST:Qhat:F:G:x:line3}. For numerical solutions we have found that all three of these formulations produce the same eigenvalues (as one would expect), and thus, in the following we will work directly with \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2}. The numerical procedure to find both the eigenvalues $\epsilon$ and eigenfunctions $f(\Theta)$ and $g(\Theta)$ \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} is essentially a \tit{shooting method}. As in the SR DE we require the boundary conditions based on physical arguments, that $f(0) = f(\pi/2)=0$ and $g(0) = g(\pi/2)=0$. Since we are looking for bound states, we search for eigenvalues in the range $-1\le\epsilon\le 1$. Thus, by choosing a value of $\epsilon$ near zero and integrating inwards from $\Theta=\pi/2$ $\Theta=0$ to one obtains solutions for $f(0)$ and $g(0)$, which are in general non-zero. One then adjusts the value of $\epsilon$ (essentially performing a line search in $\epsilon$) and re-integrates again from $\Theta=\pi/2$ to $\Theta=0$ until the desired boundary conditions on $f$ and $g$ are met at $\Theta=0$. In this fashion one can find the lowest eigenvalue of $\epsilon_1$ that yields both $f(0)=0$ and $g(0)=0$. One then repeats this procedure by searching for $\epsilon_2$ in the range $|\epsilon_1|\le\epsilon_2\le 1$ and $-1\le\epsilon_2\le-|\epsilon_1|$, and similarly for larger eigenvalues. \tit{Mathematica}\cite{Mathematica} has such a routine called \ttt{NDEigensystem} which performs this numerical procedure for coupled differential operators $\mathcal{L}_1\big(u_i(x,y,\ldots), v_i(x,y,\ldots),\ldots\big) = \lambda_i\,u_i(x,y,\ldots)$, $\mathcal{L}_2\big(u_i(x,y,\ldots), v_i(x,y,\ldots),\ldots\big) = \lambda_i\,v_i(x,y,\ldots)$, \ldots, returning the eigenvalue and eigenfunction pair $\{\lambda_i, \{u_i, v_i\}\}$ for $i=\{1,2,\ldots,78\}$ (the maximumn number of eigenvalues that \ttt{NDEigensystem} can return). We therefore write our coupled pair of radial wave equations \Eq{DE:CST:Qhat:Theta:line1} and \Eq{DE:CST:Qhat:Theta:line2} as the coupled eigenvalue equations \bea{FG:eval:oprs} \mathcal{L}_1(f,g) &=& \hspace{1.5em}{\hat{\mathcal{Q}}}(\Theta) f(\Theta) + \sin\Theta\,g(\Theta) \quad\Rightarrow\quad \mathcal{L}_1(f,g) = \epsilon\,g(\Theta), \label{FG:eval:oprs:line1} \\ \mathcal{L}_2(f,g) &=& -\left( {\hat{\mathcal{Q}}}(\Theta) g(\Theta) + \sin\Theta\,f(\Theta) \right) \;\Rightarrow\quad \mathcal{L}_2(f,g) = \epsilon\,f(\Theta). \label{FG:eval:oprs:line2} \end{eqnarray} For now we simply treat $m_s$ as a variable parameter and plot the eigenvalues from \ttt{NDEigensystem} for $m_s\in\{5, 10, 10^2, 10^3, 10^{15}\}$ as shown in \Fig{fig:DE:SST:evals:ms:5:10:100:1000:1e15}. \begin{figure}[h] \includegraphics[width=7.0in,height=2.5in]{fig_2_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s\in\{5, 10, 10^2, 10^3, 10^{15}\}$, from outside to inside, with the dashed black lines at $\pm 1$ as a guide lines for the boundary of the bound state energy region. }\label{fig:DE:SST:evals:ms:5:10:100:1000:1e15} \end{figure} The dashed black lines at $\pm 1$ denote the upper bounds for the bound state energy region for $|\epsilon_k = E_k/m\,c^2|<1$. The gray-dashed curve (working outside to inside) for $m_s=5$ shows missing values, since the these eigenvalues had imaginary components. But as the curves for $m_s=10$ (gray-dotted), $m_s=100$ (black-dashed), $m_s=1000$ (overlapping black-solid) show, the eigenvalues quickly settle down to the black-solid curve for $m_s>10^2-10^3$. In fact, this latter curve is also valid when we used $m_s=10^{15}$ appropriate for a solar mass BH. The value of the first and last five eigenvalues are found to be $\epsilon_k=\pm\,\{0.0379246, 0.0781011, 0.117483, 0.156412, 0.195074,\ldots,$ $0.980712, 0.987621, 0.992999, 0.996871, 0.999206\}$.\cite{SR:DE:Numerical:Note} \subsection{Adding numerical diffusion}\label{subsec:numerical:diffusion} \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} are \tit{convection dominated} equations, which \ttt{NDEigensystem} warns, and suggests that one add (artificial) numerical diffusion. In fact, an examination of the numerical wave function solutions reveal cusp-like behaviors in places, typical of such convection dominated problems. There is a standard numerical remedy to this problem (resulting from numerical integration instability) described in many computational physics books, but particularly lucidly (as are other numerous other topics treated similarly) in \tit{Numerical Recipes in C} \cite{NumRecinC:1992} (or in Fortran, or Fortran90, or C++, if you prefer), written by a group of numerical relativists, see Chapter 19.1, pp834-839), by Press, Teukolsky, Vetterling and Flannery. It entails adding to the righthand side of the following 1D example problem $\partial_t f(x) = v(x) \partial_x f(x)$, a second order term of the form $\alpha_d\,\partial^2_{xx}$. As discussed in Press \tit{et al}., this added diffusive term stabilizes the equation, when one performs von Neumann stability analysis, for the growth of small fluctuations (in this example, with $v(x)$ and $\alpha_d$ treated as constants). In our problem, the stability analysis is bit more complicated. As such, we have found that adding a term $\frac{\cos\Theta}{m_s}\, \partial^2_{\Theta\Theta} \,g$ to $\mathcal{L}_1$ (since it contains $\partial_{\Theta}\,g$), and similarly, a term $\frac{\cos\Theta}{m_s}\, \partial^2_{\Theta\Theta} \,f$ to $\mathcal{L}_2$ (since it contains $\partial_{\Theta} \,f$), smooths the wave functions $f(\Theta)$ and $g(\Theta)$ somewhat, without severely changing the eigenvalue spectrum. This is shown in \Fig{fig:DE:SST:evals:with:without:diffusion}. \begin{figure}[h] \includegraphics[width=5.5in,height=2.5in]{fig_3_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (black curve) without numerical diffusion, and (gray curve) with numerical diffusion. }\label{fig:DE:SST:evals:with:without:diffusion} \end{figure} The reason for the use of $\alpha_d = \cos\Theta/m_s$ is that the added diffusion term of the same order of magnitude $1/m_s$ as the coefficients in $\mathcal{L}_1$ and $\mathcal{L}_2$, and the function $\cos\Theta$ ``turns on" the diffusion as we proceed from $\Theta=\pi/2$ inwards to $\Theta=0$ where the spikiness of the wave functions without numerical diffusion are observed. The role of the diffusion term on the wave functions can be seen in \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} and \Fig{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4}, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. \begin{figure}[h] \begin{tabular}{cc} \includegraphics[width=3.0in,height=1.5in]{fig_4_top_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_4_top_right_DE_SST_AJP_26Jun2022} \\ \includegraphics[width=3.0in,height=1.5in]{fig_4_bottom_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_4_bottom_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Wave functions of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (left column) without numerical diffusion, and (right column) with numerical diffusion, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. }\label{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} \end{figure} \begin{figure}[h] \begin{tabular}{cc} \includegraphics[width=3.0in,height=1.5in]{fig_5_top_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_5_top_right_DE_SST_AJP_26Jun2022} \\ \includegraphics[width=3.0in,height=1.5in]{fig_5_bottom_left_DE_SST_AJP_26Jun2022} & \includegraphics[width=3.0in,height=1.5in]{fig_5_bottom_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Wave functions of \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2} for $m_s=25,000$, (left column) without numerical diffusion, and (right column) with numerical diffusion, for the lowest four eigenvalues $\epsilon_{k=\{1,2,3,4\}}$. }\label{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4} \end{figure} \clearpage \newpage The left column are plots without numerical diffusion, while the right column are with numerical diffusion. While in \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2}(left) the plot of $f_1(\Theta)$ appears to be suppressed, it is actually non-zero, but so small that is does not register on the plot. In \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2}(right), with diffusion, $f_1(\Theta)$ is brought out (black curve) and the sharp cusps of $g_1(\Theta)$ on the left (numerical non-smoothness due to lack of diffusion) is smoothed out (as is $f_1(\Theta)$) on the right. The smoothing for wave functions for larger (magnitude) eigenvalues is still uneven in places as can be seen \Fig{DE:SST:evecs:diff:nodiff:f1:g1:f2:g2} and \Fig{DE:SST:evecs:diff:nodiff:f3:g3:f4:g4}, but this resulted from our ``best guess" diffusion coefficient taken to be $\alpha_d = \cos\Theta/m_s$, with $m_s=25,000$. These plots indicate that a more sophisticated $\Theta$-dependent diffusion coefficient is most likely warranted. But for our purpose of exploring the student's intuition, our chose form is sufficient for illustration. There is a balancing act involved here since the choice of a $\Theta$-dependent diffusion coefficient can alter the values of the eigenvalues. For the choice of diffusion we employed, the altered eigenvalues were $\epsilon^{(\trm{diffusion})}_k=\pm\,\{0.0642606, 0.120916, 0.171706, 0.212431, 0.25394,\ldots,$ $0.981319, 0.988035, 0.993213, 0.996967, 0.999226\}$, for $m_s=25,000$, which are not too different from the eigenvalues obtained without diffusion, given by $\epsilon_k=\pm\,\{0.0379246, 0.0781011,$ $0.117483, 0.156412, 0.195074,\ldots,$ $0.980712, 0.987621, 0.992999, 0.996871, 0.999206\}$. We consider the latter eigenvalues without diffusion (with $m_s = 10^{15}$ apropos for a solar mass BH) as the true numerical eigenvalues. For a reasonable, desired degree of smoothness in the corresponding wave functions, we allow for the numerical diffusion we have chosen (after several ``numerical experiments") which does not significantly alter the eigenspectrum, as shown in \Fig{fig:DE:SST:evals:with:without:diffusion}. \subsection{Comparison of eigenvalues of DE in SST with SR DE and SE+SR} Lastly, in this section we compare the eigenvalues $\epsilon_k$ of the DE in SST (gray-curve: with diffusion; black curve, without diffusion) in \Fig{fig:DE:SST:evals:with:without:diffusion} against previous eigenvalue formulas. In \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(left) the outer cyan curve is the SR-DE for vector coupling \Eq{Greiner:p186:case:2} with $\alpha\to~\alpha_v=0.999281$ fitted to $E_1^{(vector)}/mc^2=\epsilon_1$ using $n=1$ and $j=\frac{1}{2}$. The inner purple curve is the the SR-DE for scalar coupling \Eq{Greiner:p186:case:1} with $\alpha'\to\alpha_s=26.25$ fitted to $E_1^{(scalar)}/mc^2=\epsilon_1$ using $n=1$ and $j=\frac{1}{2}$. As our intuition would suspect, the (purple) scalar coupling formula for the SR DE more closely approximates the DE in SST (black curve). \begin{figure}[h] \begin{tabular}{cc} \hspace{-0.75in} \includegraphics[width=4.0in,height=2.25in]{fig_6_left_DE_SST_AJP_26Jun2022} & \hspace{-0.1in} \includegraphics[width=4.0in,height=2.25in]{fig_6_right_DE_SST_AJP_26Jun2022} \end{tabular} \caption{Eigenvalues of (Left:, top to bottom) (cyan curve) SR DE formula with vector coupling coefficient $\alpha_v=0.999281$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (purple curve) SR DE formula with scalar coupling coefficient $\alpha_s=26.25$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (gray curve) GR DE with diffusion $m_s=25,000$, (black curve) GR DE without diffusion. (Right: top to bottom) (purple curve) SE+SR scalar coupling formula with $E_I=0.998562$ fitted to $\epsilon_1$ of DE in SST (no diffusion), (gray curve) GR DE with diffusion $m_s=25,000$, (black curve) GR DE without diffusion. (cyan curve) SE+SR vector coupling formula with $E_I=694.278$ fitted to $\epsilon_1$ of DE in SST (no diffusion), }\label{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling} \end{figure} In \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(right) we compare the DE in SST with the SE+SR formulas \Eq{E_s} and \Eq{E_v} which we had term SE ``scalar coupling" and SE ``vector coupling" respectively, since the position of the square root being either in the numerator or in the denominator, respectively, mimics that of the SR DE scalar and vector coupling formulas \Eq{Greiner:p186:case:1} and \Eq{Greiner:p186:case:2}, respectively. Here ``SE+SR" means that we ``add" $m c^2$ to the NRQM resulting SE energies $E_k$, and interpret the result as the firs order approximation to some unknown (from a NR POV) SR DE formula, involving square roots (either in the numerator or denominator). For the ``scalar coupling" case we fit $\epsilon_1 =\sqrt{1-\frac{E_I}{n^2}}$ for $n=1$ to find $E_I=0.998562$, while for the ``vector coupling" case we fit $\epsilon_1 =\frac{1}{\sqrt{1+\frac{E_I}{n^2}}}$ for $n=1$ to find $E_I=694.278$. The are plotted in \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}(right) using the same colors as the (left) plot, namely, cyan for scalar coupling and purple for vector coupling. Counter to our intuition, now it is the SE ``vector coupling" formula that most closely mimics the eigenvalues of the DE in SST (black curve). This flipping of our intuitive interpretation of the SE+SR formulas \Eq{E_s} and \Eq{E_v} is most likely due to the fact that at the values of $E^{(scalar)}_I=0.998562$ and $E^{(vector)}_I=694.278$ a first order expansion of the square root (as we assumed by adding $m\,c^2$ to $E_I$) is not warranted at these large values of $E_I$. Nonetheless, it is curious that the roles of the SE ``scalar" and ``vector" coupling flip, and nearly approximate their SR DE opposites, vector and scalar coupling, respectively, as shown in \Fig{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All}. \begin{figure}[h] \includegraphics[width=5.0in,height=2.5in]{fig_7_DE_SST_AJP_26Jun2022} \caption{Composite of the left and right plots \Fig{fig:DE:SST:evals:compare:SR:DE:and:SE:DEscalar:vector:coupling}. }\label{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All} \end{figure} \section{Solutions of the DE in SST inside the horizon?}\label{sec:DE:SST:inside:horizon} In the previous section, for solutions of the DE in SST we defined the angle $\Theta$ by $\sin\Theta~=~\sqrt{1-\frac{2 M_s}{r}}$, for $0\le\Theta\le\pi/2\leftrightarrow 2 M_s\le r \le \infty$, the region outside the horizon. Following our intuition, suppose we \tit{anlaytically continue} $\Theta\to i\,x$ so that $\sin\Theta\to i\,\sinh x$ and $\sqrt{1-\frac{2 M_s}{r}}\to i\,\sqrt{\frac{2 M_s}{r}-1}$, i.e. defining $\sinh x \overset{\textrm{def}}{=} \sqrt{\frac{2 M_s}{r}-1}$, for $0\le r\le 2 M_s \leftrightarrow \infty\ge x\ge 0$. To keep the resulting radial equations real (since the above procedure introduces explicit factor of $i$), we are led to consider redefining the ``energies" as $\epsilon\to i\,\eta$. Holding off on interpretation for now, and bolding (but not blindly) pushing forward, this leads to the following set of eigenvalue equations that we can once again use \ttt{NDEigensystems} to solve \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2}, now with $f(\Theta)\to F(x)$ and $g(\Theta)\to i\,G(x)$ % \bea{L1:L2:oprs:inside:horizon} \hspace{-0.25in} \mathcal{L}_1 &=& \frac{\cosh^4 x}{4 m_s} \left( -\tanh x \,\frac{\partial}{\partial x} + \sinh^2 x + \frac{1}{2} \right)\,G(x) - \sinh x\,F(x), \;\; \Rightarrow\; \mathcal{L}_1(F,G) = \eta\,F(x), \qquad \label{L1:opr:inside:horizon} \\ \hspace{-0.25in} \mathcal{L}_2 &=& -\frac{\cosh^4 x}{4 m_s} \left( \tanh x \,\frac{\partial}{\partial x} + \sinh^2 x + \frac{1}{2} \right)\,F(x) - \sinh x\,G(x), \;\; \Rightarrow\; \mathcal{L}_2(F,G) = \eta\,G(x). \qquad \label{L2:opr:inside:horizon} \end{eqnarray} These equations produce the ``eigenvalues" $\eta_k$ (black curve) in \Fig{fig:DE:SST:evals:inside:outside:horizon} with only negative values inside the horizon. Again, we show the eigenvalues (gray curve) $\epsilon_k$ outside the horizon, for comparison. \begin{figure}[h] \includegraphics[width=6.5in,height=2.5in]{fig_8_DE_SST_AJP_26Jun2022} \caption{Eigenvalues of (black curve) DE in SST inside the horizon, \Eq{L1:opr:inside:horizon} and \Eq{L2:opr:inside:horizon}, \\ (gray curve) DE in SST outside the horizon, \Eq{FG:eval:oprs:line1} and \Eq{FG:eval:oprs:line2}, both with $m_s=10^{15}$. } \label{fig:DE:SST:evals:inside:outside:horizon} \end{figure} We must now reconcile with imposing an interpretation on the results shown in \Fig{fig:DE:SST:evals:inside:outside:horizon}, and for the meaning of the values of $\eta_k<0$. At first glance setting $\epsilon = E/m c^2 \to i\,\eta$ seems to imply either imaginary mass, or imaginary energies, and therefore appears non-sensical. On further thought, note that the temporal and radial portion of the Schwarzschild metric $ds^2_{outside} = -\left(1-\frac{2 M_s}{r}\right)\,c^2\,dt^2 + \left(1-\frac{2 M_s}{r}\right)^{-1}\,dr^2$ for $r>2 M_s$ switches signs to $ds^2_{inside} = \left(\frac{2 M_s}{r}-1\right)\,c^2\,dt^2 - \left(\frac{2 M_s}{r}-1\right)^{-1}\,dr^2$ inside for $r<2 M_s$. The interpretation is that the timelike variable is associated with the term with the $-1$ (times a positive factor), and thus becomes $r$ inside the horizon. Thus, the variables $t$ and $r$ switch roles and become spacelike and timelike respectively, inside the horizon. Consequently, moving to the the singularity $r=0$ is inevitable, since ``time" $r$ always moves forward for inward falling particles inside the horizon. In fact, $\epsilon$ now has the interpretation of a momentum (vs an energy). Further, the stationary states, which outside had a temporal dependence of $e^{-i \epsilon_k\,t}$ now become upon replacement $\epsilon_k\to i\,\eta_k$, factors of $e^{\eta_k\,t}$, which for $\eta_k<0$ become evanescent like wave functions decaying to the singularity as $t$ increases. Therefore, the fact the numerics actually produces solutions with only $\eta_k<0$ seems to fit with our intuition. Recall that outside the horizon $e^{\pm i k r}$ are waves moving radially outward ($+$) and inward ($-$), with $\hbar k$ being the magnitude of the momentum $p_k$, with energy $p_k c = \hbar k c$. Thus it is conceivable (or at least plausibly interpretable) that $|\eta| \to \frac{(E=\hbar k c)}{m c^2}$ measures a quantized momentum inside the BH. \section{Summary and Conclusion}\label{sec:Summary:Conclusion} It has been a somewhat lengthy journey to attempt to answer a seeming simple question posed by an astute student (in 15 words or less): ``In analogy with the Coulomb problem, can one talk about bounds states in Schwarzschild spacetime?" To answer this question, we had to assemble theoretical background concepts, and analytic and numerical solution techniques for (i) the Schrodinger Equation (SE) in a central potential (the Coulomb problem), (ii) the Special Relativity (SR) generalization to the Dirac Equation (DE) in flat (Minkowski spacetime, MST), and finally (iii) the General Relativity (GR) generalization of the DE to curved spacetime (CST), specifically, the central symmetric Schwarzschild spacetime (SST). Driven by intuition, and bolstered by the courage of our convictions, we attempted to put some theoretical, analytic, and numerical terra firma beneath feet, though knowing at times we might be pushing an analogy (or two) beyond the limits of credulity. The results in \Sec{sec:bound:states:DE:SST}, and in particular, the bound state eigenvalues $|E_k|< m c^2$ numerically found in \Fig{fig:DE:SST:evals:with:without:diffusion} seem to indicate that the answer to the student's question appears to be affirmative. Further, the comparison of the the eigenvalues for the DE in SST with formulas from both the SR DE and the SE+SR, as shown in \Fig{fig:DE:SST:evals:compare:SR:DE:SE:SR:scalar:vector:coupling:All}, shows that the student's intuition, and question, had merit. The intent of this investigation was to highlight a research strategy that could be employed by the student, drawing upon knowledge obtained in upper undergraduate to first year graduate physics education, supplemented by a host of well written and highly useful expository QM and GR textbooks currently available to the student, with enough worked exercises that this whole endeavor could be the seed of an independent research project. Much more remains that could be investigated, as we have only scratched the surface here, and peaked under a few lids, on a host container of fascinating NR and SR QM, and GR topics worthy of deeper investigation (e.g. angular momentum and spin details in the SR and GR Dirac equation, solutions of massless particles in Schwarzschild and other spacetimes often used to model photons, and the role of spinor transformations under the Poincare group of rotations, boost and translations, to name just a few). It is hoped that this investigation has served as a vehicle that might peak the research interest of the inquisitive and curious student, and spur them to further guided or self study, complemented by a host available instructive textbooks, and hopefully beyond, to the research literature itself. \flushleft{\underline{\bf Extra credit (research) project:}} \newline For the ambitious student (or professor!), you can verify your understanding of this material by redoing the DE in CST calculation presented in the main text for the \tit{Reissner-N{\o}rdstrom} metric (see Ryder\cite{Ryder:2009}, p265) for a star with mass $M$ and electric charge $Q$ given by spherically symmetric metric \be{RN:metric} \hspace{-0.25in} ds^2 = -c^2 \left(1- \frac{2 G M }{c^2 r} + \frac{G Q^2}{c^4 r^2}\right) dt^2 + \left(1- \frac{2 G M }{c^2 r} + \frac{G Q^2}{c^4 r^2}\right)^{-1}\,dr^2 + r^2\,\left(d\theta^2 + \sin^2\theta\,d\phi^2\right). \end{equation} As before, $\frac{2 G M }{c^2 r} = \frac{r_s}{r}$, while $\frac{G Q^2}{c^4 r^2} = \left(\frac{e^2}{\hbar c}\right) \left(\frac{Q^2}{e^2}\right) \left(\frac{G \hbar}{c^3 r^2}\right) = \alpha_c\, \left( \frac{Q}{e} \right)^2\, \left( \frac{\ell_p}{r} \right)^2, $ where $\ell_p$ is the Planck length and $\alpha_C$ is the fine structure constant (see \App{app:Units}). In the spirit of John Wheeler (who coined the term black hole), what might one anticipate about any possible solutions even before embarking on the lengthy calculation (to back up one's intuition)? Hint: compare the magnitude and length scales of the two $r$-dependent terms in \Eq{RN:metric}. Note: Astronomically speaking, the Reissner-N{\o}rdstrom is not considered particularly physical, since any charge on a BH will most likely be rapidly neutralized by surrounding in-falling matter. However, the metric is an intuitively satisfying extension of the Schwarzschild metric, extending the sole extensive properties of the BH from mass, to mass plus charge - and it is amenable to the analysis presented in this work.
\section{Introduction} Considering that the early time inflation in the history of the universe seems to have occurred around Planck’s scale, the M/String theory inspired fields can be a possible candidate for the field responsible for inflation. Tachyon inflation is one of these fields leading to interesting cosmological results. Tachyon field, which has attained a lot of attention, is associated with the D-branes in string theory. One interesting implication of this field is that, its slow-rolling down to the potential leads to the smooth evolution of the universe from the accelerating phase of the expansion to the era dominated by the non-relativistic fluid~\cite{Sen99,Sen02a,Sen02b,Gib02}. On the other hand, we know that in the simple inflation model with canonical scalar fields, we find the primordial perturbation to be scale-invariant and gaussian~\cite{Gut81,Lin82,Alb82,Lin90,Lid00a,Lid97,Rio02,Lyt09,Mal03}. When we consider an inflation model with a canonical scalar field, we find a gaussian perturbation that the values of its tensor-to-scalar ratio for $\phi$, $\phi^2$, $\phi^{4/3}$ are not consistent with observational data~\cite{pl18b}. Therefore we have to seek some other specific potential like hilltop potential to get an observationally viable single canonical field inflation model~\cite{pl18b}. However, with every potential, the amplitude of the primordial perturbation in the single canonical field inflation is almost gaussian. Or, we should consider a nonminimal inflation model to give a better explanation of the early time cosmological inflation. However, with a non-canonical scalar field like tachyon, it is possible to get the scale-dependent and non-gaussian distribution for the amplitude of the primordial perturbations. Another interesting point about the tachyon field is corresponding to the equation of state of the tachyon field. This parameter in the tachyon field can be $-1$, which describes both very early time and very late time accelerating expansion of the universe. Also, its value can be $0$, describing the matter/dark matter dominated era. Therefore, with the tachyon field it is possible to explain the thermal history of the universe in a reasonable way~\cite{Gib02}. These are interesting issues that motivate cosmologists to consider and study the inflation models driven by a tachyon field as a non-canonical scalar field. In this regard, a lot of interesting works based on tachyon inflation have been done. For instance, the authors of Ref.~\cite{Kam18} have considered a tachyon field in the context of quantum loop gravity. In this way, they have obtained the inflation and perturbation parameters and tested the observational viability of some inflation models. In ref.~\cite{Bil19}, by considering the holographic cosmology, the tachyon inflation has been studied and the results have been compared to the observational data. The authors of Ref.~\cite{Moh20} have studied an inflation model where a tachyon field interacts with photon gas. They have shown that under some assumptions and conditions this model shows some agreement with observational data. In~\cite{Ras21}, the tachyon model with a superpotential, a potential based on supersymmetry, has been studied. It has been shown that tachyon inflation with a superpotential, at least in some ranges of the model's parameter space, is consistent with observational data. There are other works on tachyon inflation leading to interesting cosmological results~\cite{Noj03,Noz14,Bou16,Rez17,Ras18,Ras20} and all of these works show that the tachyon field can be considered as a field running the inflation. One of the important parameters in the inflation models is the sound speed of the primordial perturbation, shown by $c_{s}$. The sound speed is corresponding to the Lorentz factor $\gamma$ as $c_{s}=\frac{1}{\gamma}$. For the canonical scalar field, we get the sound speed equal to unity. However, with the non-canonical scalar fields, we have $c_{s}^{2}\neq 1$. The value of the parameter $\gamma$ and correspondingly sound speed determines the deformation of the field's kinetic energy from the canonical one. The constant sound speed is an interesting idea motivated by the authors in Refs.~\cite{Spa08,Tsu13}. Considering that in the non-canonical scalar field the sound speed is related to the time variation of the field, the constant sound speed is corresponding to the constant field's variation in times. This can lead to interesting results. Now, the question is what constant values of $c_{s}$ we should adopt, to study the model numerically. Considering that the sound speed is related to the amplitude of the non-gaussianity, by using the observational constraints on the amplitude of the non-gaussianity, we can find suitable ranges of the sound speed. Also, there are some other parameters, such as the scalar spectral index, tensor spectral index, and tensor-to-scalar ratio that help us to find some constraints on the sound speed values. To perform a numerical analysis, we use the Planck2018 constraints on the inflation parameters. Based on the $\Lambda CDM+r+\frac{dn_{s}}{d\ln k}$ model, the Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data gives the value of the scalar spectral index as $n_{s}=0.9658\pm0.0038$ and implies a constraint on the tensor-to-scalar ratio as $r<0.072$~\cite{pl18a,pl18b}. The constraint on the tensor spectral index, released by Planck2018 TT, TE, EE +lowE+lensing+BK14+BAO+LIGO and Virgo2016 data is $-0.62<n_{T}<0.53$~\cite{pl18a,pl18b}. Also, from Planck2018 TT, TE, EE+lowEB+lensing data, we have the value of the running of the scalar spectral index as $\alpha_{s}=-0.0085\pm 0.0073$~\cite{pl18a,pl18b}. Other useful constraints are corresponding to the amplitude of the non-gaussianity. The planck2018 combined temperature and polarization analysis gives the constraints on the amplitude of the non-gaussianity in the equilateral configuration as $f_{NL}^{equil}=-26\pm 47$ and in the orthogonal configuration as $f_{NL}^{ortho}=-38\pm 24$~\cite{pl19}. These several constraints determine the observational viability of every inflation model. With these preliminaries, this paper is organized as follows. In section \ref{sec2}, we review the inflation in a tachyon model with constant sound speed. In this section, we present the main equations governing the dynamics of the model in terms of the sound speed. In section \ref{sec3}, we consider an intermediate scale factor and study the perturbation and non-gaussianity parameters in this model numerically. We compare the results with several data sets to find some constraints on the model's parameter space. In section \ref{sec4} we study the power-law tachyon inflation with constant sound speed and compare the results with observational data. In section \ref{sec5}, we perform some discussion on unifying the initial inflation with late time dark energy in the tachyon model with constant sound speed. In section \ref{sec6}, we present a summary of this work. \section{\label{sec2}Review on the Tachyon Inflation with Constant Sound Speed} For the tachyon field, we have the following Dirac-Born-Infeld type effective 4-dimensional action \begin{eqnarray} \label{eq1} S=\int d^{4}x\,\sqrt{-g} \Bigg[\frac{1}{2\kappa^{4}}R-V(\phi)\,\sqrt{1-2\,X}\Bigg]\,, \end{eqnarray} with $\kappa$ being the gravitational constant, $R$ as the Ricci scalar, and $X=-\frac{1}{2}\,\partial_{\mu}\phi\,\partial^{\mu}\phi$. Also, the tachyon field $\phi$ has the potential $V(\phi)$. Cosmologists believe that the physics of the tachyon condensation can be described by such an action. Einstein's field equations, corresponding to action (\ref{eq1}), are given by \begin{eqnarray} \label{eq2} G_{\mu\nu}=\kappa^{2}\Bigg[-g_{\mu\nu}V(\phi)\sqrt{1-2\, X}+\frac{ V(\phi)\partial_{\mu}\phi\,\partial_{\nu}\phi}{\sqrt{1-2\, X}}\Bigg]\,, \end{eqnarray} which have been obtained by varying action (\ref{eq1}) with respect to the metric. Considering the FRW metric as \begin{equation} \label{eq3} ds^{2}=-dt^{2}+a^{2}(t)\delta_{ij}dx^{i}dx^{j}\,, \end{equation} we find the following Friedmann equation \begin{eqnarray} \label{eq4} 3H^{2}=\frac{\kappa^{2}\,V(\phi)}{\sqrt{1-\dot{\phi}^{2}}}\,. \end{eqnarray} corresponding to the $(0,0)$ component of Einstein's field equations. Note that, a dot on the parameter shows the derivative of the parameter with respect to the time. Also, the $(i,i)$ component of Einstein's field equations give the following second Friedmann equation \begin{eqnarray} \label{eq5} 2\dot{H}+3H^{2}=\kappa^{2}\bigg[V(\phi)\,\sqrt{1-\dot{\phi}^{2}}\,\bigg]\,. \end{eqnarray} We find the equation of motion of the tachyon field by varying the action (\ref{eq1}) with respect to the field as \begin{equation} \label{eq6}\frac{\ddot{\phi}}{1-\dot{\phi}^{2}}+3\,H\dot{\phi} +\frac{V'(\phi)}{V(\phi)}=0\,, \end{equation} where a prime shows the derivative with respect to the tachyon field. Since our purpose in this paper is to study the tachyon model with constant sound speed, we should re-write the above main equations in terms of the sound speed. To this end, we need to define the sound speed of the tachyon field. In general, the square of the sound speed is given by $c_{s}^{2}\equiv\frac{P_{,X}}{\rho_{,X}}$, that in the tachyon model takes the following form \begin{eqnarray} \label{eq7} c_{s}=\sqrt{1-\dot{\phi}^{2}}\,. \end{eqnarray} In this regard, the equations (\ref{eq4})-(\ref{eq6}) take the following forms \begin{eqnarray} \label{eq8} 3H^{2}=\frac{\kappa^{2}\,V(\phi)}{c_{s}}\,, \end{eqnarray} \begin{eqnarray} \label{eq9} 2\dot{H}+3H^{2}=\kappa^{2}\,V(\phi)\,c_{s}\,, \end{eqnarray} and \begin{equation} \label{eq10}3\,H\,\sqrt {1- c_{s}^{2}} +\frac{V'(\phi)}{V(\phi)}=0\,, \end{equation} where we have considered the sound speed as a constant parameter. To study cosmological inflation, we need the following slow-roll parameters \begin{eqnarray} \label{eq11}\epsilon=-\frac{\dot{H}}{H^{2}}\,, \end{eqnarray} \begin{eqnarray} \label{eq12}\eta=-\frac{1}{H}\frac{\ddot{H}}{\dot{H}}\,. \end{eqnarray} Note that, in the case of the constant sound speed, the third slow-roll parameter $s=\frac{1}{H}\frac{\dot{c_{s}}}{c_{s}}$ is zero. Another needed parameter to study inflation is the number of e-folds parameter defined as \begin{eqnarray} \label{eq13} N=\int H\,dt\,. \end{eqnarray} These parameters are useful to study the primordial perturbations in the model and compare the results with observational data. The observational data that we use in our work is the data released by the Planck2018 team~\cite{pl18a,pl18b,pl19}. Based on the fact that, under the assumption of statistical isotropy, the two-point correlations of the CMB anisotropies are described by the angular power spectra $C_{l}^{TT}$, $C_{l}^{TE}$, $C_{l}^{EE}$, and $C_{l}^{BB}$ (where the subscript $l$ shows the multipole moment number)~\cite{kam97,zal30,sel97,Hu97,Hu98}, the Planck team has obtained some constraints on the important perturbation parameters. To include the contributions from the scalar and tensor perturbations in the CMB angular power spectra, the Planck team has used the following expressions~\cite{pl15} \begin{eqnarray} \label{eq14} C_{l}^{ab,s}=\int_{0}^{\infty} \frac{dk}{k} \Delta_{l,a}^{s} (k) \, \Delta_{l,b}^{s} (k)\, {\cal{A}}_{s} (k)\,, \end{eqnarray} \begin{eqnarray} \label{eq15} C_{l}^{ab,T}=\int_{0}^{\infty} \frac{dk}{k} \Delta_{l,a}^{T} (k) \, \Delta_{l,b}^{T} (k)\, {\cal{A}}_{T} (k)\,. \end{eqnarray} In equations (\ref{eq14}) and (\ref{eq15}), we have shown the transfer functions by $\Delta_{l,a}^{s} (k)$ and $\Delta_{l,a}^{T} (k)$. The parameter $a$ and $b$ are given as $a,b=T,E,B$. Also, the primordial power spectrum, the parameter identified by the physics of the primordial universe~\cite{pl15}, is shown by ${\cal{A}}_{i} (k)$ (where, $i=s,T$). Model-independent forms of the scalar and tensor power spectra are given by \begin{eqnarray} \label{eq16} {\cal{A}}_{s} (k)=A_{s}\left(\frac{k}{k_{*}}\right)^{n_{s}-1+\frac{1}{2}\frac{dn_{s}}{d\ln k}\ln\big(\frac{k}{k_{*}}\big)+\frac{1}{6}\frac{d^{2}n_{s}}{d\ln k^{2}}\ln\big(\frac{k}{k_{*}}\big)^{2}+...}\,, \end{eqnarray} \begin{eqnarray} \label{eq17} {\cal{A}}_{T} (k)=A_{T}\left(\frac{k}{k_{*}}\right)^{n_{T}+\frac{1}{2}\frac{dn_{T}}{d\ln k}\ln\big(\frac{k}{k_{*}}\big)+...}\,. \end{eqnarray} These forms of the scalar and tensor power spectra have been used by the Planck team to compare the perturbation parameters with data. Note that, in equations (\ref{eq16}) and (\ref{eq17}), $A_{j}$ shows the scalar (corresponding to $j=s$) and tensor (corresponding to $j=T$) perturbations. Also, the running of the scalar or tensor spectral index and the running of the running of the scalar spectral index are given by $\frac{dn_{i}}{d\ln k}$ and $\frac{d^{2}n_{s}}{d\ln k^{2}}$, respectively. One can also find the tensor-to-scalar ratio by using the power spectra as \begin{eqnarray} \label{eq18} r=\frac{{\cal{A}}_{T}(k_{*})}{{\cal{A}}_{s}(k_{*})}\,, \end{eqnarray} which is a very important parameter in studying the inflation models. The parameters $n_{s}$ and $n_{T}$ are the scalar and tensor spectral indices showing the scale dependence of the primordial perturbations. The scalar spectral index, at the time of sound horizon exit of the physical scales, is defined as \begin{eqnarray} \label{eq19} n_{s}-1=\frac{d \ln {\cal{A}}_{s}}{d\ln k}\Bigg|_{c_{s}k=aH}\,, \end{eqnarray} where the parameter ${\cal{A}}_{s}$, the amplitude of the scalar spectral index, is given by the following definition \begin{equation} \label{eq20}{\cal{A}}_{s}=\frac{H^{2}}{8\pi^{2}{\cal{W}}_{s}c_{s}^{3}}\,, \end{equation} with \begin{equation} \label{eq21} {\cal{W}}_{s}=\frac {V\, (1-c_{s}^2)}{2{H}^{2} c_{s}^{3}}\,. \end{equation} The scalar spectral index, in terms of the slow-roll parameters, is written as \begin{eqnarray} \label{eq22} n_{s}=1-6\epsilon+2\eta\,. \end{eqnarray} This is one of the parameters that can be compared with observational data. Another important parameter in studying the inflation models is the running of the scalar spectral index, defined as \begin{eqnarray} \label{eq23} \alpha_{s}=\frac{dn_{s}}{d\ln k}=8\epsilon\,\eta-12\epsilon^{2}-2\zeta+2\eta^{2}\,, \end{eqnarray} where, the parameter $\zeta$ is given by \begin{eqnarray} \label{eq24} \zeta=\frac{\dddot{H}}{H^{2}\dot{H}}\,. \end{eqnarray} The tensor spectral index, obtained from equation (\ref{eq17}), is defined as follows \begin{eqnarray} \label{eq25} n_{T}=\frac{d \ln {\cal{A}}_{T}}{d\ln k}\Bigg|_{k=aH}\,, \end{eqnarray} with \begin{eqnarray} \label{eq26} {\cal{A}}_{T}=\frac{2\kappa^{2}H^{2}}{\pi^{2}}\,, \end{eqnarray} being the amplitude of the tensor perturbations. In terms of the slow-roll parameters, the tensor spectral index is expressed as \begin{eqnarray} \label{eq27} n_{T}=-2\epsilon\,. \end{eqnarray} Finally, we have the following expression for the tensor-to-scalar ratio \begin{eqnarray} \label{eq28} r=16\,c_{s}\,\epsilon\,. \end{eqnarray} As we see in equation (\ref{eq28}), the tensor-to-scalar ratio depends explicitly on the sound speed of the perturbations. Therefore, from the observationally viable value of $r$, we can constrain the values of the sound speed. Another important parameter that depends on the sound speed parameter is the nonlinearity parameter. The nonlinearity parameter demonstrates the non-gaussian property of the amplitude of the primordial perturbation. The Gaussian distributed primordial perturbations are characterized by a two-point correlation. To seek the additional statistical information, corresponding to the non-Gaussian distributed perturbation, it is necessary to consider the higher-order correlations. The three-point correlation function for the spatial curvature perturbation in the interaction picture is given by~\cite{Mal03} \begin{eqnarray} \label{eq29} \langle {\Phi}(\textbf{k}_{1})\,{\Phi}(\textbf{k}_{2})\,{\Phi}(\textbf{k}_{3})\rangle =(2\pi)^{3}\delta^{3}(\textbf{k}_{1}+\textbf{k}_{2}+\textbf{k}_{3}){\cal{B}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})\,, \end{eqnarray} with the following definition for the parameter ${\cal{B}}_{\Phi}$ \begin{equation} \label{eq30} {\cal{B}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})=\frac{(2\pi)^{4}{\cal{A}}_{s}^{2}}{\prod_{i=1}^{3} k_{i}^{3}}\, {\cal{D}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})\,. \end{equation} The parameter ${\cal{D}}_{\Phi}$ in equation (\ref{eq40}) is expressed as follows \begin{eqnarray} \label{eq31} {\cal{D}}_{\Phi}=\Bigg(1-\frac{1}{c_{s}^{2}}\Bigg) \Bigg[\frac{3}{4}{\cal{J}}_{1}-\frac{3}{2}{\cal{J}}_{2} -\frac{1}{4}{\cal{J}}_{3}\Bigg]\,, \end{eqnarray} where \begin{eqnarray} \label{eq32} {\cal{J}}_{1}=\frac{2\sum_{i>j}k_{i}^{2}\,k_{j}^{2}}{k_{1}+k_{2}+k_{3}}-\frac{\sum_{i\neq j}k_{i}^{2}\,k_{j}^{3}}{(k_{1}+k_{2}+k_{3})^{2}}\,, \end{eqnarray} \begin{eqnarray} \label{eq33} {\cal{J}}_{2}=\frac{\left(k_{1}\,k_{2}\,k_{3}\right)^{2}} {(k_{1}+k_{2}+k_{3})^{3}}\,, \end{eqnarray} and \begin{eqnarray} \label{eq34} {\cal{J}}_{3}=\frac{2\sum_{i>j}k_{i}^{2}\,k_{j}^{2}}{k_{1}+k_{2}+k_{3}}-\frac{\sum_{i\neq j}k_{i}^{2}\,k_{j}^{3}}{(k_{1}+k_{2}+k_{3})^{2}}+\frac{1}{2}\sum_{i}k_{i}^{3}\,. \end{eqnarray} By using the parameter ${\cal{D}}_{\Phi}$, one can find the non-linear parameter as \begin{equation} \label{eq35} f_{NL}=\frac{10}{3}\frac{{\cal{D}}_{\Phi}}{\sum_{i=1}^{3}k_{i}^{3}}\,. \end{equation} Note that, in equation (\ref{eq35}), there is a parameter $k_{i}$ which is the momentum. Depending on different values of the momenta ($k_{1}$, $k_{2}$, and $k_{3}$), we get different shapes of the non-gaussianity. In every shape, there is a maximal peak in the amplitude of the perturbations that defines the kind of the corresponding shape. We are interested in the case with a peak at $k_{1}=k_{2}=k_{3}$ limit, leading to an equilateral shape, and the case orthogonal to it~\cite{Che07,Bab04b,Fel13a,Bau12}. It is possible to write the bispectrum (\ref{eq31}) in terms of the equilateral and orthogonal shapes basis as follows~\cite{Fel13a} \begin{equation} \label{eq36} {\cal{D}}_{\Phi}={\cal{M}}_{1}\,\breve{{\cal{J}}}^{equil} + {\cal{M}}_{2} \,\breve{{\cal{J}}}^{ortho}\,, \end{equation} where, \begin{equation} \label{eq37} \breve{{\cal{J}}}^{equil}=-\frac{12}{13}\Big(3{\cal{J}}_{1}-{\cal{J}}_{2}\Big)\,. \end{equation} \begin{equation} \label{eq38} \breve{{\cal{J}}}^{ortho}=\frac{12}{14-13{\cal{C}}}\Big({\cal{C}}\big(3{\cal{J}}_{1}-{\cal{J}}_{2}\big)+3{\cal{J}}_{1}-{\cal{J}}_{2}\Big)\,, \end{equation} \begin{equation} \label{eq39} {\cal{M}}_{1}=\frac{13}{12}\Bigg[\frac{1}{24}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\,, \end{equation} and \begin{equation} \label{eq40} {\cal{M}}_{2}=\frac{14-13{\cal{C}}}{12}\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\Bigg]\,, \end{equation} with ${\cal{C}}\simeq 1.1967996$. Too see more details to obtain the non-linear parameter, see Refs.~\cite{Fel13a,Fer09,Byr14}. Now, we find the amplitudes of the non-Gaussianity in the equilateral and orthogonal configurations as \begin{equation} \label{eq41} f_{_{NL}}^{equil}=\frac{130}{36\sum_{i=1}^{3}k_{i}^{3}}\Bigg[\frac{1}{24}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\breve{\zeta}^{equil}\,, \end{equation} and \begin{equation} \label{eq42} f_{_{NL}}^{ortho}=\frac{140-130{\cal{C}}}{36\,\sum_{i=1}^{3}k_{i}^{3}}\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg) \Bigg]\breve{\zeta}^{ortho}\,. \end{equation} The equations \eqref{eq41} and \eqref{eq42} in the $k_{1}=k_{2}=k_{3}$ limit become \begin{equation} \label{eq43} f_{_{NL}}^{equil}=\frac{325}{18}\Bigg[\frac{1}{24}\bigg(\frac{1}{c_{s}^{2}}-1\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\,, \end{equation} and \begin{equation} \label{eq44} f_{_{NL}}^{ortho}=\frac{10}{9}\Big(\frac{65}{4}{\cal{C}}+\frac{7}{6}\Big)\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg) \Bigg]\,. \end{equation} Up to this point, we have presented the main equations describing the cosmological dynamics of the tachyon model. In the following, we adopt two types of scale factor (power-law and intermediate) for the tachyon model with constant sound speed and study its observational viability. \section{\label{sec3}Intermediate Tachyon Inflation with Constant-Sound Speed} In this section, we study the intermediate inflation in the tachyon model with constant sound speed. In the intermediate inflation, we have the following scale factor~\cite{Bar90,Bar93,Bar07} \begin{eqnarray} \label{eq45}a=a_{0}\,\exp\left(b\,t^{\beta}\right)\,, \end{eqnarray} where, $0<\beta<1$ and $b$ is a constant. This scale factor demonstrates that in this case, the expansion of the universe is slower than an exponential expansion and faster than a power-law expansion. From the intermediate scale factor (\ref{eq41}) we find the Hubble parameter as \begin{eqnarray} \label{eq46}H(N)=N \left( {\frac {N}{b}} \right) ^{-{\frac{1}{\beta}}}\beta\,. \end{eqnarray} To use this Hubble parameter and study the observational viability of the model, we need to reconstruct the potential and the inflation parameters in terms of the Hubble parameter and its derivatives~\cite{Bam14,Odi15}. From equation (\ref{eq8}) we obtain the potential as \begin{eqnarray} \label{eq47}V=3\,{\frac {{H}^{2}(N)\,{ cs}}{{\kappa}^{2}}}\,. \end{eqnarray} Also, the slow-roll parameters take the following form \begin{eqnarray} \label{eq48}\epsilon=-{\frac {{\frac { d}{{ d}N}}H \left( N \right) }{H \left( N \right) }} \,, \end{eqnarray} and \begin{eqnarray} \label{eq49}\eta=\frac { \left( \left( {\frac { d}{{ d}N}}H \left( N \right) \right) ^{2}\sqrt {1-c_{s}^{2}}+ \left( {\it cs}-1 \right) \left( {c_s}+1 \right) \left( H \left( N \right) {\frac {{ d}^{2}}{{ d}{N}^{2}}}H \left( N \right) +2\, \left( {\frac { d}{{ d}N}}H \left( N \right) \right) ^{2} \right) \right) }{\sqrt {1-{c_s^{2}}\,H \left( N \right) {\frac {\rm d}{{\rm d}N}}H \left( N \right) }} \,. \end{eqnarray} By using the above equations, we can rewrite equations (\ref{eq22}), (\ref{eq23}), (\ref{eq27}) and (\ref{eq28}), in terms of the Hubble parameter and its derivatives. After that, by using equation (\ref{eq46}), we can obtain the scalar spectral index, its running, tensor spectral index, and tensor-to-scalar ratio in terms of the intermediate parameters and also the sound speed. Then we perform some numerical analysis and, to obtain some constraints on the model's parameters, we compare the results with observational data. The results have been shown in several figures. Figure \ref{fig1} shows the tensor-to-scalar ratio of the intermediate tachyon model with constant sound speed versus the scalar spectral index in this model, in the background of the Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL and $95\%$ CL. As we can see from figure \ref{fig1}, this model in some ranges of the model's parameter space is consistent with observational data. Our numerical analysis shows that the intermediate tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $95\%$ CL, if $0< c_{s}\leq 1$ and $0.787\leq \beta \leq 1$. Also, this model is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL, if $0< c_{s}\leq 0.997$ and $0.833\leq \beta \leq 1$. Figure \ref{fig2} shows the running of the scalar spectral index of the intermediate tachyon model with constant sound speed versus the scalar spectral index in this model, in the background of the Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL and $95\%$ CL. By studying these parameters, we find that $\alpha_{s}-n_{s}$ of the intermediate tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL, if $0< c_{s}\leq 1$ and $0.756\leq \beta \leq 1$ and at $95\%$ CL, if $0< c_{s}\leq 1$ and $0.812\leq \beta \leq 1$. The evolution of the tensor-to-scalar ratio of the intermediate tachyon model with constant sound speed versus the tensor spectral index in this model, in the background of the Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data at $68\%$ CL and $95\%$ CL, has been shown in figure \ref{fig3}. The obtained constraints in this case are $0< c_{s}\leq 1$ and $0.081\leq \beta \leq 0.984$ at $68\%$ CL and $0< c_{s}\leq 1$ and $0.027\leq \beta \leq 1$ at $95\%$ CL. \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{1} \end{center} \caption{\small {Tensor-to-scalar ratio versus the scalar spectral index in the intermediate tachyon model with constant sound speed. }} \label{fig1} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{2} \end{center} \caption{\small {Running of the scalar spectral index versus the scalar spectral index in the intermediate tachyon model with constant sound speed.}} \label{fig2} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{3} \end{center} \caption{\small {Tensor-to-scalar ratio versus the tensor spectral index in the intermediate tachyon model with constant sound speed. }} \label{fig3} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.35]{4a} \includegraphics[scale=0.35]{4b} \end{center} \caption{\small {Left panel: ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{s}$, obtained from Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL (dark region) and $95\%$ CL (light region). Right panel: ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{T}$, obtained from Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data at $68\%$ CL (dark region) and $95\%$ CL (light region). }} \label{fig4} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{5} \end{center} \caption{\small {Ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of the amplitude of the scalar perturbation. }} \label{fig5} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.55]{6a} \includegraphics[scale=0.30]{6b} \end{center} \caption{\small {Left panel: orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity in the intermediate tachyon model with constant sound speed. Right panel: observationally viable range of the tensor-to-scalar ratio and sound speed squared, based on the observationally viable values of the orthogonal and equilateral amplitudes of the non-gaussianity.}} \label{fig6} \end{figure} We have also performed some numerical analysis to find the domain of $\beta$ and $c_{s}$, leading to observationally viable values of scalar spectral index, tensor spectral index, and tensor-to-scalar ratio. The results are shown in figure \ref{fig4}, where we have used both Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 and Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data sets at $68\%$ CL and $95\%$ CL. Both panels confirm that, depending on the value of sound speed, the observationally viable values of $\beta$ start from almost $0.75$. The interesting point is that for these ranges of parameter space, the amplitude of the scalar perturbation is consistent with observational data too. The constraint on the amplitude of the scalar perturbation, obtained from Planck2018 TT, TE, EE+lowE+lensing data, is $\ln (10^{10}{\cal_{A}}_{s})=3.044\pm 0.014$. By this observational constraint, we have obtained the observational viable domain of $c_{s}$ and $\beta$, shown in figure \ref{fig5}. Another way to check the observational viability of the model is to study the amplitudes of the non-gaussianity numerically and compare the results with observational data. To this end, we consider the amplitudes of the non-gaussianity in equilateral and orthogonal configurations, given by equations (\ref{eq43}) and (\ref{eq44}). In this way, we plot the behavior of the orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity, in the background of Planck2018 TTT, EEE, TTE and EET data. The result is shown in the left panel of figure \ref{fig6}. Our numerical analysis shows that the orthogonal and equilateral amplitudes of the non-gaussianity in the intermediate tachyon model with constant sound speed are consistent with Planck2018 TTT, EEE, TTE and EET data at $68\%$ CL if $0.276\leq c_{s}$, at $95\%$ CL if $0.213\leq c_{s}$, and at $97\%$ CL if $0.186\leq c_{s}$. These ranges of the sound speed lead to the observationally viable values of the tensor-to-scalar ratio too. In the right panel of figure \ref{fig6}, we have plotted the phase space of the tensor-to-scalar ratio and the sound speed squared, based on the domain of the sound speed leading to the observationally viable values of the amplitudes of the non-gaussianity. As the figure shows, the values of the tensor-to-scalar ratio are in the domain consistent with Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 data. We have summarized the prediction of the model for some sample values of the sound speed in table \ref{tab1}. According to our analysis, it seems that the intermediate tachyon model with constant sound speed is an observationally viable inflation model. \begin{table*} \tiny \caption{\small{\label{tab1} Ranges of the parameter $\beta$ in which the tensor-to-scalar ratio, the scalar spectral index, and the tensor spectral index of the intermediate tachyon model with constant sound speed are consistent with different data sets.}} \begin{center} \begin{tabular}{cccccc} \\ \hline \hline \\ & Planck2018 TT,TE,EE+lowE & Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE \\ & +lensing+BK14+BAO & +lensing+BK14+BAO&lensing+BK14+BAO&lensing+BK14+BAO \\ & & &+LIGO$\&$Virgo2016 &LIGO$\&$Virgo2016 \\ \hline \\$c_{s}$& $68\%$ CL & $95\%$ CL &$68\%$ CL & $95\%$ CL \\ \hline\hline \\ $0.1$& $0.981<\beta<1$ &$0.965<\beta<1$ &$0.312<\beta<0.859 $ & $0.250<\beta<1$\\ \\ \hline \\$0.4$&$0.968<\beta<1$ &$0.953<\beta<1$ &$0.630<\beta<0.960$ &$0.554<\beta<1$ \\ \\ \hline\\ $0.7$&$0.935<\beta<0.970 $&$0.920<\beta<0.995$&$0.748<\beta<0.976$ &$0.684<\beta<1$\\ \\ \hline\\ $0.9$&$0.885<\beta<0.931 $&$0.871<\beta<0.931$& $0.792<\beta<0.982 $ &$0.731<\beta<1$\\ \\ \hline \hline \end{tabular} \end{center} \end{table*} \section{\label{sec4}Power-Law Tachyon Inflation with Constant-Sound Speed} Now, we study the power-law inflation in the tachyon model with constant sound speed. In the power-law inflation, the scale factor is given by \begin{eqnarray} \label{eq50}a=a_{0}\,t^{n}\,, \end{eqnarray} leading to the following Hubble parameter \begin{eqnarray} \label{eq51}H(N)=N \,e^{-\frac{N}{n}}\,. \end{eqnarray} We use this Hubble parameter and substitute it in equations (\ref{eq47})-(\ref{eq49}) to find the inflation parameter in the power-law case. In this way, we can obtain equations (\ref{eq22}), (\ref{eq23}), (\ref{eq27}) and (\ref{eq28}) in terms of the model's parameters and perform some numerical analysis on the power-law tachyon inflation with constant sound speed. To obtain some constraints on the model's parameters, we compare the numerical results with observational data. According to our numerical analysis, the power-law tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $95\%$ CL if $0< c_{s}\leq 1$ and $222\leq n \leq 565$. Also, this model is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL if $0< c_{s}< 0.990$ and $293\leq n \leq 485$. The behavior of the tensor-to-scalar ratio has been shown in figure \ref{fig7}. The running of the scalar spectral index of the power-law tachyon model with constant sound speed versus the scalar spectral index in this model is shown in figure \ref{fig8}. By studying these parameters, we find that in this case, the model is consistent with Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL if $0< c_{s}\leq 1$ and $299\leq n \leq 473$ and at $95\%$ CL if $0< c_{s}\leq 1$ and $231\leq n \leq 551$. Figure \ref{eq9} shows the evolution of the tensor-to-scalar ratio of the power-law tachyon model with constant sound speed versus the tensor spectral index in this model, in the background of the Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data. In this case, the model has observational viability if $0< c_{s}\leq 1$ and $27.3\leq n \leq 3200$ at $68\%$ CL, and $0< c_{s}\leq 1$ and $1.67\leq n $ at $95\%$ CL. \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{7} \end{center} \caption{\small {Tensor-to-scalar ratio versus the scalar spectral index in the power-law tachyon model with constant sound speed.}} \label{fig7} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{8} \end{center} \caption{\small {Running of the scalar spectral index versus the scalar spectral index in the power-law tachyon model with constant sound speed.}} \label{fig8} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{9} \end{center} \caption{\small {Tensor-to-scalar ratio versus the tensor spectral index in the power-law tachyon model with constant sound speed.}} \label{fig9} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.35]{10a} \includegraphics[scale=0.35]{10b} \end{center} \caption{\small {Left panel: ranges of the parameters $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{s}$, obtained from Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL (dark region) and $95\%$ CL (light region). Right panel: ranges of the parameter $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{T}$, obtained from Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO, and Virgo2016 data at $68\%$ CL (dark region) and $95\%$ CL (light region).}} \label{fig10} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{11} \end{center} \caption{\small {Ranges of parameter $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of the amplitude of the scalar perturbation.}} \label{fig11} \end{figure} \begin{table*} \tiny \caption{\small{\label{tab2} Ranges of the parameter $n$ in which the tensor-to-scalar ratio, the scalar spectral index, and the tensor spectral index of the power-law tachyon model with constant sound speed are consistent with different data sets.}} \begin{center} \begin{tabular}{cccccc} \\ \hline \hline \\ & Planck2018 TT,TE,EE+lowE & Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE \\ & +lensing+BK14+BAO & +lensing+BK14+BAO&lensing+BK14+BAO&lensing+BK14+BAO \\ & & &+LIGO$\&$Virgo2016 &LIGO$\&$Virgo2016 \\ \hline \\$c_{s}$& $68\%$ CL & $95\%$ CL &$68\%$ CL & $95\%$ CL \\ \hline\hline \\ $0.1$& $368<n<492$ &$336<n<569$ &$27.3<n<300 $ & $20.1<n$\\ \\ \hline \\$0.4$&$354<n<488$ &$324<n<559$ &$102<n<1600$ &$76.0<n$ \\ \\ \hline\\ $0.7$&$330<n<447 $&$301<n<511$&$177<n<2930$ &$131<n$\\ \\ \hline\\ $0.9$&$305<n<373 $&$270<n<483$& $228<n<3130 $ &$167<n$\\ \\ \hline \hline \end{tabular} \end{center} \end{table*} The domain of $n$ and $c_{s}$, leading to observationally viable values of scalar spectral index, tensor spectral index, and the tensor-to-scalar ratio is shown in figure \ref{fig10}, where we have used both Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 and Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data sets at $68\%$ CL and $95\%$ CL. It is clear that, although from the observationally viable values of the tensor spectral index we can't find an upper bound on the parameter $n$, the observational values of the scalar spectral index imply an upper limit on it. These observationally viable ranges of $c_{s}$ and $n$ lead to the value of the amplitude of the scalar perturbations released by planck2018 data ($\ln (10^{10}{\cal_{A}}_{s})=3.044\pm 0.014$). Figure \ref{fig11} shows this issue clearly. Considering that the sound speed is constant, the behavior of the orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity in the power-law tachyon case is the same as the one in the intermediate tachyon case. This means that in this case also we have consistency with Planck2018 TTT, EEE, TTE and EET data at $68\%$ CL if $0.276\leq c_{s}$, at $95\%$ CL if $0.213\leq c_{s}$, and at $97\%$ CL if $0.186\leq c_{s}$. These ranges are compatible with the ranges obtained from $n_{s}$, $\alpha_{s}$, $n_{T}$ and $r$. Therefore, the power-law tachyon inflation with constant sound speed seems to be a fine model that in some ranges of its parameter space is consistent with observational data. In table \ref{tab2}, we have summarized some constraints obtained in studying the power-law tachyon inflation with constant sound speed. \section{\label{sec5}A Discussion on Unifying the Inflation with Dark Energy } Currently, cosmologists have been interested in the models covering a larger domain in the thermal history of the universe. In this way, the models that can describe both very early time and very late time accelerating expansion of the universe have attracted a lot of attention. In this respect, in some viable models of $f(R)$ gravity it is possible to describe both early time inflation and late time acceleration of the universe~\cite{Noj07,Noj08,Cog08,Noj11}. For instance, in Ref.~\cite{Noj08} the authors have considered a model of $f(R)$ gravity where the universe effectively starts with a large cosmological constant at the early universe (leading to inflation) and after passing the radiation/matter dominated era, reaches the small values of the cosmological constant (leading to late time accelerating expansion). In Ref.~\cite{Noj11} another model of $f(R)$ gravity that can explain the initial inflation, and at the late time can reproduce the behavior of the $\Lambda$CDM model. As we have mentioned in the ``Introduction" section, the equation of state of the tachyon field can be $-1$, corresponding to late time dark energy and early time inflationary phases of the universe. Therefore, in our model also, it seems possible to unify the inflation with dark energy. As we know, the tachyon inflation with constant sound speed is an observationally viable inflation model, at least in some domains of its parameter space. After inflation ends, and the tachyon field reaches the non-zero minimum of the potential, the universe becomes radiation dominated. During the radiation and matter dominated universe, the minimum value of the potential doesn't disturb the thermal history of the universe. With more expansion of the universe, the energy densities of the radiation and matter become small and eventually at the late time the potential become the dominant component in the energy density of the universe. In fact, this dominant component of the energy density can be considered as an effective cosmological constant, leading to late time accelerating expansion of the universe. It is also possible to consider some extension of the tachyon field so we can get both inflation and late time acceleration in the model. For instance, by considering the non-minimal coupling or non-minimal derivative coupling between the tachyon field and the gravity, depending on the values of the non-minimal parameter, it may possible to have a model covering both initial inflation and late time acceleration. \section{\label{sec6}Conclusion} In this paper, we have considered tachyon inflation where the sound speed is constant. We have obtained the Friedmann equations and equation of motion in terms of sound speed. After that, we have presented the main perturbation parameters, like the scalar spectral index, its running, tensor spectral index, and tensor-to-scalar ratio. We have also, by considering the three-point correlations, presented the amplitude of the non-gaussianity in both equilateral and orthogonal configurations. Then, we have adopted two types of scale factor: Intermediate and power-law. We have studied both cases separately. To study these cases, we have obtained the potential and inflation parameters in terms of the Hubble parameter, its derivative, and also the sound speed. In this way, we were able to study the model based on two types of scale factor. By considering the intermediate scale factor, we have studied $r-n_{s}$ behavior in the background of Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 data in both $68\%$ CL and $95\%$ CL and found some constraints on the model's parameters. The behavior of $\alpha_{s}-n_{s}$ in the background of Planck2018 TT, TE, EE+lowE+lensing data has been studied too. Another important parameter is the tensor scalar spectral index corresponding to the primordial perturbations. In this way, we have studied $r-n_{T}$ behavior in the background of Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data to find some constraints on the model's parameter space. We have also analyzed $c_{s}$ and $\beta$ phase space in both $68\%$ CL and $95\%$ CL. By studying these parameters numerically, we have found that intermediate tachyon inflation with constant sound speed is observationally viable if $0<c_{s}\leq 0.997$ and $0.833\leq\beta\leq 0.984$ at $68\%$ CL, and $0<c_{s}\leq 1$ and $0.787\leq\beta\leq 1$ at $95\%$ CL. To check the viability of the model more precisely, we have studied the non-gaussian feature of the primordial perturbations. In this regard, we have considered the Planck2018 TTT, EEE, TTE and EET constraints on the equilateral and orthogonal configurations of the non-gaussianity. In this way, we have obtained the constraints on the sound speed as $0.276\leq c_{s}\leq 1$ at $68\%$ CL, $0.213\leq c_{s}\leq 1$ at $95\%$ CL, and $0.186\leq c_{s}\leq 1$ at $97\%$ CL. We have also studied $c_{s}^{2}-r$ phase space and found that the ranges of the sound speed, obtained from the observational constraints on the non-gaussianity, lead to the observationally viable values of the tensor-to-scalar ratio. The power-law tachyon inflation with constant sound speed is another model that has been studied in this paper. Similar to the intermediate tachyon inflation case, we have studied the perturbation parameters in this model, but here, based on the power-law scale factor. In this case, by using different data sets, we have studied $r-n_{s}$, $\alpha-n_{s}$, and $r-n_{T}$ behavior. The phase space of the parameters $c_{s}$ and $n$, based on the observationally viable ranges of the scalar and tensor spectral indices has been studied too. Our numerical analysis has shown that the power-law tachyon inflation with constant sound speed is observationally viable for $0<c_{s}\leq 0.990$ and $299\leq n<\leq 473$ at $68\%$ CL, and $0<c_{s}\leq 1$ and $231\leq n <\leq 551$ at $95\%$ CL. For these obtained ranges, the amplitude of the scalar spectral index is consistent with observational data. We have also discussed the issue of unifying the inflation with dark energy in the tachyon model with constant sound speed. We have argued that the non-zero minimum value of the tachyon potential becomes important at the late time, after the energy density of the radiation/matter decreases. Based on our analysis, it seems that both intermediate and power-law tachyon models in some ranges of their parameter space are consistent with observational data.\\ {\bf Acknowledgement}\\ I thank the referee for the very insightful comments that have improved the quality of the paper considerably.\\ \section{Introduction} Considering that the early time inflation in the history of the universe seems to have occurred around Planck’s scale, the M/String theory inspired fields can be a possible candidate for the field responsible for inflation. Tachyon inflation is one of these fields leading to interesting cosmological results. Tachyon field, which has attained a lot of attention, is associated with the D-branes in string theory. One interesting implication of this field is that, its slow-rolling down to the potential leads to the smooth evolution of the universe from the accelerating phase of the expansion to the era dominated by the non-relativistic fluid~\cite{Sen99,Sen02a,Sen02b,Gib02}. On the other hand, we know that in the simple inflation model with canonical scalar fields, we find the primordial perturbation to be scale-invariant and gaussian~\cite{Gut81,Lin82,Alb82,Lin90,Lid00a,Lid97,Rio02,Lyt09,Mal03}. When we consider an inflation model with a canonical scalar field, we find a gaussian perturbation that the values of its tensor-to-scalar ratio for $\phi$, $\phi^2$, $\phi^{4/3}$ are not consistent with observational data~\cite{pl18b}. Therefore we have to seek some other specific potential like hilltop potential to get an observationally viable single canonical field inflation model~\cite{pl18b}. However, with every potential, the amplitude of the primordial perturbation in the single canonical field inflation is almost gaussian. Or, we should consider a nonminimal inflation model to give a better explanation of the early time cosmological inflation. However, with a non-canonical scalar field like tachyon, it is possible to get the scale-dependent and non-gaussian distribution for the amplitude of the primordial perturbations. Another interesting point about the tachyon field is corresponding to the equation of state of the tachyon field. This parameter in the tachyon field can be $-1$, which describes both very early time and very late time accelerating expansion of the universe. Also, its value can be $0$, describing the matter/dark matter dominated era. Therefore, with the tachyon field it is possible to explain the thermal history of the universe in a reasonable way~\cite{Gib02}. These are interesting issues that motivate cosmologists to consider and study the inflation models driven by a tachyon field as a non-canonical scalar field. In this regard, a lot of interesting works based on tachyon inflation have been done. For instance, the authors of Ref.~\cite{Kam18} have considered a tachyon field in the context of quantum loop gravity. In this way, they have obtained the inflation and perturbation parameters and tested the observational viability of some inflation models. In ref.~\cite{Bil19}, by considering the holographic cosmology, the tachyon inflation has been studied and the results have been compared to the observational data. The authors of Ref.~\cite{Moh20} have studied an inflation model where a tachyon field interacts with photon gas. They have shown that under some assumptions and conditions this model shows some agreement with observational data. In~\cite{Ras21}, the tachyon model with a superpotential, a potential based on supersymmetry, has been studied. It has been shown that tachyon inflation with a superpotential, at least in some ranges of the model's parameter space, is consistent with observational data. There are other works on tachyon inflation leading to interesting cosmological results~\cite{Noj03,Noz14,Bou16,Rez17,Ras18,Ras20} and all of these works show that the tachyon field can be considered as a field running the inflation. One of the important parameters in the inflation models is the sound speed of the primordial perturbation, shown by $c_{s}$. The sound speed is corresponding to the Lorentz factor $\gamma$ as $c_{s}=\frac{1}{\gamma}$. For the canonical scalar field, we get the sound speed equal to unity. However, with the non-canonical scalar fields, we have $c_{s}^{2}\neq 1$. The value of the parameter $\gamma$ and correspondingly sound speed determines the deformation of the field's kinetic energy from the canonical one. The constant sound speed is an interesting idea motivated by the authors in Refs.~\cite{Spa08,Tsu13}. Considering that in the non-canonical scalar field the sound speed is related to the time variation of the field, the constant sound speed is corresponding to the constant field's variation in times. This can lead to interesting results. Now, the question is what constant values of $c_{s}$ we should adopt, to study the model numerically. Considering that the sound speed is related to the amplitude of the non-gaussianity, by using the observational constraints on the amplitude of the non-gaussianity, we can find suitable ranges of the sound speed. Also, there are some other parameters, such as the scalar spectral index, tensor spectral index, and tensor-to-scalar ratio that help us to find some constraints on the sound speed values. To perform a numerical analysis, we use the Planck2018 constraints on the inflation parameters. Based on the $\Lambda CDM+r+\frac{dn_{s}}{d\ln k}$ model, the Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data gives the value of the scalar spectral index as $n_{s}=0.9658\pm0.0038$ and implies a constraint on the tensor-to-scalar ratio as $r<0.072$~\cite{pl18a,pl18b}. The constraint on the tensor spectral index, released by Planck2018 TT, TE, EE +lowE+lensing+BK14+BAO+LIGO and Virgo2016 data is $-0.62<n_{T}<0.53$~\cite{pl18a,pl18b}. Also, from Planck2018 TT, TE, EE+lowEB+lensing data, we have the value of the running of the scalar spectral index as $\alpha_{s}=-0.0085\pm 0.0073$~\cite{pl18a,pl18b}. Other useful constraints are corresponding to the amplitude of the non-gaussianity. The planck2018 combined temperature and polarization analysis gives the constraints on the amplitude of the non-gaussianity in the equilateral configuration as $f_{NL}^{equil}=-26\pm 47$ and in the orthogonal configuration as $f_{NL}^{ortho}=-38\pm 24$~\cite{pl19}. These several constraints determine the observational viability of every inflation model. With these preliminaries, this paper is organized as follows. In section \ref{sec2}, we review the inflation in a tachyon model with constant sound speed. In this section, we present the main equations governing the dynamics of the model in terms of the sound speed. In section \ref{sec3}, we consider an intermediate scale factor and study the perturbation and non-gaussianity parameters in this model numerically. We compare the results with several data sets to find some constraints on the model's parameter space. In section \ref{sec4} we study the power-law tachyon inflation with constant sound speed and compare the results with observational data. In section \ref{sec5}, we perform some discussion on unifying the initial inflation with late time dark energy in the tachyon model with constant sound speed. In section \ref{sec6}, we present a summary of this work. \section{\label{sec2}Review on the Tachyon Inflation with Constant Sound Speed} For the tachyon field, we have the following Dirac-Born-Infeld type effective 4-dimensional action \begin{eqnarray} \label{eq1} S=\int d^{4}x\,\sqrt{-g} \Bigg[\frac{1}{2\kappa^{4}}R-V(\phi)\,\sqrt{1-2\,X}\Bigg]\,, \end{eqnarray} with $\kappa$ being the gravitational constant, $R$ as the Ricci scalar, and $X=-\frac{1}{2}\,\partial_{\mu}\phi\,\partial^{\mu}\phi$. Also, the tachyon field $\phi$ has the potential $V(\phi)$. Cosmologists believe that the physics of the tachyon condensation can be described by such an action. Einstein's field equations, corresponding to action (\ref{eq1}), are given by \begin{eqnarray} \label{eq2} G_{\mu\nu}=\kappa^{2}\Bigg[-g_{\mu\nu}V(\phi)\sqrt{1-2\, X}+\frac{ V(\phi)\partial_{\mu}\phi\,\partial_{\nu}\phi}{\sqrt{1-2\, X}}\Bigg]\,, \end{eqnarray} which have been obtained by varying action (\ref{eq1}) with respect to the metric. Considering the FRW metric as \begin{equation} \label{eq3} ds^{2}=-dt^{2}+a^{2}(t)\delta_{ij}dx^{i}dx^{j}\,, \end{equation} we find the following Friedmann equation \begin{eqnarray} \label{eq4} 3H^{2}=\frac{\kappa^{2}\,V(\phi)}{\sqrt{1-\dot{\phi}^{2}}}\,. \end{eqnarray} corresponding to the $(0,0)$ component of Einstein's field equations. Note that, a dot on the parameter shows the derivative of the parameter with respect to the time. Also, the $(i,i)$ component of Einstein's field equations give the following second Friedmann equation \begin{eqnarray} \label{eq5} 2\dot{H}+3H^{2}=\kappa^{2}\bigg[V(\phi)\,\sqrt{1-\dot{\phi}^{2}}\,\bigg]\,. \end{eqnarray} We find the equation of motion of the tachyon field by varying the action (\ref{eq1}) with respect to the field as \begin{equation} \label{eq6}\frac{\ddot{\phi}}{1-\dot{\phi}^{2}}+3\,H\dot{\phi} +\frac{V'(\phi)}{V(\phi)}=0\,, \end{equation} where a prime shows the derivative with respect to the tachyon field. Since our purpose in this paper is to study the tachyon model with constant sound speed, we should re-write the above main equations in terms of the sound speed. To this end, we need to define the sound speed of the tachyon field. In general, the square of the sound speed is given by $c_{s}^{2}\equiv\frac{P_{,X}}{\rho_{,X}}$, that in the tachyon model takes the following form \begin{eqnarray} \label{eq7} c_{s}=\sqrt{1-\dot{\phi}^{2}}\,. \end{eqnarray} In this regard, the equations (\ref{eq4})-(\ref{eq6}) take the following forms \begin{eqnarray} \label{eq8} 3H^{2}=\frac{\kappa^{2}\,V(\phi)}{c_{s}}\,, \end{eqnarray} \begin{eqnarray} \label{eq9} 2\dot{H}+3H^{2}=\kappa^{2}\,V(\phi)\,c_{s}\,, \end{eqnarray} and \begin{equation} \label{eq10}3\,H\,\sqrt {1- c_{s}^{2}} +\frac{V'(\phi)}{V(\phi)}=0\,, \end{equation} where we have considered the sound speed as a constant parameter. To study cosmological inflation, we need the following slow-roll parameters \begin{eqnarray} \label{eq11}\epsilon=-\frac{\dot{H}}{H^{2}}\,, \end{eqnarray} \begin{eqnarray} \label{eq12}\eta=-\frac{1}{H}\frac{\ddot{H}}{\dot{H}}\,. \end{eqnarray} Note that, in the case of the constant sound speed, the third slow-roll parameter $s=\frac{1}{H}\frac{\dot{c_{s}}}{c_{s}}$ is zero. Another needed parameter to study inflation is the number of e-folds parameter defined as \begin{eqnarray} \label{eq13} N=\int H\,dt\,. \end{eqnarray} These parameters are useful to study the primordial perturbations in the model and compare the results with observational data. The observational data that we use in our work is the data released by the Planck2018 team~\cite{pl18a,pl18b,pl19}. Based on the fact that, under the assumption of statistical isotropy, the two-point correlations of the CMB anisotropies are described by the angular power spectra $C_{l}^{TT}$, $C_{l}^{TE}$, $C_{l}^{EE}$, and $C_{l}^{BB}$ (where the subscript $l$ shows the multipole moment number)~\cite{kam97,zal30,sel97,Hu97,Hu98}, the Planck team has obtained some constraints on the important perturbation parameters. To include the contributions from the scalar and tensor perturbations in the CMB angular power spectra, the Planck team has used the following expressions~\cite{pl15} \begin{eqnarray} \label{eq14} C_{l}^{ab,s}=\int_{0}^{\infty} \frac{dk}{k} \Delta_{l,a}^{s} (k) \, \Delta_{l,b}^{s} (k)\, {\cal{A}}_{s} (k)\,, \end{eqnarray} \begin{eqnarray} \label{eq15} C_{l}^{ab,T}=\int_{0}^{\infty} \frac{dk}{k} \Delta_{l,a}^{T} (k) \, \Delta_{l,b}^{T} (k)\, {\cal{A}}_{T} (k)\,. \end{eqnarray} In equations (\ref{eq14}) and (\ref{eq15}), we have shown the transfer functions by $\Delta_{l,a}^{s} (k)$ and $\Delta_{l,a}^{T} (k)$. The parameter $a$ and $b$ are given as $a,b=T,E,B$. Also, the primordial power spectrum, the parameter identified by the physics of the primordial universe~\cite{pl15}, is shown by ${\cal{A}}_{i} (k)$ (where, $i=s,T$). Model-independent forms of the scalar and tensor power spectra are given by \begin{eqnarray} \label{eq16} {\cal{A}}_{s} (k)=A_{s}\left(\frac{k}{k_{*}}\right)^{n_{s}-1+\frac{1}{2}\frac{dn_{s}}{d\ln k}\ln\big(\frac{k}{k_{*}}\big)+\frac{1}{6}\frac{d^{2}n_{s}}{d\ln k^{2}}\ln\big(\frac{k}{k_{*}}\big)^{2}+...}\,, \end{eqnarray} \begin{eqnarray} \label{eq17} {\cal{A}}_{T} (k)=A_{T}\left(\frac{k}{k_{*}}\right)^{n_{T}+\frac{1}{2}\frac{dn_{T}}{d\ln k}\ln\big(\frac{k}{k_{*}}\big)+...}\,. \end{eqnarray} These forms of the scalar and tensor power spectra have been used by the Planck team to compare the perturbation parameters with data. Note that, in equations (\ref{eq16}) and (\ref{eq17}), $A_{j}$ shows the scalar (corresponding to $j=s$) and tensor (corresponding to $j=T$) perturbations. Also, the running of the scalar or tensor spectral index and the running of the running of the scalar spectral index are given by $\frac{dn_{i}}{d\ln k}$ and $\frac{d^{2}n_{s}}{d\ln k^{2}}$, respectively. One can also find the tensor-to-scalar ratio by using the power spectra as \begin{eqnarray} \label{eq18} r=\frac{{\cal{A}}_{T}(k_{*})}{{\cal{A}}_{s}(k_{*})}\,, \end{eqnarray} which is a very important parameter in studying the inflation models. The parameters $n_{s}$ and $n_{T}$ are the scalar and tensor spectral indices showing the scale dependence of the primordial perturbations. The scalar spectral index, at the time of sound horizon exit of the physical scales, is defined as \begin{eqnarray} \label{eq19} n_{s}-1=\frac{d \ln {\cal{A}}_{s}}{d\ln k}\Bigg|_{c_{s}k=aH}\,, \end{eqnarray} where the parameter ${\cal{A}}_{s}$, the amplitude of the scalar spectral index, is given by the following definition \begin{equation} \label{eq20}{\cal{A}}_{s}=\frac{H^{2}}{8\pi^{2}{\cal{W}}_{s}c_{s}^{3}}\,, \end{equation} with \begin{equation} \label{eq21} {\cal{W}}_{s}=\frac {V\, (1-c_{s}^2)}{2{H}^{2} c_{s}^{3}}\,. \end{equation} The scalar spectral index, in terms of the slow-roll parameters, is written as \begin{eqnarray} \label{eq22} n_{s}=1-6\epsilon+2\eta\,. \end{eqnarray} This is one of the parameters that can be compared with observational data. Another important parameter in studying the inflation models is the running of the scalar spectral index, defined as \begin{eqnarray} \label{eq23} \alpha_{s}=\frac{dn_{s}}{d\ln k}=8\epsilon\,\eta-12\epsilon^{2}-2\zeta+2\eta^{2}\,, \end{eqnarray} where, the parameter $\zeta$ is given by \begin{eqnarray} \label{eq24} \zeta=\frac{\dddot{H}}{H^{2}\dot{H}}\,. \end{eqnarray} The tensor spectral index, obtained from equation (\ref{eq17}), is defined as follows \begin{eqnarray} \label{eq25} n_{T}=\frac{d \ln {\cal{A}}_{T}}{d\ln k}\Bigg|_{k=aH}\,, \end{eqnarray} with \begin{eqnarray} \label{eq26} {\cal{A}}_{T}=\frac{2\kappa^{2}H^{2}}{\pi^{2}}\,, \end{eqnarray} being the amplitude of the tensor perturbations. In terms of the slow-roll parameters, the tensor spectral index is expressed as \begin{eqnarray} \label{eq27} n_{T}=-2\epsilon\,. \end{eqnarray} Finally, we have the following expression for the tensor-to-scalar ratio \begin{eqnarray} \label{eq28} r=16\,c_{s}\,\epsilon\,. \end{eqnarray} As we see in equation (\ref{eq28}), the tensor-to-scalar ratio depends explicitly on the sound speed of the perturbations. Therefore, from the observationally viable value of $r$, we can constrain the values of the sound speed. Another important parameter that depends on the sound speed parameter is the nonlinearity parameter. The nonlinearity parameter demonstrates the non-gaussian property of the amplitude of the primordial perturbation. The Gaussian distributed primordial perturbations are characterized by a two-point correlation. To seek the additional statistical information, corresponding to the non-Gaussian distributed perturbation, it is necessary to consider the higher-order correlations. The three-point correlation function for the spatial curvature perturbation in the interaction picture is given by~\cite{Mal03} \begin{eqnarray} \label{eq29} \langle {\Phi}(\textbf{k}_{1})\,{\Phi}(\textbf{k}_{2})\,{\Phi}(\textbf{k}_{3})\rangle =(2\pi)^{3}\delta^{3}(\textbf{k}_{1}+\textbf{k}_{2}+\textbf{k}_{3}){\cal{B}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})\,, \end{eqnarray} with the following definition for the parameter ${\cal{B}}_{\Phi}$ \begin{equation} \label{eq30} {\cal{B}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})=\frac{(2\pi)^{4}{\cal{A}}_{s}^{2}}{\prod_{i=1}^{3} k_{i}^{3}}\, {\cal{D}}_{\Phi}(\textbf{k}_{1},\textbf{k}_{2},\textbf{k}_{3})\,. \end{equation} The parameter ${\cal{D}}_{\Phi}$ in equation (\ref{eq40}) is expressed as follows \begin{eqnarray} \label{eq31} {\cal{D}}_{\Phi}=\Bigg(1-\frac{1}{c_{s}^{2}}\Bigg) \Bigg[\frac{3}{4}{\cal{J}}_{1}-\frac{3}{2}{\cal{J}}_{2} -\frac{1}{4}{\cal{J}}_{3}\Bigg]\,, \end{eqnarray} where \begin{eqnarray} \label{eq32} {\cal{J}}_{1}=\frac{2\sum_{i>j}k_{i}^{2}\,k_{j}^{2}}{k_{1}+k_{2}+k_{3}}-\frac{\sum_{i\neq j}k_{i}^{2}\,k_{j}^{3}}{(k_{1}+k_{2}+k_{3})^{2}}\,, \end{eqnarray} \begin{eqnarray} \label{eq33} {\cal{J}}_{2}=\frac{\left(k_{1}\,k_{2}\,k_{3}\right)^{2}} {(k_{1}+k_{2}+k_{3})^{3}}\,, \end{eqnarray} and \begin{eqnarray} \label{eq34} {\cal{J}}_{3}=\frac{2\sum_{i>j}k_{i}^{2}\,k_{j}^{2}}{k_{1}+k_{2}+k_{3}}-\frac{\sum_{i\neq j}k_{i}^{2}\,k_{j}^{3}}{(k_{1}+k_{2}+k_{3})^{2}}+\frac{1}{2}\sum_{i}k_{i}^{3}\,. \end{eqnarray} By using the parameter ${\cal{D}}_{\Phi}$, one can find the non-linear parameter as \begin{equation} \label{eq35} f_{NL}=\frac{10}{3}\frac{{\cal{D}}_{\Phi}}{\sum_{i=1}^{3}k_{i}^{3}}\,. \end{equation} Note that, in equation (\ref{eq35}), there is a parameter $k_{i}$ which is the momentum. Depending on different values of the momenta ($k_{1}$, $k_{2}$, and $k_{3}$), we get different shapes of the non-gaussianity. In every shape, there is a maximal peak in the amplitude of the perturbations that defines the kind of the corresponding shape. We are interested in the case with a peak at $k_{1}=k_{2}=k_{3}$ limit, leading to an equilateral shape, and the case orthogonal to it~\cite{Che07,Bab04b,Fel13a,Bau12}. It is possible to write the bispectrum (\ref{eq31}) in terms of the equilateral and orthogonal shapes basis as follows~\cite{Fel13a} \begin{equation} \label{eq36} {\cal{D}}_{\Phi}={\cal{M}}_{1}\,\breve{{\cal{J}}}^{equil} + {\cal{M}}_{2} \,\breve{{\cal{J}}}^{ortho}\,, \end{equation} where, \begin{equation} \label{eq37} \breve{{\cal{J}}}^{equil}=-\frac{12}{13}\Big(3{\cal{J}}_{1}-{\cal{J}}_{2}\Big)\,. \end{equation} \begin{equation} \label{eq38} \breve{{\cal{J}}}^{ortho}=\frac{12}{14-13{\cal{C}}}\Big({\cal{C}}\big(3{\cal{J}}_{1}-{\cal{J}}_{2}\big)+3{\cal{J}}_{1}-{\cal{J}}_{2}\Big)\,, \end{equation} \begin{equation} \label{eq39} {\cal{M}}_{1}=\frac{13}{12}\Bigg[\frac{1}{24}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\,, \end{equation} and \begin{equation} \label{eq40} {\cal{M}}_{2}=\frac{14-13{\cal{C}}}{12}\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\Bigg]\,, \end{equation} with ${\cal{C}}\simeq 1.1967996$. Too see more details to obtain the non-linear parameter, see Refs.~\cite{Fel13a,Fer09,Byr14}. Now, we find the amplitudes of the non-Gaussianity in the equilateral and orthogonal configurations as \begin{equation} \label{eq41} f_{_{NL}}^{equil}=\frac{130}{36\sum_{i=1}^{3}k_{i}^{3}}\Bigg[\frac{1}{24}\bigg(1-\frac{1}{c_{s}^{2}}\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\breve{\zeta}^{equil}\,, \end{equation} and \begin{equation} \label{eq42} f_{_{NL}}^{ortho}=\frac{140-130{\cal{C}}}{36\,\sum_{i=1}^{3}k_{i}^{3}}\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg) \Bigg]\breve{\zeta}^{ortho}\,. \end{equation} The equations \eqref{eq41} and \eqref{eq42} in the $k_{1}=k_{2}=k_{3}$ limit become \begin{equation} \label{eq43} f_{_{NL}}^{equil}=\frac{325}{18}\Bigg[\frac{1}{24}\bigg(\frac{1}{c_{s}^{2}}-1\bigg)\bigg(2+3{\cal{C}}\bigg) \Bigg]\,, \end{equation} and \begin{equation} \label{eq44} f_{_{NL}}^{ortho}=\frac{10}{9}\Big(\frac{65}{4}{\cal{C}}+\frac{7}{6}\Big)\Bigg[\frac{1}{8}\bigg(1-\frac{1}{c_{s}^{2}}\bigg) \Bigg]\,. \end{equation} Up to this point, we have presented the main equations describing the cosmological dynamics of the tachyon model. In the following, we adopt two types of scale factor (power-law and intermediate) for the tachyon model with constant sound speed and study its observational viability. \section{\label{sec3}Intermediate Tachyon Inflation with Constant-Sound Speed} In this section, we study the intermediate inflation in the tachyon model with constant sound speed. In the intermediate inflation, we have the following scale factor~\cite{Bar90,Bar93,Bar07} \begin{eqnarray} \label{eq45}a=a_{0}\,\exp\left(b\,t^{\beta}\right)\,, \end{eqnarray} where, $0<\beta<1$ and $b$ is a constant. This scale factor demonstrates that in this case, the expansion of the universe is slower than an exponential expansion and faster than a power-law expansion. From the intermediate scale factor (\ref{eq41}) we find the Hubble parameter as \begin{eqnarray} \label{eq46}H(N)=N \left( {\frac {N}{b}} \right) ^{-{\frac{1}{\beta}}}\beta\,. \end{eqnarray} To use this Hubble parameter and study the observational viability of the model, we need to reconstruct the potential and the inflation parameters in terms of the Hubble parameter and its derivatives~\cite{Bam14,Odi15}. From equation (\ref{eq8}) we obtain the potential as \begin{eqnarray} \label{eq47}V=3\,{\frac {{H}^{2}(N)\,{ cs}}{{\kappa}^{2}}}\,. \end{eqnarray} Also, the slow-roll parameters take the following form \begin{eqnarray} \label{eq48}\epsilon=-{\frac {{\frac { d}{{ d}N}}H \left( N \right) }{H \left( N \right) }} \,, \end{eqnarray} and \begin{eqnarray} \label{eq49}\eta=\frac { \left( \left( {\frac { d}{{ d}N}}H \left( N \right) \right) ^{2}\sqrt {1-c_{s}^{2}}+ \left( {\it cs}-1 \right) \left( {c_s}+1 \right) \left( H \left( N \right) {\frac {{ d}^{2}}{{ d}{N}^{2}}}H \left( N \right) +2\, \left( {\frac { d}{{ d}N}}H \left( N \right) \right) ^{2} \right) \right) }{\sqrt {1-{c_s^{2}}\,H \left( N \right) {\frac {\rm d}{{\rm d}N}}H \left( N \right) }} \,. \end{eqnarray} By using the above equations, we can rewrite equations (\ref{eq22}), (\ref{eq23}), (\ref{eq27}) and (\ref{eq28}), in terms of the Hubble parameter and its derivatives. After that, by using equation (\ref{eq46}), we can obtain the scalar spectral index, its running, tensor spectral index, and tensor-to-scalar ratio in terms of the intermediate parameters and also the sound speed. Then we perform some numerical analysis and, to obtain some constraints on the model's parameters, we compare the results with observational data. The results have been shown in several figures. Figure \ref{fig1} shows the tensor-to-scalar ratio of the intermediate tachyon model with constant sound speed versus the scalar spectral index in this model, in the background of the Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL and $95\%$ CL. As we can see from figure \ref{fig1}, this model in some ranges of the model's parameter space is consistent with observational data. Our numerical analysis shows that the intermediate tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $95\%$ CL, if $0< c_{s}\leq 1$ and $0.787\leq \beta \leq 1$. Also, this model is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL, if $0< c_{s}\leq 0.997$ and $0.833\leq \beta \leq 1$. Figure \ref{fig2} shows the running of the scalar spectral index of the intermediate tachyon model with constant sound speed versus the scalar spectral index in this model, in the background of the Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL and $95\%$ CL. By studying these parameters, we find that $\alpha_{s}-n_{s}$ of the intermediate tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL, if $0< c_{s}\leq 1$ and $0.756\leq \beta \leq 1$ and at $95\%$ CL, if $0< c_{s}\leq 1$ and $0.812\leq \beta \leq 1$. The evolution of the tensor-to-scalar ratio of the intermediate tachyon model with constant sound speed versus the tensor spectral index in this model, in the background of the Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data at $68\%$ CL and $95\%$ CL, has been shown in figure \ref{fig3}. The obtained constraints in this case are $0< c_{s}\leq 1$ and $0.081\leq \beta \leq 0.984$ at $68\%$ CL and $0< c_{s}\leq 1$ and $0.027\leq \beta \leq 1$ at $95\%$ CL. \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{1} \end{center} \caption{\small {Tensor-to-scalar ratio versus the scalar spectral index in the intermediate tachyon model with constant sound speed. }} \label{fig1} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{2} \end{center} \caption{\small {Running of the scalar spectral index versus the scalar spectral index in the intermediate tachyon model with constant sound speed.}} \label{fig2} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{3} \end{center} \caption{\small {Tensor-to-scalar ratio versus the tensor spectral index in the intermediate tachyon model with constant sound speed. }} \label{fig3} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.35]{4a} \includegraphics[scale=0.35]{4b} \end{center} \caption{\small {Left panel: ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{s}$, obtained from Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL (dark region) and $95\%$ CL (light region). Right panel: ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{T}$, obtained from Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data at $68\%$ CL (dark region) and $95\%$ CL (light region). }} \label{fig4} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{5} \end{center} \caption{\small {Ranges of the parameters $c_{s}$ and $\beta$ in the intermediate tachyon inflation with constant sound speed, leading to observationally viable values of the amplitude of the scalar perturbation. }} \label{fig5} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.55]{6a} \includegraphics[scale=0.30]{6b} \end{center} \caption{\small {Left panel: orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity in the intermediate tachyon model with constant sound speed. Right panel: observationally viable range of the tensor-to-scalar ratio and sound speed squared, based on the observationally viable values of the orthogonal and equilateral amplitudes of the non-gaussianity.}} \label{fig6} \end{figure} We have also performed some numerical analysis to find the domain of $\beta$ and $c_{s}$, leading to observationally viable values of scalar spectral index, tensor spectral index, and tensor-to-scalar ratio. The results are shown in figure \ref{fig4}, where we have used both Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 and Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data sets at $68\%$ CL and $95\%$ CL. Both panels confirm that, depending on the value of sound speed, the observationally viable values of $\beta$ start from almost $0.75$. The interesting point is that for these ranges of parameter space, the amplitude of the scalar perturbation is consistent with observational data too. The constraint on the amplitude of the scalar perturbation, obtained from Planck2018 TT, TE, EE+lowE+lensing data, is $\ln (10^{10}{\cal_{A}}_{s})=3.044\pm 0.014$. By this observational constraint, we have obtained the observational viable domain of $c_{s}$ and $\beta$, shown in figure \ref{fig5}. Another way to check the observational viability of the model is to study the amplitudes of the non-gaussianity numerically and compare the results with observational data. To this end, we consider the amplitudes of the non-gaussianity in equilateral and orthogonal configurations, given by equations (\ref{eq43}) and (\ref{eq44}). In this way, we plot the behavior of the orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity, in the background of Planck2018 TTT, EEE, TTE and EET data. The result is shown in the left panel of figure \ref{fig6}. Our numerical analysis shows that the orthogonal and equilateral amplitudes of the non-gaussianity in the intermediate tachyon model with constant sound speed are consistent with Planck2018 TTT, EEE, TTE and EET data at $68\%$ CL if $0.276\leq c_{s}$, at $95\%$ CL if $0.213\leq c_{s}$, and at $97\%$ CL if $0.186\leq c_{s}$. These ranges of the sound speed lead to the observationally viable values of the tensor-to-scalar ratio too. In the right panel of figure \ref{fig6}, we have plotted the phase space of the tensor-to-scalar ratio and the sound speed squared, based on the domain of the sound speed leading to the observationally viable values of the amplitudes of the non-gaussianity. As the figure shows, the values of the tensor-to-scalar ratio are in the domain consistent with Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 data. We have summarized the prediction of the model for some sample values of the sound speed in table \ref{tab1}. According to our analysis, it seems that the intermediate tachyon model with constant sound speed is an observationally viable inflation model. \begin{table*} \tiny \caption{\small{\label{tab1} Ranges of the parameter $\beta$ in which the tensor-to-scalar ratio, the scalar spectral index, and the tensor spectral index of the intermediate tachyon model with constant sound speed are consistent with different data sets.}} \begin{center} \begin{tabular}{cccccc} \\ \hline \hline \\ & Planck2018 TT,TE,EE+lowE & Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE \\ & +lensing+BK14+BAO & +lensing+BK14+BAO&lensing+BK14+BAO&lensing+BK14+BAO \\ & & &+LIGO$\&$Virgo2016 &LIGO$\&$Virgo2016 \\ \hline \\$c_{s}$& $68\%$ CL & $95\%$ CL &$68\%$ CL & $95\%$ CL \\ \hline\hline \\ $0.1$& $0.981<\beta<1$ &$0.965<\beta<1$ &$0.312<\beta<0.859 $ & $0.250<\beta<1$\\ \\ \hline \\$0.4$&$0.968<\beta<1$ &$0.953<\beta<1$ &$0.630<\beta<0.960$ &$0.554<\beta<1$ \\ \\ \hline\\ $0.7$&$0.935<\beta<0.970 $&$0.920<\beta<0.995$&$0.748<\beta<0.976$ &$0.684<\beta<1$\\ \\ \hline\\ $0.9$&$0.885<\beta<0.931 $&$0.871<\beta<0.931$& $0.792<\beta<0.982 $ &$0.731<\beta<1$\\ \\ \hline \hline \end{tabular} \end{center} \end{table*} \section{\label{sec4}Power-Law Tachyon Inflation with Constant-Sound Speed} Now, we study the power-law inflation in the tachyon model with constant sound speed. In the power-law inflation, the scale factor is given by \begin{eqnarray} \label{eq50}a=a_{0}\,t^{n}\,, \end{eqnarray} leading to the following Hubble parameter \begin{eqnarray} \label{eq51}H(N)=N \,e^{-\frac{N}{n}}\,. \end{eqnarray} We use this Hubble parameter and substitute it in equations (\ref{eq47})-(\ref{eq49}) to find the inflation parameter in the power-law case. In this way, we can obtain equations (\ref{eq22}), (\ref{eq23}), (\ref{eq27}) and (\ref{eq28}) in terms of the model's parameters and perform some numerical analysis on the power-law tachyon inflation with constant sound speed. To obtain some constraints on the model's parameters, we compare the numerical results with observational data. According to our numerical analysis, the power-law tachyon model with constant sound speed is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $95\%$ CL if $0< c_{s}\leq 1$ and $222\leq n \leq 565$. Also, this model is consistent with Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL if $0< c_{s}< 0.990$ and $293\leq n \leq 485$. The behavior of the tensor-to-scalar ratio has been shown in figure \ref{fig7}. The running of the scalar spectral index of the power-law tachyon model with constant sound speed versus the scalar spectral index in this model is shown in figure \ref{fig8}. By studying these parameters, we find that in this case, the model is consistent with Planck2018 TT, TE, EE+lowE+lensing data at $68\%$ CL if $0< c_{s}\leq 1$ and $299\leq n \leq 473$ and at $95\%$ CL if $0< c_{s}\leq 1$ and $231\leq n \leq 551$. Figure \ref{eq9} shows the evolution of the tensor-to-scalar ratio of the power-law tachyon model with constant sound speed versus the tensor spectral index in this model, in the background of the Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data. In this case, the model has observational viability if $0< c_{s}\leq 1$ and $27.3\leq n \leq 3200$ at $68\%$ CL, and $0< c_{s}\leq 1$ and $1.67\leq n $ at $95\%$ CL. \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{7} \end{center} \caption{\small {Tensor-to-scalar ratio versus the scalar spectral index in the power-law tachyon model with constant sound speed.}} \label{fig7} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{8} \end{center} \caption{\small {Running of the scalar spectral index versus the scalar spectral index in the power-law tachyon model with constant sound speed.}} \label{fig8} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{9} \end{center} \caption{\small {Tensor-to-scalar ratio versus the tensor spectral index in the power-law tachyon model with constant sound speed.}} \label{fig9} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.35]{10a} \includegraphics[scale=0.35]{10b} \end{center} \caption{\small {Left panel: ranges of the parameters $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{s}$, obtained from Planck2018 TT, TE, EE+lowE+lensing+BAO+BK14 data at $68\%$ CL (dark region) and $95\%$ CL (light region). Right panel: ranges of the parameter $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of $r$ versus $n_{T}$, obtained from Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO, and Virgo2016 data at $68\%$ CL (dark region) and $95\%$ CL (light region).}} \label{fig10} \end{figure} \begin{figure}[] \begin{center} \includegraphics[scale=0.5]{11} \end{center} \caption{\small {Ranges of parameter $c_{s}$ and $n$ in the power-law tachyon inflation with constant sound speed, leading to observationally viable values of the amplitude of the scalar perturbation.}} \label{fig11} \end{figure} \begin{table*} \tiny \caption{\small{\label{tab2} Ranges of the parameter $n$ in which the tensor-to-scalar ratio, the scalar spectral index, and the tensor spectral index of the power-law tachyon model with constant sound speed are consistent with different data sets.}} \begin{center} \begin{tabular}{cccccc} \\ \hline \hline \\ & Planck2018 TT,TE,EE+lowE & Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE&Planck2018 TT,TE,EE+lowE \\ & +lensing+BK14+BAO & +lensing+BK14+BAO&lensing+BK14+BAO&lensing+BK14+BAO \\ & & &+LIGO$\&$Virgo2016 &LIGO$\&$Virgo2016 \\ \hline \\$c_{s}$& $68\%$ CL & $95\%$ CL &$68\%$ CL & $95\%$ CL \\ \hline\hline \\ $0.1$& $368<n<492$ &$336<n<569$ &$27.3<n<300 $ & $20.1<n$\\ \\ \hline \\$0.4$&$354<n<488$ &$324<n<559$ &$102<n<1600$ &$76.0<n$ \\ \\ \hline\\ $0.7$&$330<n<447 $&$301<n<511$&$177<n<2930$ &$131<n$\\ \\ \hline\\ $0.9$&$305<n<373 $&$270<n<483$& $228<n<3130 $ &$167<n$\\ \\ \hline \hline \end{tabular} \end{center} \end{table*} The domain of $n$ and $c_{s}$, leading to observationally viable values of scalar spectral index, tensor spectral index, and the tensor-to-scalar ratio is shown in figure \ref{fig10}, where we have used both Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 and Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data sets at $68\%$ CL and $95\%$ CL. It is clear that, although from the observationally viable values of the tensor spectral index we can't find an upper bound on the parameter $n$, the observational values of the scalar spectral index imply an upper limit on it. These observationally viable ranges of $c_{s}$ and $n$ lead to the value of the amplitude of the scalar perturbations released by planck2018 data ($\ln (10^{10}{\cal_{A}}_{s})=3.044\pm 0.014$). Figure \ref{fig11} shows this issue clearly. Considering that the sound speed is constant, the behavior of the orthogonal amplitude of the non-gaussianity versus the equilateral amplitude of the non-gaussianity in the power-law tachyon case is the same as the one in the intermediate tachyon case. This means that in this case also we have consistency with Planck2018 TTT, EEE, TTE and EET data at $68\%$ CL if $0.276\leq c_{s}$, at $95\%$ CL if $0.213\leq c_{s}$, and at $97\%$ CL if $0.186\leq c_{s}$. These ranges are compatible with the ranges obtained from $n_{s}$, $\alpha_{s}$, $n_{T}$ and $r$. Therefore, the power-law tachyon inflation with constant sound speed seems to be a fine model that in some ranges of its parameter space is consistent with observational data. In table \ref{tab2}, we have summarized some constraints obtained in studying the power-law tachyon inflation with constant sound speed. \section{\label{sec5}A Discussion on Unifying the Inflation with Dark Energy } Currently, cosmologists have been interested in the models covering a larger domain in the thermal history of the universe. In this way, the models that can describe both very early time and very late time accelerating expansion of the universe have attracted a lot of attention. In this respect, in some viable models of $f(R)$ gravity it is possible to describe both early time inflation and late time acceleration of the universe~\cite{Noj07,Noj08,Cog08,Noj11}. For instance, in Ref.~\cite{Noj08} the authors have considered a model of $f(R)$ gravity where the universe effectively starts with a large cosmological constant at the early universe (leading to inflation) and after passing the radiation/matter dominated era, reaches the small values of the cosmological constant (leading to late time accelerating expansion). In Ref.~\cite{Noj11} another model of $f(R)$ gravity that can explain the initial inflation, and at the late time can reproduce the behavior of the $\Lambda$CDM model. As we have mentioned in the ``Introduction" section, the equation of state of the tachyon field can be $-1$, corresponding to late time dark energy and early time inflationary phases of the universe. Therefore, in our model also, it seems possible to unify the inflation with dark energy. As we know, the tachyon inflation with constant sound speed is an observationally viable inflation model, at least in some domains of its parameter space. After inflation ends, and the tachyon field reaches the non-zero minimum of the potential, the universe becomes radiation dominated. During the radiation and matter dominated universe, the minimum value of the potential doesn't disturb the thermal history of the universe. With more expansion of the universe, the energy densities of the radiation and matter become small and eventually at the late time the potential become the dominant component in the energy density of the universe. In fact, this dominant component of the energy density can be considered as an effective cosmological constant, leading to late time accelerating expansion of the universe. It is also possible to consider some extension of the tachyon field so we can get both inflation and late time acceleration in the model. For instance, by considering the non-minimal coupling or non-minimal derivative coupling between the tachyon field and the gravity, depending on the values of the non-minimal parameter, it may possible to have a model covering both initial inflation and late time acceleration. \section{\label{sec6}Conclusion} In this paper, we have considered tachyon inflation where the sound speed is constant. We have obtained the Friedmann equations and equation of motion in terms of sound speed. After that, we have presented the main perturbation parameters, like the scalar spectral index, its running, tensor spectral index, and tensor-to-scalar ratio. We have also, by considering the three-point correlations, presented the amplitude of the non-gaussianity in both equilateral and orthogonal configurations. Then, we have adopted two types of scale factor: Intermediate and power-law. We have studied both cases separately. To study these cases, we have obtained the potential and inflation parameters in terms of the Hubble parameter, its derivative, and also the sound speed. In this way, we were able to study the model based on two types of scale factor. By considering the intermediate scale factor, we have studied $r-n_{s}$ behavior in the background of Planck2018 TT, TE, EE+lowE+lensing +BAO+BK14 data in both $68\%$ CL and $95\%$ CL and found some constraints on the model's parameters. The behavior of $\alpha_{s}-n_{s}$ in the background of Planck2018 TT, TE, EE+lowE+lensing data has been studied too. Another important parameter is the tensor scalar spectral index corresponding to the primordial perturbations. In this way, we have studied $r-n_{T}$ behavior in the background of Planck2018 TT, TE, EE +lowE+lensing+BK14 +BAO+LIGO and Virgo2016 data to find some constraints on the model's parameter space. We have also analyzed $c_{s}$ and $\beta$ phase space in both $68\%$ CL and $95\%$ CL. By studying these parameters numerically, we have found that intermediate tachyon inflation with constant sound speed is observationally viable if $0<c_{s}\leq 0.997$ and $0.833\leq\beta\leq 0.984$ at $68\%$ CL, and $0<c_{s}\leq 1$ and $0.787\leq\beta\leq 1$ at $95\%$ CL. To check the viability of the model more precisely, we have studied the non-gaussian feature of the primordial perturbations. In this regard, we have considered the Planck2018 TTT, EEE, TTE and EET constraints on the equilateral and orthogonal configurations of the non-gaussianity. In this way, we have obtained the constraints on the sound speed as $0.276\leq c_{s}\leq 1$ at $68\%$ CL, $0.213\leq c_{s}\leq 1$ at $95\%$ CL, and $0.186\leq c_{s}\leq 1$ at $97\%$ CL. We have also studied $c_{s}^{2}-r$ phase space and found that the ranges of the sound speed, obtained from the observational constraints on the non-gaussianity, lead to the observationally viable values of the tensor-to-scalar ratio. The power-law tachyon inflation with constant sound speed is another model that has been studied in this paper. Similar to the intermediate tachyon inflation case, we have studied the perturbation parameters in this model, but here, based on the power-law scale factor. In this case, by using different data sets, we have studied $r-n_{s}$, $\alpha-n_{s}$, and $r-n_{T}$ behavior. The phase space of the parameters $c_{s}$ and $n$, based on the observationally viable ranges of the scalar and tensor spectral indices has been studied too. Our numerical analysis has shown that the power-law tachyon inflation with constant sound speed is observationally viable for $0<c_{s}\leq 0.990$ and $299\leq n<\leq 473$ at $68\%$ CL, and $0<c_{s}\leq 1$ and $231\leq n <\leq 551$ at $95\%$ CL. For these obtained ranges, the amplitude of the scalar spectral index is consistent with observational data. We have also discussed the issue of unifying the inflation with dark energy in the tachyon model with constant sound speed. We have argued that the non-zero minimum value of the tachyon potential becomes important at the late time, after the energy density of the radiation/matter decreases. Based on our analysis, it seems that both intermediate and power-law tachyon models in some ranges of their parameter space are consistent with observational data.\\ {\bf Acknowledgement}\\ I thank the referee for the very insightful comments that have improved the quality of the paper considerably.\\
\section{Missing Proofs} \label{sec:missingproof} We start by giving a formal proof of the following lemma. \lemnono* To prove the above lemma, we state the following observation. \begin{observation2}\label{obs:folklore3} Suppose a fractional matching $\vec{x}$ of $G$ satisfies the odd set constraints for all odd sets of size smaller than $\nicefrac{3}{\varepsilon}+1$, then, the fractional matching $\vec{z}=\frac{\vec{x}}{1+\varepsilon}$ satisfies all odd set constraints. \end{observation2} \begin{proof} Suppose $\vec{x}$ is a fractional matching of $G$ that satisfies odd set constraints for all odd sets of size smaller than $\nicefrac{3}{\varepsilon}+1$. Let $\vec{z}$ be a fractional matching obtained by scaling down $\vec{x}$ by $1+\varepsilon$. Now, consider any odd set $B$ with $\card{B}\leqslant \nicefrac{3}{\varepsilon}$, then $\vec{x}$, and therefore $\vec{z}$ satisfies the odd set constraint corresponding to this. So, we consider $B$ such that $\card{B}\geqslant \nicefrac{3}{\varepsilon}+1$, then, \begin{align*} \frac{\card{B}}{\card{B}-1}&=1+\frac{1}{\card{B}-1}\leqslant 1+\nicefrac{\varepsilon}{3} \end{align*} The last inequality follows because of the fact that $\card{B}-1\geqslant \nicefrac{3}{\varepsilon}$. Since $\vec{x}$ satisfies the fractional matching constraints, this implies: $\sum_{e\in G[B]} x(e)\leqslant \frac{\card{B}}{2}$. Thus, we have, $\sum_{e\in G[B]} z(e)\leqslant \sum_{e\in G[B]}\frac{x(e)}{1+\varepsilon}\leqslant \frac{\card{B}}{2\cdot (1+\varepsilon)}$. By the above argument, this is upper bounded by $\frac{\card{B}-1}{2}$. Thus, $\vec{z}$ satisfies all fractional matching constraints as well as the odd set constraints. \end{proof} We now state the following lemma. \begin{proof}[Proof of \Cref{lem:folklore2}] Consider the maximum fractional matching $\vec{x}$ of $\textsc{bc}(G,\kappa)$, that obeys the capacity constraints $\kappa$. Let $\vec{z}$ be a fractional matching of $(G,\kappa)$ that is obtained from $\vec{x}$ as follows. Let $e\in E(G)$, and let $e'$, $e''$ be copies of $e$ in $\textsc{bc}(G,\kappa)$. Then, we let $z(e)=\frac{x(e')+x(e'')}{2}$. Note that $\vec{z}$ obeys fractional matching constraints, since $\vec{x}$ does. Additionally, observe that $z(e)\leqslant \kappa(e)$. Thus, for any odd set $B\subset V$ with $\card{B}\leqslant \nicefrac{3}{\varepsilon}$, we have, \begin{align*} \sum_{e\in G[B]}z(e)&=\sum_{u,v\in B}z(D(u,v))\\ &\leqslant \sum_{u,v\in B}\kappa(D(u,v))\\ &\leqslant \nicefrac{1}{\alpha_{\varepsilon}}\\ &\text{(Since }\kappa(D(e))\leqslant \nicefrac{1}{\alpha_{\varepsilon}})\\ &\leqslant \sum_{u,v\in B}\nicefrac{\varepsilon}{3}\\ &\text{(Since }\nicefrac{1}{\alpha_{\varepsilon}}\leqslant \nicefrac{\varepsilon}{3} \text{ for $\varepsilon<\nicefrac{1}{2}$)}\\ &\leqslant \frac{\varepsilon}{3}\cdot\frac{\card{B}\cdot \paren{\card{B}-1}}{2}\\ &\leqslant \frac{\card{B}-1}{2}\\ &\text{(Since }\card{B}\leqslant \nicefrac{3}{\varepsilon}) \end{align*} Thus, $\vec{z}$ satisfies the small odd set constraints, and we know from Observation \ref{obs:folklore3} that $\vec{y}=\frac{\vec{z}}{1+\varepsilon}$ satisfies all odd set constraints in addition to the fractional matching constraints. Thus, we have, $(1+\varepsilon)\cdot \sum_{e\in E(G)}y(e)=\sum_{e\in E(G)} z(e)=0.5\sum_{e\in \textsc{bc}(G)}x(e)$. Thus, if $\vec{x}$ is the optimal fractional matching of $\textsc{bc}(G)$, then we have, \begin{align*} \mu(G,\kappa)\cdot (1+\varepsilon)&\geqslant (1+\varepsilon)\cdot \sum_{e\in E(G)}y(e)\\ &=0.5\sum_{e\in \textsc{bc}(G)}x(e)\\ &\text{(From the discussion above)}\\ &=0.5\cdot \mu(\textsc{bc}(G), \kappa)\\ &\text{(Since }\vec{x}\text{ is the maximum fractional matching of }(G,\kappa)) \end{align*} This proves our claim. \end{proof} \section{Reduction from Simple to Multigraphs} In this section, we will prove the following lemma. \lemmreduc* In order to do that, we first start by proving the following simple observation. \begin{observation2}\label{obs:nummatch} Let $G$ be any simple graph with maximum matching size $\mu(G)$, then $G$ contains at most $2^{O(\mu(G)\cdot \log n)}$ matchings of any size. \end{observation2} \begin{proof} Since the total number of edges is at most $n^2$, there are at most $n^2\choose j$ ways of choosing a matching of size $j$. Thus, we have, the total number of matchings possible is at most: \begin{align*} \sum_{j=1}^{\mu(G)} {n^2\choose j}\leqslant \sum_{j=1}^{\mu(G)}{2^{2\cdot j\log n}}\leqslant \mu(G)\cdot 2^{2\cdot \mu(G)\log n}\leqslant 2^{3\cdot \mu(G)\log n} \end{align*} This proves our claim. \end{proof} Now, in order to prove \Cref{lem:reduction}, we first give a procedure that takes as input a simple graph, and outputs a multigraph with $O(\nicefrac{\mu(G)}{\varepsilon})$ vertices that preserves a fixed matching $M$ of $G$ approximately with probability $1-\exp\paren{-O\paren{\card{M}\cdot\varepsilon^3}}$. \begin{algorithm}[H] \algorithmicrequire{ Graph $G$, a sparsification parameter $\tau$, and a parameter $\varepsilon>0$}\\ \algorithmicensure{ A multigraph $\mathcal{G}(\mathcal{V},\mathcal{E})$ with $\card{\mathcal{V}}=\tau$} \caption{\textsc{Vertex-Red-Basic}($G,\tau,\varepsilon$)} \begin{algorithmic}[1] \State Partition $V$ into $\tau$ bins, $\mathcal{V}\coloneqq (B_1,B_2\cdots, B_{\tau})$, by assigning every vertex to one of the $\tau$ bins uniformly at random. \State For a vertex $u$, let $B(u)$ denote the bin chosen for $u$. For any edge $(u,v)\in E$, if $B(u)\neq B(v)$, then add an edge $e_{u,v}$ between $B(u)$ and $B(v)$. \State Return the multigraph $\mathcal{G}(\mathcal{V},\mathcal{E})$. \end{algorithmic} \label{alg:vertexred} \end{algorithm} \begin{lemma2}\label{lem:subred2} Let $G$ be a simple graph, let $\varepsilon\in (0,1)$, and let $\tau\geqslant \frac{4\cdot \mu(G)}{\varepsilon}$. Let $\mathcal{G}=\textsc{Vertex-Red-Basic}(G,\tau,\varepsilon)$, and let $M$ be any fixed matching of $G$. Then with probability $1-2^{-\frac{\card{M}\cdot \varepsilon^3}{32}}$, there exists a matching $\mathcal{M}$ in $\mathcal{G}$ such that if $\mathcal{M}\subset M$, and $\card{\mathcal{M}}\geqslant (1-\varepsilon)\cdot \card{M}$. \end{lemma2} \begin{proof} Let $\delta=\nicefrac{\varepsilon}{4}$, and we define $t\coloneqq 2\card{M}$. Since $\varepsilon<1$, this implies that $\delta<\nicefrac{1}{2}$. Recall that we have $\tau=\frac{4\mu(G)}{\varepsilon}$ bins, and we combine these bins arbitrarily so that we end up with $\frac{t}{\delta}$ groups. Call these groups $Z_1,\cdots, Z_{\nicefrac{t}{\delta}}$. Note that every vertex $v$ lands in bin $Z_i$ with probability $\nicefrac{\delta}{t}$. Call a group $Z_i$ bad if it doesn't contain even one of the $2\card{M}$ vertices of $M$. Associate a random variable $\textsf{X}_i$ with $Z_{i}$ that takes value $1$ if $Z_i$ is bad, and zero otherwise. Let $\textsf{X}=\sum_{i=1}^{\nicefrac{t}{\delta}}\textsf{X}_i$ denote the number of bad groups. We have, \begin{align*} \prob{\textsf{X}_i=1}=\paren{1-\frac{\delta}{t}}^t=e^{-\delta}\leqslant 1-\delta+\nicefrac{\delta^2}{2} \end{align*} Thus, we have, $\expect{\textsf{X}}\leqslant \nicefrac{t}{\delta}\paren{1-\delta+\nicefrac{\delta^2}{2}}$. Note that $\textsf{X}$ is a sum of negatively correlated variables. This is because if $Z_i$ is empty, then $Z_j$ has an increased likelihood of being not empty for $i\neq j$. So, using \Cref{lem:chernoff}, we have, \begin{align*} \prob{\textsf{X}\geqslant \paren{1+\delta^2}\cdot \frac{t}{\delta}\cdot \paren{1-\delta+\nicefrac{\delta^2}{2}}}&\leqslant \exp\paren{-\delta^4\cdot \frac{t}{\delta}\cdot \paren{1-\delta+\frac{\delta^2}{2}}}\\ &\leqslant \exp\paren{-\delta^3\cdot t+\delta^4\cdot t}\\ &\leqslant \exp\paren{\frac{-\delta^3}{2}\cdot t}\\ &\text{(Since $\delta<\frac{1}{2}$)} \end{align*} Thus, with probability at least $1-2^{-\frac{\card{M}\cdot \varepsilon^3}{128}}$, $\textsf{X}\leqslant \paren{\frac{t}{\delta}-t+2\delta t}$. Thus, with probability at least $1-2^{-\frac{\card{M}\cdot \varepsilon^3}{128}}$, we have that $t-2\delta \cdot t$ groups are good, which means that they contain at least one vertex of $M$. Now, we need to show that if at least $t-2\delta \cdot t$ bins are good, then the edges of $M$ form a matching $\mathcal{M}$ of $\mathcal{G}$ such that $\card{\mathcal{M}}\geqslant \card{M}\cdot(1-\varepsilon)$. For every bin, we fix one vertex of $M$, and remove the rest. Since $t-2\delta \cdot t$ of the bins are good, this implies that we lost at most $2\delta \cdot t$ vertices of the matching $M$. Consequently, we deleted at most $2\delta \cdot t$ edges of $M$. The remaining edges have their endpoints in different bins, and therefore, they form a matching in $\mathcal{G}$. The size of this matching is $\card{M}-2\delta \cdot t$, which is equal to $\card{M}-\varepsilon \card{M}$. Thus, we have our claim. \end{proof} \begin{lemma2}\label{lem:subred1} Let \textsc{Vertex-Red}() be an algorithm that does independent runs of \textsc{Vertex-Red-Basic}(). Let $H_1,H_2\cdots, H_{\lambda}$ be multigraphs on independent runs of \textsc{Vertex-Red-Basic}() on input $G,\tau\geqslant \frac{4\cdot \mu(G)}{\varepsilon}$ and $\varepsilon\in (0,\nicefrac{1}{2})$, for $\lambda \geqslant \frac{1600\cdot \log n}{\varepsilon^4(1-\varepsilon)}$. Then, with probability at least $1-\exp(-O(\nicefrac{\mu(G)\cdot \log n}{\varepsilon}))$, for every matching $M$ with $\card{M}\geqslant (1-\varepsilon)\cdot \mu(G)$, there is a matching $M'$ in some $H_i$ such that $M'\subset M$, and $\card{M'}\geqslant (1-\varepsilon)\card{M}$. \end{lemma2} \begin{proof} Consider a fixed matching $M$, of size at least $(1-\varepsilon)\cdot \mu(G)$, then with probability at least $1-\exp\paren{-\frac{\mu(G)\cdot (1-\varepsilon)\cdot \varepsilon^3}{128}}$, a fixed $H_i$ contains a matching $M'\subset M$ with $\card{M'}\geqslant (1-\varepsilon)\card{M}$. This is implied by \Cref{lem:subred2}. Since each of the $H_j$'s are independent runs of $\textsc{Vertex-Red}()$, with probability at least $1-\exp\paren{-\frac{12\cdot\mu(G)\cdot \log n}{\varepsilon}}$, some $H_i$ contains a matching $M'\subset M$ with $\card{M'}\geqslant (1-\varepsilon)\card{M}$. Taking a union bound over all possible matchings of size at least $(1-\varepsilon)\cdot \mu(G)$, from Observation \ref{obs:nummatch}, we have that with probability at least $1-\exp\paren{-\frac{9\cdot \mu(G)\cdot \log n}{\varepsilon}}$, for every matching $M$ with $\card{M}\geqslant (1-\varepsilon)\cdot \mu(G)$, there is a matching $M'$ in some $H_i$ such that $M'\subset M$ and $\card{M'}\geqslant (1-\varepsilon)\cdot \card{M}$. \end{proof} We now show the following. \begin{lemma2}\label{lem:mainvertexred} Let $G$ be a simple graph, and let $H_1,\cdots,H_{\lambda}$ be the output of independent runs of \textsc{Static-Match}() with $G$ and $\varepsilon$ as input, where $\lambda\geqslant \frac{1600\cdot \log n}{\varepsilon^4\cdot (1-\varepsilon)}$. Consider an adversary that deletes edges from $G$. We simulate this deletion process in $H_i$'s, that is, if $(u,v)$ is deleted from $G$, then $e_{u,v}$ (if present) is deleted from each of the $H_i$'s. Let $\mu_{i}^t$ denote the size of the maximum matching of $H_i$ after $t$ deletions, and similarly, let $\mu^{t}$ denote the size of the maximum matching of $G$ after $t$ deletions. Suppose $\mu^t>(1-\varepsilon)\mu^0$, then, $\mu^t_{i}>(1-2\varepsilon)\mu^0$ for some $i\in [\lambda]$. \end{lemma2} \begin{proof} Let $M$ be a matching in $G$ satisfying the statement of the lemma. That is, suppose at time $t$, $\card{M}>(1-\varepsilon)\mu^0$. Since the adversary is only performing deletions, $\card{M}>(1-\varepsilon)\mu^0$ initially, when no deletions had been performed. So, by \Cref{lem:subred1}, before any deletions are performed, there is a matching $M'$ in one of the $H_i$'s such that $M'\subset M$, and $\card{M'}\geqslant (1-\varepsilon)\cdot \card{M}>(1-2\varepsilon+\varepsilon^2)\mu^0$. Since all of the edges of $M$ survive at time $t$, this is true of $M'$ as well. Thus, $\mu^t_{i}>(1-2\varepsilon)\mu^0$ for some $i\in [\lambda]$. This proves our claim. \end{proof} We now show that \Cref{lem:reduction} follows from the lemmas proved in this section. \begin{proof}[Proof of \Cref{lem:reduction}] Consider \Cref{lem:reduction}\ref{item:first}. This is implied by the properties of \textsc{Vertex-Red}(), \Cref{lem:subred1}, and \Cref{lem:subred2}. Property \Cref{lem:reduction}\ref{item:third} is on the other hand implied by the contrapositive of \Cref{lem:mainvertexred}. \end{proof} \section{Properties of \textsc{Static-Match}()} \label{app:statmatch} In this section we will show \Cref{lem:propstatmatch}, which we first restate. \statmatch* We first note that the properties \ref{item:a}-\ref{item:b} and \ref{item:d}-\ref{item:folklore} are evident in Property 3.1 and the proof of Lemma 2.3 of \cite{DP14}. We now show that \ref{item:c} also holds. \begin{proof}[Proof of \Cref{lem:propstatmatch}] Let $\textsc{Basic-Static-Match}()$ be the algorithm satisfying of \cite{DP14} satisfying \ref{item:a}-\ref{item:b} and \ref{item:d}-\ref{item:folklore}. We now show how to obtain \textsc{Static-Match}() from \textsc{Basic-Static-Match}(). We run \textsc{Basic-Static-Match}() with input $G$ and $\delta=\nicefrac{\varepsilon}{3}$. Thus, we get a matching $M$ of size at least $(1-\nicefrac{\varepsilon}{3})\cdot \mu(G)$. Moreover, the duals $\vec{y}$ and $\vec{z}$ output by the algorithm have $yz(V)\leqslant (1+\nicefrac{\varepsilon}{3})\cdot \mu(G)$. We consider the following procedure: for every $B\in \Omega$ with $z(B)>0$ and $\card{B}\geqslant \frac{3}{\varepsilon}+1$, we increase $y(v)$ by $\nicefrac{z(B)}{2}$ for every $v\in B$, and we decrease $z(B)$ to $0$. Thus, each $y(v)$ is still a multiple of $\varepsilon$ and each $z(B)$ is still a multiple of $\varepsilon$. We conclude that \ref{item:d} still holds. Moreover, this transformation keeps the value of the dual constraint for every edge the same. Thus, we still satisfy \ref{item:e}. We update $\Omega$ by removing $B$ from it. The set $\Omega$ still remains laminar. Thus, \ref{item:b} is still satisfied. We first observe that the time taken to do this is linear in the sum of the sizes of the odd sets $B$ with $z(B)>0$. Note that since $yz(V)$ sums to at most $(1+\nicefrac{\varepsilon}{3})\mu(G)$, and non-zero $z(B)$ have value at least $\nicefrac{\varepsilon}{3}$, this implies that the sums of the sizes of the odd sets is at most $\nicefrac{3}{\varepsilon}\cdot (1+\nicefrac{\varepsilon}{3})\cdot \mu(G)$. Thus, the time taken for the procedure is $O(\nicefrac{m}{\varepsilon})$. Next, we observe that the value of $yz(V)$ changes by at most $\sum_{B:\card{B}\geqslant \nicefrac{3}{\varepsilon}+1}\frac{z(B)}{2}$. This value is at most $\nicefrac{\varepsilon}{3}\cdot (1+\nicefrac{\varepsilon}{3})\cdot \mu(G)$, as shown by the following calculation. \begin{equation*} \begin{split} \paren{\frac{3}{\varepsilon}}\cdot \sum_{B:\card{B}\geqslant \nicefrac{3}{\varepsilon}+1}\frac{z(B)}{2}\leqslant \sum_{B:\card{B}\geqslant \nicefrac{3}{\varepsilon}+1}\paren{\frac{\card{B}-1}{2}}\cdot z(B)&\leqslant (1+\nicefrac{\varepsilon}{3})\cdot \mu(G) \end{split} \end{equation*} Thus, the new $yz(V)$ has value at most $(1+\nicefrac{\varepsilon}{3})^2\cdot \mu(G)$. This implies that \ref{item:folklore} holds. Finally, since every blossom $B$ of size at least $\nicefrac{3}{\varepsilon}+1$ has $z(B)=0$, thus, \ref{item:c} holds. \end{proof} \section{Rounding Fractional Matchings} \label{sec:rounding} In this section, we will prove \Cref{thm:sparse}. More concretely, we give state the procedure \textsc{Sparsification}() that takes as input a fractional matching $\vec{x}$ of a simple graph $G$, and outputs a graph $H$, of size $\Tilde{O}(\mu(G))$. If $x(e)\leqslant \varepsilon^6$ for all $e\in E(G)$, then, $H$ contains an integral matching of size at least $(1-10\varepsilon)\cdot \sum_{e\in E}x(e)$ in its support. The proof of this theorem is implicit in the work of \cite{Wajc2020}, but we describe it here for completeness. We first state the algorithm, and then describe some of its properties. The algorithm uses a dynamic edge coloring algorithm as a subroutine. The update time of the subroutine is $O(\log n)$ in the worst case. The sparsification algorithm takes as input a fractional matching $\vec{x}$ and a parameter $\varepsilon>0$. Then, it classifies $E(G)$ into classes as follows: $E_i=\set{e\mid x(e)\in [(1+\varepsilon)^{-i}, (1+\varepsilon)^{-i+1})}$. Note that we only consider edges $e\in E(G)$ with $x(e)\geqslant \paren{\nicefrac{\varepsilon}{n}}^2$, since the total contribution of these edges to fractional matching is at most $\varepsilon^2$, and we can afford to ignore them if we want to compute a $(1+\varepsilon)$ approximation. We now state the algorithm. \begin{algorithm} \caption{\textsc{Sparsification}($\vec{x},\varepsilon$)} \begin{algorithmic}[1] \State $d\leftarrow \frac{4\cdot \log\paren{\nicefrac{2}{\varepsilon}}}{\varepsilon^2}$ \For{$i\in \set{1,2,\cdots, 2\log_{1+\varepsilon}(\nicefrac{n}{\varepsilon})}$} \State Compute a $3\lceil (1+\varepsilon)^i\rceil$-edge colouring $\Phi_i$ of $E_i$. \State Let $S_i$ be a sample of $3\cdot \min\set{\lceil d\rceil,\lceil (1+\varepsilon)^i\rceil}$ colours without replacement in $\Phi_i$. \label{line:blah} \State Return $K=(V,\cup_{i}\cup_{M\in S_i}M)$ \EndFor \end{algorithmic} \end{algorithm} We state the guarantees of the edge coloring subroutine. \begin{observation2} The size of $K$ output by \textsc{Sparsification}($\vec{x},\varepsilon$) is, \begin{align*} \card{E(H)}=O\paren{\frac{\log (\nicefrac{n}{\varepsilon})}{\varepsilon}\cdot d\cdot \mu(\textup{supp}(H))} \end{align*} \end{observation2} \begin{lemma2}\label{lem:edgecolour}\cite{BDHN18} There is a deterministic dynamic algorithm that maintains a $2\Delta-1$ edge coloring of a graph $G$ in $O(\log n)$ worst case update time, where $\Delta$ is the maximum degree of the graph $G$. \end{lemma2} \begin{observation2}\label{obs:sampling} Let $\varepsilon\in (0,\nicefrac{1}{2})$ and suppose the input to \textsc{Sparsification}($\vec{x},\varepsilon$) is a matching $\vec{x}$ with $x(e)\leqslant \varepsilon^6$, then, $\card{S_i}=3\cdot \lceil d\rceil$. \end{observation2} To show that $K$ contains a matching of size at least $(1-\varepsilon)\cdot \sum_{e\in E(G)}x(e)$ in its support, we will show a random fractional matching $\vec{y}$ in $K$ that sends flow at most $\varepsilon$ through each of its edges. Moreover, $\expect{\sum_{e\in E}y(e)}\geqslant (1-6\varepsilon)\cdot \sum_{e\in E(G)} x(e)$. This will show the existence of a large matching that sends flow at most $\varepsilon$ through each of its edges. To show this, we give some properties of the algorithm. \begin{lemma2}\label{lem:probbounds} Let $\varepsilon\in (0,\nicefrac{1}{2})$ and let $\vec{x}$ be the input to \textsc{Sparsification}($\vec{x},\varepsilon$) such that $x(e)\leqslant \varepsilon^6$ for all $e\in E(G)$. Then, for every edge $e$, $\prob{e\in K}\in \bracket{\nicefrac{x(e)\cdot d}{(1+\varepsilon)^2},x(e)\cdot d\cdot (1+\varepsilon)}$. \end{lemma2} \begin{proof} Since $\vec{x}$ has the property that for every $e\in E$, $x(e)\leqslant \varepsilon^6$, this implies that we sample $3\cdot \lceil d\rceil$ from $\Phi_i$ colors for each $i$ (from Observation \ref{obs:sampling}). Thus, if we consider an edge $e\in E_i$, then, we have, \begin{align*} \prob{e\in K}=\frac{\lceil d\rceil}{\lceil(1+\varepsilon)^i\rceil}\leqslant \frac{d\cdot (1+\varepsilon)}{(1+\varepsilon)^i}\leqslant d\cdot (1+\varepsilon)\cdot x(e) \end{align*} The last inequality follows from the fact that $x(e)\in \bracket{(1+\varepsilon)^{-i},(1+\varepsilon)^{-i+1}}$. Moreover, we have, \begin{align*} \frac{\lceil d\rceil}{\lceil (1+\varepsilon)^i\rceil}\geqslant \frac{d}{(1+\varepsilon)^i+1} \geqslant \frac{d}{(1+\varepsilon)^{i+1}} \geqslant \frac{d\cdot x(e)}{(1+\varepsilon)^2} \end{align*} The first inequality follows from the fact that $\lceil d\rceil \geqslant d$, and $\lceil (1+\varepsilon)^i\rceil\leqslant (1+\varepsilon)^i+1$. The second inequality follows from the following reasoning. \begin{align*} (1+\varepsilon)^{-i}\leqslant x(e)\leqslant \varepsilon^6\leqslant \nicefrac{1}{d}\leqslant \varepsilon \end{align*} Thus, $\nicefrac{1}{\varepsilon}\leqslant (1+\varepsilon)^i$, and $1\leqslant \varepsilon\cdot (1+\varepsilon)^i$, so the second inequality follows. The last inequality follows from the fact that $x(e)\in \bracket{(1+\varepsilon)^{-i},(1+\varepsilon)^{-i+1}}$. \end{proof} For an edge $e\in E(G)$, we define $X_e$ to be an indicator random variable that takes value $1$ if $e\in K$, and $0$ otherwise. \begin{observation2}\label{obs:negatcorr} For a vertex $v$, the variables $\set{X_e\mid e\text{ incident on }v}$ are negatively associated. \end{observation2} \begin{proof} If the edges $e$ and $e'$ incident on $v$ belong in $E_i$ and $E_j$ where $i\neq j$, then $X_e$ and $X_{e'}$ are independent. On the other hand, if they belong in the same $E_i$, then recall we did an edge colouring on $E_i$, so $e'$ and $e$ got different colours. So, if the colour corresponding to $e$ is picked into $K$, then this reduces the probability of the colour corresponding to $e'$ being picked into $K$. \end{proof} \begin{lemma2}\label{cor:cond} From Observation \ref{obs:negatcorr} and \Cref{lem:probbounds}, we can conclude that for edges $e$ and $e'$ incident on $v$, \begin{align*} \prob{X_e=1\mid X_{e'}=1}\leqslant \prob{X_e=1}\leqslant x(e)\cdot d\cdot (1+\varepsilon) \end{align*} \end{lemma2} \begin{lemma2} For any vertex $v$ and edge $e'$ incident on $v$, the variables $\set{[X_e\mid X_{e'}]\mid e\text{ incident on }v}$ are negatively associated. \end{lemma2} \begin{lemma2}[Concentration Bound]\label{lem:concentrate} Let $X$ be the sum of negatively associated random variables $X_1,\cdots, X_k$ with $X_i\in [0,M]$ for each $i\in [k]$. Then, for $\sigma^2=\sum_{i=1}^k \var{X_i}$ and all $a>0$, \begin{align*} \prob{X>\expect{X}+a}\leqslant \exp\paren{\frac{-a^2}{2\paren{\sigma^2+\nicefrac{a\cdot M}{3}}}} \end{align*} \end{lemma2} Following theorem directly implies \Cref{thm:sparse}. \begin{theorem2}\label{thm:largeintegral} Let $K$ be the subgraph of $G$ output by \textsc{Sparsification}(), when run on the matching $\vec{x}$ with parameters $\varepsilon\in (0,\nicefrac{1}{2})$. If $x(e)\leqslant \varepsilon^6$ for all $e\in E$, then $K$ supports a fractional matching $\vec{y}$ such that $y(e)\leqslant \varepsilon$, and $\sum_{e\in E}y(e)\geqslant (1-6\varepsilon)\cdot \sum_{e\in E}x(e)$. \end{theorem2} \begin{proof} To show the theorem, we will describe a process that simulates \textsc{Sparsification}($\vec{x},\varepsilon$), and outputs a random matching $\vec{y}$ such that $y(e)\leqslant \varepsilon$ for every $e\in E$. Additionally, $\expect{\sum_{e\in E}y(e)}\geq(1-6\varepsilon)\sum_{e\in E}x(e)$. As an intermediate step, let $z(e)=\frac{(1-4\varepsilon)}{d}\cdot X_e$. So, we have, \begin{align*} \expect{z(e)}&\geqslant \expect{z(e)\mid X_e=1}\cdot \prob{X_e=1}\geqslant \frac{1-4\varepsilon}{d}\cdot \frac{x(e)\cdot d}{(1+\varepsilon)^2}\geqslant (1-6\varepsilon)\cdot x(e) \end{align*} The second to last inequality follows from \Cref{lem:probbounds}. Next define $\vec{y}$ as follows: \begin{align*} y(e)=\begin{cases} 0 \text{ if for one of the endpoints }v\text{ of }e,\ \sum_{e'\in v}z(e')>1\\ z(e) \text{ otherwise.} \end{cases} \end{align*} Essentially, our procedure is creating matching $\vec{z}$ as follows: it assigns value $\frac{1-4\varepsilon}{d}$ to $z(e)$ if $e$ was included in $K$, and $0$ otherwise. The value $y(e)$ is the same $z(e)$ in all cases except when $e$ is picked into $K$, but at one of the endpoints $v$, $\sum_{e'\ni v}z(e')>1$ , that is, the fractional matching constraint is violated at $v$. We show that it is unlikely that an edge (when picked) has one of its endpoints violated. Let $e'$ be an edge incident on $v$, we to bound the probability that $z(e')\neq y(e')$. \begin{align*} \expect{\sum_{e\in v}z(e)\mid X_{e'}=1}&\leqslant \frac{1-4\varepsilon}{d}+\expect{\sum_{e'\neq e, e\in v}z(e)\mid X_{e'}}\\ &\leqslant \varepsilon+\sum_{e\in v}x(e)\cdot (1+\varepsilon)\cdot d\cdot\paren{\frac{1-4\varepsilon}{d}}\\ &\text{(Since $\nicefrac{1}{d}\leqslant \varepsilon$ and from \Cref{cor:cond})}\\ &\leqslant (1-\varepsilon) \end{align*} Note that at any end point of $e'$, conditioned on $e'$ being sampled, the expected sum of $z(e)$'s at that endpoint is upper bounded by $(1-\varepsilon)$. Now, $z(e)$ is assigned value $0$ only if the sum of the $z(e)$'s deviates from the expected value by $\varepsilon$. To see this, we want to compute $\var{[z(e)\mid X_{e'}]}$, where $e$ and $e'$ share an end point. Note that $[z(e)\mid X_{e'}]$ takes value $\frac{1-4\varepsilon}{d}$ with probability $\prob{X_e\mid X_{e'}}$, otherwise it takes value $0$. \begin{align*} \var{\bracket{z(e)\mid X_{e'}}}&\leqslant \expect{\bracket{z(e)\mid X_{e'}}^2}\\ &\leqslant\paren{\frac{1-4\varepsilon}{d}}^2\cdot \prob{X_e\mid X_{e'}}\\ &\leqslant \paren{\frac{1-4\varepsilon}{d}}^2\cdot x(e)\cdot d\cdot (1+\varepsilon)\\ &\text{(From \Cref{cor:cond})}\\ &\leqslant \frac{x(e)}{d} \end{align*} This implies that $\sum_{e\in v}\var{\bracket{z(e)\mid X_{e'}}}\leqslant \frac{1}{d}$. So, we want to compute the probability that the sum of the random variables $\set{\bracket{z(e)\mid X_{e'}}}$ deviates from the expected value by $\varepsilon$. Applying \Cref{lem:concentrate}, we have, \begin{align*} \prob{\sum_{e\in v}\bracket{z(e)\mid X_{e'}}\geqslant \expect{\sum_{e\in v}\bracket{z(e)\mid X_{e'}}}+\varepsilon}&\leqslant \exp\paren{\frac{-\varepsilon^2}{2\paren{\frac{1}{d}+\frac{\varepsilon}{3\cdot d}}}}\\ &\paren{\text{Since }z(e)\in \bracket{0,\nicefrac{1-4\varepsilon}{d}}}\\ &\leqslant \exp\paren{-\varepsilon^2\cdot 0.25\cdot d}\\ &\text{(Since }\varepsilon\in (0,1))\\ &\leqslant \frac{\varepsilon}{2}\\ &\text{(Since }d=\frac{4\cdot \log(\nicefrac{2}{\varepsilon})}{\varepsilon^2}) \end{align*} Taking union bound over both endpoints, we know that $\prob{y(e)=z(e)\mid X_e=1}\geqslant (1-\varepsilon)$. Thus, we have: \begin{align*} \expect{y(e)}&=\paren{\frac{1-4\varepsilon}{d}}\cdot \prob{y(e)=z(e)}\\ &=\paren{\frac{1-4\varepsilon}{d}}\cdot \prob{y(e)=z(e)\mid X_e=1}\prob{X_e=1}\\ &\geqslant \paren{\frac{1-4\varepsilon}{d}}\cdot (1-\varepsilon)\cdot \frac{x(e)\cdot d}{(1+\varepsilon)^2}\\ &\geqslant (1-7\varepsilon)\cdot x(e) \end{align*} Thus, $\expect{\sum_{e\in E}y(e)}\geqslant (1-7\varepsilon)\cdot \sum_{e\in E}x(e)$. Moreover, for all $e\in E$, $y(e)\leqslant \varepsilon^6$. This proves our claim. \end{proof} We now restate the lemma, and then show its proof. \lemuno* \begin{proof} Note that \ref{lem:sparseb} is implied by \Cref{thm:largeintegral}, and the fact that any fractional matching $\vec{y}$ with $y(e)\leqslant \varepsilon$ for all $e\in E$, satisfies odd set constraints for all odd sets of size at most $\nicefrac{1}{\varepsilon}$. To see \ref{lem:sparsea}, note that deleting $e$ from supp($\vec{x}$) corresponds to just deleting $e$ from $G_i$, and since our edge coloring algorithm is able to handle edge insertions and deletions in $\Tilde{O}_{\varepsilon}(1)$ time, such updates can be handled in $\Tilde{O}_{\varepsilon}(1)$ update time (see \Cref{lem:edgecolour}). Finally, if an update reduces $x(e)$ for some $e\in E$, then this corresponds to deleting $e$ from some $G_i$ and adding it to $G_j$ for some $j<i$. Thus, the edge colouring algorithms running on $G_i$ and $G_j$ have to handle an edge insertion and deletion respectively, and this can be done in $\Tilde{O}_{\varepsilon}(1)$ time (see \Cref{lem:edgecolour}). \end{proof} \section{Introduction} In dynamic graph algorithms, the main goal is to maintain a key property of the graph while an adversary makes changes to the edges of the graph. An algorithm is called \emph{incremental} if it handles only insertions, \emph{decremental} if it handles only deletions and \emph{fully dynamic} if it handles both insertions as well as deletions. The goal is to minimize the update time of the algorithm, which is the time taken by the algorithm to adapt to a single adversarial edge insertion or deletion and output accordingly. For incremental/decremental algorithms, one typically seeks to minimize the \emph{total update time}, which is the aggregate sum of update times over the \emph{entire} sequence of edge insertions/deletions.\\ \\ We consider the problem of maintaining a $(1+\varepsilon)$-approximation to the maximum matching in a dynamic graph. In the fully dynamic setting, the best known update time for this problem is $O(\sqrt{m})$ (see \cite{GP13}), and the conditional lower bounds proved in the works of \cite{HKNS15} and \cite{KPP16} suggest that $O(\sqrt{m})$ is a hard barrier to break through. For this reason, several relaxations of this problem have been studied. For example, one line of research has shown that we can get considerably faster update times if we settle for large approximation factors (see for example \cite{BHI15,BHN16,BK21,BS16,Wajc2020,RSW22,BK22}). Another research direction has been to consider the more relaxed incremental or decremental models. In the incremental (insertion-only) setting, there have been a series of upper and lower bound results \cite{BLSZ14,Dahlgaard16,Gupta2014}, culminating in the result of \cite{GLSSS19}, who gave an optimal $O_{\varepsilon}(m)$ \emph{total} update time (amortized $O_{\varepsilon}(1)$) for $(1+\varepsilon)$-approximate maximum matching. \\ \\ The decremental (deletion-only) setting requires an entirely different set of techniques. In fact, $O(\sqrt{m})$ update time ($O(m^{1.5})$ total time) remained the best known until recently, when \cite{BPT20} gave an $\mbox{\rm poly}(\nicefrac{\log n}{\varepsilon})$ amortized update time algorithm for bipartite graphs. However, achieving a similar result for non-bipartite graphs remained an open problem. Our main theorem essentially closes the gap between bipartite and general graphs. \begin{restatable}{theorem2}{thmmain}\label{thm:mainintegral} Let $G$ be an unweighted graph and let $\varepsilon\in (0,\nicefrac{1}{2})$. Then, there exists a decremental algorithm with total update time $\Tilde{O}_{\varepsilon}(m)$ (amortized $\Tilde{O}(1)$) that maintains an integral matching $M$ of value at least $(1-\varepsilon)\cdot \mu(G)$, with high probability, where $G$ refers to the current version of the graph. The algorithm is randomized but works against an adaptive adversary. The dependence on $\varepsilon$ is $2^{O(\nicefrac{1}{\varepsilon^2})}$. \end{restatable} Our result largely completes the picture for partially dynamic matching by showing that in general graphs one can achieve $\mbox{\rm poly}(\log n)$ update time in both incremental and decremental settings. But there are a few secondary considerations that remain. Firstly, our update time is $O(\mbox{\rm poly}(\log n))$, rather than the $O(1)$ for the incremental setting (see \cite{GLSSS19}). Secondly, both the incremental of \cite{GLSSS19} and our decremental result for general graphs have an exponential dependence on $\nicefrac{1}{\varepsilon}$, whereas incremental/decremental algorithms for bipartite graphs have a polynomial dependence on $\nicefrac{1}{\varepsilon}$ (see \cite{GLSSS19,Gupta2014}). \\ \\ In algorithms literature, it has been the case that efficient matching algorithms for bipartite graphs do not easily extend to general graphs. Existence of blossoms (among other things), poses a technical challenge to obtaining analogous results for the general case. Consider the polynomial time algorithms for maximum matching for bipartite graphs, the most efficient algorithm, using alternating BFS, was discovered by Hopcroft and Karp; Karzanov (see \cite{HK73,Karzanov1973}) in 1973. However, several new structural facts and algorithmic insights were used by Micali and Vazirani to get the same runtime for general graphs (see \cite{MV80}). This is also a feature of recent work in different models, such as the streaming (see \cite{AG11,EKMS11} and \cite{MMU22}), fully dynamic (see \cite{BS15,BS16} and \cite{BHN16,BLM20}), and parallel models (see \cite{FGT16,ST17} and \cite{MV00,NV20}). We refer the reader to Section 1.3 of \cite{AV20} for a detailed discussion of this phenomenon. \section{High-Level Overview} \label{sec:overview} Our algorithm for \Cref{thm:mainintegral} follows the high-level framework of \emph{congestion balancing} introduced by Bernstein, Probst-Gutenberg, and Saranurak \cite{BPT20}. They used this framework to solve approximate decremental matching in bipartite graphs, and also to solve more general flow problems. But the framework as they used it was entirely limited to cut/flow problems. As we discuss below, extending this framework to non-bipartite graphs introduces significant technical challenges. Moreover, our result shows how the key subroutine of congestion balancing is naturally amenable to a primal-dual analysis, which we hope can pave the way for this technique to be applied to other decremental problems. \subsection{Previous Techniques} A key observation of \cite{BPT20} is that it is sufficient to develop an $\Tilde{O}_{\varepsilon}(m)$ algorithm that does the following: it either maintains a fractional matching of size at least $(1-2 \varepsilon)\cdot \mu(G)$ or certifies that $\mu(G)$ has dropped by a $(1-\varepsilon)$ factor because of adversarial deletions. Since a result of~\cite{Wajc2020} enables us to round any bipartite fractional matching to an integral matching of almost the same value, their observation gives an algorithm to maintain an integral matching of size $(1-3\varepsilon)\cdot \mu(G)$ in $\Tilde{O}_{\varepsilon}(m)$ time under adversarial deletions. To motivate why \cite{BPT20} consider computing a fractional matching, consider the following ``lazy'' algorithm that works with an integral matching: compute an $(1+\varepsilon)$ approximate integral matching $M$ of $G$ using a static $O(\nicefrac{m}{\varepsilon})$ algorithm, wait for $\varepsilon\cdot \mu(G)$ edges of $M$ to be deleted and then recompute the matching. Since we assume an adaptive adversary, the update time could be as large as $\Omega(\nicefrac{m^2}{\varepsilon\cdot n})$, this is because the adversary could proceed by only deleting edges of $M$. As a result, the goal should be to maintain a \emph{robust} matching that can survive many deletions. Thus, \cite{BPT20} aim to maintain a ``balanced'' fractional matching $\vec{x}$ that attempts to put a low value on every edge. In doing so, the adversary will have to delete a lot of edges of $G$ to reduce the value of $\vec{x}$ by $\varepsilon\cdot \mu(G)$. \paragraph*{Balanced Fractional Matching in Bipartite Graphs} In order to ensure that the fractional matching is spread out and robust, the authors of \cite{BPT20} impose a capacity function $\kappa$ on the edges of the graph (initially, all edges have low capacity) and compute a fractional matching obeying these capacities. The main ingredient of the algorithm is the subroutine $\textsc{M-or-E*}(G,\varepsilon,\kappa)$ which returns one of the following in $\Tilde{O}_{\varepsilon}(m)$ time: \begin{enumerate} \item A fractional matching $\vec{x}$ such that $\sum_{e\in E}x(e)\geqslant (1-\varepsilon)\cdot \mu(G)$ and $x(e)\leqslant \kappa(e)$ for all $e\in E$, or, \item\label{item:one1} A set of edges $E^*$ such that have the following two properties. \begin{enumerate} \item\label{item:onea} The total capacity through $E^*$ must be small: $\kappa(E^*)=O(\mu(G)\log n)$ and, \item\label{item:oneb} For all $\card{M}\geqslant (1-3\varepsilon)\cdot \mu(G)$, $\card{M\cap E^*}\geqslant \varepsilon\cdot \mu(G)$. \end{enumerate} \end{enumerate} Property \ref{item:onea} ensures that the total capacity increase is small, while Property \ref{item:oneb} ensures that we only increase capacity on important edges that are actually needed to form a large matching. The authors of \cite{BPT20} show that $\textsc{M-or-E*}()$ can be used as a black-box to solve decremental matching: at each step, \textsc{M-or-E*}() is used to find a large fractional matching $\vec{x}$ (this matching is then rounded using \cite{Wajc2020} to get an integral matching), or to output the set $E^*$ along which we increase capacities. They are able to show that because of Properties \ref{item:onea} and \ref{item:oneb}, the edge capacities remain small on average.\\ \\ The \emph{congestion balancing} framework of \cite{BPT20} thus, consists of an outer algorithm that uses \textsc{M-or-E*}() as a subroutine. The outer algorithm for bipartite graphs carries over to general graphs as well. But, \textsc{M-or-E*}() is significantly more challenging to implement for the case of general graphs, so this subroutine will be our focus for the rest of the high level review. For the case of bipartite graphs the algorithm \textsc{M-or-E*}() is easier to implement because maximum fractional matchings correspond to maximum flows in bipartite graphs. Hence, existing algorithms for approximate maximum flows can be used to find the approximate maximum fractional matching obeying capacity $\kappa$. Moreover, if such a fractional matching is not large, then in bipartite graphs, the set of bottleneck edges is exactly a minimum cut of the graph. For general graphs, due to the odd set constraints, max flow, which was the key analytic and algorithmic tool in \cite{BPT20}, no longer corresponds to a maximum fractional matching that avoids the integrality gap. \subsection{Our Contribution: Implementing \textsc{M-or-E*()} in General Graphs} At a high-level, there are several structural and computational challenges to implementing \textsc{M-or-E*}() in the case of general graphs. We explain what the potential impediments are, and detail how our techniques circumvent these. \subparagraph*{Fractional Matchings in General Graphs} In general graphs, not all fractional matchings have a large integral matching in their support and therefore, cannot be rounded to give a large matching. While fractional matchings that obey odd set constraints do avoid the integrality gap, it seems hard to compute such a matching that also obeys capacity function $\kappa$. In order to get past this, we define a candidate fractional matching that is both easy to compute as well as contains a large integral matching in its support. More concretely, our fractional matching either puts flow one through an edge, or a flow of value at most $\varepsilon$. It can be proved that such a fractional matching obeys all small odd set constraints, and avoids the integrality gap. Our main contribution are two structural lemmas which show that we can find our candidate matching efficiently. \begin{enumerate} \item First, given a graph $G$ with capacity $\kappa$, we want to determine if the \emph{value} of the maximum fractional matching obeying $\kappa$ and odd set constraints (denoted $\mu(G,\kappa)$) is at least $(1-\varepsilon)\cdot \mu(G)$. In general graphs, we do this by giving a sampling theorem: let $G_s$ be the graph created by sampling edge $e$ with probability proportional to $\kappa(e)$, then $\mu(G_s)\geqslant \mu(G,\kappa)-\varepsilon\cdot n$ with high probability. Thus, $\mu(G_s)$ is a good proxy for $\mu(G,\kappa)$ and it can be estimated efficiently by running any integral matching algorithm on $G_s$. \item Suppose we have determined at some point that $\mu(G,\kappa)$ is large, we are still left with the task of finding a fractional matching. Our next contribution is a structural theorem that enables us to deploy existing flow algorithms to find such a matching. Let $M$ be the approximate maximum matching of $G_s$. Let $M_{L}=\set{e\in M\mid \kappa(e)\leqslant \beta}$ and $M_{H}=\set{e\in M\mid \kappa(e)> \beta}$, where $\beta=O(\mbox{\rm poly}(\log n))$, and $L$ and $H$ are for low and high respectively. Let $V_{L}=V(M_L)$ and $V_{H}=V(M_H)$. Intuitively, $M$ breaks up our vertex set into two parts: vertices matched by low capacity edges (denoted $V_L$) and those that are matched by high capacity edges (denoted $V_{H}$). By adding some slack to our capacity constraints (we show that some slack can be incorporated in congestion balancing framework), we are able to treat the high-capacity edges as integral and compute a matching on $V_H$ using a black box for integral matching in general graphs. Additionally, we show that the maximum fractional matching on low capacity edges of $G[V_L]$ has value at least as much as $\card{M_L}$ (the low capacity edges of $M$), up to an additive error of $\varepsilon\cdot n$. To compute this fractional matching, we show that because we are only considering edges of small capacity, small odd set constraints are automatically satisfied, so we can transform $G[V_L]$ into a \emph{bipartite graph} and then use an existing flow algorithm. \end{enumerate} The second obstacle is finding the set $E^*$. As mentioned before, for the case of bipartite graphs, the max flow-min cut theorem gives us an easy characterization of the bottleneck edges. However, for the case of general graphs, this characterization is not clear. To get around this, we consider the dual of the matching LP of $G_s$, and show that the bottleneck edges can be identified by considering the dual constraints associated with the edges. This generalizes the cut-or-matching approach of \cite{BPT20}. Since $G_s$ is integral, we can compute the approximate dual by using an existing primal-dual algorithm of \cite{DP14} for integral matching in general graphs.\\ \\ Additionally, there are some secondary technical challenges as well. As mentioned before, our structural theorems only guarantee preservation of matching sizes up to an additive error of $\varepsilon\cdot n$. When $\mu(G)=o(n)$, then the results, applied directly are insufficient for us. To get around this, we use a vertex sparsification technique to get $O_{\varepsilon}(\log n)$ multigraphs which preserve all matchings of $G$, but contain only $O(\nicefrac{\mu(G)}{\varepsilon})$ vertices. However, we now have to show that all of our ideas work for multigraphs as well. Finally, the rounding scheme of \cite{Wajc2020} cannot be applied as a black-box to any fractional matching in a general graph. Thus, unlike in \cite{BPT20}, we cannot use \cite{Wajc2020} as a black box and instead have to embed its techniques into the congestion balancing framework. \subparagraph*{Assumption on Matching Size} Let $G$ be the input graph with vertex set $V$, note that if $\mu(G)\leqslant 100\log \card{V}$ at any time, then we can maintain a $(1+\varepsilon)$ approximate matching using Definition 3.2 and Lemma 3.3 from \cite{GP13} to solve the problem in $\Tilde{O}(\nicefrac{m}{\varepsilon})$ time. Thus, we only run our algorithm while $\mu(G)\geqslant 100\log \card{V}$. As mentioned before, our structural theorems, \Cref{lem:matchingGs} and \Cref{lem:sampling2} are proved for multigraphs $H$ with matching size at least $\mu(H)=\Omega(\varepsilon\cdot \card{V(H)})$. This is because \Cref{lem:mainvertexred} allows us to reduce to the problem of decremental matching in multigraphs with large matching. Now, suppose $H$ is the multigraph obtained by running the reduction of \Cref{lem:mainvertexred} on $G$, the input graph. Then, the structural theorems, \Cref{lem:matchingGs} and \Cref{lem:sampling2} hold with probability at least $1-\exp(-\mu(H))$. Since $\mu(H)= \Omega(\mu(G))$ and $\mu(G)\geqslant 100\log \card{V}$, these structural theorems hold with high probability. \section{Preliminaries} \label{sec:prelimsmain} We consider the problem of maintaining an approximate maximum integral matching in a graph $G$ in the decremental setting. In this setting, we are given a graph (possibly non-bipartite) $G_0=(V_0,E_0)$ with $\card{V_0}=n$ and $\card{E_0}=m$, and the adversary deletes edges from the graph one at a time. The goal is to maintain an approximate maximum matching of the graph $G$ as edges are deleted, while minimizing the \emph{total update time}, that is the aggregate sum of update times over the \emph{entire} sequence of deletions. \paragraph*{Notation.} Throughout the paper, we will use $G$ to refer to the current version of the graph, and let $V$ and $E$ be the vertex and edge sets of $G$ respectively. Additionally, let $\mu(G)$ denote the size of the maximum integral matching of $G$ (the current graph). During the course of the algorithm, we will maintain a fractional matching which corresponds to a non-negative vector $\vec{x}\in \mathbb{R}_{\geqslant 0}$ satisfying fractional matching constraints: $\sum_{e\ni v} x(e)\leqslant 1$. For a set $S\subseteq E$, then we let $x(S)=\sum_{e\in S} x(e)$. Given a capacity function $\kappa(e)$, we say that $x$ obeys $\kappa$ if $x(e) \leqslant \kappa(e)$ for all $e \in E$. For a vector $\vec{x}$, we use $\text{supp}(\vec{x})$ to be set of edges that are in the support of $\vec{x}$. For a fractional matching $\vec{x}$, we say that $\vec{x}$ satisfies odd set constraints if for every odd-sized $B\subseteq V$, $\sum_{e\in G[B]}x(e)\leqslant \frac{\card{B}-1}{2}$. Throughout this paper, for any $\varepsilon\in (0,1)$, we will let $\alpha_{\varepsilon}= \log n\cdot 2^{\nicefrac{60}{\varepsilon^2}}$ and $\rho_{\varepsilon}=\log n\cdot 2^{\nicefrac{40}{\varepsilon^2}}$. We now restate our main result. \thmmain* In order to solve this problem, we reduce to the case where it is sufficient to solve the same problem in multigraphs which have a large matching. More concretely, we prove the following theorem, and then show how that implies an algorithm for \Cref{thm:mainintegral}. \begin{restatable}{theorem2}{mainfracthm}\label{thm:mainfrac} We are given an unweighted multi-graph $G_0=(V,E_0)$. Let $\card{V}=n$, and $\varepsilon\in (0,\nicefrac{1}{2})$. Suppose $\mu(G_0)\geqslant \nicefrac{\varepsilon\cdot n}{100}$ and $\mu\geqslant (1-\varepsilon)\cdot \mu(G)$. Suppose $G$ is subject to adversarial deletions. There is an algorithm \textsc{Dec-Matching}($G,\mu,\varepsilon$), which processes deletions in $\Tilde{O}_{\varepsilon}(m)$ total update time and has the following guarantees. \begin{enumerate}[label=(\alph*)] \item \label{item:onethm} When the algorithm terminates, $\mu(G)<(1-2\varepsilon)\cdot \mu$. Upon termination the algorithm outputs ``\textsc{no}''. \item \label{item:twothm}Until the algorithm terminates, it maintains an integral matching with size at least $(1-20\varepsilon)\cdot \mu$. \end{enumerate} \end{restatable} The main contribution of our paper is to give an algorithm that proves \Cref{thm:mainfrac}. We first show how an algorithm for \Cref{thm:mainfrac} gives us an algorithm for \Cref{thm:mainintegral}. For this, we need the following reduction, which has been used previously in other settings such as for stochastic matchings (see for example \cite{AKL19} and \cite{CCEHMMV16}). Here, we use it in the decremental setting. \begin{restatable}{lemma2}{lemmreduc}\label{lem:reduction}\cite{AKL19} Let $\delta\in (0,1)$ and let $G$ be a graph with maximum matching size at least $\mu$ and at most $(1+\delta)\cdot \mu$. There is an algorithm \textsc{Vertex-Red}($G,\mu,\delta$) that takes as input $G$, $\mu$, and $\delta$, and in $O(m\cdot \nicefrac{\log n}{\delta^4})$ time returns $\lambda=\frac{100\cdot \log n}{\delta^4}$ multi-graphs $H_1,\cdots, H_{\lambda}$ that have the following properties, where the second property holds with probability at least $1-\exp\paren{\nicefrac{-\mu \cdot \log n}{\delta^4}}$. \begin{enumerate}[label=(\alph*)] \item \label{item:first} For all $i\in [\lambda]$, $\card{V(H_i)}= \frac{4\cdot (1+\delta)\cdot \mu}{\delta}$ and $E(H_i)\subseteq E(G)$. \item\label{item:third} Suppose $G$ is subject to deletions, which we simulate on each of the $H_i$'s. More concretely, if an edge $e$ is deleted from $G$, then its copy in each of the $H_i$'s (if present) is also deleted. Suppose at the end of the deletion process, $\mu(G)\geqslant \tilde{\mu}$ for some $\tilde{\mu}\geqslant (1-2\delta)\cdot \mu$. Then, $\mu(H_i)\geqslant (1-\delta)\cdot \tilde{\mu}$ for some $i\in [\lambda]$. \end{enumerate} \end{restatable} Additionally, we will need the following algorithm to compute an integral matching in the static setting in general graphs (see \cite{DP14}). \begin{lemma2}\label{lem:static}\cite{DP14} There is an $O(\nicefrac{m}{\varepsilon})$ time algorithm \textsc{Static-Match}() that takes as input a graph $G$, and returns an integral matching $M$ of $G$ with $\card{M}\geqslant (1-\varepsilon)\cdot \mu(G)$. \end{lemma2} We now show how \Cref{thm:mainfrac} in combination with \Cref{lem:reduction} and \Cref{lem:static} implies \Cref{thm:mainintegral}. We first give an informal description of the algorithm. \paragraph*{Description of Algorithm for Theorem \ref{thm:mainintegral} (\Cref{alg:proofthm1})} The algorithm takes as input a graph $G$, and a parameter $\varepsilon>0$. It first estimates the size of the maximum matching of $G$ by running \textsc{Static-Match}($G,\varepsilon$) on input $G$ and $\varepsilon$. We denote this estimate by $\mu$. The algorithm then initiates a procedure to create multiple parallel instances of the algorithm \textsc{Dec-Matching}(). As mentioned in \Cref{thm:mainfrac}, the algorithm \textsc{Dec-Matching}() can maintain a $(1+\varepsilon)$-approximate maximum matching in a multi-graph, provided the size of the matching in the multi-graph is large. So, we use the reduction mentioned in \Cref{lem:reduction} which creates multigraphs $H_1,H_2,\cdots, H_{\lambda}$, which have have $O(\nicefrac{\mu(G)}{\varepsilon})$ vertices as opposed to $n$ vertices, and we run parallel instances of \textsc{Dec-Matching}($H_i,\mu\cdot (1-\varepsilon),\varepsilon$) for all $i\in [\lambda]$. This procedure also instantiates a pointer \textsc{cur}, which points to the least indexed $H_i$, with a matching of size at least $(1-\varepsilon)\cdot \mu$. The algorithm will output the matching indexed by \textsc{cur} (denoted $M_{\textsc{cur}}$). The algorithm then moves on to a procedure for handling deletions of edges. If an edge $e\in G$ is deleted, then it is deleted from each of the $H_i$'s. If at this point, the size of the matching of any $H_\textsc{cur}$ by a large value, then we increment the value of \textsc{cur}. If the sizes of the maximum matchings of all $H_i$ have dropped by a significant amount, then we can conclude that with high probability, $\mu(G)$ has also dropped by a significant amount (by \Cref{lem:reduction}\ref{item:third}), and we restart the algorithm. \begin{algorithm} \algorithmicrequire{ Graph $G$ and a parameter $\varepsilon>0$}\\ \algorithmicensure{ A matching $M$ with $\card{M}\geqslant (1-8\varepsilon)\cdot \mu(G)$} \caption{} \begin{algorithmic}[1] \State $M\leftarrow\textsc{Static-Match}(G,\varepsilon)$. \label{item:line1} \State $\mu\leftarrow \card{M}$ \Procedure{Instantiating \textsc{Dec-Matching}()}{} \State $i\leftarrow 1$. \State \textsc{cur}$\leftarrow \lambda +1$. \State Let $H_1,H_2,\cdots ,H_{\lambda}$ be the multi-graphs returned by $\textsc{Vertex-Red}(G,\mu,\varepsilon)$. \State Label $H_1,\cdots, H_{\lambda}$ as \textsc{active}. \For{$i\leqslant \lambda$} \If{$\card{\textsc{Static-Match}(H_i,\varepsilon)}<(1-\varepsilon)\cdot \mu$} \State Label $H_i$ as \textsc{inactive}. \EndIf \EndFor \For{$H_i$ labelled \textsc{active}} \State Initialize $\textsc{Dec-Matching}(H_i,\mu\cdot (1-\varepsilon),\varepsilon)$ (denoted $\mathcal{A}_i$). \State Let $M_i$ be the matching maintained by $\mathcal{A}_i$. \EndFor \State Let \textsc{cur} be the least index $i$ such that $H_i$ is \textsc{active}. \If{\textsc{cur} $>\lambda$} \State \textsc{terminate}. \EndIf \EndProcedure \Procedure{Handling deletion of edge $e$}{} \For{$H_i$ labelled \textsc{active}} \State Feed the deletion of $e$ to $\mathcal{A}_i$. \label{alg1:dec-matching} \If{$\mathcal{A}_i$ returns \textsc{no}} \State Label $H_i$ as \textsc{inactive}. \EndIf \EndFor \If{$H_{\textsc{cur}}$ labelled \textsc{active}}\label{line:curactive} \Comment{Label may changed in \Cref{alg1:dec-matching}}. \State Continue to output $M_{\textsc{cur}}$. \Else \If{$\textsc{cur}>\lambda$} \label{line:restart} \State Return to \Cref{item:line1}. \Comment{Start over with new estimate for $\mu$}. \Else \State $\textsc{cur}\leftarrow \textsc{cur}+1$ \State Goto \Cref{line:curactive}. \EndIf \EndIf \EndProcedure \end{algorithmic} \label{alg:proofthm1} \end{algorithm} \begin{proof}[Proof of \Cref{thm:mainintegral}] We show that \Cref{alg:proofthm1} proves \Cref{thm:mainintegral}. Suppose we run \Cref{alg:proofthm1} on $G,\varepsilon$. We first argue that it returns an integral matching of $G$ of size at least $(1-10\varepsilon)\cdot \mu(G)$, where $\mu\geqslant \mu(G)\cdot (1-\varepsilon)$. To see this, first observe that the multi-graphs $H_1,\cdots, H_{\lambda}$ obtained as an output of \textsc{Vertex-Red}($G,\mu\cdot (1+\varepsilon),\varepsilon$) have $\frac{4\cdot (1+\varepsilon)\cdot \mu}{\varepsilon}\leqslant \nicefrac{8\cdot \mu}{\varepsilon}$ vertices, and $\mu(H_i)\geqslant (1-\varepsilon)\cdot \mu$. Thus, $H_i$ satisfy the requirements of \textsc{Dec-Matching}() mentioned in \Cref{thm:mainfrac}. Hence, each instance of $\textsc{Dec-Matching}(H_i,\mu,\varepsilon)$, until it returns \textsc{no}, maintains a matching of size at least $(1-9\varepsilon)\cdot \mu$, which is at least $(1-10\varepsilon)\cdot \mu(G)$. \\ \\ We now want to bound the runtime of the algorithm. To do this, we first make the following claim. \begin{claim2}\label{claim:restart} Each time \Cref{alg:proofthm1} executes \Cref{line:restart}, $\mu(G)$ has dropped by a factor of $(1-2\varepsilon)$. \end{claim2} \begin{proof}First observe that every time the algorithm returns to \Cref{item:line1}, it must be the case that all the \textsc{Dec-Matching}($H_i,\mu\cdot(1-\varepsilon),\varepsilon$) returned \textsc{no}, which implies by \Cref{thm:mainfrac}\ref{item:onethm} that for all $i\in [\lambda]$, $\mu(H_i) < (1-2\varepsilon)\cdot (1-\varepsilon)\cdot \mu$. This in turn implies that $\mu(G)<(1-2\varepsilon)\cdot\mu$. Suppose for contradiction that this is not the case, then from \Cref{lem:reduction}\ref{item:third} we can conclude that there is a matching of size at least $(1-3\varepsilon)\cdot \mu$ in some $H_i$. More specifically, if $\mu(G)\geqslant (1-2\varepsilon)\cdot \mu$, then it satisfies the hypothesis of \Cref{lem:reduction}\ref{item:third}, and we know that $\mu(H_i)\geqslant (1-3\varepsilon)\cdot\mu$ for some $i\in \bracket{\lambda}$. This implies that \textsc{Dec-Matching}($H_i,(1-\varepsilon)\cdot \mu,\varepsilon$) wouldn't have terminated for some $i\in \bracket{\lambda}$ (by \Cref{thm:mainfrac}\ref{item:onethm}). Consequently, \Cref{line:restart} wouldn't have been executed, which is a contradiction. Thus, each time we execute \Cref{line:restart}, $\mu(G)$ has dropped from at least $\mu$ to at most $(1-2\varepsilon)\cdot \mu$. This proves our claim. \end{proof} From Claim \ref{claim:restart}, we can conclude that we return to \Cref{line:recomp} at most $\log_{1+\varepsilon} n$ times. Thus, to upper bound the runtime of \Cref{alg:proofthm1}, it is sufficient to upper bound the runtime of each of the procedures. The runtime of the first procedure, which instantiates $\lambda$ instances of \textsc{Dec-Matching}($H_i,\mu\cdot(1-\varepsilon),\varepsilon$) is dominated by the runtime of \textsc{Vertex-Red}($G,\mu,\varepsilon$), which is run once in the procedure, and the runtime of \textsc{Static-Match}($H_i,\varepsilon$) which is run $\lambda$ times (once per multi-graph $H_i$). Thus, the total time of this procedure is $O(\nicefrac{m}{\varepsilon}\cdot \log\nicefrac{1}{\varepsilon}\cdot \lambda)$, which is $O(\nicefrac{m}{\varepsilon^5}\cdot \log n\cdot \log \nicefrac{1}{\varepsilon})$.\\ \\ Next, we upper bound the runtime of the second procedure. If an edge $e$ is deleted from $G$, then the time to delete it from each of the multigraphs $H_1,\cdots, H_{\lambda}$ is $\lambda$. Finally, we are also maintaining at most $\lambda$ active instances of $\mathcal{A}_i$, and their total update time is $O_{\varepsilon}(m\cdot \mbox{\rm poly}(\log n))$. Consequently, we have that the total update time of \Cref{alg:proofthm1} is $O_{\varepsilon}(m\cdot \mbox{\rm poly}(\log n))$. \end{proof} \section{Algorithm for \Cref{thm:mainfrac}} We use $\mu(G,\kappa)$ to denote the value of the maximum fractional matching of $G$ obeying the capacity function $\kappa$ and the odd set constraints. To give an algorithm for \Cref{thm:mainfrac}, we need two ingredients. We first maintain a balanced fractional matching of the multigraph. To do this, we will show an algorithm \textsc{M-or-E*}() that given as input a capacitated graph $G$ either outputs a fractional matching nearly obeying these capacities if $\mu(G,\kappa)\geqslant (1-\varepsilon)\cdot \mu(G)$, or if $\mu(G,\kappa)<(1-\varepsilon)\cdot \mu(G)$, it outputs a set of edges $E^*$ along which capacities must be increased. When we say that the capacities are nearly obeyed, we mean that the can be exceeded only by a small factor. Then, we round the fractional matching using a known algorithm. We start with some definitions, and then go on to state the guarantees of the two algorithms, and show how it gives us an algorithm for \Cref{thm:mainfrac}. \begin{definition2} Given a multi-graph $G$, for a pair of vertices $u$ and $v$, define $D(u,v)$ to be the set of edges between $u$ and $v$. Similarly, if $e$ is an edge between $u,v$, then $D(e)\coloneqq D(u,v)$. \end{definition2} \begin{definition2}\label{def:dist} Let $G$ be a multigraph with $n$ vertices and $m$ edges. Let $\kappa$ be a capacity function on the edges. Suppose $\vec{x}$ is a fractional matching of $G$ ($\vec{x}$ is a vector of length $m$). Then, we define $\vec{x}^C$ to be a vector of support size at most $\min\set{m,{n\choose 2}}$, where for a pair of vertices, $u$ and $v$, $x^C{(u,v)}\coloneqq\sum_{e\in D(u,v)}x(e)$. Essentially, if $\vec{x}$ is a fractional matching on a multigraph, then $\vec{x}^C$ is a fractional matching obtained by ``collapsing'' all edges together. Similarly, if $\vec{y}$ is a vector of support at most $n\choose 2$, then we define $\vec{y}^{D}$ to be an $m$ length vector such that for every $e\in E$ between a pair of vertices $u$ and $v$, $y^D(e)\coloneqq\frac{y(u,v)\cdot \kappa(e)}{\kappa(D(e))}$ ($D$ is for distributed, and we distribute the flow among the edges in proportion to their capacity). \end{definition2} \begin{remark2} Note that in doing the transformations in \Cref{def:dist}, the support size of the transformed vector is always at most $m$. Thus, it doesn't negatively affect our runtime. \end{remark2} We now state our core ingredient, which either finds a balanced fractional matching of the multigraph $G$, or gives a set of edges $E^*$ along which we can increase capacity. \begin{restatable}{lemma2}{lemtrio}\label{lem:MorE} Let $G$ be a multi-graph with $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$. Let $\kappa$ be a capacity function on the edges of $G$. There is an algorithm \textsc{M-or-E*}(), that takes as input $G,\kappa, \varepsilon\in (0,\nicefrac{1}{2})$ and $\mu\geqslant (1-\varepsilon)\cdot \mu(G)$ and in time $O(\nicefrac{m\cdot \log n}{\varepsilon})$ returns one of the following. \begin{enumerate}[label=(\alph*)] \item\label{item:aa} A fractional matching $\vec{x}$ of value at least $(1-10\varepsilon)\cdot \mu$ with the following properties. \begin{enumerate}[label=(\roman*)] \item\label{item:ai} For any $e\in \text{supp}(\vec{x})$ with $\kappa(D(e)) > \nicefrac{1}{\alpha_{\varepsilon}^2}$, $x(e)=\frac{\kappa(e)}{\kappa(D(e))}$, and $x(D(e))=1$. \item\label{item:aii} For any $e\in \text{supp}(\vec{x})$ with $\kappa(D(e)) \leqslant \nicefrac{1}{\alpha_{\varepsilon}^2}$, $x(e)\leqslant \kappa(e)\cdot \alpha_{\varepsilon}$ and $x(D(e))\leqslant \kappa(D(e))\cdot \alpha_{\varepsilon}$. \end{enumerate} \item\label{item:bi} \label{item:edgeE} A set $E^*$ of edges such that $\kappa(E^*)=O(\mu \log n)$ such that for any integral matching $M$ with $\card{M}\geqslant (1-3\varepsilon)\cdot \mu$, we have $\card{M\cap E^*}\geqslant \varepsilon \mu$. Moreover, $\kappa(e)<1$ for all $e\in E^*$. Additionally, for every pair of vertices $u,v\in V$, either $D(u,v)\cap E^*=\emptyset$ or $D(u,v)\subseteq E^*$. \end{enumerate} \end{restatable} We give some intuition for \Cref{lem:MorE}. Recall that we need a balanced fractional matching that contains a large integral matching in its support. However, as mentioned before, in the case of general graphs, finding such a balanced fractional matching is not straightforward. In order to get past this obstacle, we define a balanced fractional matching that is easy to find and also avoids the integrality gap. We will explain how to find it in the subsequent sections. For now, we explain at a high-level, why the fractional matching $\vec{x}$ found by \Cref{lem:MorE} avoids the integrality gap. Consider $\vec{x}^C$. Observe from \Cref{lem:MorE}\ref{item:aa}, that for any pair of vertices $u,v\in G$, either $x^C((u,v))=1$ or $x^C((u,v))\leqslant \varepsilon$. Thus, $\vec{x}$ satisfies odd-set constraints for all odd sets of size at most $\nicefrac{1}{\varepsilon}$. By a folklore lemma, we can then argue that $\vec{x}$ contains an integral matching of size at least $(1+\varepsilon)^{-1}\cdot \sum_{u\neq v}x^C((u,v))$. \\ \\ We will use \textsc{M-or-E*}() as a subroutine in the algorithm \textsc{Dec-Matching}(). The fractional matching output by \textsc{M-or-E*}() will have certain properties, we state these properties now, since it will be helpful in visualizing the fractional matching. We give a proof for these later on. \begin{property} We will use \textsc{M-or-E*}($G,\mu,\kappa,\varepsilon$) as a subroutine in \textsc{Dec-Matching}($G,\mu,\varepsilon$) to get a matching $\vec{x}$ with the following properties. \begin{enumerate} \item Each time \textsc{M-or-E*}() returns the set $E^*$, we increase capacity along $E^*$ by multiplying $\kappa(e)$ for each $e\in E^*$ by the same factor. \item Consider $u,v\in V$ and let $e$ and $e'$ be edges in between $u,v$. Then, $\kappa(e')=\kappa(e)$ at all times during the run of the algorithm. \end{enumerate} \end{property} The next property follows immediately from \Cref{lem:MorE}\ref{item:a}. \begin{property}\label{prop:capexceed} Let $\vec{x}$ be the matching output by \textsc{M-or-E*}($G,\mu,\kappa,\varepsilon$), then $x(e)\leqslant \kappa(e)\cdot \alpha_{\varepsilon}^2$ for all $e\in E$. \end{property} \begin{definition2} Let $G$ be a multi-graph, let $\kappa$ be a capacity function on the edges of $G$ and let $\vec{x}$ be a fractional matching obeying $\kappa$. Let $\varepsilon\in (0,1)$. Then, we split $\vec{x}$ into two parts, $\vec{x}^f$ and $\vec{x}^i$, where $\vec{x}=\vec{x}^f+\vec{x}^i$, and $\text{supp}(\vec{x}^f)=\set{e\in E\mid \kappa(D(e))< \nicefrac{1}{\alpha_{\varepsilon}^2}}$ and $\text{supp}(\vec{x}^i)=\set{e\in E\mid \kappa(D(e))\geqslant \nicefrac{1}{\alpha_{\varepsilon}^2}}$ (here, $\vec{x}^f$ stands for fractional and $\vec{x}^i$ stands for integral. Although the edges in $\vec{x}^i$ are not technically integral, they are large enough for us to round them). \end{definition2} We briefly state the implications of this definition. We do not use these properties in our proof, but it will be useful to state it nevertheless. \begin{property}\label{prop:disjsupp} Let $G$ be any multigraph, and let $\vec{x}$ be a fractional matching of $G$. Then, for any pair of vertices $u,v$, either $D(u,v)\subseteq \text{supp}(\vec{x}^i)$ and $D(u,v)\cap \text{supp}(\vec{x}^f)=\emptyset$ or, $D(u,v)\subseteq \text{supp}(\vec{x}^f)$ and $D(u,v)\cap \text{supp}(\vec{x}^i)=\emptyset$. \end{property} \begin{observation2}\label{obs:suppmatch} Suppose $\vec{x}$ is a fractional matching returned by \textsc{M-or-E*}() and consider $\vec{z}=\vec{x}^i$. Then supp($\vec{z}^C$) is a matching. This is implied by \Cref{lem:MorE}\ref{item:a}\ref{item:ai}. \end{observation2} The second ingredient is a method to round fractional matchings. This theorem is implicit in \cite{Wajc2020}, however, we prove it in \Cref{sec:rounding} for completeness. \begin{restatable}{lemma2}{lemuno}\label{thm:sparse}\cite{Wajc2020} Suppose $G$ is an unweighted simple graph, let $\varepsilon\in(0,\nicefrac{1}{2})$, and let $\vec{x}$ be a fractional matching of $G$ such that $x(e)\leqslant \varepsilon^6$. Then, there is a dynamic algorithm \textsc{Sparsification}($\vec{x},\varepsilon$), that has the following properties. \begin{enumerate}[label=(\alph*)] \item \label{lem:sparseb} The algorithm maintains a subgraph $H\subseteq\text{supp}(\vec{x})$ such that $\card{E(H)}=O_{\varepsilon}(\mu(\text{supp}(\vec{x}))\cdot \mbox{\rm poly}(\log n))$, and with high probability $\mu(H)\geqslant (1-\varepsilon)\cdot \sum_{e\in E}x(e)$. \item \label{lem:sparsea} The algorithm handles the following updates to $\vec{x}$: the adversary can either remove an edge from supp($\vec{x}$) or for any edge $e$, the adversary can reduce $x(e)$ to some new value $x(e)\geqslant 0$. \item \label{lem:sparsec} The algorithm handles the above-mentioned updates in $\Tilde{O}_{\varepsilon}(1)$ worst-case time. \end{enumerate} \end{restatable} We now show how \textsc{M-or-E*}() and \textsc{Sparsification}() give us an algorithm for \Cref{thm:mainfrac}. This algorithm, called \textsc{Dec-Matching}() is stated in \Cref{alg:decmatching}. Before proving \Cref{thm:mainfrac}, we give a brief description of \Cref{alg:decmatching}. \paragraph*{Description of \textsc{Dec-Matching}().} The algorithm takes as input a multigraph $H$, a parameter $\varepsilon$, an estimate $\mu$ on the size of the maximum matching of $H$. It then instantiates the capacities of the graph $H$. The algorithm then starts a new phase by running \textsc{M-or-E*}() on this capacitated graph. The algorithm returns either a large fractional matching or if the matching is small, then it outputs a set of edges $E^*$. The algorithm \textsc{Dec-Matching}() in the latter case, increases capacities along $E^*$. Suppose at some point the algorithm finds a large fractional matching $\vec{x}$. Recall that $\vec{x}$ output by \textsc{M-or-E*}() is a matching with the following property: consider $\vec{x}^C$, then either $x^C((u,v))=1$ or $x^C((u,v))\leqslant \nicefrac{1}{\alpha_\varepsilon}$ (this is guaranteed by \Cref{lem:MorE}\ref{item:aa}). The algorithm extracts the latter part, and sparsifies it using \textsc{Sparsification}(). Since the fractional matching input to \textsc{Sparsification}() has flow at most $\nicefrac{1}{\alpha_{\varepsilon}}$ between any pair of vertices, it satisfies the hypothesis of \Cref{thm:sparse} and obtain a graph $S$ containing $\Tilde{O}(\mu(H))$ edges and a large integral matching. Thus, \textsc{Stat-Match}($S,\varepsilon$) is run and this matching is combined with the ``integral'' part of $\vec{x}$ and is output. \begin{algorithm}[H] \caption{\textsc{Dec-Matching}($H,\mu,\varepsilon$)} \label{alg:decmatching} \begin{algorithmic}[1] \State \label{line:initcap}For all $e\in E(H)$, $\kappa(e)\leftarrow \nicefrac{1}{\alpha_{\varepsilon}^{\lceil \log_{\alpha_{\varepsilon}}n \rceil}}$ \Procedure{Starting a new phase}{} \State $\mu'\leftarrow$ \textsc{Static-Match}($H,\varepsilon$)\label{line:recomp}\Comment{$\mu'\geqslant (1-\varepsilon)\cdot \mu(H)$} \If{$\mu'\leqslant (1-3\varepsilon)\cdot \mu$} \State Return \textsc{no}, and terminate.\label{line:funcinfty} \Else \While{\textsc{M-or-E*}($H,\kappa,\varepsilon,\mu'$) returns $E^*$} \State \label{line:capchange} For all $e\in E^*$, $\kappa(e)\leftarrow \kappa(e)\cdot \alpha_{\varepsilon}$ \EndWhile \EndIf \EndProcedure \State Let $\vec{x}$ be the matching returned by \textsc{M-or-E*}(). \Comment{This is a matching in a multigraph.} \State $\vec{y}\leftarrow \vec{x}^f$, $\vec{z}\leftarrow \vec{x}^i$ \label{line:recomp2}\Comment{We will update $\vec{y},\vec{z}$ as edges are deleted.}\label{line:yz} \State $S\leftarrow \textsc{Sparsification}(\vec{y}^C,\varepsilon)$ \Comment{Recall that Sparsification is dynamic} \label{line:sparsification} \State \label{line:rounding}$M\leftarrow \textsc{Static-Match}(S,\varepsilon)\cup \text{supp}(\vec{z}^C)$. \Comment{The matching $M$ that is output.} \State $\textsc{CounterM}\leftarrow 0$ \Comment{Counter for deletions to the integral matching.} \State $\textsc{Counterx}\leftarrow 0$ \Comment{Counter for deletions to the fractional matching.} \Procedure{For deletion of edge $e$ from $H$}{} \If{$e\in \text{supp}(\vec{x})$} \State Delete $e$ from $\text{supp}(\vec{x})$ \State Update $\vec{z}^C$ accordingly. \Comment{See \Cref{line:yz}.} \State \label{line:sparsification-update}Update $\vec{y}^C$ accordingly; \textsc{Sparsification} from Line \ref{line:sparsification} then updates $S$. \State \textsc{Counterx} $\leftarrow$ \textsc{Counterx} $+\vec{x}(e)$. \EndIf \If{\textsc{Counterx} $>\varepsilon\cdot \mu$} \State Goto \Cref{line:recomp} \Comment{End current phase and start new one.} \EndIf \If{$e\in M$} \State Delete $e$ from $M$ \State \textsc{CounterM} $\leftarrow$ \textsc{CounterM} $+1$ \EndIf \If{$\textsc{CounterM}>\varepsilon\cdot \mu$} \Comment{The matching $M$ is out of date} \State $\textsc{CounterM}\leftarrow 0$ \State \label{line:rebuild} $M\leftarrow \textsc{Static-Match}(S,\varepsilon)\cup \text{supp}(\vec{z}^C)$ \Comment{Recompute outputted matching $M$} \EndIf \EndProcedure \end{algorithmic} \end{algorithm} The algorithm then begins a procedure to handle deletions. If an edge $e$ is deleted, then it is removed from \text{supp}($\vec{x}$), and \textsc{Sparsification}() algorithm updates $S$ in $\Tilde{O}_{\varepsilon}(1)$ worst case time (\Cref{thm:sparse}\ref{lem:sparsec}). The algorithm also maintains counters for deletions to $\vec{x}$ and $M$. If the total value deleted from $\vec{x}$ exceeds $\varepsilon\cdot \mu$, then it ends the current phase, and starts a new one. On the other hand, if the number of edges deleted from $M$ exceeds $\varepsilon\cdot \mu$, then, the algorithm recomputes $M$ by running \textsc{Static-Match}($S,\varepsilon$) again (since we know that $\vec{x}$ still contains a large integral matching in its support). \\ \\ We also state the following lemma which we prove later in \Cref{sec:lemma7}. \begin{lemma2}\label{lem:runtime} The subroutine \textsc{M-or-E*}() and \Cref{line:recomp} are called at most $O(\alpha_{\varepsilon}^3\log n)$ times in \Cref{alg:decmatching} during the course of deletion of $m$ edges. \end{lemma2} We restate our theorem and give a proof. \mainfracthm* \begin{proof}[Proof of \Cref{thm:mainfrac}] We first argue that \Cref{alg:decmatching} satisfies the properties of \Cref{thm:mainfrac}. First observe that when the algorithm executes \Cref{line:funcinfty}, $\mu'\leqslant (1-3\varepsilon)\cdot \mu$. Since $\mu'\geqslant (1-\varepsilon)\cdot\mu(H)$, this implies that $\mu(H)\leqslant (1-2\varepsilon)\cdot \mu$. Thus, when the algorithm terminates, $\mu(H)\leqslant (1-2\varepsilon)\cdot \mu$, which proves \Cref{thm:mainfrac}\ref{item:onethm}. \\ \\ Next, we want to argue that while the algorithm does not terminate, it outputs an integral matching of size at least $\mu\cdot(1-20\varepsilon)$. This corresponds to \Cref{thm:mainfrac}\ref{item:twothm}. We first argue that the algorithm indeed outputs matching. Note that the output matching $M$ is the union of \text{supp}($\vec{z}^C$) and the output of \textsc{Sparsification}($\vec{y}^C,\varepsilon$). Note that $M$ is a union of two matchings, is implied by Observation \ref{obs:suppmatch} and \Cref{thm:sparse}. Moreover, these matchings are vertex disjoint and this follows from Property \ref{prop:disjsupp}. Thus, $M$ is a matching. \\ \\ Next,we want to argue that $\card{M}\geqslant (1-20\varepsilon)\cdot \mu$. First, observe that the fractional matching output by \textsc{M-or-E*}() has value at least $(1-10\varepsilon)\cdot\mu$. Next, observe that because of \Cref{lem:MorE}\ref{item:a}\ref{item:aii}, the matching output by \textsc{M-or-E*}() satisfies the conditions of \Cref{thm:sparse}: that is, $\vec{y}^C$ has the property that the flow through any edge in the support, $\vec{y}^C(e) \leqslant \nicefrac{1}{\alpha_{\varepsilon}}\leqslant \varepsilon^3$. Thus, $\mu(S)\geqslant (1-\varepsilon)\cdot \sum_{e\in \text{supp}(\vec{y}^C)}y^C(e)$. Therefore, the matching output by \textsc{Sparsification}($\vec{y}^C,\varepsilon$) is an integral matching of size at least $(1-2\varepsilon)\cdot \sum_{e\in \text{supp}(\vec{y}^C)}\vec{y}^C(e)$. Thus, we can conclude, \begin{align*} \card{M}& \geqslant \sum_{e\in\text{supp}(\vec{x}^i)}x(e)+(1-2\varepsilon)\cdot\sum_{e\in\text{supp}(\vec{x}^f)}x(e)\geqslant (1-2\varepsilon)\cdot \sum_{e\in \text{supp}(\vec{x})}x(e)\\ &\geqslant (1-2\varepsilon)\cdot (1-10\varepsilon)\cdot \mu\geqslant (1-12\varepsilon)\cdot \mu. \end{align*} Thus, the algorithm maintains an integral matching of size at least $(1-20\varepsilon)\cdot \mu$ at all times. \\ \\ {\bf Running Time:} We now bound the runtime of the algorithm. Let us first consider the time to initialize a phase. From \Cref{lem:runtime} we know that \textsc{M-or-E*}() is called $O(\alpha_{\varepsilon}^3\log n)$ times, so this is also an upper bound on the number of phases. Thus, \textsc{Static-Match}() in Lines \ref{line:recomp} and \ref{line:rounding} is also called at most $O(\alpha_{\varepsilon}^3\log n)$ times, since these are only called once per phase. All of these subroutines have runtime $\tilde{O}_{\varepsilon}(m)$, so since each is executed $\tilde{O}_{\varepsilon}(1)$ times, the total runtime of all of them is $\tilde{O}_{\varepsilon}(m)$.\\ \\ We also need to bound the runtime of \textsc{Sparsification}($\vec{y}^C,\varepsilon$) and the total contribution of \textsc{Static-Match}() in \Cref{line:rebuild}. Let us start by bounding the runtime of \textsc{Sparsification}(). The time to initialize this procedure in \Cref{line:sparsification} is $\tilde{O}(m)$, and this is only done once per phase, so again the total time spent on initialization is $\tilde{O}(m)$. We also spend time updating the output sparsifier $S$ in Line \ref{line:sparsification-update}. The worst-case runtime time of \textsc{Sparsification}() is $O(\log n)$ per update to $\vec{x}$ (see \Cref{thm:sparse}\ref{lem:sparsea} and \ref{lem:sparsec}). But each update to $\vec{x}$ corresponds to an adversarial deletion of some edge $e$, so there at most $m$ adversarial deletions, so the total overall time (among all phases) to update \textsc{Sparsification} is $\tilde{O}(m)$. \\ \\ Finally, we want to count the contribution of \textsc{Static-Match}() in \Cref{line:rebuild}. Observe that by \Cref{thm:sparse}, the subgraph $S$ has size $\Tilde{O}(\mu)$. Thus, the runtime of \textsc{Static-Match}($S,\varepsilon$) is $\Tilde{O}(\nicefrac{\mu}{\varepsilon})$. Now, observe that \Cref{line:rebuild} is only executed when \textsc{CounterM} has increased from $0$ to $\varepsilon\cdot \mu$, and the counter increases by at most $1$ per adversarial deletion, so there are at least $\varepsilon\cdot \mu$ adversarial deletions per execution of \Cref{line:rebuild}. Thus, \textsc{Stat-Match}($S,\varepsilon$) is called at most $\nicefrac{m}{\varepsilon\cdot \mu}$ times in \Cref{line:rebuild}, and the total contribution here is $O(\nicefrac{m}{\varepsilon^2})$. \end{proof} \section{Proof of \Cref{lem:runtime}} \label{sec:lemma7} To prove \Cref{lem:runtime}, which gives an upper bound on the number of times the subroutine \textsc{M-or-E*}() is called in \Cref{alg:decmatching}, we will follow the framework of \cite{BPT20}. We will define a potential function, and show that it increases as $\kappa$ is increased, and edges are deleted. While their framework applies to simple graphs, we verify that it is possible to extend it to multigraphs as well. Towards this, we give a few definitions. \begin{definition2}\label{def:mathcalM} Let $G$ be a multigraph, and let $\kappa$ be the capacity function on the edges of the graph (if the adversary deletes an edge $e$, then let $\kappa(e)$ be the capacity of the edge right before it is deleted). Let $\mu$ be the input to \textsc{Dec-Matching}(), when it is run on $G,\varepsilon$. Let $\mathcal{M}$ be the set of all integral matchings of $G$ of size at least $(1-3\varepsilon)\cdot \mu$. Define the cost of an edge $e$ to be $c(e)=\log\paren{n\cdot \kappa(e)}$. For any integral matching $M$, define $c(M)=\sum_{e\in M}c(e)$. Define $\Pi(G,\kappa)=\min_{M\in \mathcal{M}}c(M)$. If $\mathcal{M}=\emptyset$, then $\Pi(G,\kappa)=\infty$. \end{definition2} \begin{observation2} Initially, $\kappa(e)=\nicefrac{1}{\alpha_{\varepsilon}^{\lceil \log_{\alpha_{\varepsilon}}n \rceil}}$, so $\Pi(G,\kappa)=0$. Note that capacities $\kappa$ only change (increase) in \Cref{line:capchange}. Consequently, the capacity function $\kappa$ is monotonically increasing. Moreover, edges of $G$ are only deleted. Thus, $\Pi(G,\kappa)$ is monotonically increasing as well. \end{observation2} \begin{observation2} When $\Pi(G,\kappa)=\infty$, then \Cref{line:funcinfty} of \Cref{alg:decmatching} causes the algorithm to terminate. \end{observation2} \begin{proof} Suppose \Cref{line:funcinfty} doesn't cause \Cref{alg:proofthm1} to terminate, then, $\mu'\geqslant (1-3\varepsilon)\cdot \mu$. Thus, $\mu(G)\geqslant \mu'\geqslant (1-3\varepsilon)\cdot \mu$, and $\mathcal{M}\neq \emptyset$. \end{proof} \begin{lemma2} In \textsc{Dec-Matching}(), we say that we begin a new phase when $\varepsilon\cdot \mu$ value of the fractional matching $\vec{x}$ has been deleted (that is, the value of \textsc{Counterx} has increased to $\varepsilon\cdot \mu$). Suppose we are in a phase when the algorithm does not terminate. Let $\kappa$ be the final capacities right before we process deletions. Then, $\Pi(G,\kappa)=O(\mu\log n)$. \end{lemma2} \begin{proof} First observe that for any edge $e$, $\kappa(e)\leqslant 1$, so $c(e)=O(\log n)$. This is implied by the fact that $E^*$ only consists of edges that have $\kappa(e)<1$ and $\kappa(e)$ is a power of $\alpha_{\varepsilon}$ (see \Cref{lem:MorE}\ref{item:bi}). This is because, we start with $\kappa(e)$ to be a power of $\alpha_{\varepsilon}$ initially (see \Cref{line:initcap}), and each time we increase $\kappa(e)$, we multiply it by $\alpha_{\varepsilon}$. Moreover, for any matching $M$, we know that $\card{M}\leqslant \mu\cdot (1+\varepsilon)$, since, $\mu\geqslant \mu(G)\cdot (1-\varepsilon)$. Thus, $c(M)\leqslant (1+\varepsilon)\cdot\mu\cdot \log n$. This implies that $\Pi(G,\kappa)=O(\mu\log n)$. \end{proof} \begin{definition2} Let $E_0$ be the initial set of edges and let $\kappa(E_0)=\sum_{e\in E_0}\kappa(e)$. \end{definition2} \begin{lemma2}\label{lem:congestbal} Suppose a call to \textsc{M-or-E*}($G,\mu,\kappa,\varepsilon$) returns the set $E^*$ instead of a matching. Let $\kappa'$ denote the new edge capacities after increasing capacities along $E^*$. Then, \begin{enumerate}[label=(\alph*)] \item\label{item:congesta} $\kappa'(E_0)\leqslant \kappa(E_0)+\alpha_{\varepsilon} \cdot \mu\cdot \log n$, and \item\label{item:congestb} $\Pi(G,\kappa')\geqslant \Pi(G,\kappa)+\nicefrac{\mu}{\varepsilon}$. \end{enumerate} \end{lemma2} \begin{proof} We begin by recalling \Cref{lem:MorE}\ref{item:edgeE} that $\kappa(E^*)\leqslant \mu\log n$. Thus, $\kappa'(E^*)=O(\alpha_{\varepsilon}\cdot\mu\cdot \log n)$, since it is obtained by scaling up $\kappa(E^*)$ by a $\alpha_{\varepsilon}$. Thus, we have, \begin{align*} \kappa'(E_0)&=\kappa(E_0\setminus E^*)+\kappa'(E^*) \end{align*} This implies that: \begin{align*} \kappa'(E_0)-\kappa(E_0\setminus E^*)&= \kappa'(E^*) = \alpha_{\varepsilon}\cdot \mu\cdot \log n \end{align*} The LHS is lower bounded by $\kappa'(E_0)-\kappa(E_0)$, since $\kappa(E_0)\geqslant \kappa(E_0\setminus E^*)$. This proves the first part of our claim. To prove the second part, first notice that $E^*$ contains at least $\varepsilon\cdot\mu$ edges of every matching of size at least $(1-3\varepsilon)\cdot \mu$ (implied by \Cref{lem:MorE}\ref{item:edgeE}). Let $M,M'\in \mathcal{M}$ be the matchings that minimize $\Pi(G,\kappa)$ and $\Pi(G,\kappa')$ respectively, and let $c$ and $c'$ be the cost functions associated with $\kappa$ and $\kappa'$ respectively. Observe that $\card{M}\geqslant (1-3\varepsilon)\cdot \mu$ and $\card{M'}\geqslant (1-3\varepsilon)\cdot \mu$ (by definition of $\mathcal{M}$, see \Cref{def:mathcalM}). Then, we have the following. \begin{align*} \Pi(G,\kappa')&=c'(M')=\sum_{e\in M'} \log\paren{n\cdot \kappa'(e)}=\sum_{e\in M'\setminus E^*}\log \paren{n\cdot \kappa(e)}+ \sum_{e\in M'\cap E^*} \log\paren{n\cdot \kappa(e)\cdot \alpha_{\varepsilon}}\\ &=\sum_{e\in M'}\log \paren{n\cdot \kappa(e)}+\card{M'\cap E^*}\cdot \log \alpha_{\varepsilon} \\ &= c(M')+ \nicefrac{\mu}{\varepsilon}\geqslant \Pi(G,\kappa)+ \nicefrac{\mu}{\varepsilon}\\ &\text{(Since }\card{M'\cap E^*}\geqslant \varepsilon\cdot \mu\text{ and }\alpha_{\varepsilon}=2^{\nicefrac{32}{\varepsilon^2}}) \\ \end{align*} This proves our claim. \end{proof} \begin{observation2}\label{obs:calls1} The total number of calls to \textsc{M-or-E*}() (till \textsc{Dec-Matching}() terminates) that return $E^*$ are upper bounded by $O(\varepsilon\cdot \log n)$. This is because the potential function $\Pi(G,\kappa)$ is upper bounded by $O(\mu\log n)$, and each call to \textsc{M-or-E*}() that returns $E^*$, increments $\Pi(G,\kappa)$ by $\nicefrac{\mu}{\varepsilon}$ (see \Cref{lem:congestbal}\ref{item:congestb}). Thus, we can deduce that $\kappa(E_0)\leqslant \alpha_{\varepsilon} \cdot \mu \cdot \varepsilon \cdot \log n$. This is because each call to \textsc{M-or-E*}() that returns $E^*$ increases $\kappa(E_0)$ by at most $\alpha_{\varepsilon}\cdot \mu\cdot \log n$ (see \Cref{lem:congestbal}\ref{item:congesta}), and there are at most $\varepsilon\cdot \log n$ such calls. \end{observation2} We now want to upper bound the number of calls made to \textsc{M-or-E*}() that return a matching. This is upper bounded by the number of phases. \begin{lemma2}\label{lem:phases} The total number of phases is at most $O(\alpha_{\varepsilon}^3\cdot \log n)$. \end{lemma2} \begin{proof} We define $\Phi_{\textsf{del}}\coloneqq\sum_{e\in E_{\textsf{del}}}\kappa(e)$ to be the capacity of deleted edges. Observe that $\Phi_{\textsf{del}}\leqslant \kappa(E_0)\leqslant \alpha_{\varepsilon}\cdot \varepsilon\cdot \mu\cdot \log n$. This is implied by the definition of $\kappa(E_0)$ and Observation \ref{obs:calls1}. Moreover, each time we start a new phase, at least $\varepsilon\cdot \mu$ value of the fractional matching has been deleted. That is $\sum_{e\in E}x(e)$ has dropped by $\varepsilon\cdot \mu$. From Property \ref{prop:capexceed}, one can conclude that for any edge $e$, $x(e)\leqslant \alpha_{\varepsilon}^2\cdot \kappa(e)$. This implies that a phase contributes at least $\nicefrac{\varepsilon\cdot \mu}{\alpha^2_{\varepsilon}}$ to $\Phi_{\textsf{del}}$. Using the fact that $\Phi_{\textsf{del}}$ is upper bounded by $\alpha_{\varepsilon}\cdot \varepsilon\cdot \mu\cdot \log n$, we can conclude that the total number of phases is at most $\alpha_{\varepsilon}^3\cdot \log n$. \end{proof} \begin{proof}[Proof of \Cref{lem:runtime}] From \Cref{lem:phases} and Observation \ref{obs:calls1}, we conclude that the total number of calls to \textsc{M-or-E*}() are upper bounded by $O(\alpha_{\varepsilon}^3\cdot \log n)$. Finally, the number of calls to \textsc{Static-Match}() due to \Cref{line:recomp} are upper bounded by the number of phases, which are at most $O(\alpha_{\varepsilon}^3\cdot \log n)$ by \Cref{lem:phases}. \end{proof} \section{Ingredients for Algorithm \textsc{M-or-E*}()} Recall that we use $\mu(G,\kappa)$ to denote the value of the maximum fractional matching of $G$ obeying capacity function $\kappa$ and the odd set constraints. As in the congestion balancing setup of \cite{BPT20}, we want to check if $\mu(G,\kappa)\geqslant (1-\varepsilon)\cdot \mu(G)$. However, unlike in bipartite graphs, where we can use flows to find fractional matching, there is no simple way to check if $\mu(G,\kappa)\geqslant (1-\varepsilon)\cdot \mu(G)$ in general graphs. Our first structural result circumvents this issue. Let $G_s$ be the graph obtained by sampling every edge $e$ with probability $p(e)=\min\set{1,\kappa(e)\cdot \rho_{\varepsilon}}$. We show that $\mu(G_s)\geqslant \mu(G,\kappa)-\varepsilon\cdot \mu(G)$. Thus, we can run \textsc{Static-Match}($G_s,\varepsilon$) to estimate $\mu(G,\kappa)$. \\ \\ At a high level, \textsc{M-or-E*}() proceeds in three phases. In Phase 1, it creates $G_s$ and computes $\mu(G_s)$. If this matching is large, it proceeds to Phase 2, where it finds a fractional matching $\vec{x}$ such that $\sum_{e\in E}x(e)\geqslant (1-\varepsilon)\cdot \mu(G)$. On the other hand, if $\mu(G_s)$ is small, then it proceeds to Phase 3, where it finds the set of edges $E^*$ along which it increases capacity. In the subsequent sections, we will state the main structural properties we use in each of the phases. Finally, in \Cref{sec:lemma7}, we put together these ingredients to give \textsc{M-or-E*}(), and prove \Cref{lem:MorE}. \subsection{Phase 1 of \textsc{M-or-E*()}} Before we formally state the main guarantees of Phase 1, we will state some standard results in matching theory, that we will use in our main result for Phase 1. \subsubsection{Some Standard Ingredients For Phase 1} The first ingredient we use is the Tutte-Berge formula. \begin{definition2} Let $G$ be any graph (possibly containing multiedges), and let $U\subseteq V$, then \text{odd}$_{G}(V\setminus U)$ refers to the number of odd components in $G[V\setminus U]$. \end{definition2} \begin{lemma2}[Tutte-Berge Formula]\cite{Schrijver2003CombinatorialOP}\label{lem:tuttewitn} The size of a maximum matching in a graph $G=(V,E)$ is equal to $\frac{1}{2}\min\limits_{U\subseteq V}\paren{\card{U}+\card{V}-\text{odd}_{G}(V\setminus U)}$. \end{lemma2} Additionally, we will use some properties of the matching polytope. \begin{lemma2}\cite{Schrijver2003CombinatorialOP}\label{lem:folklore} Let $G$ be any graph, and let $\vec{x}$ be a fractional matching that in addition to the fractional matching constraints, also satisfies the following for all odd-sized $U\subseteq V$: $\sum\limits_{e\in G[U]} x(e)\leqslant \frac{\card{U}-1}{2}$. Then, there is an integral matching $M\subseteq \text{supp}(x)$ with $\card{M}=\sum_{e\in E}x(e)$. \end{lemma2} \begin{definition2} Let $G$ be any multigraph, and let $S,T\subseteq V$, then $\delta_{G}(S,T)$ is defined as the set of edges that have one endpoint in $S$ and the other in $T$. Additionally, for $S\subseteq V$, we define $\delta_G(S)$ to be the set of edges that have one end point in $S$, and the other in $V\setminus S$. \end{definition2} \begin{figure} \centering \includegraphics[scale=0.25]{tutte.png} \caption{The figure shows the graph $G$, and a partition $\mathcal{P}=\set{U,E_1,\cdots, E_{q},O_1,\cdots, O_{t}}$ satisfying \ref{item:part1} and \ref{item:part2}. The thick edges (in red and purple), are the edges in supp($\vec{x}$), where $\vec{x}$ is the fractional matching realizing $\mu(G,\kappa)$. The purple edges (edges between the odd components, or between odd and even components) correspond to $E_{\textsf{miss}}^{\mathcal{P}}$, and $\sum_{e\in E _{\textsf{miss}}^{\mathcal{P}}}x(e)\geqslant 2\cdot \varepsilon\cdot \mu(G)$. } \label{fig:tutte} \end{figure} \subsubsection{Main Lemma for Phase 1} As mentioned earlier, in Phase 1 of \textsc{M-or-E*}(), we first create a sampled graph $G_s$. In the following lemma, we show that $\mu(G_s)$ is a good estimate for $\mu(G,\kappa)$ with high probability. \begin{lemma2}\label{lem:matchingGs} Let $G$ be a multigraph with $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$ where $\varepsilon\in (0,\nicefrac{1}{2})$. Let $\kappa$ be a capacity function on the edges of $G$, and let $G_{s}$ be obtained by sampling every edge $e\in G$ with probability $p(e)=\min\set{\kappa(e)\cdot \rho_{\varepsilon},1}$. Let $\mu(G,\kappa)$ be the value of the maximum fractional matching of $G$ obeying the capacities $\kappa$, and the odd set constraints. Then, with high probability, $\mu(G_s)\geqslant \mu(G,\kappa)-\varepsilon\cdot\mu(G)$. \end{lemma2} \begin{proof} We want to show that with probability at least $1-\nicefrac{1}{n^2}$, $\mu(G_s)\geqslant \mu(G,\kappa)-\varepsilon\cdot \mu(G)$. In order to do this, by \Cref{lem:folklore}, it is sufficient to show that with probability at least $1-\nicefrac{1}{n^2}$, $\frac{1}{2}\min\limits_{U\subseteq V}\paren{\card{U}+\card{V}-\text{odd}_{G_s}(V-U)}\geqslant \mu(G,\kappa)-\varepsilon\cdot \mu(G)$. \\ \\ Towards this, we consider a fixed partition $\mathcal{P}$ of $V$ into sets $U,O_1,\cdots,O_{t},E_1,\cdots,E_{q}$ with the following properties (see \Cref{fig:tutte}). \begin{enumerate}[label=(\alph*)] \item \label{item:part1} We have, $q\geqslant 0$ and $t>n-2\cdot \mu(G,\kappa)+2\varepsilon\cdot \mu(G)+\card{U}$. \item \label{item:part2} Sets $O_i$ for $i\in \bracket{t}$ are odd-sized sets and sets $E_l$ for $l\in [q]$ are even-sized sets. \end{enumerate} Note that if $\mu(G_s)< \mu(G,\kappa)-\varepsilon\cdot \mu(G)$, then there is a partition $\mathcal{P}=\set{U,O_1,\cdots,O_t,E_1,\cdots,E_{q}}$ of $G_s$, satisfying \ref{item:part1} and \ref{item:part2} such that $V\setminus U$ is the union of components $O_1,\cdots, O_t,E_1,\cdots, E_{q}$. Note that if $V\setminus U$ is the union of disjoint components $O_1,\cdots, O_{t},E_1,\cdots,E_{q}$ then $\delta_{G_s}(O_i,O_l)=\emptyset$ for all $i\neq l$ and $\delta_{G_s}(O_i,E_l)=\emptyset$ for all $i\neq l$ (this is evident from \Cref{lem:tuttewitn}). Thus, to upper bound the probability that $\mu(G_s)<\mu(G,\kappa)-\varepsilon\cdot \mu(G)$, it is sufficient to upper bound the probability that for all partitions $\mathcal{P}=\set{U,O_1,\cdots,O_{t},E_1,\cdots,E_{q}}$ satisfying \ref{item:part1} and \ref{item:part2}, \emph{none} of the edges $E^{\mathcal{P}}_{\textsf{miss}}=\set{e\mid e\in \delta_{G}(O_i,O_l)\text{ for }i\neq l}\cup\set{e\mid e\in \delta_{G}(O_i,E_l)\text{ for }i\in [t],l\in[q]}$ are sampled in $G_s$. In order to bound this probability, we make the following claim. \begin{claim2} \label{claim:emiss} For a partition $\mathcal{P}$ satisfying \ref{item:part1} and \ref{item:part2}, $\kappa(E^{\mathcal{P}}_\textsf{miss})\geqslant 2\cdot \varepsilon\cdot \mu(G)$. \end{claim2} \begin{proof} Let $\vec{x}$ be a fractional matching obeying odd set constraints and capacity function $\kappa$ such that $\sum_{e\in E}x(e)=\mu(G,\kappa)$. In order to prove this claim, we show that if $\kappa(E_{\textsf{miss}}^{\mathcal{P}})<2\cdot \varepsilon\cdot \mu(G)$, then, $x(E_{\textsf{miss}}^{\mathcal{P}})>\kappa(E_{\textsf{miss}}^{\mathcal{P}})$, which will contradict the fact that $\vec{x}$ is a fractional matching obeying $\kappa$.\\ \\ With this proof strategy in mind, for contradiction assume that $\kappa(E^{\mathcal{P}}_{\textsf{miss}})< 2\cdot \varepsilon\cdot \mu(G)\leqslant n-2\cdot \mu(G,\kappa)+2\cdot \varepsilon\cdot \mu(G)$. The last inequality follows from the fact that $\mu(G,\kappa)$ corresponds to the value of the maximum fractional matching, so, $\mu(G,\kappa)\leqslant \frac{n}{2}$. Since $\vec{x}$ obeys odd set constraints, $\sum_{l\leqslant t}x(\delta_G(O_l))\geqslant t$ (from \Cref{lem:folklore}). Note that $\sum_{l\leqslant t}x(\delta_G(O_{l},U))\leqslant \card{U}$, otherwise for some $v\in U$, $\sum_{e\ni v}x(e)>1$, violating the fact that $\vec{x}$ is a fractional matching. Next, we observe that $\sum_{l\leqslant t}x(\delta_G(O_l))=x(E^{\mathcal{P}}_{\textsf{miss}})+\sum_{l\leqslant t}x(\delta_G(O_l,U))$. This follows from the fact that all edges emanating out of $O_i$ in $G$, are either incident on other $O_j$, or $E_k$ or $U$. We have the following set of inequalities. \begin{align*} x(E_{\textsf{miss}}^{\mathcal{P}})&=\sum_{l\leqslant t} x(\delta_{G}(O_l))-\sum_{l\leqslant t}x(\delta_{G}(O_l,U))\\ &\geqslant t-\card{U}\\ &> n-2\cdot \mu(G,\kappa)+2\cdot \varepsilon\cdot \mu(G)+\card{U}-\card{U} \end{align*} Thus, that $x(E^{\mathcal{P}}_{\textsf{miss}})\geqslant n-2\mu(G,\kappa)+2\varepsilon\mu(G)>\kappa(E^{\mathcal{P}}_{\textsf{miss}})$. This is a contradiction. This concludes our proof, and we know that $\kappa(E^{\mathcal{P}}_{\textsf{miss}})\leqslant 2\cdot \varepsilon\cdot \mu(G)$. \end{proof} Additionally, we have the following claim, which allows us to only focus on $\mathcal{P}$ for which all $e\in E^{\mathcal{P}}_{\textsf{miss}}$ have $\kappa(e)<\nicefrac{1}{\rho_{\varepsilon}}$. \begin{observation2}\label{claim:emiss2} Suppose $\kappa(e)\geqslant \nicefrac{1}{\rho_{\varepsilon}}$ for any $e\in E^{\mathcal{P}}_{\textsf{miss}}$, then, $$\prob{\text{None of the edges in }E^{\mathcal{P}}_{\textsf{miss}}\text{ are sampled in }G_s}=0.$$ \end{observation2} The above observation follows from the fact that an edge $e$ is sampled with probability $\min\set{1,\kappa(e)\cdot \rho_{\varepsilon}}$. From the above claim, we can deduce that if for any partition $\mathcal{P}$, if $E_{\textsf{miss}}^{\mathcal{P}}$ has an edge $e$ with $\kappa(e)\geqslant \nicefrac{1}{\rho_{\varepsilon}}$, then the contribution of $E_{\textsf{miss}}^{\mathcal{P}}$ to our probability bound will be $0$. Hence, it is sufficient to focus on $E_{\textsf{miss}}^{\mathcal{P}}$ where all edges $e$ have $\kappa(e)< \nicefrac{1}{\rho_{\varepsilon}}$. We now bound the following probability. \begin{align*} \Prob\paren{\text{None of the edges in }{E^{\mathcal{P}}_{\textsf{miss}}\text{ are sampled in }G_s}}&\leqslant \prod_{e\in E^{\mathcal{P}}_{\textsf{miss}}}(1-p(e))\\ &\leqslant \exp\paren{-\sum_{e\in E^{\mathcal{P}}_{\textsf{miss}}}p(e)}\\ &=\exp\paren{-\sum_{e\in E^{\mathcal{P}}_{\textsf{miss}}}\kappa(e)\cdot \rho_{\varepsilon}}\\ &\text{(Since }p(e)=\kappa(e)\cdot \rho_{\varepsilon}\text{ (from Claim \ref{claim:emiss2} and discussion above))}\\ &=\exp\paren{-\varepsilon\cdot 2^{\nicefrac{16}{\varepsilon^2}}\cdot \mu(G)\cdot \log n}\\ &\text{(From Claim \ref{claim:emiss} and the fact that }\rho_{\varepsilon}=2^{\nicefrac{16}{\varepsilon^2}}\cdot \log n)\\ &\leqslant \exp\paren{- 2^{\nicefrac{15}{\varepsilon^2}}\cdot \mu(G)\cdot \log n}. \end{align*} Note that the total number of partitions of graph $G$ are upper bounded by $2^{n\cdot \log n}$. Hence, this also upper bounds the total number of partitions of $G$ satisfying \ref{item:part1} and \ref{item:part2}. Since $n\leqslant \nicefrac{16\cdot\mu(G)}{\varepsilon}$ and $\varepsilon<\nicefrac{1}{2}$, the bound on the number of partitions is at most $2^{\nicefrac{16\mu}{\varepsilon}\cdot \log n}$ (by assumption), taking a union bound over all the partitions, we know that the with probability at least $1-\exp\paren{-\nicefrac{\mu(G)\cdot \log n}{\varepsilon}}$, in $G_s$, we have no partition $\mathcal{P}=\set{U,O_1,\cdots, O_t,E_{1},\cdots, E_{q}}$ satisfying \ref{item:part1} and \ref{item:part2}. Thus, by \Cref{lem:tuttewitn}, we have that with high probability, $\mu(G_s)\geqslant \mu(G,\kappa)-\varepsilon\cdot \mu(G)$. \end{proof} \subsection{Phase 2 of \textsc{M-or-E*}()} The algorithm proceeds to Phase 2 only if the integral matching $M_s$ found in $G_s$ is close to $\mu(G,\kappa)$. Recall that our goal is to compute a fractional matching so that we can apply the congestion balancing framework. However, as mentioned before there is no straightforward way of computing a maximum fractional matching in a general graph obeying capacity $\kappa$. To overcome this, we give a candidate fractional matching which is easy to compute, and is sufficient for our purposes (that is, it avoids the integrality gap). At a high level, this is what Phase 2 does, and in this section we describe our candidate fractional matching and show that it is close in value to $\mu(G_s)$ with high probability, and therefore it is close to $\mu(G,\kappa)$ as well (by \Cref{lem:matchingGs}). \subsubsection{Preliminaries for Phase 2} Phase 2 starts by computing $M_s = \textsc{Static-Match}(G_s, \varepsilon)$ and then uses $M_s$ to compute the desired fractional matching. We will split $M_s$ into low capacity edges and high capacity edges, and as a result split $V$ into vertices matched using high capacity edges, and low capacity edges. We begin by giving a formal definition of low capacity edges. \begin{definition2}\label{def:lowedges} Let $G$ be any multigraph, and let $\kappa$ be a capacity function on the edges of $G$. Let $\varepsilon\in (0,\nicefrac{1}{2})$. Define $E_{L}(G,\kappa)=\set{e\in E\mid e\in D(u,v)\text{ and }\kappa(D(u,v))\leqslant \nicefrac{1}{\alpha^2_{\varepsilon}}}$. Intuitively, $E_{L}(G,\kappa)$ is the set of low total capacity edges. \end{definition2} As mentioned in the high-level overview, in order to prove our probabilistic claims, we will give some slack to the capacities. This motivates our next definition. \begin{definition2}\label{def:kappaplus} Let $G$ be a multi-graph, and let $\kappa$ be a capacity function on the edges of $G$. Let $\varepsilon\in (0,\nicefrac{1}{2})$. We define the capacity function $\kappa^+$ as follows. \begin{enumerate}[label=(\alph*)] \item For all $e\in E_{L}(G,\kappa)$, $\kappa^{+}(e)=\kappa(e)\cdot \alpha_{\varepsilon}$. \item For all $e\in E\setminus E_{L}(G,\kappa)$, $\kappa^+(e)=\kappa(e)$. \end{enumerate} \end{definition2} To make our analysis easier to follow, we need the following definition of a bipartite double cover of $G$. \begin{definition2} Let $G$ be a multi-graph and let $\kappa$ be a capacity function on the edges of $G$. We define the bipartite double cover $\textsc{bc}(G)$ to be a bipartite graph with capacity function $\kappa_{\textsc{bc}}$ as follows. \begin{enumerate}[label=(\alph*)] \item For every vertex $v\in V(G)$, make two copies $v$ and $v'$ in $V(\textsc{bc}(G))$. \item If $e$ is an edge between $u,v\in V(G)$, then for each such $e$ we add two edges $e'$ and $e''$, one between $u$ and $v'$ and the other between $v$ and $u'$. We let $\kappa_{\textsc{bc}}(e')=\kappa_{\textsc{bc}}(e'')=\kappa(e)$. \end{enumerate} \end{definition2} We have the following standard claim relating $\mu(G)$ and $\mu(\textsc{bc}(G))$. \begin{claim2}\label{claim:matchbc} For any multi-graph $G$, $\mu(\textsc{bc}(G))\geqslant 2\cdot \mu(G)$. \end{claim2} Next, we state the following lemma, which follows from standard techniques, and we give a formal proof of it in \Cref{sec:missingproof}. The lemma essentially states that a fractional matching which has low flow on all edges has a very small integrality gap. \begin{restatable}{lemma2}{lemnono}\label{lem:folklore2} Let $G$ be a multigraph, and let $\varepsilon\in (0,1)$. Let $\kappa$ be a capacity function on the edges of $G$, with $\kappa(D(e))\leqslant \nicefrac{1}{\alpha_{\varepsilon}}$ for all $e\in E(G)$. Then, $\mu(\textsc{bc}(G),\kappa_{\textsc{bc}})\leqslant 2\cdot\paren{1+\varepsilon}\cdot \mu(G,\kappa)$, where $\mu(G,\kappa)$ is the maximum fractional matching of $G$ obeying $\kappa$ and the odd set constraints, and $\mu(\textsc{bc}(G),\kappa_{\textsc{bc}})$ is the maximum fractional matching of $\textsc{bc}(G)$ obeying $\kappa_{\textsc{bc}}$. \end{restatable} Additionally, we will need the following version of Hall's theorem, which follows as a corollary of \Cref{lem:tuttewitn}. \begin{proposition}[Extended Hall's Theorem]\label{prop:halls} Let $G=(L\cup R, E)$ be a bipartite graph with $n=\card{L}=\card{R}$, then $\mu(G)=n-\max_{S\subseteq L}\paren{\card{S}-\card{N_G(S)}}$, where $N_G(S)$ refers to the neighbourhood of $S$ in $G$. \end{proposition} In order to prove \Cref{lem:sampling2}, we will use the following version of Chernoff bound. \begin{lemma2}[Chernoff Bound]\label{lem:chernoff} Let $X_1,\cdots, X_k$ be negtively correlated random variables, and let $X$ denote their sum, and let $\mu=\expect{X}$. Suppose $\mu_{\text{min}}\leqslant \mu\leqslant \mu_{\text{max}}$, then for all $\delta>0$, $\Prob\paren{X\geqslant (1+\delta)\mu_{\text{max}}}\leqslant \paren{\frac{e^{\delta}}{(1+\delta)^{\delta}}}^{\mu_{\text{max}}}$. \end{lemma2} Additionally, we state the following observation, we refer the readers to \cite{BPT20} for a proof of this observation. The proof follows from a standard application of the max-flow min-cut theorem, and we refer the reader to \cite{BPT20} for a proof sketch. \begin{observation2}\label{obs:maxflow} Let $G$ be any bipartite multigraph, with vertex bipartitions $S$ and $T$. Let $\kappa$ be the capacity function on the edges of the graph. Then, for any $C\subseteq S$, and $D\subseteq T$, we have, $\card{S}-\card{C}+\card{D}+\kappa(C,T\setminus D)\geqslant\mu(G,\kappa)$. Moreover, there are sets $C\subseteq S$ and $D\subseteq T$ such that equality holds. \end{observation2} \subsubsection{Main Result for Phase 2} We briefly give some intuition about the statement of our claim. Recall in the high level review, we mentioned that $M$, the integral matching of $G_s$ can split into two parts $M_H$, which is the high capacity part, and $M_L$, the low capacity part, and we defined $V_H=V(M_H)$ and $V_L=V(M_L)$. We said that congestion balancing allows us to give slack to capacities, and therefore, we can round up the capacities of $M_H$ to $1$. However, we still want to compute a fractional matching $G[V_L]$. In order to do this, we observe that if the fractional matching in $G[V_L]$ is only on low capacity edges, then we can use flow algorithms to compute such a matching. Therefore, our main structural result for Phase 2 states that if $\vec{y}$ is a fractional matching on the low capacity edges of $G[V_L]$, then with high probability $\sum_{e\in E}y(e)\geqslant \card{M_L}-\varepsilon\cdot \mu(G)$. We now state this result formally. \begin{figure} \centering \includegraphics[scale=0.25]{sampling.png} \caption{In the left panel, we consider the graph $G$, and $W=\set{a,b,c,d,e}$. The red edges correspond to the low capacity edges of $G[W]$, denoted $G[W]\cap E_{L}$. On the right panel, we have the bipartite graph $\textsc{bc}(G[W]\cap E_L)$, which has bipartitions $W_1$ and $W_2$. The solid red edges are the edges going between $C$ and $\bar{D}$, and these edges have capacity $\kappa(C,\bar{D})$. } \label{fig:sampling} \end{figure} \begin{lemma2}\label{lem:sampling2} Let $G$ be a multigraph and let $\varepsilon\in (0,\nicefrac{1}{2})$ and suppose $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$. Let $\kappa$ be a capacity function on the edges, and let $G_s$ be the graph obtained from $G$ by sampling each edge $e$ with probability $p(e)=\kappa(e)\cdot \rho_{\varepsilon}$. Let $E_{L}\coloneqq E_{L}(G,\kappa)$. Then, for all $W\subseteq V$, we have, with high probability, \begin{align*} \mu(\textsc{bc}\paren{G_{s}[W]\cap E_{L}})\leqslant \mu(\textsc{bc}\paren{G[W]\cap E_{L}},\kappa^{+}_{\textsc{bc}})+\varepsilon\mu(G). \end{align*} \end{lemma2} \begin{remark2} Note that by definition of $E_{L}$, all edges $e\in G[W]\cap E_{L}$ have $\kappa(e)\leqslant \nicefrac{1}{\alpha_{\varepsilon}^2}$. Thus, $\kappa^+_{\textsc{bc}}(e)=\kappa_{\textsc{bc}}(e)\cdot \alpha_{\varepsilon}$ for all $e\in \textsc{bc}(G[W]\cap E_{L})$. \end{remark2} Before we prove it, we have the following statement as a corollary of \Cref{lem:sampling2}. \begin{corollary2}\label{cor:sampling} Let $G$ be a multi-graph, and let $\varepsilon\in (0,1)$. Let $\kappa$ be a capacity function on the edges of $G$ with $\kappa(e)\leqslant \nicefrac{1}{\alpha_{\varepsilon}}$ for all $e\in E(G)$. Then, with high probability, \textbf{for all} $W\subseteq V$, $\mu(G_{s}[W]\cap E_{L})\leqslant \mu(G[W]\cap E_{L},\kappa^+)+2\varepsilon\mu(G)$. \end{corollary2} \begin{proof} The inequality in the statement follows due to the following line of reasoning. \begin{equation*} \begin{split} 2\cdot \mu(G_{s}[W]\cap E_{L})&\leqslant \mu(\textsc{bc}\paren{G_{s}[W]\cap E_{L}})\\ &\text{(since $2\cdot \mu(H)\leqslant \mu(\textsc{bc}(H))$, see Claim \ref{claim:matchbc})}\\ &\leqslant \mu(\textsc{bc}(G[W]\cap E_{L}),\kappa^+_{\textsc{bc}})+\varepsilon\mu(G)\\ &\text{(by \Cref{lem:sampling2})}\\ &\leqslant 2\cdot(1+\varepsilon)\cdot \mu(G[W]\cap E_{L},\kappa^+)+\varepsilon\mu(G)\\ &\text{(by \Cref{lem:folklore2})}\\ &\leqslant 2\cdot \mu(G[W]\cap E_{L},\kappa^+) +3\varepsilon\mu(G). \end{split} \end{equation*} The last inequality follows from the fact that any fractional matching in $G$ obeying $\kappa^+$ and odd set constraints is upper bounded by $\mu(G)$ (by \Cref{lem:folklore}). Dividing by two on both sides, we have our claim. \end{proof} \begin{proof}[Proof of \Cref{lem:sampling2}] Consider a fixed $W\subseteq V$. From now, we will use $H$ to denote $\textsc{bc}(G[W]\cap E_{L})$ and let $H_s$ denote $\textsc{bc}(G_s[W]\cap E_{L})$. For the bipartite graph $H$, we will use $W_1$ and $W_2$ to denote the two bipartitions of $H$ corresponding to $W$ (see \Cref{fig:sampling} for an illustration). We now want to prove that $\mu(H_s)\leqslant \mu(H,\kappa^+_{\textsc{bc}})+\varepsilon\mu(G)$. To prove this, it is sufficient to show a set $C\subseteq W_1$ such that $\card{C}-\card{N_{H_s}(C)}\geqslant \card{W_1}-\mu(H,\kappa^+_{\textsc{bc}})-\varepsilon\mu(G)$ with high probability. Then, by \Cref{prop:halls}, we have the following inequality. \begin{align*} \card{W_1}-\mu(H,\kappa^+_{\textsc{bc}})-\varepsilon\mu(G) \leqslant \card{C}-\card{N_{H_s}(C)}\leqslant \card{W_1}-\mu(H_{s}). \end{align*} This would prove our claim. Towards this, we consider the set $C\subseteq W$ satisfying the following equation (applying Observation \ref{obs:maxflow} to $H$ with $\kappa_{\textsc{bc}}^+$ on edges), and show that this is the required set. \begin{align}\label{eqn:upper} \card{W_1}-\card{C}+\card{D}+\kappa_{\textsc{bc}}^+(C,W_2\setminus D)=\mu(H,\kappa^+_{\textsc{bc}}). \end{align} We want to show that $\card{C}-\card{N_{H_s}(C)}\geqslant\card{W_1}-\mu(H,\kappa^+_{\textsc{bc}})-\varepsilon\mu(G)$ with high probability. Let $L$ be the set of vertices in $N_{H_s}(C)\cap W_2\setminus D$. We know that $\card{N_{H_s}(C)}\leqslant \card{D}+\card{L}$. \\ \\ Let $E_{H}(C,W_2\setminus D)$ be the set of edges in $H$ between $C$ and $W_2\setminus D$. Let $X$ be the random variable that denotes the number of edges in $E_{H}(C,W_2\setminus D)$ that are sampled in $H_s$. Note that $X$ is not a sum of independent random variables. Recall that $H$ is a subgraph of $\textsc{bc}(G)$, and suppose $e\in G[W]\cap E_{L}$ is included in $G_s$, then $e'$ and $e''$ (recall $e'$ and $e''$ are copies of $e$ in \textsc{bc}($G$)) are both included in $H_s$ else both are excluded. Thus, the random variables associated with $e'$ and $e''$ are correlated with each other. We instead consider an arbitrary subset of $E^*_{H}(C,W_2\setminus D)$ of $E_{H}(C,W_2\setminus D)$ that satisfies the following properties. \begin{enumerate}[label=(\alph*)] \item For $e\in G[W]$, if $\set{e',e''}\subset E_{H}(C,W_2\setminus D)$, then exactly one of $e'$ or $e''$ is included in $E^*_{H}(C,W_2\setminus D)$. \item For $e\in G[W]$, if $\set{e',e''}\cap E_{H}(C,W_2\setminus D)=\set{e'}$, then only $e'$ is included in $E^*_{H}(C,W_2\setminus D)$. \item For $e\in G[W]$, if $\set{e',e''}\cap E_{H}(C,W_2\setminus D)=\set{e''}$, then only $e''$ is included in $E^*_{H}(C,W_2\setminus D)$. \end{enumerate} We consider the random variable $Y$ that denotes the number of edges in $E^*_{H}(C,W_2\setminus D)$ that are included in $H_s$. Observe that $Y$ is a sum of independent random variables satisfying the condition of \Cref{lem:chernoff}. Moreover, $X\leqslant 2Y$. Thus, it is sufficient to upper bound the value $Y$ can take with high probability. Note that for any $e\in E_{H}(C,W_2\setminus D)$, using the definition of $H$, $\kappa_{\textsc{bc}}(e)<\nicefrac{1}{\alpha^2_{\varepsilon}}$. Since $\rho_{\varepsilon}<\alpha_{\varepsilon}$, $p(e)=\rho_{\varepsilon}\cdot \kappa_{\textsc{bc}}(e)<1$. So, we have, \begin{align*} \expect{Y}&\leqslant \sum_{e\in E_{H}^*(C,W_2\setminus D)} p(e)\leqslant \rho_{\varepsilon}\cdot \kappa_{\textsc{bc}}(C,W_2\setminus D)\leqslant \frac{\kappa^+_{\textsc{bc}}(C,W_2\setminus D)}{2^{\nicefrac{16}{\varepsilon^2}}}. \end{align*} The last inequality follows from the fact that in $H$, $\kappa^+_{\textsc{bc}}(e)=\kappa_{\textsc{bc}}(e)\cdot \alpha_{\varepsilon}$ for all $e\in H$. By definition of $H$ and \Cref{def:kappaplus}, for all edges $e\in H$, the corresponding original edge in $G$ is in $E_{L}(G,\kappa)$. We want to bound the following probability. \begin{align*} \prob{Y\geqslant \frac{\kappa^+_{\textsc{bc}}(C,W_2\setminus D)}{2^{\nicefrac{16}{\varepsilon^2}}}+4\cdot\varepsilon\mu(G)} \end{align*} Applying \Cref{lem:chernoff} with $\delta=\frac{4\cdot\varepsilon\mu(G)\cdot 2^{\nicefrac{16}{\varepsilon^2}}}{\kappa^+_{\textsc{bc}}(C,W_2\setminus D)}$, we have, \begin{align*} \prob{Y\geqslant \frac{\kappa^+_{\textsc{bc}}(C,W_2\setminus D)}{2^{\nicefrac{16}{\varepsilon^2}}}+4\cdot\varepsilon\mu(G)}&=\exp\paren{\varepsilon\mu(G)-\varepsilon\mu(G)\log\paren{1+\frac{4\cdot\varepsilon\mu(G)\cdot 2^{\nicefrac{16}{\varepsilon^2}}}{\kappa^{+}_{\textsc{bc}}(C,W_2\setminus D)}}}\\ &\text{(Since }\kappa^{+}_{\textsc{bc}}(C,W_2\setminus D)\leqslant 4\cdot \mu(G)\text{ as proved below.)} \\ &\leqslant \exp\paren{\varepsilon\mu(G)-\varepsilon\mu(G)\log\paren{1+\varepsilon\cdot 2^{\nicefrac{16}{\varepsilon^2}}}}\\ &\text{(From \Cref{eqn:upper}, we have $\kappa^+_{\textsc{bc}}(C,W_2\setminus D)\leqslant 4\cdot\mu(G)$)}\\ &\leqslant\exp\paren{\varepsilon\mu(G)-\varepsilon\mu(G)\log\paren{1+ 2^{\nicefrac{16}{\varepsilon^2}}}}\\ &\text{(Using the fact that }2^{\nicefrac{1}{\varepsilon^2}}\geqslant \nicefrac{1}{\varepsilon})\\ &=\exp\paren{\varepsilon\mu(G)-\nicefrac{16\mu(G)}{\varepsilon}}. \\ &\text{(Using the fact that }2^{\nicefrac{16}{\varepsilon^2}}\leqslant 2^{\nicefrac{16}{\varepsilon^2}}+1) \end{align*} To see why $\kappa^+_{\textsc{bc}}(C,W_2\setminus D)\leqslant 4\cdot\mu(G)$, consider \Cref{eqn:upper}, \begin{align*} \kappa^+_{\textsc{bc}}(C,W_2\setminus D)&=\mu(H,\kappa^+_{\textsc{bc}})-\card{W_1}+\card{C}-\card{D}\\ &\leqslant 2\cdot (1+\varepsilon)\cdot\mu(G[W]\cap E_{L}(G,\kappa),\kappa^+_{\textsc{bc}})-\card{W_1}+\card{W_1}\\ &\text{(Using \Cref{lem:folklore2}, the fact that }H=\textsc{bc}(G[W]\cap E_{L}(G,\kappa)),\text{ and $\kappa^+$ satisfies the hypothesis.})\\ &\leqslant 4\cdot \mu(G). \end{align*} Taking a union bound over all $W$, which are at most $2^{\nicefrac{16\cdot \mu(G)}{\varepsilon}}$ many, we have our bound (this is because by statement of the lemma, $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$). \end{proof} \subsection{Phase 3: Finding set $E^*$} The algorithm \textsc{M-or-E*}() proceeds to Phase 3 if the matching found in $G_s$ in Phase 1 is small, and hence $\mu(G,\kappa)$ is too small. In this case, we need to find a set $E^*$ satisfying the properties of \Cref{lem:MorE}. In particular, we need to find a set $E^*$ with $\kappa(E^*)=O(\mu(G)\log n)$ such that for every large matching $M$, there are a lot of edges going through $E^*$. In order to do this, we rely on the properties of the dual variables associated with the matching problem. The algorithm \textsc{Static-Match}(), luckily for us, solves both the primal as well as the dual solution. We first begin by stating the Dual Matching Program, and then we state the properties of dual variables guaranteed by \textsc{Static-Match}(). \paragraph*{Dual Matching Program} The dual of the maximum cardinality matching linear program is the following (see \cite{Schrijver2003CombinatorialOP}). \begin{mini*} {}{\sum_{v\in V} y(u)+\sum\limits_{B\subseteq V_{\text{odd}}}\frac{\card{B}-1}{2}z(B)}{}{} \addConstraint{yz(e)}{\geqslant 1, }{}{\quad \quad e\in E(G)} \addConstraint{y(u)}{\geqslant 0}{}{\quad \quad \forall u\in U} \addConstraint{z(B)}{\geqslant 0}{}{\quad \quad \forall B\subseteq V_{\text{odd}}} \end{mini*} where, \begin{enumerate} \item We define $yz((u,v))\coloneqq y(u)+y(v)+\sum\limits_{\substack{B\in V_{\text{odd}},\\ (u,v)\in G[B]}} z(B)$ and, \item We define $f(y,z)\coloneqq \sum_{v\in V} y(u)+\sum_{B\subseteq V_{\text{odd}}}\frac{\card{B}-1}{2}z(B)$. \end{enumerate} We now describe some properties of the \textsc{Static-Match}(). We state them without proof for now, and postpone the full proof to the \Cref{app:statmatch}. \begin{restatable}{lemma2}{statmatch}\label{lem:propstatmatch} There is an $O(\nicefrac{m}{\varepsilon})$ time algorithm \textsc{Static-Match}() that takes as input a simple graph $G$ with $m$ edges and a parameter $\varepsilon>0$, and returns a matching $M$, and dual vectors $\vec{y}$ and $\vec{z}$ that have the following properties. \begin{enumerate}[label=(\alph*)] \item \label{item:a}It returns an integral matching $M$ such that $\card{M}\geqslant (1-\varepsilon)\cdot \mu(G)$. \item \label{item:b} A set $\Omega$ of laminar odd-sized sets, such that $\set{B\mid z(B)>0}\subseteq \Omega$. \item \label{item:c} For all odd-sized $B$ with $\card{B}\geqslant \nicefrac{1}{\varepsilon}+1$, $z(B)=0$. \item \label{item:d} Each $y(v)$ is a multiple of $\varepsilon$ and $z(B)$ is a multiple of $\varepsilon$. \item \label{item:e} For every edge $e\in E$, $yz(e)\geqslant 1-\varepsilon$. We say that such an edge $e$ is \emph{approximately covered} by $\vec{y}$ and $\vec{z}$. \item \label{item:folklore}The value of the dual objective, $f(y,z)$ is at most $(1+\varepsilon)\cdot \mu(G)$. \end{enumerate} \end{restatable} \subsubsection{Main Guarantees of Phase 3} We now state the following helper lemma which will be instrumental in proving one of the two main properties of $E^*$, namely, that every large matching has a lot of edges passing through $E^*$. \begin{lemma2}\label{lem:guaran2} Suppose $G$ is a graph, and let $H\subset G$ be a subgraph of $G$. Let $\vec{y},\vec{z}$ be the dual variables returned on execution of \textsc{Static-Match}() on $H$ and $\varepsilon$. Let $E_{H}=\set{e\in G\mid yz(e)\geqslant 1-\varepsilon}$. Let $M$ be any matching of $G$, then $E\setminus E_{H}$ contains at least $\card{M}-(1+\varepsilon)^2\cdot \mu(H)$ edges of $M$. \end{lemma2} \begin{proof} Suppose we scale up the dual variables $\vec{y}$ and $\vec{z}$ by a factor of $1+\varepsilon$. Then, $\vec{y}$ and $\vec{z}$ is a feasible solution for the dual matching program for the graph $E_{H}$. Thus, using weak duality, we have that $\mu(E_{H})\leqslant (1+\varepsilon)^2\cdot\mu(H)$ (this inequality follows from \Cref{lem:propstatmatch}\ref{item:folklore}). Suppose $M$ is a matching of $G$, and suppose $E\setminus E_{H}$ contains fewer than $\card{M}-(1+\varepsilon)^2\cdot \mu(H)$ edges of $M$, and therefore, fewer than $\card{M}-\mu(E_H)$ edges of $M$, then this implies that $\card{M}=\card{M\cap E_{H}}+\card{M\cap E\setminus E_H}<\mu(E_{H})+\card{M}-\mu(E_{H})=\card{M}$, which is a contradiction. Thus, $E\setminus E_H$ contains at least $\card{M}-\mu(E_H)\geqslant \card{M}-(1+\varepsilon)^2\cdot \mu(H)$ edges of $M$. \end{proof} We now state set $E^*$, and show that it has the property that $\kappa(E^*)=O(\mu(G)\log n)$. \begin{lemma2}\label{lem:propE*} Let $G$ be a graph multi-graph such that $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$, and let $\kappa$ be a capacity function on the edges of the graph. Suppose $G_s$ is the graph obtained by sampling every edge $e\in E$ with probability $p(e)=\kappa(e)\cdot \rho_{\varepsilon}$. Let $\vec{y}, \vec{z}$ be the duals returned by \textsc{Static-Match}($G_s,\varepsilon$). Let $E^*=\set{e\in E\mid yz(e)<1-\varepsilon}$. Then, with high probability, $\kappa(E^*)=O(\mu(G)\log n)$. \end{lemma2} \begin{proof} We consider the set $\mathcal{D}$ of assignments $\vec{y},\vec{z}$ to the vertices and odd components that satisfy the following properties. \begin{enumerate}[label=(\alph*)] \item \label{item:one} For all $v\in V$, $y(v)$ is a multiple of $\varepsilon$, and for all $B\subset V$, $z(B)$ is a multiple of $\varepsilon$. \item Let $\Omega=\set{B\subset V\mid z(B)>0}$, then $\Omega$ is laminar. \item\label{item:three} If $z(B)>0$ for some $B\subseteq V$, then $\card{B}\leqslant \nicefrac{1}{\varepsilon}$. \end{enumerate} Observe that, \begin{align*} \card{\mathcal{D}}&\leqslant \sum_{i=0}^{2n} {n^{\nicefrac{1}{\varepsilon}}\choose i}\cdot \paren{\frac{1}{\varepsilon}}^n\cdot \paren{\frac{1}{\varepsilon}}^{2n}\\ &\leqslant n\cdot {n^{\nicefrac{1}{\varepsilon}}\choose 2n}\cdot \paren{\frac{1}{\varepsilon}}^n\cdot \paren{\frac{1}{\varepsilon}}^{2n}\\ &\leqslant 2^{\paren{\nicefrac{4}{\varepsilon}}\cdot n\log n}\\ &\leqslant 2^{\paren{\nicefrac{16}{\varepsilon^2}}\cdot \mu(G)\log n}\\ &\text{(Since }\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}) \end{align*} This number follows from the following argument. Since $\Omega$ is laminar, it can contain at most $2n$ sets. Moreover, from \ref{item:three}, we deduce that these $2n$ sets are chosen from among $n^{\nicefrac{1}{\varepsilon}}$ sets. Further, from \ref{item:one}, we deduce that every vertex can be assigned at most $\nicefrac{1}{\varepsilon}$ values, and every $B\in \Omega$ can be assigned $\nicefrac{1}{\varepsilon}$ values. Therefore, for a given choice of $\Omega$, there are at most $\paren{\nicefrac{1}{\varepsilon}}^n\cdot \paren{\nicefrac{1}{2\varepsilon}}^{2n}$ choices for $\vec{y}$ and $\vec{z}$. Moreover, the above number also upper bounds the number of possible duals that can be a returned by the algorithm \textsc{Static-Match}(), since the duals in $\mathcal{D}$ satisfy a subset of the properties satisfied by the duals returned by \textsc{Static-Match}(). \\ \\ Now consider a fixed $\vec{y},\vec{z}$ that satisfies the above-mentioned properties, and let $E^*$ be the set of edges that are not approximately covered by $\vec{y},\vec{z}$. Note that for any $e\in E^*$, $\kappa(e)\leqslant \nicefrac{1}{\alpha_{\varepsilon}}$. If not, then, $\kappa(e)=1$, then it would be sampled in $G_s$ with probability $1$, and be approximately covered by $\vec{y},\vec{z}$. Thus, for any $e\in E^*$, $p(e)=\kappa(e)\cdot \rho_{\varepsilon}$. Now, suppose $\kappa(E^*)>17\mu(G)\log n$. Then, observe that if $\vec{y},\vec{z}$ is output by \textsc{Static-Match}() (denote this event by $\mathcal{E}_{y,z}$) then none of the edges in $E^*$ were sampled (denote this event by $\mathcal{E}_2$). Thus, we have, \begin{align*} \prob{\mathcal{E}_{y,z}}\leqslant \prob{\mathcal{E}_2}&\leqslant \prod_{e\in E^*}(1-p(e))\leqslant \exp\paren{-\sum_{e\in E^*}\kappa(e)\cdot \rho_{\varepsilon}}\leqslant \exp\paren{-17\cdot 2^{\nicefrac{1}{\varepsilon^2}}\cdot \mu(G)\log n} \end{align*} Now, observe that if we are able to upper bound $\prob{\bigcup\limits_{\vec{y},\vec{z}\in \mathcal{D}}\mathcal{E}_{y,z}}$, then our lemma follows. This upper bound follows by taking a union bound over the set $\mathcal{D}$. \begin{align*} \prob{\bigcup\limits_{\vec{y},\vec{z}\in \mathcal{D}}\mathcal{E}_{y,z}}&\leqslant \exp\paren{-17\cdot 2^{\nicefrac{1}{\varepsilon^2}}\cdot \mu\log n}\cdot 2^{\paren{\nicefrac{16}{\varepsilon^2}}\cdot \mu\log n}=O\paren{\exp\paren{-2^{\nicefrac{1}{\varepsilon^2}\cdot \mu \log n}}} \end{align*} \end{proof} \section{Algorithm \textsc{M-or-E*}()} \label{sec:more} In this section, we give the main subroutine \textsc{M-or-E*}() (see \Cref{lem:MorE}). We first define a few terms, and state a known result about finding an approximate fractional matching. \begin{definition2} Recall \Cref{def:lowedges}, and consider an integral matching $M$, define $E^M_{L}(G,\kappa)=E_{L}(G,\kappa)\cap M$, and let $V_{M}^{L}$ be the endpoints of $E_{L}^M(G,\kappa)$. \end{definition2} \begin{lemma2}\label{lem:fracmatch} Given a multigraph $G$ (possibly non-bipartite), with edge capacities $\kappa$ and $\varepsilon\in (0,1)$, such that $\kappa(D(e))\leqslant \nicefrac{1}{\alpha_{\varepsilon}}$ for all $e\in E$, then there is an algorithm \textsc{Frac-Match}() that takes as input $G$, $\kappa$ and $\varepsilon$, and returns a fractional matching $\vec{x}$ such that $\sum_{e\in E}x(e)\geqslant (1-\varepsilon)\cdot \mu(G,\kappa)$, obeying the capacities $\kappa$ and the odd set constraints. The runtime of this algorithm is $O(\nicefrac{m\cdot \log n}{\varepsilon})$. \end{lemma2} \begin{proof} For the case of bipartite graphs, there is an algorithm that takes as input a graph $G$, an edge capacity function $\kappa$ and $\varepsilon\in (0,1)$, and returns in $O(\nicefrac{m}{\varepsilon})$ time, a $(1+\varepsilon)$-approximate fractional matching of $G$, obeying capacities $\kappa$ (see \cite{BPT20}). We use this algorithm for a non-bipartite graph as follows: we run the algorithm on $\textsc{bc}(G,\kappa)$, to obtain a matching $\vec{x}$. To get a fractional matching in $G$ that obeys $\kappa$, we do the following: let $e\in E(G)$ and let $e',e''\in E(\textsc{bc}(G))$ be copies of $e$. We let $z(e)=\frac{x(e')+x(e'')}{2}$. Then, the matching $\vec{z}$ obeys $\kappa$ as well as the fractional matching constraints since $\vec{x}$ obeys them. \end{proof} For the purpose of the algorithm recall, definition of $\kappa^+$ in \Cref{def:kappaplus} given $\kappa$ and $\varepsilon$. We now formalize the algorithm \textsc{M-or-E*}(). Recall that it takes as input a multigraph $G$ with $\mu(G)\geqslant \nicefrac{\varepsilon\cdot n}{16}$. \begin{algorithm} \caption{\textsc{M-or-E*}($G,\kappa,\varepsilon,\mu$)} \label{alg:MorE} \begin{algorithmic}[1] \State Include each $e\in E(G)$ independently with probability $p(e)=\kappa(e)\cdot \rho_{\varepsilon}$ into graph $G_s$. \State Let $M$ and $\vec{y},\vec{z}$ be the output of $\textsc{Static-Match}(G_s,\varepsilon)$. \Comment{Phase 1} \If{$\card{M}<\mu-6\varepsilon\mu$}\Comment{Phase 3} \State Return $E^*=\set{e\in E(G)\mid \kappa(e)<1\text{ and }yz(e)<1-\varepsilon}$. \Else \Comment{Phase 2} \State $M_{I}\leftarrow M\setminus E_{L}(G,\kappa)$ \State $\vec{y}\leftarrow M_{I}^D$\Comment{Converting $M_I$ into a matching on a multigraph.} \State\label{line:fracmatch}$\vec{x}\leftarrow \textsc{Frac-Match}(G[V_{L}^M]\cap E_{L}(G,\kappa),\kappa^+,\varepsilon)$ \EndIf \State Return $\vec{y}+ \vec{x}$. \end{algorithmic} \end{algorithm} We now show \Cref{lem:MorE} holds, we first restate it. \lemtrio* \begin{proof}[Proof of \Cref{lem:MorE}] We first show the runtime of the algorithm. Graph $G_s$ can be computed in time $O(m)$. Using \Cref{lem:propstatmatch}, we conclude that we can compute $E^*$ in order $O(\nicefrac{m}{\varepsilon})$ time by running \textsc{Static-Match}($G_s,\varepsilon$) and \Cref{lem:fracmatch} implies that \Cref{line:fracmatch} takes $O(\nicefrac{m\cdot \log n}{\varepsilon})$ time. \\ \\ We show \Cref{lem:MorE}\ref{item:aa}. First observe that $V_{L}^M$ and $V(M_I)$ are disjoint, and since $\vec{y}$ and $\vec{x}$ are fractional matchings, $\vec{z}$ is also a fractional matching. Note that $\card{M}\geqslant \mu-7\varepsilon\mu$. Moreover, $\sum_{e\in E}x(e)\geqslant (1-\varepsilon)\cdot \mu(G[V_L^M]\cap E_{L},\kappa^+)$. This follows from applying \Cref{lem:fracmatch} to $G[V_L^M]\cap E_{L}(G,\kappa)$ with capacity function $\kappa^+$. Recall \Cref{def:kappaplus} and \Cref{def:lowedges} to see that $\kappa^+(D(e))\leqslant \nicefrac{1}{\alpha_{\varepsilon}}$ for $e\in G[V_L^M]\cap E_{L}(G,\kappa)$, thus satisfying the requirements of \Cref{lem:fracmatch}. Next, applying \Cref{cor:sampling}, we have, $\mu(G_{s}[V_{L}^M]\cap E_{L})\leqslant \mu(G[V_{L}^M]\cap E_{L},\kappa^+)+5\cdot \varepsilon\mu(G)$. Thus, we have $\sum_{e\in E}x(e)\geqslant (1-\varepsilon)\cdot \paren{\mu(G_{s}[V_L^M]\cap E_L)-5\varepsilon\mu}\geqslant (1-\varepsilon)\cdot \paren{\card{M\setminus M_I}-5\varepsilon\mu}$. This is because $M\setminus M_I$ is a matching of $G_{s}[V_{L}^M]\cap E_{L}$. So, $\sum_{e\in E}z(e)\geqslant \card{M_I}+(1-\varepsilon)\cdot (\card{M\setminus M_I}-5\varepsilon\mu)\geqslant (1-\varepsilon)\cdot (\mu-12\varepsilon\mu)\geqslant \mu-13\varepsilon\mu$.\\ \\ We now show \Cref{lem:MorE}\ref{item:aa}\ref{item:ai} and \ref{item:aii}. Consider any edge $e\in \text{supp}(\vec{z})$ with $\kappa(D(e))\leqslant \nicefrac{1}{\alpha_{\varepsilon}^2}$, $e\in G[V_{L}^M]\cap E_{L}(G,\kappa)$. Thus, $e\in \text{supp}(\vec{x})$, and therefore, from \Cref{lem:fracmatch}, $z(e)=x(e)\leqslant \kappa^+(e)\leqslant \kappa(e)\cdot \alpha_{\varepsilon}$ and $z(D(e)) =x(D(e))\leqslant \kappa(D(e))\cdot \alpha_{\varepsilon}$. Similarly, for any $e\in \text{supp}(\vec{z})$ with $\kappa(D(e))>\nicefrac{1}{\alpha_{\varepsilon}^2}$, $e\in \text{supp}(\vec{y})$. By definition of $\vec{y}$, $z(e)=y(e)=\nicefrac{\kappa(e)}{\kappa(D(e))}$ (recall \Cref{def:dist}) and $z(D(e))=1$. This proves our claim. \\ \\ Next, we show \Cref{lem:MorE}\ref{item:bi}. First recall from the assumption of \Cref{lem:MorE} that $\mu\geqslant (1-\varepsilon)\cdot \mu(G)$. From this fact and \Cref{lem:propE*}, we can conclude that $\kappa(E^*)=O(\mu\log n)$ and that for all $e\in E^*$, $\kappa(e)<1$. Next, observe that $\mu(G_s)\leqslant (1+\varepsilon)\cdot \card{M}\leqslant (1-6\varepsilon)\cdot \mu$. Applying \Cref{lem:guaran2} with $H=G_s$, we have, that for any matching $M'$ of $G$, $\card{E^*\cap M'}\geqslant \card{M'}-(1+\varepsilon)^2\cdot (1-6\varepsilon)\cdot \mu$. If $\card{M'}\geqslant (1-3\varepsilon)\cdot \mu$, then we have $\card{E^*\cap M'}\geqslant \varepsilon\cdot \mu$. Finally, consider any pair of vertices $u,v$ and let $e',e''\in D(u,v)$. Then, either both $e',e''$ are both approximately covered by $\vec{y},\vec{z}$ or neither of them are. This implies that either $D(u,v)\subseteq E^*$ or $D(u,v)\cap E^*=\emptyset$. \end{proof}
\section{Introduction} Estimating entropy -- that is, the measure of uncertainty \cite{shannon1948mathematical,cover2012elements} -- of a random variable from its samples is often a key question in analysis of complex systems. This estimation from a finite (and often small) set of samples is a hard problem, especially for high dimensional systems, where the number of states that a variable can take quickly overwhelms the number of samples $N$. Then many of the states, hereafter called {\em low probability states}, have probability $<1/N$. Collectively, we refer to all of these states as the {\em tail} of the probability distribution. While there may be a lot of samples in the tail, each low probability state will not be sampled typically, or will be sampled at most once. Because of the tail, the entropy estimator that replaces probabilities of states by their empirical frequencies (the so called {\em naive} or {\em Maximum Likelihood} estimator \cite{strong1998entropy}) has a large sample size dependent bias \cite{paninski2003estimation}. Corrections have been derived to overcome this bias \cite{miller1955note, grassberger2003entropy, berry2013simple}, but these tend to be valid only in the well-sampled regime. Outside of this regime, Bayesian \cite{wolpert1995estimating, nemenman2002entropy, archer2014bayesian} and some non-parametric \cite{chao2003nonparametric, chao2013entropy, cerquetti2019exact} estimators may still result in low bias estimates by imposing {\em a priori} assumptions on the probabilities of the low-probability states. Although these Bayesian and non-parametric estimators perform well on some data sets, it is known that no estimator can be universally unbiased in this regime \cite{paninski2003estimation,Antos01}. Thus it is crucial to understand how these estimators extract information about entropy from data, and hence when they will fail. Unfortunately, such theoretical understanding is missing for many estimators. Ma was the first to point out that estimation of entropy is possible for poorly-sampled uniform distributions by analysing a particular statistics of the data: {\em coincidences} \cite{ma1981calculation}. Nemenman extended the theoretical idea that coincidences determine entropy to non-uniform distributions obeying some Bayesian priors \cite{nemenman2011coincidences}. However, a similar theoretical understanding is still missing in a broader context, and it remains unclear which statistics of data, in addition to the number of coincidences, may contribute to entropy estimation and why. In this paper, we analytically investigate two Bayesian estimators: that of Nemenman, Shafee and Bialek \cite{nemenman2002entropy,nemenman2004entropy} and of Archer and Pillow \cite{archer2014bayesian}. We focus on the regime, which is arguably the most important for real life applications, where the number of states with at least one sample, $K_1$, is similar to the total number of samples, $K_1 \sim N \gg 1$, and yet $K_1<N$, so that there are coincidences in the data. Outside of this regime, the probability distribution is either well-sampled (so that many different methods for entropy estimation would work), or there are no coincidences at all (so that entropy estimation is impossible). In our regime of interest, we show that the result of the estimation by the studied estimators depends on the Maximum Likelihood entropy estimate $S_0$, the number of coincidences, and also on two measures of {\em dispersion of coincidences}. The first of these, $K_2$, is the number of states with at least two samples. The second, which we call $Q_1$, characterizes the spread of coincidences over states with three or more samples. We show that values of these statistics are related to the structure of the tails of the probability distribution that is assumed by the estimators. Specifically, a short, exponential, tail is more likely to be inferred by the estimators when there many coincidences or they are dispersed. If the number of coincidences is intermediate, and the coincidences are concentrated, then the estimators infer a long tail. In between these two regions, a mixed tail dominates. We show that the studied estimators correct Maximum Likelihood, and that the correction is larger when there are fewer coincidences and they are concentrated, which in turn happens with a large exponential tail or a slowly-decaying long tail. This understanding relates the observable data statistics to assumptions that Bayesian estimators make about the underlying probability distributions (see Fig.~\ref{fig1}), and hence provides an intuitive explanation for how these estimators work and, crucially, when they fail. \section{Overview of Bayesian entropy estimation} Given a probability distribution $\{q_x\}=\bm{q}$ for a discrete one-dimensional random variable $X$, its entropy is defined as \cite{shannon1948mathematical} \begin{equation} S(\bm{q})=-\sum_{x} q_x \log q_x. \label{s2e1} \end{equation Note that we use the natural logarithm throughout this paper, and hence entropy is measured in {\em nats}. One is often faced with a problem when $S$ must be estimated for unknown $q_x$ from a set of $N$ samples $\{x_1, \dots, x_N\}$ from the probability distribution. The Maximum Likelihood estimator of entropy, $S_0$, is then defined by replacing the probabilities with frequencies $q_x\to \hat{q}_x= n_x/N$, \begin{equation} S_0=S(\hat{\bm{q}})=-\sum_{x} \frac{n_x}{N} \log \frac{n_x}{N}. \label{MLEstimator} \end{equation} States with zero frequencies in the sample do not contribute to $S_0$ resulting typically in underestimation of the entropy \cite{paninski2003estimation}. In general, because of this low probability tail, estimation of entropy from data is very hard when the number of samples is smaller than the number of effective states of the variable, $N \ll \exp(S)$. \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{fig1_v1c.pdf} \caption{Relation between assumptions about the tail structure and the statistics that determine entropy estimation. The set of unsampled states, $q_i \leq 1/N$, which we refer to as the {\em tail}, may contribute substantially to the entropy. However, the Maximum Likelihood estimation overlooks this contribution. If the rank ordered plot of the tail is exponential with the scale $\alpha$ (top panel), then the tail has effectively $\alpha$ states, which contribute $\delta S\sim \log\alpha$ to the entropy. While the tail cannot be observed directly, it pulls samples from the head of the distribution, so that the number of coincidences, $\Delta$, in the head decreases as $\alpha$ grows. Thus one can estimate $\alpha$ and hence the entropy itself from $\Delta$. Alternatively, if the rank-ordered plot of the tail has a power law structure with the exponent $-1/d$, then the tail does not have a finite effective size (bottom panels). Then its contribution to entropy depends on $d$ as $\delta S \sim (1-d)^{-1}$. In this case, one can estimate $d$, and hence the entropy, from the dispersion of the coincidences, which depends, in part, on how many samples happen once or more, $K_1$, or twice or more, $K_2$, in the dataset.} \label{fig1} \end{figure*} Bayesian estimators address the problem by imposing various {\em a priori} assumptions $p(\bm{q})$. One then uses Bayes theorem to infer the {\em a posteriori} distribution of $\bm{q}$, and finally integrates over $\bm{q}$ to get the {\em a posteriori} distribution or moments of entropy. Specifically, the mean posterior entropy $\hat{S}=\langle S|\bm{n}\rangle$ given the counts $\bm{n}=\{n_x\}$ of how many times state $x$ was sampled is given by \begin{align} \hat{S}&=\langle S|\bm{n}\rangle = \int S(\bm{q})p(S|\bm{q})p(\bm{q}|\bm{n})d\bm{q} \nonumber \\ &= \int S(\bm{q}) \delta\left(S+\sum_{x}q_x \log q_x\right) p(\bm{q}|\bm{n})d\bm{q}, \end{align} where $p(\bm{q}|\bm{n})$ is the posterior over $\bm{q}$ under some prior $p(\bm{q})$, \begin{equation} p(\bm{q}|\bm{n})=\frac{p(\bm{n}|\bm{q})p(\bm{q})}{p(\bm{n})} = \frac{\prod_x q_x^{n_x}p(\bm{q})}{p(\bm{n})}. \label{eq:posterior} \end{equation} For distributions with known finite size $\mathcal{A}$ of the space of the possible outcomes (aka the {\em alphabet size}), the Dirichlet distribution is often chosen as a prior due to its conjugacy with the categorical distribution: \begin{equation} p(\bm{q})= \text{Dirichlet}(\bm{q}|\lambda) \propto \prod_{i=1}^{\mathcal{A}}q_i^\lambda, \end{equation} where $\lambda$ is known as the concentration parameter. Note that any chosen prior $p(\bm{q})$ implicitly imposes assumptions on the structure of the low probability tail (and hence its contribution to the entropy) based on the observed statistics of the well-sampled part of the probability distribution. However, these implicit assumptions usually are not made explicit, and they remain mysterious even for most commonly used Bayesian estimators. Lifting this veil is the goal of this work. \subsection{The Nemenman-Shafee-Bialek (NSB) Estimator} Nemenman et al.\ \cite{nemenman2002entropy} showed that, for variables with the finite alphabet size $\mathcal{A}$, Dirichlet priors on $\bm{q}$ with a fixed value for the concentration parameter $\lambda$ correspond to highly concentrated \textit{a priori} distribution on entropy, which persists for large sample sizes. This bias induces incorrect entropy estimates, which nonetheless have low variance and hence are certain about their outputs. To address this issue, Ref.~\cite{nemenman2002entropy} suggested a Dirichlet-mixture prior \begin{equation} p_{\rm NSB}(\bm{q}) = \int \text{Dirichlet}(q|\lambda)p_{\rm prior}(\lambda)d\lambda, \end{equation} where $p(\lambda)$ are the mixture weights determined by \begin{equation} p_{\rm prior}(\lambda) \propto \partial_\lambda \langle S|\lambda\rangle= \mathcal{A} \psi_1(\mathcal{A}\lambda+1)-\psi_1(\lambda+1), \end{equation} and where $\langle S|\lambda\rangle$ is the {\em a priori} expected entropy under the $\text{Dirichlet}(\bm{q}|\lambda)$ prior, and $\psi_1(\cdot)$ is the tri-gamma function \cite{abramowitz+stegun}. This choice of weights implies a nearly uniform {\em a priori} distribution for the entropy $S$ on the interval $[0,\log \mathcal{A}].$ The resulting entropy estimate is then \begin{align} \hat{S}_{\text{NSB}} &= \langle S|\bm{n} \rangle = \int \int S(\bm{q})p(\bm{q}|\bm{n},\lambda)p(\lambda|\bm{n})d\bm{q}d\lambda \nonumber \\ &= \int \langle S|\bm{n},\lambda\rangle \frac{p(\bm{n}|\lambda)p_{\rm prior}(\lambda)}{p(\bm{n})}d\lambda. \end{align} Here $ \langle S|\bm{n},\lambda\rangle$ is the posterior mean entropy under the prior $\text{Dirichlet}(\bm{q}|\lambda)$, and $p(\bm{n}|\lambda)$ is the evidence (which has a Polya distribution) \cite{minka00}, \begin{align} p(\bm{n}|\lambda) &= \int p(\bm{n}|\bm{q})p(\bm{q}|\lambda)d\bm{q} \nonumber \\ &= \frac{N! \Gamma(\mathcal{A}\lambda)}{\Gamma(\lambda)^{\mathcal{A}}\Gamma(N+\mathcal{A}\lambda)} \prod_{i=1}^{\mathcal{A}} \frac{\Gamma(n_i+\lambda)}{n_i!} \end{align} where $\Gamma(\cdot)$ is the gamma function \cite{abramowitz+stegun}. Using the analytical expressions for the first two moments of posterior mean entropy $ \langle S|\bm{n},\lambda\rangle$ (available from Refs.~\cite{wolpert1995estimating,nemenman2002entropy}), one then uses one-dimensional numerical integration over $\lambda$ to obtain $\hat{S}_{\text{NSB}}$. \subsection{The Dirichlet and the Pitman-Yor Processes} \label{DP-PYP} When the size of the state space is unknown or infinite, the standard NSB construction does not work. Then one commonly uses one of the following two stochastic processes to construct a prior $p(\bm{q})$ over a countably infinite state space: the Pitman-Yor Process (PYP) \cite{pitman1997two} and its special case, the Dirichlet Process (DP) \cite{Ferguson73}. To specify these processes, one requires two inputs: a parameter vector and a base distribution. Parameters of the Pitman-Yor process are known as the discount parameter $d$, $0\le d < 1$, and the concentration parameter $\alpha$. The parameters control the shape of typical distributions generated by the process. Specifically, $d$ controls the structure of the low probability tail of $\bm{q}$, so that the tail typically decays as $q_x\propto x^{-1/d}$. The concentration parameter $\alpha$ control the probability mass near the head of the distribution. In the limit $d\rightarrow 0$, $\text{PYP}(d,\alpha)$ becomes the Dirichlet Process, $\text{DP}(\alpha)$. In other words, the Dirichlet Process generates distributions with short tails. When the base distribution is the Beta distribution, one draws samples $q_x\sim \text{PYP}(d,\alpha)$ via the so called {\em stick-breaking process} \cite{Ishwaran01}, which uses an infinite sequence of independent Beta-distributed random variables $\beta_x \sim \text{Beta}(1-d,\alpha+ x d)$, so that \begin{equation} \Tilde{q}_x = \beta_x\prod_{y=1}^{x-1}(1-\beta_y). \end{equation} Thus obtained $\Tilde{\bm{q}}$ are not strictly decreasing with $x$, and so one obtains a strictly non-increasing distribution $\bm{q}$ from them by rank ordering. \subsection{Expectations over DP and PYP Posteriors} Previous studies \cite{Ishwaran03} showed that PYP priors (for multinomial observations) yield a posterior $p(\bm{q}|\bm{n},\alpha , d)$, which consists of two parts: probability of $K_1$ states that exist in the sample with the counts of, at least, one, and probability of states that are not sampled. We will denote the set of states with nonzero counts as $\mathbb{K}$, and its cardinality is $K_1=||\mathbb{K}||$. Then the first term of the posterior is given by the Dirichlet distribution, $p(\bm{q}\in\mathbb{K}|\bm{\mu})\propto \prod_x q_x^{\mu_x}$, where $\bm{\mu}$ is a concentration vector $\bm{\mu}=(n_1-d,\cdots,n_{K_1}-d,\alpha+K_1 d)$. This leaves the probability of $q_* =1-\sum_{x\in\mathbb{K}}q_x$ for the unobserved states. In other words, the states with nonzero counts contribute the following to the posterior: \begin{align} p(\bm{q}\in \mathbb{K}|\bm{n})&=p(q_1,\cdots,q_{K_1},q_*|\bm{n}) \nonumber \\ &= \text{Dirichlet}(n_1-d,\cdots,n_{K_1}-d,\alpha+K_1 d) \nonumber \\ &\propto q_*^{\alpha+K_1d}\prod_{i=1}^{K_1} q_i^{n_i-d}. \label{DirDistribution} \end{align} For the states that have no samples, the posterior is equal to the prior. Thus their contribution to the posterior is the Pitman-Yor Process, normalized by their total probability being $q^*$: \begin{equation} p(\bm{q}\not\in \mathbb{K})= p(q_{K_1+1},q_{K_1+2},\cdots)= q^* \text{PYP}(d,\alpha+K_1 d). \end{equation} Overall, this yields a closed form solution for the posterior mean and variance of the entropy $S$. Specifically, the resulting posterior mean $\langle S|\bm{n},\alpha , d\rangle$ is \begin{equation} \begin{split} \langle S|\bm{n},\alpha , d\rangle = \psi(\alpha+N+1)-\frac{\alpha+K_1d}{\alpha+N}\psi(1-d) \\ -\frac{1}{\alpha+N}\left(\sum_{x=1}^{K_1} (n_x-d)\psi(n_x-d+1)\right), \end{split} \label{PYPEntropy} \end{equation} where $\psi(x)=\partial_{x}\log\Gamma(x)$ is the di-gamma function \cite{abramowitz+stegun}. Unfortunately, this is usually not a good estimate of entropy since, for fixed $\alpha$ and $d$, the prior $\text{PYP}(d, \alpha)$ on $\bm{q}$ corresponds to a highly concentrated {\em a priori} distribution on entropy, just like was noted before in the context of the NSB estimator. To counter this, Archer and Pillow \cite{archer2014bayesian} followed the NSB prescription and introduced a prior (mixture) over the parameters of $PYP(d,\alpha )$, $p_{\rm prior}(\alpha , d)$, which uniformized the induced prior over entropy (with the caveat that, for a distribution on a countable alphabet, the entropy may be infinite, and hence strict uniform distribution over entropy is impossible). Specifically, they used \begin{align} &p_{\rm prior}(\alpha , d) = p(\gamma) = e^{-10/(1-\gamma)}, \quad \quad {\rm where}\\ &\gamma= (\psi(1)-\psi(1-d))/(\psi(\alpha+1)-\psi(1-d)) \label{gamma}, \end{align} and then they confirmed numerically that this choice of the prior leads to good estimates of entropy for various test data sets. In other words, they proposed a new estimate of entropy, the Pitman-Yor Mixture (PYM): \begin{align} \hat{S}_{PYM} &= \langle S|\bm{n}\rangle = \int \langle S|\bm{n},\alpha , d\rangle p_{\rm posterior}(\alpha , d|\bm{n})d(\alpha , d) \nonumber \\ &= \int \langle S|\bm{n},\alpha , d\rangle \frac{p(\bm{n}|\alpha , d)p_{\rm prior}(\alpha , d)}{p(\bm{n})}d(\alpha , d), \label{estimator} \end{align} where $\langle S|\bm{n},\alpha , d\rangle$ is given in Eq.~(\ref{PYPEntropy}). The evidence $p(\bm{n}|\alpha , d)$ is then given by (see Ref.~\cite{archer2014bayesian} for a detailed derivation) \begin{equation} p(\bm{n}|\alpha , d) = \frac{\Gamma(1+\alpha)\prod_{l=1}^{K_1} (\alpha+l d)\prod_{x=1}^{K_1}\Gamma(n_x-d)}{\Gamma(1-d)^{K_1} \Gamma(\alpha+N)}. \label{evidence} \end{equation} Note that taking $d\to 0$ in Eqs.~(\ref{estimator} and \ref{evidence}) and making the identification $\alpha = \mathcal{A}\lambda$ in the limits $\lambda\rightarrow 0$ and $\mathcal{A}\rightarrow \infty$ such that $\alpha$ is finite, result in a countably-infinite analogue of the NSB estimator. \section{Determining data statistics that define entropy estimates} In the section, we approximate the likelihood function of the Pitman-Yor process, Eq.~(\ref{evidence}), analytically in terms of coincidence-based data statistics. We then numerically show that the resulting analytical entropy estimates are close to the exact Pitman-Yor Mixture estimator. We focus on the regime where the Maximum Likelihood entropy estimator fails dramatically. For this, we study random variables with many accessible states in the regime where the number of unique samples, $K_1$, is of the order of the total sample size $N$. This regime corresponds to $K_1 \lesssim N\le\exp(S)$, where $N$ is the number of samples and $S$ is the true entropy. We start by considering the log-likelihood function, which is the logarithm of the evidence $p(\bm{n}|\alpha , d)$ in Eq.~(\ref{evidence}): \begin{multline} \mathcal{L}({\bf n}|\alpha , d)=\log\Gamma(1+\alpha)-\log\Gamma(N+\alpha)+\log\Gamma\left(\frac{\alpha}{d}+K_1\right)\\ -\log\Gamma\left(\frac{\alpha}{d}+1\right) +\sum_{i=1}^{K_1}\log\Gamma(n_i-d)-K_1 \log\Gamma(1-d). \label{logLikelihood} \end{multline} We now define $K_m$ as the number of states with at least $m$ counts in the total sample of size $N$, $K_m=\sum_{n_i \ge m}1$. We denote by $m_f$ the largest occupancy of any state in the sample. Further, we define $\mathcal{K}$ as the vector, whose $m$th element is $K_m$. We note that, for any function $f(n)$, \begin{equation} \sum_i f(n_i) = \sum_m (K_m-K_{m+1}) f(m). \label{identity} \end{equation} Thus, in particular, the log-likelihood $\mathcal{L}({\bf n}|\alpha , d)$ can be viewed as $\mathcal{L}({\mathcal{K}}|\alpha , d)$. With this, we can expand Eq.~(\ref{logLikelihood}) around $d=0$ to get (see Appendix \ref{A1:approxL} for details): \begin{multline} \mathcal{L}(\bm{n}|\alpha , d)\approx \mathcal{L}_a(\mathcal{K}|\alpha , d)\equiv \log\Gamma(1+\alpha)\\-\log\Gamma(N+\alpha)+\log\Gamma\left(\frac{\alpha}{d}+K_1\right)-\log\Gamma\left(\frac{\alpha}{d}+1\right)\\ +(K_1-1)\log d+K_2\log(1-d)-Q_1 d+\mathcal{O}(d^2), \label{eq02:L} \end{multline} where \begin{equation} Q_1=\sum_{m=3}^{m_f}\frac{K_m}{m-1}, \end{equation} and the subscript $a$ denotes the $d\to0$ asymptotic nature of the expression. \begin{figure*}[!t] \includegraphics[width=\textwidth]{fig2_v12c.pdf} \caption{Comparison between PYM and related estimators and their approximations for distributions with different tails. The upper panels ({\bf a}-{\bf c}) show the distributions, whose entropy is being estimated. The lower panels ({\bf d}-{\bf f}) show the corresponding entropy estimates as a function of the number of samples, averaged over ten sets of samples. The full estimators, PYM and NSB (with a large alphabet size $\mathcal{A} = 20K_1$), almost overlap with our approximations, aPYM and aNSB. In all panels, we show results for Maximum Likelihood (black), NSB (blue), aNSB (dashed blue), PYM (orange), aPYM (dashed orange), and $\hat{S}_{\text{long}}$ (green) estimators. The dashed gray line represents the true value of entropy for each of the studied distributions.} \label{fig2} \end{figure*} By rewriting the Maximum Likelihood estimate $S_0$ of Eq.~(\ref{MLEstimator}) in terms of coincidences (see Appendix~\ref{A3:approxS}), using the identity Eq.~(\ref{identity}), and approximating certain terms that are finite in the limit $d\rightarrow 1$ via a Taylor expansion around $d\ll 1$, the mean posterior entropy, Eq.~(\ref{PYPEntropy}), results in (see Appendix~\ref{A2:approxS}): \begin{multline} \langle S|{\bf n},\alpha , d \rangle \approx \langle S|{\mathcal{K}}, \alpha , d \rangle_a \equiv \psi(N+\alpha+1)\\- \left(\frac{\alpha+K_1}{\alpha+N}\right)\psi(1-d) +\frac{1}{\alpha+N}\Bigg[ .N(S_{0}-\log N)-K_1 \\ +K_2(\log 4-1-\psi(2-d))+Q_1 d \\ +\mathcal{O}\left(d^2,\sum_{m=3} \frac{K_m}{(m-1)^2}\right) \Bigg], \label{eq03:S} \end{multline} where $\mathcal{O}(d^2,\sum_{m=3}K_m/m^2)$ means that we kept terms that are at most linear in $d$ and at most proportional to $\sum_{m=3} \frac{K_m}{(m-1)}$. Interestingly, within this approximation, the log-likelihood and the posterior mean entropy depend on the sample size $N$, the Maximum Likelihood entropy estimate $S_0$, and the three characteristics of the coincidence vector: $K_1,K_2$ and $Q_1$. The final step in approximating the estimator $\hat{S}_{PYM}$, Eq.~(\ref{estimator}), is to integrate the expected entropy for fixed hyper-parameters $\langle S|{\mathcal{K}},\alpha,d\rangle_a$ over the posterior $p_{\rm posterior}(\alpha,d|{\bf n})\propto p({\bf n}|\alpha,d)p_{\rm prior}(\alpha,d)$ to form the Pitman-Yor mixture. Then the variance of the resulting estimator is dominated by the contribution from the uncertainty in the posterior distribution of the parameters $\alpha,d$, which is about $80\%$ of the total variance in our simulations. This procedure of replacing $\langle S|\bm{n},\alpha,d\rangle$ with the asymptotic expression $\langle S|{\mathcal{K}},\alpha,d\rangle_a$ in Eq.~(\ref{estimator}) leads to a new estimator of entropy, which we call {\em approximate PYM} estimator, or aPYM. This estimator is fully determined by just few data statistics, $N$, $S_0$, $K_1$, $K_2$, and $Q_1$. There are also two limiting cases of this estimator. First, by taking $d\to0$ in Eqs.~(\ref{eq02:L}, \ref{eq03:S}), we define the approximate version of the NSB limit of the PYM estimator on a countably infinite number of possible outcomes, which we denote as aNSB. At the other extreme, taking $\alpha\to0$ in Eqs.~(\ref{eq02:L}, \ref{eq03:S}), corresponds to a prior that favors distributions with long tails. We denote the corresponding estimator as $\hat{S}_{\text{long}}$. \begin{figure*}[!t] \includegraphics[width=\textwidth]{fig3_v5.pdf} \caption{{\bf a}: Phase diagram of the dominant tail hypothesis selected by the PYM estimator as a function of various statistics of the data sample. The explored statistics are the fraction of coincidences in the sample, $\Delta/N$, and dispersion of the coincidences, $K_2/K_1$. This diagram is evaluated at the third crucial data statistics set at $Q_1=0.3\,Q_{\text{max}}= 0.3 (\Delta-K_2)/2$. {\bf b}: Schematic diagram that illustrates how sample sets with different $\Delta$, $K_1$, and $K_2$ may look like. An empty or gray circle above a state $x_i$ represent a single sample for that state. Gray circles denote coincidences. } \label{fig3} \end{figure*} The above observation that, in the undersampled regime where $\exp(S/2)<N<\exp(S)$, the PYM entropy estimator and its relatives are determined approximately by just few statistics of the data, $\{N,S_0,K_1,K_2,Q_1\}$, is the main result of our paper. To corroborate this, we explore the quality of the approximation numerically for different distributions $\bm{q}$. Figure~\ref{fig2} presents results for three distributions with different structures of tails, generated from the Pitman-Yor Process: a distribution with an exponential tail (Fig.~\ref{fig2}{\bf a}, ${\rm PYP}(d=0,\alpha=400)={\rm DP}(400)$), one with a mixed tail (Fig.~\ref{fig2}{\bf b}: ${\rm PYP}(d=0.4,\alpha=100)$), and one with a long tail (Fig.~\ref{fig2}{\bf c}: ${\rm PYP}(d=0.6,\alpha=0)$). In the lower panels we show the results of estimating entropy for different dataset sizes using the ML estimator, the PYM estimator, the NSB estimator with a large alphabet size $\mathcal{A} = 20K_1$, and the three approximations: aPYM, aNSB, and $\hat{S}_{\text{long}}$. All results are averaged over ten sets of random samples. In all cases, the differences between NSB and aNSB on the one hand, and PYM and aPYM on the other are negligible, supporting the accuracy of the approximation. All four of these estimators produce high quality estimates for all sample sizes. Further, we also checked that the approximation of the posterior error of the estimators is close to that of the full versions (not shown). In contrast, $\hat{S}_{\text{long}}$ only performs well when the distribution has a long tail, and the Maximum Likelihood never works well. \section{Tail-hypothesis and entropy estimation phase diagrams} \label{phasediagram} The above discussion shows that the PYM estimator and its relatives work by first estimating the most likely $\alpha$ and $d$ from the sampled data, and then using these estimated parameters to approximate the structure of the low probability tail (from short, to long) and hence of its contributions to the entropy. We further showed that, in the regime of interest, the log-likelihood of $\alpha$ and $d$ is dominated by just few statistics: $N$, $S_0$, $K_1$, $K_2$, and $Q_1$. It is thus illustrative to understand, which combinations of these statistics select which hypothesis on the structure of the tail. Building the corresponding phase diagram of the selected tail structure as a function of the data statistics is the goal of this Section. We will consider three classes of tails: exponential ($d=0$ selected, denoted as hypothesis $H=1$), long tail ($\alpha=0$ selected, denoted as hypothesis $H=2$), and a mixed tails (arbitrary $\alpha$ and $d$, denoted as $H=3$). Our goal is then to evaluate which of the three tail hypotheses has a higher probability given the data. Long and short tail hypotheses have one parameter each, while the mixed tail hypothesis has two parameters and contains the other two hypotheses as special cases. Thus when evaluating the log-likelihoods of each of the hypotheses, we must penalize them for having a different number of parameters, which we do using Bayesian Information Criterion \cite{Schwarz78}. To do this, we evaluate the likelihoods \begin{equation} \mathcal{L}_H = \log p({\mathcal{K}}|\hat \alpha,\hat d)+\log p_{\rm prior}(\hat \alpha,\hat d)-\frac{n_H}{2}\log N, \label{eq03:aL} \end{equation} where $\hat\alpha$ and $\hat d$ are the maximum likelihood values of the parameters within each hypothesis, and $n_H$ is the number of parameters for the hypothesis ($n_H=2$ for $H=3$, and $n_H=1$ otherwise). We remind the reader that, by construction, $\hat\alpha = 0$ for the long tail hypothesis, $H=2$, and $\hat d =0$ for the short tailed hypothesis, $H=1$. We determine the regions of the $N,S_0,K_1,K_2,Q_1$ space, where one of the three $\mathcal{L}_H$ dominates, and plot the slice of this phase diagram in Fig.~\ref{fig3}. Specifically, in the Figure, we vary the total number of {\em coincidences}, $\Delta=N-K_1$, and the number of {\em states with coincidences}, that is, the number of states with more than two counts, $K_2$. By sampling many distributions, we empirically observe that the value $Q_1\sim 0.6 (\Delta-K_2)/2$ is when the rest of the $\Delta-K_2$ counts are uniformly dispersed, and $Q_1$ tends to zero when the rest of the counts are concentrated in a single state. Note that the maximum value $Q_1$ can take is $Q_{\text{max}} = \frac{\Delta-K_2}{2}$. For this reason, we choose the intermediate representative value $Q_1=0.3 Q_{\max}=0.3\frac{\Delta-K_2}{2}$. To simplify the presentation, we plot the winning tail hypothesis as a function of $\Delta/N$ and $K_2/K_1$. Normalized in this way, the diagram is constrained to a square of size 1, as $0\le \Delta/N, K_2/K_1\le 1$. In addition, $K_2\le \Delta$, which means that the upper left corner is not accessible. The ratio $\Delta/N$ determines how common are the coincidences, and the ratio $K_2/K_1$ describes whether the coincidences in the data are concentrates in a few states, or dispersed over many states (see Figure~\ref{fig3}b). Figure~\ref{fig3}{\bf a} show that the exponential tail hypothesis dominates when there are many coincidences, $\Delta/N\sim 1$, or when the coincidences are dispersed, that is $K_2/K_1 \sim 1$ or $K_2/\Delta \sim 1$. Both cases can be explained as corresponding to distributions that are relatively uniform on some fixed number of states, and have zero probability elsewhere. A long tail only dominates when the fraction of coincidences has an intermediate value, but the coincidences are highly concentrated, $K_2/K_1\ll 1$. In other words, in this case, there are dominant states, but a lot of samples still fall outside of them. For other values of $\Delta/N$ and $K_2/K_1$, the mixed tail hypothesis dominates. Equipped with this picture of which tail hypothesis is selected by the PYM estimator as a function of data statistics, we now can calculate how the estimator corrects the ML entropy value $S_0$ for different data statistics. Integrating the mean posterior entropy $\langle S|\mathcal{K},\alpha,d \rangle_a$, Eq.~(\ref{eq03:S}), over our approximation of the posterior, $p_a (\alpha,d|{\mathcal{K}})$, which we obtain by exponentiating Eq.~(\ref{eq02:L}), we get the approximate PYM estimator $\hat{S}_{PYM,a}$. The Maximum Likelihood estimate $S_0$ enters linearly in the posterior mean entropy, Eq.~(\ref{eq03:S}). Thus we write \begin{equation} \langle S|\mathcal{K},\alpha,d \rangle_a= b_{\alpha,d} \,S_0 +\delta S_{\alpha,d}, \end{equation} where $b_{\alpha,d}$ and $\delta S_{\alpha,d}$ can be read off from Eq.~(\ref{eq03:S}). Performing the integral over the approximate posterior, this becomes: \begin{equation} \hat{S}=\delta S+b \,S_0, \label{S-hat} \end{equation} where $\delta S$ and $b$ are averages of the corresponding $\alpha$- and $d$-dependent quantities. Thus {\em independent} of the Maximum Likelihood entropy value, within our approximation, the PYM estimator obtains the entropy estimate by decreasing the ML contribution from the well-sampled head of the distribution and adding an offset that comes from the low probability tail. This is similar to so-called partition-based entropy estimators, \cite{nemenman2004entropy, chao2013entropy,Srivastava17,Nemenman15}, which divide the state space into sub-spaces, estimate entropy in each sub-space, and then add the estimates weighted by the probability of being in a corresponding sub-space. However, here this partitioning arises naturally from the Bayesian framework within our approximations. \begin{figure*}[!t] \includegraphics[width=\textwidth]{fig4_v6b.pdf} \caption{Corrections to entropy estimation as a function of determining data statistics. We break down the final estimation for entropy in two parts, as $\hat{S}= \delta S(\sfrac{\Delta}{N},\sfrac{K_2}{K_1})+b(\sfrac{\Delta}{N},\sfrac{K_2}{K_1})\,S_0$, where $\delta S$ is the additive correction and $b$ is scaling factor or weight for the Maximum Likelihood estimate. Well-sampled distributions are located in the upper-right corner where $\delta S=0$ and $b=1$. As in the previous plots, we leave $Q_1=0.3\,(\Delta-K_2)/2$. {\bf a}: Additive correction to entropy. {\bf b}: Scaling correction to entropy.} \label{fig4} \end{figure*} Both the scale factor and the offset depend on the dominant $\alpha$ and $d$ contributing to the estimator, and hence on the usual statistics of the data, $\Delta$, $K_1$, $K_2$, and $Q_1$. Specifically, we numerically observe that the value of $b$ obtained from Eq.~(\ref{S-hat}) satisfies \begin{equation} b=\langle N/(\alpha+N)\rangle\le 1, \end{equation} where the average is over the product of the approximate posterior obtained by exponentiating Eq.~(\ref{eq02:L}) and the prior $p(\gamma) = e^{-7\gamma/100}$ with $\gamma$ defined in Eq.~\ref{gamma}. Note that $\alpha$ is a measure of how much probability is concentrated in the tail. Thus the ratio $N/(\alpha+N)$ approximates the overall weight of the the well-sampled head of the distribution, requiring to decrease the contribution to the entropy from the head by this factor. This matches our assertion that the aPYM estimator is a partition-based estimator, separating the head from the tail. In Figure~\ref{fig4} we show results of numerical estimation of the offset $\delta S$ and the scaling factor $b$ as a function of the fraction of coincidences, $\Delta/N$, and the dispersion of coincidences, $K_2/K_1$. As in the previous case, we keep $Q_1=0.3Q_{\rm max}$. We also set $N=10^4$. Figure~\ref{fig4}(a) shows that the additive term grows when the fraction of coincidences $\Delta/N$ decreases, and when $K_2/K_1$ is small, so that coincidences are concentrated. Both of these cases correspond to a lot of mass in the tail (see corresponding long tail region in Figure~\ref{fig3}(a). The largest values of $\delta S$ occur along the boundary strip $(\Delta/N,K_2/K_1\ll 1)$ and the boundary $K_2=\Delta$. Panel {\bf b} shows that the scaling factor $b$ is close to 1 in most areas, except near the boundary edge $K_2=\Delta$. Along this boundary, the scaling factor becomes the largest when the number of coincidences decreases, $\Delta/N\ll1$. Figure~\ref{fig4} clearly highlights when Bayesian corrections to the ML estimation of entropy are essential: regions of few and concentrated coincidences. \section{Discussion} The major finding of this work is an excellent approximation for the PYM estimator, one of the best Bayesian entropy estimators, and its various relatives (such as NSB). The approximation simplifies the numerics considerably. Crucially, the approximation also shows that the output of the PYM entropy estimator depends on just a few statistics of the data, namely the maximum likelihood (ML) entropy estimate, the fraction of coincidences $\Delta/N$, and the dispersion of coincidences $K_1/K_2$, and $Q_1$. We showed that that workflow of the estimator can be interpreted as first estimating the parameters $d$ and $\alpha$ based on the aforementioned statistics, and with them the tail structure and the total weight of the tail. Then the estimator rescales the ML entropy estimate by the weight of the well-sampled head of the distribution, and adds to it the estimated entropy of the tail. The phase diagrams of which tail structure the estimator selects, Fig.~\ref{fig3}, and how it corrects the ML estimate, Fig.~\ref{fig4}, illustrate these points. Early work of Ma \cite{ma1981calculation} showed that when states are equiprobable, in the under-sampled regime, the coincidences in counts can help with the inference of the entropy of a system. Later Nemenman~\cite{nemenman2011coincidences} showed that in the severely under-sampled regime ($K_1$ close to $N$), entropy estimation depends on the number of coincidences $K_1$. Further, he pointed out how reliable entropy estimates may be obtained by partitioning the overall state space of the variable into sub-spaces with similar sampling properties \cite{Nemenman15}. Here we extend these results to the whole regime where entropy estimation is challenging for multinomial observations, $\exp(S/2)<N <\exp(S)$, by approximating the more general PYM estimator. Our identification of the small set of statistics, which define the output of the estimator, lifts the veil from its inner workings, allowing for a simple, semi-analytical estimation procedure. In particular, this allows us to predict if a particular estimator will be biased simply by looking at the values of the select statistics of the data. How to match {\em a priori} assumptions about the underlying distributions to the data to allow for an unbiased estimation of quantities of interest---such as entropy \cite{nemenman2004entropy, archer2014bayesian} or the mutual information \cite{hernandez2019estimating}--- is an open problem \cite{hernandez2022inferring}. It requires understanding the relation between the {\em a priori} assumptions and the data features that influence the inference. In this work, we build such a link for entropy estimation, and we hope that similar links might exist for other difficult estimation problems. \begin{acknowledgments} This work was funded, in part, by the Simons Investigator award (IN), the Simons-Emory International Consortium (AR and IN), NSF Grant 1822677 (DGH and IN), NIH Grant 2R01NS084844 (AR, IN), the International Physics of Living Systems Network (NSF Grant 1806833, AR). IN would like to acknowledge hospitality of the Aspen Center for Physics, funded in part by NSF Grant 1607611. \end{acknowledgments}
\section{Related Work}\label{sec:related-work} \input{sections/prior.tex} \section{Polarized Spiral Point Spread Function}\label{sec:proposed-method} \input{sections/methods.tex} \section{Simulations}\label{sec:simulations} \input{sections/simulations.tex} \section{Experimental Results}\label{sec:experiments} \input{sections/experiments.tex} \section{Conclusion}\label{sec:conclusions} \input{sections/conclusions.tex} \ifpeerreview \else \section*{Acknowledgments} The authors would like to thank Dong Yan for his help in preparing the fluorescent bead and fluorescent strands samples. This work was supported by NSF awards IIS-1730574, IIS-1730147, CCF-1652569, IIS-1652633, and EEC-1648451. A.-K.G. acknowledges partial financial support from the National Institute of General Medical Sciences of the National Institutes of Health (Grant No. R00GM134187), the Welch Foundation (Grant No. C-2064-20210327), and startup funds from the Cancer Prevention and Research Institute of Texas (Grant No. RR200025). \fi \bibliographystyle{IEEEtranN} \section{Conclusion and Further Scope}\label{sec:conclusions} We presented PS$^2$F for single shot, monocular depth estimation of extended (linear) structures where we separated the two lobes of the DHPSF into the two orthogonal states of polarization. This separation enabled us to remove ambiguities when estimating depth of line segments by breaking the inherent symmetry of DHPSF with respect to depth. The snapshot capabilities of PS$^2$F will enable faster microscopic imaging, including high resolution light sheet microscopy at real-time rates with fewer captures. We designed and demonstrated a compact physical realization of PS$^2$F with a single, polarization-sensitive camera, and showed that PS$^2$F results in $2\times$ or higher accuracy compared to current state-of-the-art phase masks such as DHPSF and the Tetrapod PSF. We believe that our approach of leveraging polarization multiplexed phase masks combined with polarization-sensitive cameras opens new avenues for high resolution snapshot depth imaging at micro and macro scales. PS$^2$F is inherently designed for extended/linear structures, and imaging such structures will involve high photon counts than when imaging single molecules or point-like emitters. Hence, we do not specifically deal with a low-photon count scenario in this work. Nevertheless, PS$^2$F can be adapted to low-photon count scenarios by using advanced recovery techniques including deep network-based approaches. Improvements such as modeling spatially varying PSF, as well as accounting for polarized input light will further improve PS$^2$F results. \section{Experimental Results}\label{sec:experiments} \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-experiments/fig11_v2.pdf} \caption{\textbf{Lab prototype.} (A) Our lab prototype is a $4f$ system consists of a polarization sensor (FLIR BFS-U3-51S5P-C) equipped with a $50$~mm F-mount lens. The phase mask and polarizers are on an XY stage. An achromat (AC254-050-A) forms the other half of the $4f$ system. We used an additional z-axis motorized translation stage to capture the PSF stack. (B) Two orthogonal polarizer pieces. (C) Images of the polarizers and phase mask with a camera focused on the pupil-plane.} \label{fig:4f-setup} \end{figure} We perform real-world imaging with PS$^2$F to verify our simulations. In this section we detail out the experiments and compare the PS$^2$F with the corresponding DHPSF. \subsection{Imaging setup description} We built a $4f$ imaging setup with the first lens being an achromatic lens with focal length $50$~mm, and the second lens being a Canon camera lens with effective focal length $=50$~mm. The polarizer-phase mask is added to the pupil plane of the $4f$ system. Fig. \ref{fig:4f-setup} shows the prototype. \subsubsection{Polarizer-Mask design} We fabricated a $3$~mm diameter DHPSF mask using a 3-D printer employing two-photon lithography. See Supplementary for more details on fabrication. The fabricated mask was the same mask used in the Section \ref{sec:simulations} to perform simulations of the VascuSynth dataset. To create a single-mask, single-sensor design for the PS$^2$F as outlined in Section \ref{sec:proposed-method}, two rectangular pieces of thin-film polarizers were laser cut and put side-by-side in a 3D printed holder. Both the polarizer pieces were laser cut such that their polarization orientation was $0\deg$ and $90\deg$ respectively. The fabricated DHPSF mask was attached to the back side of the 3D printed holder, carefully aligning the two mask halves w.r.t. to the polarizer edges under a bright-field 4x microscope. This polarizer-mask aperture is illustrated in Fig. \ref{fig:4f-setup}(B),(C). Further details can be found in the Supplementary figures. \subsubsection{PSF capture and calibration} To reconstruct 3D scenes, we use an experimentally captured PSF stack for the PS$^2$F. Using the polarization sensor, we image a $5~\mu$m pinhole over a range of [$-2.5$~mm, $+2.5$~mm] on either side of the focus plane. The experimentally-captured PSFs can be seen in Fig.~\ref{fig:psfstack-vis}. \subsection{Imaging experiments} With the fabricated mask and optical setup, we perform a series of experiments with 3D skew planar targets (using transmissive illumination) and with 3D fluorescence targets. We image using a polarization camera sensor (FLIR BFS-U3-51S5P-C) to obtain the individual PS$^2$F lobes in two polarization channels. By averaging across the Bayer pattern, we can readily obtain the image of the scene as imaged by a DHPSF, which we use for comparison purposes. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-experiments/psf_expt_v2.pdf} \caption{\textbf{PSF stacks captured with a lab prototype.} We achieved high quality separation between the two lobes in 0 and 90 degree images with our setup. We noticed no significant effect of any misalignment between the phase mask and the two polarizer halves. Scale bars indicate $100~\mu$m.} \label{fig:psfstack-vis} \end{figure} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/usaf-expt-compare3-v2.pdf} \caption{\textbf{Skew USAF target reconstruction.} Showing reconstructed depth map, YZ maximum intensity projection, zoom-in inset of highlighted ROI for (A) PS$^2$F, and for (B) DHPSF. Column (C) shows slice plots for marked yellow lines in the zoom-in insets, corresponding to Group (5,3) elements for PS$^2$F (top plot) and DHPSF (bottom plot). We were able to resolve up to Group (5,3) elements (12.40 $\mu$m linewidth) using the PS$^2$F, whereas only up to Group (4,5) elements (19.69 $\mu$m linewidth) using the DHPSF. Text below YZ MIP plots indicate the depth RMS error obtained for reconstruction when compared to ground truth, showing that the PS$^2$F performs $\sim2\times$ better. All scale bars indicate $1$~mm.} \label{fig:expt-usaf} \end{figure*} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/beads-expt-compare2-v2.pdf} \caption{\textbf{3D reconstructed fluorescent bead volumes.} The scatter plots show 3D reconstructed views with PS$^2$F and the DHPSF for bead densities approximately corresponding to (A) 3 beads/mm$^3$, (B) 5 beads/mm$^3$, and (C) 8 beads/mm$^3$. We captured ground truth information with a semi-aligned confocal ground truth (hence an approximate ground truth). Highlighted regions indicate various errors: incorrect estimation by DHPSF \textcolor{mygreen}{(solid green)}, missing estimation by DHPSF \textcolor{mygreen}{(dashed green)}, incorrect estimation by PS$^2$F \textcolor{myblue}{(solid blue)}, and incorrect estimation by both PSFs \textcolor{myyellow}{(solid yellow)}. Axes values are in units of $mm$.} \label{fig:expt-fluorescence} \end{figure*} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/vasc-expt-compare1-v2.pdf} \caption{\textbf{Experimental results on fluorescent cotton strands}. (A) PS$^2$F and (B) DHPSF. First column in (A), (B) show the captured scene images (polarization channels in the PS$^2$F image are shown in red and green colors). Second column in (A), (B) shows the reconstructions. Highlight ROIs (in yellow) are shown in the last column of (A), (B) -- depicting the inaccuracy of DHPSF, as well as the ability of PS$^2$F to reconstruct scenes with complex linear structures. In the last row (C), columns (from left-to-right) show color-coded depth maps obtained from confocal imaging (treated as ground truth), PS$^2$F reconstructions, and DHPSF reconstructions respectively. Note the confocal GT depth map is an approximate depth map, due to imperfect registration between GT and the imaged volumes.} \label{fig:expt=vasc} \end{figure*} \subsubsection{3D skew planar targets} We imaged a USAF planar target placed at a skew angle, and then performed single-shot 3D reconstruction of the target. Furthermore, we obtain a focal stack of the target so as to obtain ground truth depth map. The ground truth depth map was obtained from the focal stack using the Shape-from-focus method~\cite{peruz}. For the FOV being imaged, the relative intensity of each of the PS$^2$F lobes changed across the FOV, causing a spatial variance in the PSF. This was readily accounted for by adding a per-pixel, per-channel weight term to the 3D scene variable \textbf{x} in Eq. \ref{eq:optim_func}, and jointly optimizing for the 3D scene \textbf{x} and the weights. The results are depicted in Fig. \ref{fig:expt-usaf}. After modelling of the relative strengths of the PS$^2$F lobes as variables to be estimated during the optimization, we achieved an accurate reconstruction result matching the simulation results in terms of resolving Group (5,3) elements for the PS$^2$F. The YZ maximum intensity projections demonstrate that the skew plane of the USAF has been correctly estimated using the PS$^2$F, as compared to using the DHPSF. When judged against the ground truth, we see a $\sim50\%$ lower RMS error achieved with the PS$^2$F as compared to the DHPSF. \subsubsection{3D fluorescence imaging} We, next, imaged a set of prepared fluorescent samples, consisting of $10~\mu$m fluorescent beads in a PDMS substrate (for more information, see Supplementary). The bead concentration was varied from approximately $3$ to $8$ beads/mm$^3$. The resulting samples were $\sim2$~mm in thickness. We imaged these samples using the same mask as before, but in a $4f$ optical system using lenses of focal length $30$~mm and $50$~mm resulting in a $1.6\times$ magnification. We reconstructed using the same optimization algorithm as specified in Eq. \ref{eq:optim_func}, and compared against ground truth data obtained with a confocal microscope. The results are shown in Fig.\ \ref{fig:expt-fluorescence}. Bead localization performance using the DHPSF seems to get relatively worse as the bead concentration increases. In contrast, PS$^2$F enabled accurate results with increasing bead concentration. It is important to note that in this experiment, we use an optimization objective to perform bead reconstruction. More sophisticated and accurate methods (such as likelihood-based methods) exist for reconstruction of beads or point-like objects. Moreover, the ground truth (GT) data was captured using confocal imaging method. However, due to the complexity of matching/registering imaging volumes, we only obtain an approximated alignment or approximate registration between the GT and the imaging volume captured. Using the same optical setup as above, we also image fluorescent cotton strands suspended in a PDMS sample (see Supplementary for more information about sample preparation). Fig.~\ref{fig:expt=vasc} shows the results for the same. PS$^2$F method is better able to capture the 3D structure of the fluorescent strands - especially in parts with highly complex geometry. Imaging with PS$^2$F results in sharper reconstructions, which helps in reconstructing the fine, complex structure of the strands. The estimated depth map from the PS$^2$F is also more accurate as compared to the DHPSF case. \section{Introduction}\label{sec:introduction} } \IEEEPARstart{3}{D} scanning is crucial to a wide range of applications, including microscopy~\cite{fischer2011microscopy}, autonomous driving~\cite{arnold2019survey}, and robot-assisted surgeries~\cite{reiter2014surgical}. Among the multitude of approaches to measure depth, perhaps the hardest are those that involve passive, monocular, and snapshot measurements---a scenario that prioritizes compactness and time resolution. Estimating 3D information in this context relies on cues such as shading~\cite{ping1994shape} or defocus~\cite{levin2007image, zhou2009coded}; however, the underlying inverse problem is challenging and ill-posed. A key property of real cameras is that the defocus blur changes with the depth of scene points, which has been leveraged by prior work to obtain depth. However, the blur produced by conventional pupils is not conducive to robust depth estimation; to mitigate this, there has been significant interest in the design of engineered pupil plane masks in the form of an amplitude ~\cite{levin2007image}, or phase mask~\cite{pavani2009three,greengard2006depth,pavani2008high, shechtman2014optimal,badieirostami2010three,nehme2020deepstorm3d,nehme2020learning}. In particular, the seminal work of Pavani \textit{et al.} \cite{pavani2008high} has demonstrated that the so-called Double-Helix PSF (DHPSF), consisting of two lobes rotating about a center, can provide high spatial and depth resolutions, at least in the context of super-resolution localization microscopy. The majority of prior work on engineered PSFs focus on isolated point light sources; this is a consequence of their intended application---namely super-resolved localization of fluorescent particles---and the simplification that this assumption provides. However, real-world scenes often consist of more complex geometric primitives such as lines, edges, and curves~\cite{cho2011blur,joshi2008psf}; a PSF optimized for point sources is inadequate at recovering such complex geometries due to ambiguities in depth estimation. This is all the more important in applications involving linear structures including blood vasculature, neuronal networks (of the biological kind), and microtubules. We propose and evaluate polarized spiral PSF (PS$^2$F), a novel engineered PSF that is well-suited for linear structures. Our key enabling observation is that asymmetric PSFs achieve higher depth accuracy for scenes comprising of linear structures. We achieve this asymmetry by observing that the DHPSF is comprised of two lobes that rotate about a common center; by capturing images with the individual lobes, we can reliably estimate depth. We propose a compact realization of our approach with a novel polarization-based imaging system. PS$^2$F is generated by combining a DHPSF mask with orthogonal linear polarizers on the two halves of the mask. We then use a polarization sensor, capable of measuring images along four polarization angles in a snapshot by using a Bayer-like tiling. Finally, the 3D scene is estimated by solving a depth-dependent deconvolution problem. An overview of depth estimation with PS$^2$F is illustrated in Fig.~\ref{fig:fig1}. {\flushleft \textbf{Contributions.}} We propose a new engineered PSF for imaging linear structures and make the following contributions. \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-simulations/teaser-v2.pdf} \caption{\textbf{Polarized Spiral PSF (PS$^2$F) for single-shot 3D scene estimation.} We propose a novel monocular depth imager based on an engineered PSF. Numerous real-world scenes can be modeled as linear structures. We observe that asymmetric PSFs, such as a single lobe of the DHPSF, enable us to accurately estimate the 3D geometry of such linear structures. We leverage this observation and propose a compact polarization-based setup that produces accurate depth maps in a snapshot manner.} \label{fig:fig1} \end{figure*} \begin{itemize} \item We theoretically demonstrate that the DHPSF is ill-suited for 3D imaging of linear structures by analysing its Cram\'er-Rao lower bound (CRLB). \item We propose a compact polarization-based setup, using a novel polarizer-phase mask encoding that is capable of better 3D estimation in a snapshot manner. \item We demonstrate the advantages of PS$^2$F through several experiments with our lab prototype over a wide range of scene geometries. \end{itemize} {\flushleft \textbf{Limitations.}} At its core, PS$^2$F assumes that the scene intensity is unpolarized; if the scene is polarized, the resultant depth estimate may not be accurate. Our technique, like most techniques that rely on defocus blur, performs poorly for regions with little or no textures or structures. \section{Proposed Method}\label{sec:proposed-method} We begin with a brief background on the DHPSF, an analysis of its properties, and its performance for linear/extended structures. Subsequently, we introduce our proposed PSF and its associated reconstruction technique. \subsection{Background: DHPSF and mask design} This subsection outlines the background on the generation of the DHPSF \cite{pavani2009three} and other GL-based rotating PSFs \cite{lew2011corkscrew}. \subsubsection{Phase mask for GL-based rotating PSFs} Any propagating paraxial wave/beam can be expressed in terms of the orthogonal basis of GL modes, which are characterized by two integers $(n,m)$. Different combinations of the GL modes lead to varying properties of paraxial beams. In \cite{piestun2000propagation}, it was shown that beams that constituted of GL modes that are in arithmetic progression continuously rotated and scaled upon propagation. Using a beam corresponding to GL modes of (1,1), (5,3), (9,5), (13,7), it was shown in \cite{greengard2006depth} that the amplitude and phase profile of the beam at $z=0$ can be used as a mask in front of a lens (or in the Fourier plane of a $4f$ lens system) to generate a PSF that rotates with defocus. Subsequently, this mask was implemented in a phase-only manner using a variant of Gerchberg-Saxton optimization procedure with constraints included for the GL modal domain~\cite{pavani2008high}. The resulting phase function creates a PSF that rotates with changing point source depth, and is popularly known as the DHPSF due to its double helical structure. This PSF has been successfully used in 3D localization of fluorescent particles \cite{pavani2009three}\cite{grover2011photon}. \subsubsection{Properties of GL-based rotating PSFs} Piestun \textit{et al.}~\cite{piestun2000propagation} present a theory on paraxial rotating beams using GL modes, and characterizing the resulting rotation range, rotation rates, and beam scaling rates. Specifically, given a $4f$ system with lenses of focal length $f$, a GL-based mask designed with beam width $w_0$ for wavelength $\lambda$, and having GL modes lying on a single line with slope $V_1$, the following can be established: \begin{itemize} \item The total rotation in one direction of defocus is $(V_1\pi/2)$ \item The angle of the rotating PSF $$\phi(z) = \phi_0 + V_1\ \text{tan}^{-1}\left(\frac{z}{\lambda f^2/\pi w_0^2}\right)$$ \item The PSF will rotate by $(V_1\pi/4)$ over a depth of $\frac{\lambda f^2}{\pi w_0^2}$ . \end{itemize} Even though phase-only masks for rotating PSFs are generated after an optimization procedure~\cite{pavani2008high}, the above rates and ranges are still a good approximation to understand the properties of any GL-based rotating PSF. \subsection{Challenges in imaging linear/extended structures} Phase masks such as the DHPSF~\cite{pavani2009three}, or the Tetrapod PSF~\cite{shechtman2014optimal}, were designed for 3D estimation of \textit{point sources}. To image scenes with extreme fluorophore density, the concept of a scene made up of point-like emitters is not accurate. With increasing fluorophore/light source density, scenes consist of edge-like and linear structures. Using a DHPSF for recovering depth of a line then leads to an ambiguity between depth and orientation of the line. This is illustrated in Fig. \ref{fig:depth-ambiguity-illustration}. Given a line image formed using the DHPSF, there will always be a global depth ambiguity---two different depths can give rise to the same line image. This makes the DHPSF ill-suited for imaging edge/linear structures. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/dhpsf_depth_ambiguity.pdf} \caption{\textbf{Ambiguities caused by DHPSF for lines}. Estimation of depth and orientation of a line using the DHPSF is an ill-conditioned problem. There is an inherent global ambiguity in depth estimation, since for every line image, two possible candidate depths are possible.} \label{fig:depth-ambiguity-illustration} \end{figure} \subsection{Designing a PSF for 3D lines} A line with constant intensity blurred with an arbitrary PSF produces another line -- thereby leading to loss of information along the line direction. In case of the DHPSF, the loss of information implies that two lines at two different depths ($-\Delta z$ and $+\Delta z$) produce the same measurement. This ambiguity can however be resolved if we measure two separate images with each lobe of the PSF. Imaging with the individual DHPSF lobes separately will remove the global degeneracy issues (up to a $2\pi$ lobe rotation range). Next, we provide a theoretical analysis of the advantages to be obtained by such a separation. For estimation from data (image(s) in our case), we can use theoretical tools to understand what ultimate precision is possible. The Fisher information and Cram\'er-Rao Lower bound (CRLB) give the best possible precision that can exist by any estimator. In the case of an image of an isolated point source, the Fisher information matrix for estimating parameters $\theta = (x,y,z)$ is given by \cite{shechtman2014optimal}: \begin{equation} \label{eq:FI-point-source} [FI(\theta)]_{ij} = \sum_{k=1}^{N_p}\frac{1}{\mu_\theta(k) + \beta}\left(\frac{\partial \mu_\theta}{\partial \theta}\right)_i\left(\frac{\partial \mu_\theta}{\partial \theta}\right)^T_j, \end{equation} where $N_p$ is the number of pixels, $k$ is a variable for summing over all pixels, $T$ corresponds to the transpose operator, $\beta$ is the background Poisson noise level, and $\mu_\theta$ is the scaled PSF image for a point source at $\theta=(x,y,z)$ (scaled by the number of photons $N$). The diagonal elements of the inverse of the Fisher information matrix provides the CRLB for estimation of each of the parameters $\theta = (x,y,z)$. $$[\textrm{CRLB}(\theta)]_i \equiv [FI(\theta)]^{-1}_{ii} $$ We next extend the Cram\'er-Rao analysis to imaging of lines. Given a line image (or a one-sided edge image), the parameters to be estimated are the associated depth $z$ of the line/edge (at center of the patch), the associated orientation $\phi$, i.e. $\theta=(z,\phi)$. Thus, given a line image $\psi$, the CRLB estimates for depth $z$ is given by \begin{align} [FI(\theta)]_{ij} = \sum_{k=1}^{N_p}\frac{1}{\psi_\theta(k) + \beta}\left(\frac{\partial \psi_\theta}{\partial \theta}\right)_i\left(\frac{\partial \psi_\theta}{\partial \theta}\right)^T_j \label{eq:fi}\\ [\textrm{CRLB}(\theta)]_z \equiv [FI(\theta)]^{-1}_{zz} \label{eq:fiz} \end{align} The estimation of line orientation from a line path is easier than depth estimation, as orientation can be computed from the global structure of the line image. Hence, for simplicity of exposition, we focus on the analysis of the CRLB$_z$ values here. Analysis regarding the CRLB$_\phi$ values is shown in the Supplementary. Using Eqns~\ref{eq:fi}, \ref{eq:fiz} we can calculate the $\sqrt{\textrm{CRLB}_z}$ values for a given line image for all $(z,\phi)$. This calculation can be readily extended to line images captured using two separate PSFs. In such a case, the final Fisher information matrix is the sum of the individual Fisher matrices. Assuming a $4f$ system with 50~mm focal length lenses, and $N=100,000$ photons, $\beta=5$ photons/pixel, we compute the $\sqrt{\textrm{CRLB}_z}$ plots (as a function of $z$, $\phi$) for the DHPSF~\cite{pavani2008high}, the Tetrapod PSF~\cite{shechtman2014optimal}, and the individual DHPSF lobes with $25\%$ photons in each lobe. Note that $\sqrt{\textrm{CRLB}_z}$ obtained with a mask pair should be compared with a single mask having 2x SNR (2x total photons), as shown in \cite{nehme2020learning}. However, our polarizer-phase mask setup causes a $50\%$ light loss (see further in Section 3). Hence, for an appropriate comparison, we compare the individual DHPSF lobe pair, with other PSFs having $4\times$ the SNR. There are possible ways to remove this loss, with a setup demonstrated in \cite{nehme2020learning}, which will further lower the estimated $\sqrt{\textrm{CRLB}_z}$ values. Fig.\ \ref{fig:crlb-plots} illustrates the $\sqrt{\textrm{CRLB}_z}$ values (log-scale) for a given line at depth $z$ and having orientation $\phi$. The mean and standard deviation of the $\sqrt{\textrm{CRLB}_z}$ values are also shown in the individual insets. The DHPSF plot in Fig. \ref{fig:crlb-plots} shows several peaks with large magnitude. This occurs at specific $(z,\phi)$ where the line connecting the two lobes of the PSF and the scene line are either perpendicular or parallel to each other. The ability to discern lines at slightly differing depths is greatly reduced at these points due to symmetry. For the Tetrapod PSF~\cite{shechtman2014optimal} and the PhaseCam3D PSF~\cite{wu2019phasecam3d}, the plots in Fig. \ref{fig:crlb-plots} also show more peaks and ridges, but with lower magnitude compared to the DHPSF. For the PSF pair of the individual DHPSF lobes, we see that most of the peaks are removed, except for a central peak in the plot. The peak corresponds to the case when the PSF is in focus and the the line orientation is perpendicular to the PSF lobes. This central peak does not affect the effectiveness of separating lobes, because it is a narrow peak over a only small area in the parameter space of $(z,\phi)$. The average $\sqrt{\textrm{CRLB}_z}$ values for the individual DHPSF lobe pair is similar to the DHPSF with 4x SNR. Thus, with 4x lower SNR, the PS$^2$F is able to achieve a comparable $\sqrt{\textrm{CRLB}_z}$ value with other PSFs, and does not show consistent peaks and ridges as seen in the plots. Note that CRLB calculations only provide insight about local ambiguities or precision levels. However, we also observe that a PSF pair created out of separating the DHPSF lobes removes global ambiguities as well. Thus overall, a PSF pair created from the individual DHPSF lobes is better for estimating depths of line/edge patches. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/crlb_z_plots_v1.pdf} \caption{\textbf{CRLB comparison for various phase masks.} The figure shows $\sqrt{CRLB_z}$ log-intensity (log-mm) plots for line images, as a function of line depth $(z)$ and line orientation $(\theta)$. PS$^2$F achieves a comparable CRLB over the whole range of depth and orientations with fewer peaks, even with 4x lower SNR. In contrast, competing approaches have curves with large CRLB values, which introduce ambiguities in the depth estimate.} \label{fig:crlb-plots} \end{figure} \subsection{Using polarizers to separate out the DHPSF lobes in a single $4f$ optical system} Imaging with a PSF pair is possible by having two separate parallel $4f$ imaging systems with separate sensors, as demonstrated in \cite{nehme2020learning}. However, the DHPSF mask (and the Fresnel ring-based two-lobe masks) enjoy the property of PSF lobe separability. Partitioning such a mask into two halves along a particular axis, and allowing light through only one half leads to the creation of only a single rotating PSF lobe, as depicted in Fig.~\ref{fig:psf-partition-illustration}. The primary reason is that the light falling on one half is modulated to form one lobe, while the light falling on the other half is modulated to form the other lobe. Such a partitioning leads to partitioning of the DHPSF into two separate, distinct lobes. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/fig4.pdf} \caption{\textbf{Effect of separating the phase mask.} The phase mask for generating the DHPSF has a special property that separating the phase profile along the highlighted red line yields a separation of the two lobes. We leverage this to implement a compact realization of Polarized Spiral PSF (PS$^2$F) using a polarization camera sensor and two polarizers in the pupil plane.} \label{fig:psf-partition-illustration} \end{figure} The separability of DHPSF lobes can be leveraged to build a compact imaging system. In PS$^2$F, we add two linear polarizers on each mask half, one oriented along $s$-polarization and the other oriented along $p$-polarization. This ensures that the $s$-polarization component is modulated using one mask half, and consequently the $p$-polarized component is modulated using the other mask half. The PSF created out of such a polarizer-phase mask encoding contains two lobes, but with each lobe being in a different polarization (either $s$ or $p$). This addition of polarizers causes a $50\%$ light loss, which was taken into consideration while estimating the $\sqrt{\textrm{CRLB}_z}$ values. Such a \textit{polarized spiral PSF} can be captured efficiently using a polarized camera sensor. A polarized camera sensor contains a 2D array of pixels with a Bayer pattern that has 90, 45, 135, 0 degree polarizers. After addition of the polarizers in the mask-plane, the orthogonally polarized DHPSF lobes can be imaged in their respective channels of a polarization sensor. This allows for a compact $4f$ system, allowing the imaging of a PSF pair using just a single $4f$ system channel, making the proposed method a single-shot, single-sensor method. \subsection{Imaging model and reconstruction procedure} We assume that the light from the scene is unpolarized in our imaging model. The imaging of a 3D scene with a depth-dependent PSF can be approximated as a sum of 2D convolutions between the depth-dependent PSF and the per-plane scene intensity: \begin{equation} \label{eq:imaging-model} I_c(x,y) = \sum_{z=z_{start}}^{z_{end}} h_c(x,y;z) * s(x,y,z) \end{equation} where $I_c(x,y)$ is the image intensity at $(x,y)$ in polarization channel $c$, $h_c(x,y;z)$ is the 2D PSF corresponding to depth $z$ and polarization channel $c$, and $s(x,y,z)$ is the scene intensity at point $(x,y,z)$. The goal of 3D reconstruction is to estimate a 3D matrix $s(x,y,z)$ from the noisy measurements of $I(x,y)$. There are numerous ways to solve this problem including regularized least-squares~\cite{boominathan2020phlatcam,yanny2020miniscope3d, xue2020single}, data learning-based~\cite{wu2019phasecam3d,chang2019deep}, and more recently, ones based on deep network-based regularizer~\cite{zhang20203d}. To keep the exposition simple, we demonstrate recovery by modelling the 3D reconstruction problem as a regularized least squares optimization problem. Specifically, we formulate the 3D scene estimation problem as: \begin{equation} \label{eq:optim_func} \underset{\textbf{x}}{\text{argmin}} \left\lVert \begin{bmatrix}I_0\\I_{90}\end{bmatrix} - S\begin{bmatrix}H_0\\H_{90}\end{bmatrix}\textbf{x}\right\rVert^2_2 + \lambda_{TV}\lVert\Psi(\textbf{x})\rVert_1 + \lambda_{L1}\lVert \textbf{x} \rVert_1 \end{equation} where $\textbf{x}$ is a 3D matrix of scene intensities, $H_0, H_{90}$ are the depth-specific PSF operators corresponding to the individual two lobes that are polarized to $0^\circ$ and $90^\circ$ states respectively, and $I_0, I_{90}$ are captured images in the $0^\circ$ and $90^\circ$ polarization channels respectively. $S$ is the summing operator that sums across the depth channels of $H_0\textbf{x}$ and $H_{90}\textbf{x}$ separately. We employ TV and L1 regularizers as scene priors, with $\lambda_{TV}, \lambda_{L1}$ as hyperparameters to control their regularization effects respectively. We solve the optimization problem using autograd functionality in PyTorch and the Adam optimizer~\cite{kingma2014adam} to leverage the speed of graphical processing units (GPUs). A key requirement for our technique is that the scene emits unpolarized light. This is required so as to ensure that both the lobes have produce PSFs with similar intensity levels that we can calibrate a priori. When the incident light is polarized, the blur kernels in Eq.~\ref{eq:imaging-model} will have an unknown scaling, which leads to a model mismatch; in an extreme scenario, we lose all the information in one of the lobes if the polarization angle of the incident light is orthogonal to the corresponding polarizer in the pupil plane. While this is a limitation of our technique, we largely encountered unpolarized light in our intended application of fluorescence microscopy. \section{Related Work}\label{sec:related-work} PS$^2$F draws motivation from the prior work on engineered PSFs. We discuss the relevant ones in this section. \subsection{Mask-based PSF encoding} Encoding PSFs by designing masks has been used in several works, including extended depth-of-field imaging~\cite{castro2004asymmetric, yang2007optimized,zammit2014extended}, 3D sensing in the macroscopic~\cite{zhou2009coded} and microscopic~\cite{pavani2009three,shechtman2014optimal} regimes, and in a lensless imaging setting~\cite{boominathan2020phlatcam}. The designed masks can be placed in front of the imaging lens or in the pupil plane of the imaging system. Earlier works include designing amplitude masks or coded apertures~\cite{levin2007image,zhou2009coded} -- such amplitude masks however lead to poor light efficiency and very coarse depth information. Several follow up works utilized phase-only modulation masks which include the rotating PSFs~\cite{pavani2009three,lew2011corkscrew,prasad2013rotating}, and PSFs obtained by optimizing the Fisher information~\cite{shechtman2014optimal}. Phase masks have also been designed in an end-to-end optimization setting, both for macroscopic~\cite{wu2019phasecam3d,chang2019deep} and microscopy~\cite{nehme2020deepstorm3d,nehme2020learning} tasks. \subsection{3D super-resolution microscopy using phase masks} Coupled with STORM~\cite{huang2008three} and PALM~\cite{betzig2006imaging, hess2006ultra}, masks that result in rotating PSFs with depth have been able to perform super-resolution microscopy~\cite{pavani2009three,lew2011corkscrew} of individual fluorophore emitters to nanometer-scale accuracy. The most successful of such designs is the DHPSF~\cite{pavani2009three}, a popular choice for 3D localization of point sources~\cite{huang2008three,juette2008three,badieirostami2010three}. 3D localization performance have been further improved by designing Tetrapod PSFs~\cite{shechtman2014optimal}, which were generated by directly optimizing the Fisher information of the phase mask for $(x,y,z)$ localization of point sources. Recently, deep learning techniques have enabled an end-to-end design framework that simultaneously learn phase masks as well as neural networks for obtaining 3D localizations from raw captures~\cite{nehme2020deepstorm3d, nehme2020learning}. Of particular interest is the work by Nehme \textit{et al.}~\cite{nehme2020learning}, where a pair of masks were designed in an end-to-end learning process. The two masks are placed in two parallel $4f$ imaging channel. The mask pair, coupled with a learnt deep network, allows for high accuracy 3D localization of single fluorophores over a large depth range and for a high density of point-like emitters. Their work focuses on super-resolution microscopy and on 3D localization of \textit{point-like} emitters, while our result tackles extended linear structures. \subsection{3D sensing of linear and extended structures} Human/animal bodies have a rich, dense 3D network of blood vessels running through them. Capturing the 3D structure of vasculature is usually done using variants of light-sheet microscopy (LSM)~\cite{girkin2018light,di2018whole,lugo20173d}, confocal microscope (CM)~\cite{st2005confocal,kelch2015organ}, optical projection tomography~\cite{bassi2011vivo}, and optical coherence tomographic angiography~\cite{Makita:06}. However, confocal and LSM methods involves scanning individual points, lines, or planes which makes the systems bulky and decreases the achievable imaging rate, making them non-ideal for fast dynamic 3D imaging. \subsection{Depth sensing using rotating PSFs} There is a rich body of work devoted to PSFs with lobes that rotate with defocus. These can be clustered into two groups based on their use of Gauss-Laguerre (GL) modes~\cite{pavani2008high} or Fresnel rings~\cite{prasad2013rotating}. GL-based rotating PSFs~\cite{pavani2009three,lew2011corkscrew} have been used for super-resolving and localizing single molecules in 3D~\cite{pavani2009three,grover2011photon,gustavsson20183d, bennett2020novel}. Both GL-based~\cite{quirin2013depth} and Fresnel-ring-based~\cite{berlich2016single,wang2017single} rotating PSFs have also been used for single-shot depth estimation of scenes. However, the first work employs a two phase-mask solution. The latter two employ a patch-based technique involving processing in the cepstrum domain. They only recover very simple scenes at low spatial and depth resolutions. Such recovery approaches do not extend to more complex geometry such as linear structures. \subsection{Using polarization channels for PSF encoding} There have been a few works that leverage polarization to help in 3D super-resolution imaging~\cite{roider2014axial,nehme2020learning,ikoma2021deep}. These works typically use a beam-splitter to split the incoming light into two separate $4f$ optical system channels which enabled individual channels to be encoded with a different mask. The images were then captured over two non-overlapping regions of a single sensor, or on two separate sensors. Such solutions however call for bulkier optics, and more importantly, a need for sub-pixel alignment to achieve high spatial resolution. We instead leverage snapshot polarimetric cameras that are naturally well-aligned, and hence easier to work with, and more compact. \section{Simulations}\label{sec:simulations} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-simulations/fig7_v2.pdf} \caption{\textbf{Simulations on vascusynth dataset with varying complexity.} The figure shows reconstructions with various PSF designs on scenes with medium complexity (top row) and high complexity (bottom row). PS$^2$F is superior in depth estimation, especially in regions where there is higher vasculature (scene) complexity (see insets). Scale bars indicate $0.25$~mm.} \label{fig:vasc-scenes-dmaps} \end{figure*} To evaluate the performance of our proposed PS$^2$F, we perform extensive simulations on the \href{http://vascusynth.cs.sfu.ca/Welcome.html}{VascuSynth 2013} dataset~\cite{hamarneh2010vascusynth, jassi2011vascusynth}, which provides 10 simulated 3D volumes of vascular trees for each of 12 levels of complexity. Scene complexity is set by the number of bifurcations in the vascular tree (ranging from 1 to 56 in steps of 5). \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/rebut_RMSE_compare.pdf} \caption{\textbf{Accuracy vs.\ scene complexity}. Depth estimation performance, for the proposed PS$^2$F and other comparing PSFs (averaged over 10 scenes for each complexity level, shaded regions show variance across the 10 scenes). This demonstrates the suitability of PS$^2$F, and that the separation of the DHPSF lobes helps in 3D reconstruction.} \label{fig:sim-compare} \end{figure} Each 3D vasculature volume in the dataset consists of $101\times101\times101$ voxels. These volumes were converted into 3D scenes of size $256\times256\times256$ through interpolation methods. This new volume was assumed to cover a volume of $1.76\times1.76\times5.00$ mm$^3$ with each voxel occupying a $6.9\times6.9\times19.5~\mu$m$^3$. We further assume a $4f$ system with both lenses having focal lengths $f=50$~mm. We render images of the vascular structures using several different masks of diameter $3$~mm in the Fourier plane. We assume that the light from the vasculature is monochromatic of wavelength $\lambda=532$~nm. We render images assuming occlusion of vasculature and that light is received only from the surface of the vasculature. We assume that the center of the 3D scene is $f=50$~mm away from the first lens, thus, the scene spans a defocus range of [$-2.5$~mm, $+2.5$~mm]. The masks (and their corresponding PSFs) used in rendering the simulated images were: \begin{itemize} \item DHPSF: We optimize for a GL-based DHPSF mask (following procedure in~\cite{pavani2008high}), based on the rotating paraxial beam with $w_0=0.4$~mm and having GL modes corresponding to $(1,1), (3,5), (5,9), (7,13)$. Full 180 degree lobe rotation was achievable over $2\frac{\lambda f^2}{\pi w_0^2}=5.3$~mm. \item PS$^2$F: Using the above mask, we construct an equivalent mask pair corresponding to PS$^2$F by partitioning the mask into two halves appropriately. \item Tetrapod PSF: We obtain a Tetrapod PSF mask designed for $550$~nm wavelength from~\cite{gustavsson20183d} and repurpose it for imaging over our specified $5$~mm depth range. \end{itemize} We then added Poisson and Gaussian read-out noise to the final measurements. For an appropriate comparison, we perform 3D reconstructions with individual single lobes (PS$^2$F), two-shot capture with DHPSF, and a two-shot capture using the Tetrapod PSF. To accurately account for the $50\%$ light loss in the PS$^2$F case, we also render out images with half the signal level as compared to the signal levels in DHPSF and Tetrapod PSF case. After obtaining the 3D volumetric estimate, we filter out points whose sum across the z-stack is lower than a fixed threshold. We then estimate depth using the index corresponding to a maximum-intensity-projection (MIP) of the 3D estimate. \subsection{Comparison with DHPSF and Tetrapod PSF} Fig. \ref{fig:sim-compare} shows the depth estimation performance for the proposed PS$^2$F, DHPSF, and the Tetrapod PSF. We compare the PSFs depth estimation using the Mean Absolute Error (MAE), Root Mean Square Error (RMSE), and the multi-scale Structural Similarity Index Measure~\cite{wang2003multiscale} (MS-SSIM) metrics. The entire set of results can be seen in the Supplementary. The proposed PS$^2$F method performs $\sim2\times$ better in terms of the RMS error for depth estimation. The depth estimation performance improvement is even higher ($\sim3\times$) for vasculature with more bifurcations, i.e. greater complexity. Depth estimation results for two example vasculature scenes can be seen in Fig. \ref{fig:vasc-scenes-dmaps}. In Fig. \ref{fig:vasc-scenes-dmaps} (top row), the vasculature scene is not very complex. The depth map estimation results are fairly similar for all the three PSFs (PS$^2$F, DHPSF, Tetrapod PSF), but the zoom-in insets highlight the errors seen in the DHPSF and Tetrapod PSF depth estimation. In Fig. \ref{fig:vasc-scenes-dmaps} (bottom row), the scene has greater complexity, and the PS$^2$F produces much better depth estimation than the other two PSFs. The zoom-in insets especially highlight the same. \begin{figure}[!t] \centering \centering \includegraphics[width=\columnwidth]{figs-simulations/usaf-sim-compare-part1-v2.pdf} \caption{\textbf{Simulated reconstruction results of a USAF target at a skew angle.} (A) Shows the depth map estimates for PS$^2$F, DHPSF, and Tetrapod PSF respectively. The XZ maximum intensity projections are shown in (B) in their respective columns, with ground truth shown in the inset. (C) shows zoom-in insets of depth maps (after thresholding to remove small artifacts) of highlighted regions in (A). PS$^2$F has a better reconstruction in lateral and axial dimensions as compared to the DHPSF and the Tetrapod PSF. Scale bars for (A), (B), (C) indicate $0.5$~mm, $0.5$~mm, and $0.1$~mm respectively.} \label{fig:sims-usaf} \end{figure} \subsection{Resolution performance} Furthermore, we analyse the reconstruction resolution performance by reconstructing a simulated skew USAF target. We assume a USAF target places at the focus plane, but tilted with respect to the optical axis so that its depth changing horizontally across the target; we render using the same three masks in the same $4f$ optical system as above. The results are shown in Fig.\ \ref{fig:sims-usaf}. The proposed PS$^2$F generates a better reconstruction than the DHPSF and Tetrapod PSF reconstructions, both in lateral and axial dimensions. As seen from the XZ MIP plots in Fig.\ \ref{fig:sims-usaf}, the PS$^2$F and Tetrapod PSF reconstructions are able to obtain the skew angle of the target, with the former being much better. The DHPSF fails to get the skew angle, due to the global depth ambiguity issues explained in Section \ref{sec:proposed-method}. In the lateral dimensions, the Tetrapod PSF is unable to reconstruct Group 4 and 5. This could be attributed to the fact that the PSF has a larger support. On the other hand, the PS$^2$F and DHPSF have concentrated lobes, and thus are able to reconstruct elements of Group (5,3) (12.40~$\mu$m linewidth) and (5,2) (13.92~$\mu$m linewidth) respectively. The obtained $xy$-resolution can be theoretically justified. The separation of lobes in the PS$^2$F method allows for resolution of finer scene features due to the compactness of the PSF. The resolution will be roughly determined by the width of the individual PSF lobes. Fitting a 2D Gaussian to each individual lobe, the average $2\sigma$-value for the Gaussian fit is $\sim1.74~\text{pixels} = 12~\mu$m in the object space. This readily matches with the fact that the PS$^2$F can resolve elements of Group (5,3), which has line elements that are spaced at $12.40~\mu$m distance. The diffraction-limited resolution of an Airy disk PSF for the same optical system configuration is given by $1.22\lambda f/D = 10.81~\mu$m, indicating that the $xy$-resolution obtained by the PS$^2$F is close to the diffraction-limited resolution ($1.15\times$) as well. For PS$^2$F, we also analysed the spread of reconstruction signal across the z-channel. For pixels with a significant signal level, we fit a 1D Gaussian curve to estimate this spread. The median $2\sigma$-value across z-channel obtained was $2\times304=608~\mu$m, which is an estimate of the axial resolution performance of the proposed PS$^2$F. Note that the diffraction-limited axial performance of an Airy disk PSF for the same optical system configuration is given by $4\lambda(f/D)^2 = 591~\mu$m. We are able to achieve $\sim1.03\times$ the diffraction-limited axial resolution with the proposed PS$^2$F. \section{Supplementary: Methods} \subsection{Properties of GL-based rotating PSFs} \cite{piestun2000propagation} presented a theory on paraxial rotating beams using GL modes, along with a detailed estimation of rotation range, rotation rates, and beam scaling rates. These theoretical values could be extended to give some theoretical description about GL-based rotating PSFs. However, to the best of our knowledge, this has not been shown. Here we present a calculation for the rotation range and rates for GL-based rotating PSFs. Consider a monochromatic paraxial beam with width $w_0$ and wavelength $\lambda$, which is generated from GL modes that lie on a single line in the GL modal plane with a slope $V_1$, expressed as: \begin{equation} n_j = V_1m_j + V_2 \ \ \ \ j=1,2,... \end{equation} where $V_1, V_2$ are integer constants, and $\{m_j\}$ are non-negative numbers in arithmetic progression. \cite{piestun2000propagation} showed that such a paraxial beam rotates at a rate \begin{equation} \frac{d\phi}{dz} = \frac{V_1}{1 + (z/z_R)^2} \end{equation} where $z_R = \frac{\pi w_0^2}{\lambda}$ is the Rayleigh length. A point with coordinates $(\rho_0,\phi_0)$ at $z=0$, follows the trajectory \cite{piestun2000propagation}: \begin{align} \rho &= \rho_0 \sqrt{1 + (z/z_R)^2} \\ \phi &= \phi_0 + V_1(arctan(z/z_R)) \end{align} which implies that the maximum rotation possible (in one direction) is $V_1(\pi/2)$. Thus, for the rotating beam case, over a distance of $z_R$, the beam rotates $V_1(\pi/4)$. The way to find out rotation and scaling rates for a GL-based rotating PSF is to find the equivalent Rayleigh length $z_R$ for it in a $4f$ optical system. Note that the GL modes arise from modal solutions of the Fresnel diffraction operator\cite{piestun2000propagation}. The Fresnel diffraction integral is given as follows \cite{goodman2005introduction}: \begin{equation} \label{eq:fresnel} U(u,v) = \frac{e^{jkz}}{j\lambda z}e^{j\frac{k}{2z}(u^2+v^2)}\mathcal{F}\{U(x,y)e^{j\frac{k}{2z}(x^2+y^2)}\}_{\frac{u}{\lambda z},\frac{v}{\lambda z}} \end{equation} where $j^2=-1$ and $k$ is the wave number. Eqn \ref{eq:fresnel} determines the wave profile $U(u,v)$ at $z$ distance away from an input wave profile $U(x,y)$. Thus, the phase term that determines the \emph{defocus} or \emph{propagation} by a distance $z$ is \begin{equation}\label{eq:rl-beam} \frac{k}{2z}(u^2+v^2) \end{equation} Similarly, upon solving the propagation integrals for a $4f$ system with defocus, the phase term that affects the defocus is given by \begin{equation}\label{eq:rl-4f} \text{exp}\left(j\frac{k}{2f}(\frac{\Delta z}{f})(u^2+v^2)\right) \end{equation} where $\Delta z$ is the distance from the in-focus plane. Comparing Eqns \ref{eq:rl-beam} and \ref{eq:rl-4f}, we can obtain the equivalent \textit{Rayleigh length} in the $4f$ system (say $z'_R$) \begin{align} \frac{1}{z_R} = \frac{z'_R}{f^2} \implies z'_R = f^2/z_R = \frac{\lambda f^2}{\pi w_0^2} \end{align} Thus, given a GL-based rotating PSF mask designed with beam width $w_0$, for wavelength $\lambda$ and $V_1$ slope in the GL modal plane: \begin{itemize} \item The total rotation amount in one direction of defocus is $V_1(\pi/2)$ \item The angle of the rotating PSF $\phi(z) = \phi_0 + V_1(\text{arctan}(\frac{z}{\lambda f^2/\pi w_0^2}))$ \item And over a depth of $\frac{\lambda f^2}{\pi w_0^2}$ the PSF will rotate by $V_1(\pi/4)$ radians. \end{itemize} \subsection{Analysis of CRLB$_{\phi}$} \begin{figure}[h!] \centering \includegraphics{suppl_crlbphi_plots.pdf} \caption{\textbf{CRLB$_{\phi}$ comparison for various phase masks.} The figure shows $\sqrt{CRLB_{\phi}}$ log-scale plots for line images, as a function of line depth $(z)$ and line orientation $(\theta)$. Individual insets depict the mean and standard deviation of the $\sqrt{\textrm{CRLB}_{\phi}}$ values, which are all very low.} \label{fig:crlb-phi} \end{figure} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{fig5_v2.pdf} \caption{\textbf{Simulation methodology to evaluate PS$^2$F}. Given a 3D volume, we used a simulated stack of depth-dependent PSFs to obtain the two polarization images. We then computed the depth at each spatial point with the two polarization images as inputs, resulting in the point cloud shown on the right.} \label{fig:sims-vasc} \end{figure*} \begin{table*}[!t] \setlength{\tabcolsep}{.25cm} \renewcommand{\arraystretch}{1.3} \caption{Depth Estimation performance, averaged over 10 scenes for each scene complexity (noise level$=0.02$). $B$ refers to the number of bifurcations (B) in the 3D vascular structure. Showing mean absolute error (MAE) and RMS error (RMSE) values (all units in mm), and multi-scale SSIM~\cite{wang2003multiscale} (MS-SSIM) values. For the first two metrics, lower is better (indicated by $\downarrow$), and for MS-SSIM, higher is better (indicated by $\uparrow$).} \centering \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline & \multicolumn{3}{c|}{PS$^2$F(proposed)} & \multicolumn{3}{c|}{DHPSF(x2)~\cite{pavani2009three}} & \multicolumn{3}{c|}{TetrapodPSF(x2)~\cite{shechtman2014optimal}}\\ \hline\hline & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$ & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$ & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$\\ \hline B=6 & \textbf{0.198} & \textbf{0.292} & \textbf{0.990} & 0.284 & 0.46 & \textbf{0.990} & 0.352 & 0.541 & \textbf{0.990}\\ \hline B=16 & \textbf{0.225} & \textbf{0.334} & \textbf{0.980} & 0.494 & 0.781 & 0.979 & 0.532 & 0.811 & 0.979\\ \hline B=26 & \textbf{0.211} & \textbf{0.321} & \textbf{0.972} & 0.507 & 0.839 & 0.971 & 0.574 & 0.872 & 0.970\\ \hline B=36 & \textbf{0.224} & \textbf{0.344} & \textbf{0.969} & 0.548 & 0.904 & 0.967 & 0.585 & 0.906 & 0.966\\ \hline B=46 & \textbf{0.261} & \textbf{0.413} & \textbf{0.966} & 0.564 & 0.933 & 0.964 & 0.769 & 1.129 & 0.963\\ \hline \end{tabular} \label{tab:sim-compare} \end{table*} In the CRLB analysis performed in Sec 3.3, we consider an analysis for CRLB$_z$ only. Intuitively, the estimation of line orientation from the image of a line patch is an easier task as compared to estimating line depth. This is because line orientation can be readily estimated from the global structure present in the line image. To verify this, we perform CRLB$_{\phi}$ calculations with the same parameters ($N=100,000$ photon, $\beta=5$ photons/pixel). The $\sqrt{\textrm{CRLB}_{\phi}}$ plots are shown in Supplementary Fig.~\ref{fig:crlb-phi}. The $\sqrt{\textrm{CRLB}_{\phi}}$ values are very low, indicating that our intuition about line orientation estimation being an easier problem is correct. \section{Supplementary: Simulations} \subsection{VascuSynth dataset simulations} The VascuSynth dataset simulations pipeline is best illustrated by Fig.~\ref{fig:sims-vasc}. Table~\ref{tab:sim-compare} shows the comparative results for the proposed PS$^2$F, DHPSF, and the Tetrapod PSF over three metrics - Mean Absolute Depth Error, RMS Depth Error, and multi-scale SSIM~\cite{wang2003multiscale}. We also perform simulations under varying levels of noise, to judge the robustness of the proposed PS$^2$F. We simulate vasculature scene renderings with Poisson noise, and also Gaussian read-out noise of levels $0.02$, $0.05$, and $0.1$ - which correspond to PSNRs of 34dB, 26dB, and 20dB. Fig. \ref{fig:sim-noise-analysis} shows that with increasing noise, the proposed PS$^2$F reconstruction only worsens slightly, showing the robustness of the proposed PSF and method to noise. \begin{figure}[h!] \centering \includegraphics[width=0.8\columnwidth]{RMSE_noise.pdf} \caption{Depth Estimation performance, for different noise levels (averaged over 5 scenes for each complexity level, shaded regions show variance across the 5 scenes). This shows the robustness of the proposed PSF and technique.} \label{fig:sim-noise-analysis} \end{figure} \subsection{Ablation study: Comparison with polarized stereo pair} The proposed PS$^2$F has been created using a novel polarizer-phase mask design and the usage of a polarization-based camera sensor. We perform an ablation study, comparing our proposed PS$^2$F (polarizer-phase mask design) with a PSF created using just polarizer halves in the Fourier (pupil) plane. This creates images in the $s$, $p$-polarized channels that correspond to different half apertures (taking inspiration from \cite{bando2008extracting}). With such a polarized stereo-vision effect, where each polarization channel sees a slightly different perspective, we attempt reconstruction of a 3D USAF skew target in simulation. The reconstructed intensity plots (depth color-coded) for the PS$^2$F and the Polarized-Stereo PSF are shown in Fig.~\ref{fig:ablation}. A polarizer-phase mask design gives better spatial resolution, and a $\sim2\times$ lower depth RMS error in reconstruction as compared to the polarizer-only design. This can possibly be attributed to the fact that the synthetic polarized-stereo PSF pair has a narrow baseline length, and also that the PSFs are no longer compact, leading to worse spatial resolution. \begin{figure}[t!] \centering \includegraphics[width=\columnwidth]{suppl_ablation_v1.pdf} \caption{\textbf{Comparing polarizer-phase mask design vs. polarizer-only design:} Simulated 3D USAF target reconstructions shown for the PS$^2$F case and for the polarized-stereo (polarizers-only) case. Showing depth color-coded reconstructed intensity plots. A polarizer-phase mask design gives a $\sim2\times$ lower depth RMS error in reconstruction as compared to the polarizer-only design, and poorer spatial resolution. Scale bars indicate $0.1$~mm.} \label{fig:ablation} \end{figure} \section{Supplementary: Experiments} \subsection{Pupil plane encoding using polarizers and phase mask} \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{fig10.pdf} \caption{\textbf{Polarizer-mask setup in the pupil plane.} (A) shows the 3D printed holder, in which two polarizer films (B) aligned to orthogonal polarizations (0 and 90 degree) will be placed, along with the fabricated DHPSF mask (C). (D) shows the entire setup kept in front of an LCD screen with polarized illumination. This demonstrates how such a setup allows for light from only one half of the mask to pass through. (E) shows an image of the polarizer-phase mask after keeping it in the pupil plane (imaged using a polarization camera).} \label{fig:polarizer-mask-setup} \end{figure} Figure~\ref{fig:polarizer-mask-setup} illustrates how the polarizer-phase mask encoding was achieved. \subsection{Phase mask fabrication} We used a two-photon photolithography 3D printer (Photonic Professional \textit{GT}, Nanoscribe GmbH~\cite{nanoscribe}) to print the phase mask structures using IP-Dip photoresist on a $700~\mu$m thick fused silica substrate. A $3$~mm diameter phase mask corresponding to the DHPSF phase mask profile was fabricated, with $2~\mu$m discretization in the $xy$-dimensions. For reliable fabrication, the phase profile was also quantized to 5 different levels. \subsection{Fluorescent sample preparation} The fluorescent bead sample was prepared by adding 10$\mu$m fluorescent beads solutions with varying concentrations in polydimethylsiloxane (PDMS) (Sylgard, Dow Corning; 10:1 elastomer:cross-linker weight ratio). The fluorescent cotton strand 3D sample was prepared by using a fluorescent highlighter on a tiny piece of a cotton ball, which consisted of strands with thicknesses ranging from $10$-$30~\mu$m. The fluorescent-highlighted sample was then placed in a PDMS solution. All samples were cured at room temperature for a minimum of 24 hours.\\ \subsection{Hyperparameter selection} In reconstruction, the optimization objective (Eq 5 in the main paper) has two hyperparameters - $\lambda_{TV}$ and $\lambda_{L1}$ - these control total variation (TV) and L1 regularization. The three broad types of experimental reconstructions shown in the submission (USAF target, fluorescent beads, fluorescent cotton strands) had different scene settings and properties, such as transmission vs. fluorescence illumination (different background signal statistics), extended vs. point-like scenes (different sparsity) and thus the hyperparameters were set accordingly to - \begin{itemize} \item USAF (planar) target: $(\lambda_{L1}, \lambda_{TV})=(0.02, 0.005)$ \item Fluorescent beads: $(\lambda_{L1}, \lambda_{TV})=(0.05, 0.00)$ \item Fluorescent strands: $(\lambda_{L1}, \lambda_{TV})=(0.002, 0.002)$ \end{itemize} For the same type of scene, high sensitivity to hyperparameter weights was not observed in our reconstructions. \section{Related Work}\label{sec:related-work} \input{sections/prior.tex} \section{Polarized Spiral Point Spread Function}\label{sec:proposed-method} \input{sections/methods.tex} \section{Simulations}\label{sec:simulations} \input{sections/simulations.tex} \section{Experimental Results}\label{sec:experiments} \input{sections/experiments.tex} \section{Conclusion}\label{sec:conclusions} \input{sections/conclusions.tex} \ifpeerreview \else \section*{Acknowledgments} The authors would like to thank Dong Yan for his help in preparing the fluorescent bead and fluorescent strands samples. This work was supported by NSF awards IIS-1730574, IIS-1730147, CCF-1652569, IIS-1652633, and EEC-1648451. A.-K.G. acknowledges partial financial support from the National Institute of General Medical Sciences of the National Institutes of Health (Grant No. R00GM134187), the Welch Foundation (Grant No. C-2064-20210327), and startup funds from the Cancer Prevention and Research Institute of Texas (Grant No. RR200025). \fi \bibliographystyle{IEEEtranN} \section{Conclusion and Further Scope}\label{sec:conclusions} We presented PS$^2$F for single shot, monocular depth estimation of extended (linear) structures where we separated the two lobes of the DHPSF into the two orthogonal states of polarization. This separation enabled us to remove ambiguities when estimating depth of line segments by breaking the inherent symmetry of DHPSF with respect to depth. The snapshot capabilities of PS$^2$F will enable faster microscopic imaging, including high resolution light sheet microscopy at real-time rates with fewer captures. We designed and demonstrated a compact physical realization of PS$^2$F with a single, polarization-sensitive camera, and showed that PS$^2$F results in $2\times$ or higher accuracy compared to current state-of-the-art phase masks such as DHPSF and the Tetrapod PSF. We believe that our approach of leveraging polarization multiplexed phase masks combined with polarization-sensitive cameras opens new avenues for high resolution snapshot depth imaging at micro and macro scales. PS$^2$F is inherently designed for extended/linear structures, and imaging such structures will involve high photon counts than when imaging single molecules or point-like emitters. Hence, we do not specifically deal with a low-photon count scenario in this work. Nevertheless, PS$^2$F can be adapted to low-photon count scenarios by using advanced recovery techniques including deep network-based approaches. Improvements such as modeling spatially varying PSF, as well as accounting for polarized input light will further improve PS$^2$F results. \section{Experimental Results}\label{sec:experiments} \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-experiments/fig11_v2.pdf} \caption{\textbf{Lab prototype.} (A) Our lab prototype is a $4f$ system consists of a polarization sensor (FLIR BFS-U3-51S5P-C) equipped with a $50$~mm F-mount lens. The phase mask and polarizers are on an XY stage. An achromat (AC254-050-A) forms the other half of the $4f$ system. We used an additional z-axis motorized translation stage to capture the PSF stack. (B) Two orthogonal polarizer pieces. (C) Images of the polarizers and phase mask with a camera focused on the pupil-plane.} \label{fig:4f-setup} \end{figure} We perform real-world imaging with PS$^2$F to verify our simulations. In this section we detail out the experiments and compare the PS$^2$F with the corresponding DHPSF. \subsection{Imaging setup description} We built a $4f$ imaging setup with the first lens being an achromatic lens with focal length $50$~mm, and the second lens being a Canon camera lens with effective focal length $=50$~mm. The polarizer-phase mask is added to the pupil plane of the $4f$ system. Fig. \ref{fig:4f-setup} shows the prototype. \subsubsection{Polarizer-Mask design} We fabricated a $3$~mm diameter DHPSF mask using a 3-D printer employing two-photon lithography. See Supplementary for more details on fabrication. The fabricated mask was the same mask used in the Section \ref{sec:simulations} to perform simulations of the VascuSynth dataset. To create a single-mask, single-sensor design for the PS$^2$F as outlined in Section \ref{sec:proposed-method}, two rectangular pieces of thin-film polarizers were laser cut and put side-by-side in a 3D printed holder. Both the polarizer pieces were laser cut such that their polarization orientation was $0\deg$ and $90\deg$ respectively. The fabricated DHPSF mask was attached to the back side of the 3D printed holder, carefully aligning the two mask halves w.r.t. to the polarizer edges under a bright-field 4x microscope. This polarizer-mask aperture is illustrated in Fig. \ref{fig:4f-setup}(B),(C). Further details can be found in the Supplementary figures. \subsubsection{PSF capture and calibration} To reconstruct 3D scenes, we use an experimentally captured PSF stack for the PS$^2$F. Using the polarization sensor, we image a $5~\mu$m pinhole over a range of [$-2.5$~mm, $+2.5$~mm] on either side of the focus plane. The experimentally-captured PSFs can be seen in Fig.~\ref{fig:psfstack-vis}. \subsection{Imaging experiments} With the fabricated mask and optical setup, we perform a series of experiments with 3D skew planar targets (using transmissive illumination) and with 3D fluorescence targets. We image using a polarization camera sensor (FLIR BFS-U3-51S5P-C) to obtain the individual PS$^2$F lobes in two polarization channels. By averaging across the Bayer pattern, we can readily obtain the image of the scene as imaged by a DHPSF, which we use for comparison purposes. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-experiments/psf_expt_v2.pdf} \caption{\textbf{PSF stacks captured with a lab prototype.} We achieved high quality separation between the two lobes in 0 and 90 degree images with our setup. We noticed no significant effect of any misalignment between the phase mask and the two polarizer halves. Scale bars indicate $100~\mu$m.} \label{fig:psfstack-vis} \end{figure} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/usaf-expt-compare3-v2.pdf} \caption{\textbf{Skew USAF target reconstruction.} Showing reconstructed depth map, YZ maximum intensity projection, zoom-in inset of highlighted ROI for (A) PS$^2$F, and for (B) DHPSF. Column (C) shows slice plots for marked yellow lines in the zoom-in insets, corresponding to Group (5,3) elements for PS$^2$F (top plot) and DHPSF (bottom plot). We were able to resolve up to Group (5,3) elements (12.40 $\mu$m linewidth) using the PS$^2$F, whereas only up to Group (4,5) elements (19.69 $\mu$m linewidth) using the DHPSF. Text below YZ MIP plots indicate the depth RMS error obtained for reconstruction when compared to ground truth, showing that the PS$^2$F performs $\sim2\times$ better. All scale bars indicate $1$~mm.} \label{fig:expt-usaf} \end{figure*} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/beads-expt-compare2-v2.pdf} \caption{\textbf{3D reconstructed fluorescent bead volumes.} The scatter plots show 3D reconstructed views with PS$^2$F and the DHPSF for bead densities approximately corresponding to (A) 3 beads/mm$^3$, (B) 5 beads/mm$^3$, and (C) 8 beads/mm$^3$. We captured ground truth information with a semi-aligned confocal ground truth (hence an approximate ground truth). Highlighted regions indicate various errors: incorrect estimation by DHPSF \textcolor{mygreen}{(solid green)}, missing estimation by DHPSF \textcolor{mygreen}{(dashed green)}, incorrect estimation by PS$^2$F \textcolor{myblue}{(solid blue)}, and incorrect estimation by both PSFs \textcolor{myyellow}{(solid yellow)}. Axes values are in units of $mm$.} \label{fig:expt-fluorescence} \end{figure*} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-experiments/vasc-expt-compare1-v2.pdf} \caption{\textbf{Experimental results on fluorescent cotton strands}. (A) PS$^2$F and (B) DHPSF. First column in (A), (B) show the captured scene images (polarization channels in the PS$^2$F image are shown in red and green colors). Second column in (A), (B) shows the reconstructions. Highlight ROIs (in yellow) are shown in the last column of (A), (B) -- depicting the inaccuracy of DHPSF, as well as the ability of PS$^2$F to reconstruct scenes with complex linear structures. In the last row (C), columns (from left-to-right) show color-coded depth maps obtained from confocal imaging (treated as ground truth), PS$^2$F reconstructions, and DHPSF reconstructions respectively. Note the confocal GT depth map is an approximate depth map, due to imperfect registration between GT and the imaged volumes.} \label{fig:expt=vasc} \end{figure*} \subsubsection{3D skew planar targets} We imaged a USAF planar target placed at a skew angle, and then performed single-shot 3D reconstruction of the target. Furthermore, we obtain a focal stack of the target so as to obtain ground truth depth map. The ground truth depth map was obtained from the focal stack using the Shape-from-focus method~\cite{peruz}. For the FOV being imaged, the relative intensity of each of the PS$^2$F lobes changed across the FOV, causing a spatial variance in the PSF. This was readily accounted for by adding a per-pixel, per-channel weight term to the 3D scene variable \textbf{x} in Eq. \ref{eq:optim_func}, and jointly optimizing for the 3D scene \textbf{x} and the weights. The results are depicted in Fig. \ref{fig:expt-usaf}. After modelling of the relative strengths of the PS$^2$F lobes as variables to be estimated during the optimization, we achieved an accurate reconstruction result matching the simulation results in terms of resolving Group (5,3) elements for the PS$^2$F. The YZ maximum intensity projections demonstrate that the skew plane of the USAF has been correctly estimated using the PS$^2$F, as compared to using the DHPSF. When judged against the ground truth, we see a $\sim50\%$ lower RMS error achieved with the PS$^2$F as compared to the DHPSF. \subsubsection{3D fluorescence imaging} We, next, imaged a set of prepared fluorescent samples, consisting of $10~\mu$m fluorescent beads in a PDMS substrate (for more information, see Supplementary). The bead concentration was varied from approximately $3$ to $8$ beads/mm$^3$. The resulting samples were $\sim2$~mm in thickness. We imaged these samples using the same mask as before, but in a $4f$ optical system using lenses of focal length $30$~mm and $50$~mm resulting in a $1.6\times$ magnification. We reconstructed using the same optimization algorithm as specified in Eq. \ref{eq:optim_func}, and compared against ground truth data obtained with a confocal microscope. The results are shown in Fig.\ \ref{fig:expt-fluorescence}. Bead localization performance using the DHPSF seems to get relatively worse as the bead concentration increases. In contrast, PS$^2$F enabled accurate results with increasing bead concentration. It is important to note that in this experiment, we use an optimization objective to perform bead reconstruction. More sophisticated and accurate methods (such as likelihood-based methods) exist for reconstruction of beads or point-like objects. Moreover, the ground truth (GT) data was captured using confocal imaging method. However, due to the complexity of matching/registering imaging volumes, we only obtain an approximated alignment or approximate registration between the GT and the imaging volume captured. Using the same optical setup as above, we also image fluorescent cotton strands suspended in a PDMS sample (see Supplementary for more information about sample preparation). Fig.~\ref{fig:expt=vasc} shows the results for the same. PS$^2$F method is better able to capture the 3D structure of the fluorescent strands - especially in parts with highly complex geometry. Imaging with PS$^2$F results in sharper reconstructions, which helps in reconstructing the fine, complex structure of the strands. The estimated depth map from the PS$^2$F is also more accurate as compared to the DHPSF case. \section{Introduction}\label{sec:introduction} } \IEEEPARstart{3}{D} scanning is crucial to a wide range of applications, including microscopy~\cite{fischer2011microscopy}, autonomous driving~\cite{arnold2019survey}, and robot-assisted surgeries~\cite{reiter2014surgical}. Among the multitude of approaches to measure depth, perhaps the hardest are those that involve passive, monocular, and snapshot measurements---a scenario that prioritizes compactness and time resolution. Estimating 3D information in this context relies on cues such as shading~\cite{ping1994shape} or defocus~\cite{levin2007image, zhou2009coded}; however, the underlying inverse problem is challenging and ill-posed. A key property of real cameras is that the defocus blur changes with the depth of scene points, which has been leveraged by prior work to obtain depth. However, the blur produced by conventional pupils is not conducive to robust depth estimation; to mitigate this, there has been significant interest in the design of engineered pupil plane masks in the form of an amplitude ~\cite{levin2007image}, or phase mask~\cite{pavani2009three,greengard2006depth,pavani2008high, shechtman2014optimal,badieirostami2010three,nehme2020deepstorm3d,nehme2020learning}. In particular, the seminal work of Pavani \textit{et al.} \cite{pavani2008high} has demonstrated that the so-called Double-Helix PSF (DHPSF), consisting of two lobes rotating about a center, can provide high spatial and depth resolutions, at least in the context of super-resolution localization microscopy. The majority of prior work on engineered PSFs focus on isolated point light sources; this is a consequence of their intended application---namely super-resolved localization of fluorescent particles---and the simplification that this assumption provides. However, real-world scenes often consist of more complex geometric primitives such as lines, edges, and curves~\cite{cho2011blur,joshi2008psf}; a PSF optimized for point sources is inadequate at recovering such complex geometries due to ambiguities in depth estimation. This is all the more important in applications involving linear structures including blood vasculature, neuronal networks (of the biological kind), and microtubules. We propose and evaluate polarized spiral PSF (PS$^2$F), a novel engineered PSF that is well-suited for linear structures. Our key enabling observation is that asymmetric PSFs achieve higher depth accuracy for scenes comprising of linear structures. We achieve this asymmetry by observing that the DHPSF is comprised of two lobes that rotate about a common center; by capturing images with the individual lobes, we can reliably estimate depth. We propose a compact realization of our approach with a novel polarization-based imaging system. PS$^2$F is generated by combining a DHPSF mask with orthogonal linear polarizers on the two halves of the mask. We then use a polarization sensor, capable of measuring images along four polarization angles in a snapshot by using a Bayer-like tiling. Finally, the 3D scene is estimated by solving a depth-dependent deconvolution problem. An overview of depth estimation with PS$^2$F is illustrated in Fig.~\ref{fig:fig1}. {\flushleft \textbf{Contributions.}} We propose a new engineered PSF for imaging linear structures and make the following contributions. \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-simulations/teaser-v2.pdf} \caption{\textbf{Polarized Spiral PSF (PS$^2$F) for single-shot 3D scene estimation.} We propose a novel monocular depth imager based on an engineered PSF. Numerous real-world scenes can be modeled as linear structures. We observe that asymmetric PSFs, such as a single lobe of the DHPSF, enable us to accurately estimate the 3D geometry of such linear structures. We leverage this observation and propose a compact polarization-based setup that produces accurate depth maps in a snapshot manner.} \label{fig:fig1} \end{figure*} \begin{itemize} \item We theoretically demonstrate that the DHPSF is ill-suited for 3D imaging of linear structures by analysing its Cram\'er-Rao lower bound (CRLB). \item We propose a compact polarization-based setup, using a novel polarizer-phase mask encoding that is capable of better 3D estimation in a snapshot manner. \item We demonstrate the advantages of PS$^2$F through several experiments with our lab prototype over a wide range of scene geometries. \end{itemize} {\flushleft \textbf{Limitations.}} At its core, PS$^2$F assumes that the scene intensity is unpolarized; if the scene is polarized, the resultant depth estimate may not be accurate. Our technique, like most techniques that rely on defocus blur, performs poorly for regions with little or no textures or structures. \section{Proposed Method}\label{sec:proposed-method} We begin with a brief background on the DHPSF, an analysis of its properties, and its performance for linear/extended structures. Subsequently, we introduce our proposed PSF and its associated reconstruction technique. \subsection{Background: DHPSF and mask design} This subsection outlines the background on the generation of the DHPSF \cite{pavani2009three} and other GL-based rotating PSFs \cite{lew2011corkscrew}. \subsubsection{Phase mask for GL-based rotating PSFs} Any propagating paraxial wave/beam can be expressed in terms of the orthogonal basis of GL modes, which are characterized by two integers $(n,m)$. Different combinations of the GL modes lead to varying properties of paraxial beams. In \cite{piestun2000propagation}, it was shown that beams that constituted of GL modes that are in arithmetic progression continuously rotated and scaled upon propagation. Using a beam corresponding to GL modes of (1,1), (5,3), (9,5), (13,7), it was shown in \cite{greengard2006depth} that the amplitude and phase profile of the beam at $z=0$ can be used as a mask in front of a lens (or in the Fourier plane of a $4f$ lens system) to generate a PSF that rotates with defocus. Subsequently, this mask was implemented in a phase-only manner using a variant of Gerchberg-Saxton optimization procedure with constraints included for the GL modal domain~\cite{pavani2008high}. The resulting phase function creates a PSF that rotates with changing point source depth, and is popularly known as the DHPSF due to its double helical structure. This PSF has been successfully used in 3D localization of fluorescent particles \cite{pavani2009three}\cite{grover2011photon}. \subsubsection{Properties of GL-based rotating PSFs} Piestun \textit{et al.}~\cite{piestun2000propagation} present a theory on paraxial rotating beams using GL modes, and characterizing the resulting rotation range, rotation rates, and beam scaling rates. Specifically, given a $4f$ system with lenses of focal length $f$, a GL-based mask designed with beam width $w_0$ for wavelength $\lambda$, and having GL modes lying on a single line with slope $V_1$, the following can be established: \begin{itemize} \item The total rotation in one direction of defocus is $(V_1\pi/2)$ \item The angle of the rotating PSF $$\phi(z) = \phi_0 + V_1\ \text{tan}^{-1}\left(\frac{z}{\lambda f^2/\pi w_0^2}\right)$$ \item The PSF will rotate by $(V_1\pi/4)$ over a depth of $\frac{\lambda f^2}{\pi w_0^2}$ . \end{itemize} Even though phase-only masks for rotating PSFs are generated after an optimization procedure~\cite{pavani2008high}, the above rates and ranges are still a good approximation to understand the properties of any GL-based rotating PSF. \subsection{Challenges in imaging linear/extended structures} Phase masks such as the DHPSF~\cite{pavani2009three}, or the Tetrapod PSF~\cite{shechtman2014optimal}, were designed for 3D estimation of \textit{point sources}. To image scenes with extreme fluorophore density, the concept of a scene made up of point-like emitters is not accurate. With increasing fluorophore/light source density, scenes consist of edge-like and linear structures. Using a DHPSF for recovering depth of a line then leads to an ambiguity between depth and orientation of the line. This is illustrated in Fig. \ref{fig:depth-ambiguity-illustration}. Given a line image formed using the DHPSF, there will always be a global depth ambiguity---two different depths can give rise to the same line image. This makes the DHPSF ill-suited for imaging edge/linear structures. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/dhpsf_depth_ambiguity.pdf} \caption{\textbf{Ambiguities caused by DHPSF for lines}. Estimation of depth and orientation of a line using the DHPSF is an ill-conditioned problem. There is an inherent global ambiguity in depth estimation, since for every line image, two possible candidate depths are possible.} \label{fig:depth-ambiguity-illustration} \end{figure} \subsection{Designing a PSF for 3D lines} A line with constant intensity blurred with an arbitrary PSF produces another line -- thereby leading to loss of information along the line direction. In case of the DHPSF, the loss of information implies that two lines at two different depths ($-\Delta z$ and $+\Delta z$) produce the same measurement. This ambiguity can however be resolved if we measure two separate images with each lobe of the PSF. Imaging with the individual DHPSF lobes separately will remove the global degeneracy issues (up to a $2\pi$ lobe rotation range). Next, we provide a theoretical analysis of the advantages to be obtained by such a separation. For estimation from data (image(s) in our case), we can use theoretical tools to understand what ultimate precision is possible. The Fisher information and Cram\'er-Rao Lower bound (CRLB) give the best possible precision that can exist by any estimator. In the case of an image of an isolated point source, the Fisher information matrix for estimating parameters $\theta = (x,y,z)$ is given by \cite{shechtman2014optimal}: \begin{equation} \label{eq:FI-point-source} [FI(\theta)]_{ij} = \sum_{k=1}^{N_p}\frac{1}{\mu_\theta(k) + \beta}\left(\frac{\partial \mu_\theta}{\partial \theta}\right)_i\left(\frac{\partial \mu_\theta}{\partial \theta}\right)^T_j, \end{equation} where $N_p$ is the number of pixels, $k$ is a variable for summing over all pixels, $T$ corresponds to the transpose operator, $\beta$ is the background Poisson noise level, and $\mu_\theta$ is the scaled PSF image for a point source at $\theta=(x,y,z)$ (scaled by the number of photons $N$). The diagonal elements of the inverse of the Fisher information matrix provides the CRLB for estimation of each of the parameters $\theta = (x,y,z)$. $$[\textrm{CRLB}(\theta)]_i \equiv [FI(\theta)]^{-1}_{ii} $$ We next extend the Cram\'er-Rao analysis to imaging of lines. Given a line image (or a one-sided edge image), the parameters to be estimated are the associated depth $z$ of the line/edge (at center of the patch), the associated orientation $\phi$, i.e. $\theta=(z,\phi)$. Thus, given a line image $\psi$, the CRLB estimates for depth $z$ is given by \begin{align} [FI(\theta)]_{ij} = \sum_{k=1}^{N_p}\frac{1}{\psi_\theta(k) + \beta}\left(\frac{\partial \psi_\theta}{\partial \theta}\right)_i\left(\frac{\partial \psi_\theta}{\partial \theta}\right)^T_j \label{eq:fi}\\ [\textrm{CRLB}(\theta)]_z \equiv [FI(\theta)]^{-1}_{zz} \label{eq:fiz} \end{align} The estimation of line orientation from a line path is easier than depth estimation, as orientation can be computed from the global structure of the line image. Hence, for simplicity of exposition, we focus on the analysis of the CRLB$_z$ values here. Analysis regarding the CRLB$_\phi$ values is shown in the Supplementary. Using Eqns~\ref{eq:fi}, \ref{eq:fiz} we can calculate the $\sqrt{\textrm{CRLB}_z}$ values for a given line image for all $(z,\phi)$. This calculation can be readily extended to line images captured using two separate PSFs. In such a case, the final Fisher information matrix is the sum of the individual Fisher matrices. Assuming a $4f$ system with 50~mm focal length lenses, and $N=100,000$ photons, $\beta=5$ photons/pixel, we compute the $\sqrt{\textrm{CRLB}_z}$ plots (as a function of $z$, $\phi$) for the DHPSF~\cite{pavani2008high}, the Tetrapod PSF~\cite{shechtman2014optimal}, and the individual DHPSF lobes with $25\%$ photons in each lobe. Note that $\sqrt{\textrm{CRLB}_z}$ obtained with a mask pair should be compared with a single mask having 2x SNR (2x total photons), as shown in \cite{nehme2020learning}. However, our polarizer-phase mask setup causes a $50\%$ light loss (see further in Section 3). Hence, for an appropriate comparison, we compare the individual DHPSF lobe pair, with other PSFs having $4\times$ the SNR. There are possible ways to remove this loss, with a setup demonstrated in \cite{nehme2020learning}, which will further lower the estimated $\sqrt{\textrm{CRLB}_z}$ values. Fig.\ \ref{fig:crlb-plots} illustrates the $\sqrt{\textrm{CRLB}_z}$ values (log-scale) for a given line at depth $z$ and having orientation $\phi$. The mean and standard deviation of the $\sqrt{\textrm{CRLB}_z}$ values are also shown in the individual insets. The DHPSF plot in Fig. \ref{fig:crlb-plots} shows several peaks with large magnitude. This occurs at specific $(z,\phi)$ where the line connecting the two lobes of the PSF and the scene line are either perpendicular or parallel to each other. The ability to discern lines at slightly differing depths is greatly reduced at these points due to symmetry. For the Tetrapod PSF~\cite{shechtman2014optimal} and the PhaseCam3D PSF~\cite{wu2019phasecam3d}, the plots in Fig. \ref{fig:crlb-plots} also show more peaks and ridges, but with lower magnitude compared to the DHPSF. For the PSF pair of the individual DHPSF lobes, we see that most of the peaks are removed, except for a central peak in the plot. The peak corresponds to the case when the PSF is in focus and the the line orientation is perpendicular to the PSF lobes. This central peak does not affect the effectiveness of separating lobes, because it is a narrow peak over a only small area in the parameter space of $(z,\phi)$. The average $\sqrt{\textrm{CRLB}_z}$ values for the individual DHPSF lobe pair is similar to the DHPSF with 4x SNR. Thus, with 4x lower SNR, the PS$^2$F is able to achieve a comparable $\sqrt{\textrm{CRLB}_z}$ value with other PSFs, and does not show consistent peaks and ridges as seen in the plots. Note that CRLB calculations only provide insight about local ambiguities or precision levels. However, we also observe that a PSF pair created out of separating the DHPSF lobes removes global ambiguities as well. Thus overall, a PSF pair created from the individual DHPSF lobes is better for estimating depths of line/edge patches. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/crlb_z_plots_v1.pdf} \caption{\textbf{CRLB comparison for various phase masks.} The figure shows $\sqrt{CRLB_z}$ log-intensity (log-mm) plots for line images, as a function of line depth $(z)$ and line orientation $(\theta)$. PS$^2$F achieves a comparable CRLB over the whole range of depth and orientations with fewer peaks, even with 4x lower SNR. In contrast, competing approaches have curves with large CRLB values, which introduce ambiguities in the depth estimate.} \label{fig:crlb-plots} \end{figure} \subsection{Using polarizers to separate out the DHPSF lobes in a single $4f$ optical system} Imaging with a PSF pair is possible by having two separate parallel $4f$ imaging systems with separate sensors, as demonstrated in \cite{nehme2020learning}. However, the DHPSF mask (and the Fresnel ring-based two-lobe masks) enjoy the property of PSF lobe separability. Partitioning such a mask into two halves along a particular axis, and allowing light through only one half leads to the creation of only a single rotating PSF lobe, as depicted in Fig.~\ref{fig:psf-partition-illustration}. The primary reason is that the light falling on one half is modulated to form one lobe, while the light falling on the other half is modulated to form the other lobe. Such a partitioning leads to partitioning of the DHPSF into two separate, distinct lobes. \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/fig4.pdf} \caption{\textbf{Effect of separating the phase mask.} The phase mask for generating the DHPSF has a special property that separating the phase profile along the highlighted red line yields a separation of the two lobes. We leverage this to implement a compact realization of Polarized Spiral PSF (PS$^2$F) using a polarization camera sensor and two polarizers in the pupil plane.} \label{fig:psf-partition-illustration} \end{figure} The separability of DHPSF lobes can be leveraged to build a compact imaging system. In PS$^2$F, we add two linear polarizers on each mask half, one oriented along $s$-polarization and the other oriented along $p$-polarization. This ensures that the $s$-polarization component is modulated using one mask half, and consequently the $p$-polarized component is modulated using the other mask half. The PSF created out of such a polarizer-phase mask encoding contains two lobes, but with each lobe being in a different polarization (either $s$ or $p$). This addition of polarizers causes a $50\%$ light loss, which was taken into consideration while estimating the $\sqrt{\textrm{CRLB}_z}$ values. Such a \textit{polarized spiral PSF} can be captured efficiently using a polarized camera sensor. A polarized camera sensor contains a 2D array of pixels with a Bayer pattern that has 90, 45, 135, 0 degree polarizers. After addition of the polarizers in the mask-plane, the orthogonally polarized DHPSF lobes can be imaged in their respective channels of a polarization sensor. This allows for a compact $4f$ system, allowing the imaging of a PSF pair using just a single $4f$ system channel, making the proposed method a single-shot, single-sensor method. \subsection{Imaging model and reconstruction procedure} We assume that the light from the scene is unpolarized in our imaging model. The imaging of a 3D scene with a depth-dependent PSF can be approximated as a sum of 2D convolutions between the depth-dependent PSF and the per-plane scene intensity: \begin{equation} \label{eq:imaging-model} I_c(x,y) = \sum_{z=z_{start}}^{z_{end}} h_c(x,y;z) * s(x,y,z) \end{equation} where $I_c(x,y)$ is the image intensity at $(x,y)$ in polarization channel $c$, $h_c(x,y;z)$ is the 2D PSF corresponding to depth $z$ and polarization channel $c$, and $s(x,y,z)$ is the scene intensity at point $(x,y,z)$. The goal of 3D reconstruction is to estimate a 3D matrix $s(x,y,z)$ from the noisy measurements of $I(x,y)$. There are numerous ways to solve this problem including regularized least-squares~\cite{boominathan2020phlatcam,yanny2020miniscope3d, xue2020single}, data learning-based~\cite{wu2019phasecam3d,chang2019deep}, and more recently, ones based on deep network-based regularizer~\cite{zhang20203d}. To keep the exposition simple, we demonstrate recovery by modelling the 3D reconstruction problem as a regularized least squares optimization problem. Specifically, we formulate the 3D scene estimation problem as: \begin{equation} \label{eq:optim_func} \underset{\textbf{x}}{\text{argmin}} \left\lVert \begin{bmatrix}I_0\\I_{90}\end{bmatrix} - S\begin{bmatrix}H_0\\H_{90}\end{bmatrix}\textbf{x}\right\rVert^2_2 + \lambda_{TV}\lVert\Psi(\textbf{x})\rVert_1 + \lambda_{L1}\lVert \textbf{x} \rVert_1 \end{equation} where $\textbf{x}$ is a 3D matrix of scene intensities, $H_0, H_{90}$ are the depth-specific PSF operators corresponding to the individual two lobes that are polarized to $0^\circ$ and $90^\circ$ states respectively, and $I_0, I_{90}$ are captured images in the $0^\circ$ and $90^\circ$ polarization channels respectively. $S$ is the summing operator that sums across the depth channels of $H_0\textbf{x}$ and $H_{90}\textbf{x}$ separately. We employ TV and L1 regularizers as scene priors, with $\lambda_{TV}, \lambda_{L1}$ as hyperparameters to control their regularization effects respectively. We solve the optimization problem using autograd functionality in PyTorch and the Adam optimizer~\cite{kingma2014adam} to leverage the speed of graphical processing units (GPUs). A key requirement for our technique is that the scene emits unpolarized light. This is required so as to ensure that both the lobes have produce PSFs with similar intensity levels that we can calibrate a priori. When the incident light is polarized, the blur kernels in Eq.~\ref{eq:imaging-model} will have an unknown scaling, which leads to a model mismatch; in an extreme scenario, we lose all the information in one of the lobes if the polarization angle of the incident light is orthogonal to the corresponding polarizer in the pupil plane. While this is a limitation of our technique, we largely encountered unpolarized light in our intended application of fluorescence microscopy. \section{Related Work}\label{sec:related-work} PS$^2$F draws motivation from the prior work on engineered PSFs. We discuss the relevant ones in this section. \subsection{Mask-based PSF encoding} Encoding PSFs by designing masks has been used in several works, including extended depth-of-field imaging~\cite{castro2004asymmetric, yang2007optimized,zammit2014extended}, 3D sensing in the macroscopic~\cite{zhou2009coded} and microscopic~\cite{pavani2009three,shechtman2014optimal} regimes, and in a lensless imaging setting~\cite{boominathan2020phlatcam}. The designed masks can be placed in front of the imaging lens or in the pupil plane of the imaging system. Earlier works include designing amplitude masks or coded apertures~\cite{levin2007image,zhou2009coded} -- such amplitude masks however lead to poor light efficiency and very coarse depth information. Several follow up works utilized phase-only modulation masks which include the rotating PSFs~\cite{pavani2009three,lew2011corkscrew,prasad2013rotating}, and PSFs obtained by optimizing the Fisher information~\cite{shechtman2014optimal}. Phase masks have also been designed in an end-to-end optimization setting, both for macroscopic~\cite{wu2019phasecam3d,chang2019deep} and microscopy~\cite{nehme2020deepstorm3d,nehme2020learning} tasks. \subsection{3D super-resolution microscopy using phase masks} Coupled with STORM~\cite{huang2008three} and PALM~\cite{betzig2006imaging, hess2006ultra}, masks that result in rotating PSFs with depth have been able to perform super-resolution microscopy~\cite{pavani2009three,lew2011corkscrew} of individual fluorophore emitters to nanometer-scale accuracy. The most successful of such designs is the DHPSF~\cite{pavani2009three}, a popular choice for 3D localization of point sources~\cite{huang2008three,juette2008three,badieirostami2010three}. 3D localization performance have been further improved by designing Tetrapod PSFs~\cite{shechtman2014optimal}, which were generated by directly optimizing the Fisher information of the phase mask for $(x,y,z)$ localization of point sources. Recently, deep learning techniques have enabled an end-to-end design framework that simultaneously learn phase masks as well as neural networks for obtaining 3D localizations from raw captures~\cite{nehme2020deepstorm3d, nehme2020learning}. Of particular interest is the work by Nehme \textit{et al.}~\cite{nehme2020learning}, where a pair of masks were designed in an end-to-end learning process. The two masks are placed in two parallel $4f$ imaging channel. The mask pair, coupled with a learnt deep network, allows for high accuracy 3D localization of single fluorophores over a large depth range and for a high density of point-like emitters. Their work focuses on super-resolution microscopy and on 3D localization of \textit{point-like} emitters, while our result tackles extended linear structures. \subsection{3D sensing of linear and extended structures} Human/animal bodies have a rich, dense 3D network of blood vessels running through them. Capturing the 3D structure of vasculature is usually done using variants of light-sheet microscopy (LSM)~\cite{girkin2018light,di2018whole,lugo20173d}, confocal microscope (CM)~\cite{st2005confocal,kelch2015organ}, optical projection tomography~\cite{bassi2011vivo}, and optical coherence tomographic angiography~\cite{Makita:06}. However, confocal and LSM methods involves scanning individual points, lines, or planes which makes the systems bulky and decreases the achievable imaging rate, making them non-ideal for fast dynamic 3D imaging. \subsection{Depth sensing using rotating PSFs} There is a rich body of work devoted to PSFs with lobes that rotate with defocus. These can be clustered into two groups based on their use of Gauss-Laguerre (GL) modes~\cite{pavani2008high} or Fresnel rings~\cite{prasad2013rotating}. GL-based rotating PSFs~\cite{pavani2009three,lew2011corkscrew} have been used for super-resolving and localizing single molecules in 3D~\cite{pavani2009three,grover2011photon,gustavsson20183d, bennett2020novel}. Both GL-based~\cite{quirin2013depth} and Fresnel-ring-based~\cite{berlich2016single,wang2017single} rotating PSFs have also been used for single-shot depth estimation of scenes. However, the first work employs a two phase-mask solution. The latter two employ a patch-based technique involving processing in the cepstrum domain. They only recover very simple scenes at low spatial and depth resolutions. Such recovery approaches do not extend to more complex geometry such as linear structures. \subsection{Using polarization channels for PSF encoding} There have been a few works that leverage polarization to help in 3D super-resolution imaging~\cite{roider2014axial,nehme2020learning,ikoma2021deep}. These works typically use a beam-splitter to split the incoming light into two separate $4f$ optical system channels which enabled individual channels to be encoded with a different mask. The images were then captured over two non-overlapping regions of a single sensor, or on two separate sensors. Such solutions however call for bulkier optics, and more importantly, a need for sub-pixel alignment to achieve high spatial resolution. We instead leverage snapshot polarimetric cameras that are naturally well-aligned, and hence easier to work with, and more compact. \section{Simulations}\label{sec:simulations} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{figs-simulations/fig7_v2.pdf} \caption{\textbf{Simulations on vascusynth dataset with varying complexity.} The figure shows reconstructions with various PSF designs on scenes with medium complexity (top row) and high complexity (bottom row). PS$^2$F is superior in depth estimation, especially in regions where there is higher vasculature (scene) complexity (see insets). Scale bars indicate $0.25$~mm.} \label{fig:vasc-scenes-dmaps} \end{figure*} To evaluate the performance of our proposed PS$^2$F, we perform extensive simulations on the \href{http://vascusynth.cs.sfu.ca/Welcome.html}{VascuSynth 2013} dataset~\cite{hamarneh2010vascusynth, jassi2011vascusynth}, which provides 10 simulated 3D volumes of vascular trees for each of 12 levels of complexity. Scene complexity is set by the number of bifurcations in the vascular tree (ranging from 1 to 56 in steps of 5). \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{figs-simulations/rebut_RMSE_compare.pdf} \caption{\textbf{Accuracy vs.\ scene complexity}. Depth estimation performance, for the proposed PS$^2$F and other comparing PSFs (averaged over 10 scenes for each complexity level, shaded regions show variance across the 10 scenes). This demonstrates the suitability of PS$^2$F, and that the separation of the DHPSF lobes helps in 3D reconstruction.} \label{fig:sim-compare} \end{figure} Each 3D vasculature volume in the dataset consists of $101\times101\times101$ voxels. These volumes were converted into 3D scenes of size $256\times256\times256$ through interpolation methods. This new volume was assumed to cover a volume of $1.76\times1.76\times5.00$ mm$^3$ with each voxel occupying a $6.9\times6.9\times19.5~\mu$m$^3$. We further assume a $4f$ system with both lenses having focal lengths $f=50$~mm. We render images of the vascular structures using several different masks of diameter $3$~mm in the Fourier plane. We assume that the light from the vasculature is monochromatic of wavelength $\lambda=532$~nm. We render images assuming occlusion of vasculature and that light is received only from the surface of the vasculature. We assume that the center of the 3D scene is $f=50$~mm away from the first lens, thus, the scene spans a defocus range of [$-2.5$~mm, $+2.5$~mm]. The masks (and their corresponding PSFs) used in rendering the simulated images were: \begin{itemize} \item DHPSF: We optimize for a GL-based DHPSF mask (following procedure in~\cite{pavani2008high}), based on the rotating paraxial beam with $w_0=0.4$~mm and having GL modes corresponding to $(1,1), (3,5), (5,9), (7,13)$. Full 180 degree lobe rotation was achievable over $2\frac{\lambda f^2}{\pi w_0^2}=5.3$~mm. \item PS$^2$F: Using the above mask, we construct an equivalent mask pair corresponding to PS$^2$F by partitioning the mask into two halves appropriately. \item Tetrapod PSF: We obtain a Tetrapod PSF mask designed for $550$~nm wavelength from~\cite{gustavsson20183d} and repurpose it for imaging over our specified $5$~mm depth range. \end{itemize} We then added Poisson and Gaussian read-out noise to the final measurements. For an appropriate comparison, we perform 3D reconstructions with individual single lobes (PS$^2$F), two-shot capture with DHPSF, and a two-shot capture using the Tetrapod PSF. To accurately account for the $50\%$ light loss in the PS$^2$F case, we also render out images with half the signal level as compared to the signal levels in DHPSF and Tetrapod PSF case. After obtaining the 3D volumetric estimate, we filter out points whose sum across the z-stack is lower than a fixed threshold. We then estimate depth using the index corresponding to a maximum-intensity-projection (MIP) of the 3D estimate. \subsection{Comparison with DHPSF and Tetrapod PSF} Fig. \ref{fig:sim-compare} shows the depth estimation performance for the proposed PS$^2$F, DHPSF, and the Tetrapod PSF. We compare the PSFs depth estimation using the Mean Absolute Error (MAE), Root Mean Square Error (RMSE), and the multi-scale Structural Similarity Index Measure~\cite{wang2003multiscale} (MS-SSIM) metrics. The entire set of results can be seen in the Supplementary. The proposed PS$^2$F method performs $\sim2\times$ better in terms of the RMS error for depth estimation. The depth estimation performance improvement is even higher ($\sim3\times$) for vasculature with more bifurcations, i.e. greater complexity. Depth estimation results for two example vasculature scenes can be seen in Fig. \ref{fig:vasc-scenes-dmaps}. In Fig. \ref{fig:vasc-scenes-dmaps} (top row), the vasculature scene is not very complex. The depth map estimation results are fairly similar for all the three PSFs (PS$^2$F, DHPSF, Tetrapod PSF), but the zoom-in insets highlight the errors seen in the DHPSF and Tetrapod PSF depth estimation. In Fig. \ref{fig:vasc-scenes-dmaps} (bottom row), the scene has greater complexity, and the PS$^2$F produces much better depth estimation than the other two PSFs. The zoom-in insets especially highlight the same. \begin{figure}[!t] \centering \centering \includegraphics[width=\columnwidth]{figs-simulations/usaf-sim-compare-part1-v2.pdf} \caption{\textbf{Simulated reconstruction results of a USAF target at a skew angle.} (A) Shows the depth map estimates for PS$^2$F, DHPSF, and Tetrapod PSF respectively. The XZ maximum intensity projections are shown in (B) in their respective columns, with ground truth shown in the inset. (C) shows zoom-in insets of depth maps (after thresholding to remove small artifacts) of highlighted regions in (A). PS$^2$F has a better reconstruction in lateral and axial dimensions as compared to the DHPSF and the Tetrapod PSF. Scale bars for (A), (B), (C) indicate $0.5$~mm, $0.5$~mm, and $0.1$~mm respectively.} \label{fig:sims-usaf} \end{figure} \subsection{Resolution performance} Furthermore, we analyse the reconstruction resolution performance by reconstructing a simulated skew USAF target. We assume a USAF target places at the focus plane, but tilted with respect to the optical axis so that its depth changing horizontally across the target; we render using the same three masks in the same $4f$ optical system as above. The results are shown in Fig.\ \ref{fig:sims-usaf}. The proposed PS$^2$F generates a better reconstruction than the DHPSF and Tetrapod PSF reconstructions, both in lateral and axial dimensions. As seen from the XZ MIP plots in Fig.\ \ref{fig:sims-usaf}, the PS$^2$F and Tetrapod PSF reconstructions are able to obtain the skew angle of the target, with the former being much better. The DHPSF fails to get the skew angle, due to the global depth ambiguity issues explained in Section \ref{sec:proposed-method}. In the lateral dimensions, the Tetrapod PSF is unable to reconstruct Group 4 and 5. This could be attributed to the fact that the PSF has a larger support. On the other hand, the PS$^2$F and DHPSF have concentrated lobes, and thus are able to reconstruct elements of Group (5,3) (12.40~$\mu$m linewidth) and (5,2) (13.92~$\mu$m linewidth) respectively. The obtained $xy$-resolution can be theoretically justified. The separation of lobes in the PS$^2$F method allows for resolution of finer scene features due to the compactness of the PSF. The resolution will be roughly determined by the width of the individual PSF lobes. Fitting a 2D Gaussian to each individual lobe, the average $2\sigma$-value for the Gaussian fit is $\sim1.74~\text{pixels} = 12~\mu$m in the object space. This readily matches with the fact that the PS$^2$F can resolve elements of Group (5,3), which has line elements that are spaced at $12.40~\mu$m distance. The diffraction-limited resolution of an Airy disk PSF for the same optical system configuration is given by $1.22\lambda f/D = 10.81~\mu$m, indicating that the $xy$-resolution obtained by the PS$^2$F is close to the diffraction-limited resolution ($1.15\times$) as well. For PS$^2$F, we also analysed the spread of reconstruction signal across the z-channel. For pixels with a significant signal level, we fit a 1D Gaussian curve to estimate this spread. The median $2\sigma$-value across z-channel obtained was $2\times304=608~\mu$m, which is an estimate of the axial resolution performance of the proposed PS$^2$F. Note that the diffraction-limited axial performance of an Airy disk PSF for the same optical system configuration is given by $4\lambda(f/D)^2 = 591~\mu$m. We are able to achieve $\sim1.03\times$ the diffraction-limited axial resolution with the proposed PS$^2$F. \section{Supplementary: Methods} \subsection{Properties of GL-based rotating PSFs} \cite{piestun2000propagation} presented a theory on paraxial rotating beams using GL modes, along with a detailed estimation of rotation range, rotation rates, and beam scaling rates. These theoretical values could be extended to give some theoretical description about GL-based rotating PSFs. However, to the best of our knowledge, this has not been shown. Here we present a calculation for the rotation range and rates for GL-based rotating PSFs. Consider a monochromatic paraxial beam with width $w_0$ and wavelength $\lambda$, which is generated from GL modes that lie on a single line in the GL modal plane with a slope $V_1$, expressed as: \begin{equation} n_j = V_1m_j + V_2 \ \ \ \ j=1,2,... \end{equation} where $V_1, V_2$ are integer constants, and $\{m_j\}$ are non-negative numbers in arithmetic progression. \cite{piestun2000propagation} showed that such a paraxial beam rotates at a rate \begin{equation} \frac{d\phi}{dz} = \frac{V_1}{1 + (z/z_R)^2} \end{equation} where $z_R = \frac{\pi w_0^2}{\lambda}$ is the Rayleigh length. A point with coordinates $(\rho_0,\phi_0)$ at $z=0$, follows the trajectory \cite{piestun2000propagation}: \begin{align} \rho &= \rho_0 \sqrt{1 + (z/z_R)^2} \\ \phi &= \phi_0 + V_1(arctan(z/z_R)) \end{align} which implies that the maximum rotation possible (in one direction) is $V_1(\pi/2)$. Thus, for the rotating beam case, over a distance of $z_R$, the beam rotates $V_1(\pi/4)$. The way to find out rotation and scaling rates for a GL-based rotating PSF is to find the equivalent Rayleigh length $z_R$ for it in a $4f$ optical system. Note that the GL modes arise from modal solutions of the Fresnel diffraction operator\cite{piestun2000propagation}. The Fresnel diffraction integral is given as follows \cite{goodman2005introduction}: \begin{equation} \label{eq:fresnel} U(u,v) = \frac{e^{jkz}}{j\lambda z}e^{j\frac{k}{2z}(u^2+v^2)}\mathcal{F}\{U(x,y)e^{j\frac{k}{2z}(x^2+y^2)}\}_{\frac{u}{\lambda z},\frac{v}{\lambda z}} \end{equation} where $j^2=-1$ and $k$ is the wave number. Eqn \ref{eq:fresnel} determines the wave profile $U(u,v)$ at $z$ distance away from an input wave profile $U(x,y)$. Thus, the phase term that determines the \emph{defocus} or \emph{propagation} by a distance $z$ is \begin{equation}\label{eq:rl-beam} \frac{k}{2z}(u^2+v^2) \end{equation} Similarly, upon solving the propagation integrals for a $4f$ system with defocus, the phase term that affects the defocus is given by \begin{equation}\label{eq:rl-4f} \text{exp}\left(j\frac{k}{2f}(\frac{\Delta z}{f})(u^2+v^2)\right) \end{equation} where $\Delta z$ is the distance from the in-focus plane. Comparing Eqns \ref{eq:rl-beam} and \ref{eq:rl-4f}, we can obtain the equivalent \textit{Rayleigh length} in the $4f$ system (say $z'_R$) \begin{align} \frac{1}{z_R} = \frac{z'_R}{f^2} \implies z'_R = f^2/z_R = \frac{\lambda f^2}{\pi w_0^2} \end{align} Thus, given a GL-based rotating PSF mask designed with beam width $w_0$, for wavelength $\lambda$ and $V_1$ slope in the GL modal plane: \begin{itemize} \item The total rotation amount in one direction of defocus is $V_1(\pi/2)$ \item The angle of the rotating PSF $\phi(z) = \phi_0 + V_1(\text{arctan}(\frac{z}{\lambda f^2/\pi w_0^2}))$ \item And over a depth of $\frac{\lambda f^2}{\pi w_0^2}$ the PSF will rotate by $V_1(\pi/4)$ radians. \end{itemize} \subsection{Analysis of CRLB$_{\phi}$} \begin{figure}[h!] \centering \includegraphics{suppl_crlbphi_plots.pdf} \caption{\textbf{CRLB$_{\phi}$ comparison for various phase masks.} The figure shows $\sqrt{CRLB_{\phi}}$ log-scale plots for line images, as a function of line depth $(z)$ and line orientation $(\theta)$. Individual insets depict the mean and standard deviation of the $\sqrt{\textrm{CRLB}_{\phi}}$ values, which are all very low.} \label{fig:crlb-phi} \end{figure} \begin{figure*}[!t] \centering \includegraphics[width=\textwidth]{fig5_v2.pdf} \caption{\textbf{Simulation methodology to evaluate PS$^2$F}. Given a 3D volume, we used a simulated stack of depth-dependent PSFs to obtain the two polarization images. We then computed the depth at each spatial point with the two polarization images as inputs, resulting in the point cloud shown on the right.} \label{fig:sims-vasc} \end{figure*} \begin{table*}[!t] \setlength{\tabcolsep}{.25cm} \renewcommand{\arraystretch}{1.3} \caption{Depth Estimation performance, averaged over 10 scenes for each scene complexity (noise level$=0.02$). $B$ refers to the number of bifurcations (B) in the 3D vascular structure. Showing mean absolute error (MAE) and RMS error (RMSE) values (all units in mm), and multi-scale SSIM~\cite{wang2003multiscale} (MS-SSIM) values. For the first two metrics, lower is better (indicated by $\downarrow$), and for MS-SSIM, higher is better (indicated by $\uparrow$).} \centering \begin{tabular}{|c||c|c|c|c|c|c|c|c|c|} \hline & \multicolumn{3}{c|}{PS$^2$F(proposed)} & \multicolumn{3}{c|}{DHPSF(x2)~\cite{pavani2009three}} & \multicolumn{3}{c|}{TetrapodPSF(x2)~\cite{shechtman2014optimal}}\\ \hline\hline & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$ & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$ & MAE $\downarrow$ & RMSE $\downarrow$ & MS-SSIM $\uparrow$\\ \hline B=6 & \textbf{0.198} & \textbf{0.292} & \textbf{0.990} & 0.284 & 0.46 & \textbf{0.990} & 0.352 & 0.541 & \textbf{0.990}\\ \hline B=16 & \textbf{0.225} & \textbf{0.334} & \textbf{0.980} & 0.494 & 0.781 & 0.979 & 0.532 & 0.811 & 0.979\\ \hline B=26 & \textbf{0.211} & \textbf{0.321} & \textbf{0.972} & 0.507 & 0.839 & 0.971 & 0.574 & 0.872 & 0.970\\ \hline B=36 & \textbf{0.224} & \textbf{0.344} & \textbf{0.969} & 0.548 & 0.904 & 0.967 & 0.585 & 0.906 & 0.966\\ \hline B=46 & \textbf{0.261} & \textbf{0.413} & \textbf{0.966} & 0.564 & 0.933 & 0.964 & 0.769 & 1.129 & 0.963\\ \hline \end{tabular} \label{tab:sim-compare} \end{table*} In the CRLB analysis performed in Sec 3.3, we consider an analysis for CRLB$_z$ only. Intuitively, the estimation of line orientation from the image of a line patch is an easier task as compared to estimating line depth. This is because line orientation can be readily estimated from the global structure present in the line image. To verify this, we perform CRLB$_{\phi}$ calculations with the same parameters ($N=100,000$ photon, $\beta=5$ photons/pixel). The $\sqrt{\textrm{CRLB}_{\phi}}$ plots are shown in Supplementary Fig.~\ref{fig:crlb-phi}. The $\sqrt{\textrm{CRLB}_{\phi}}$ values are very low, indicating that our intuition about line orientation estimation being an easier problem is correct. \section{Supplementary: Simulations} \subsection{VascuSynth dataset simulations} The VascuSynth dataset simulations pipeline is best illustrated by Fig.~\ref{fig:sims-vasc}. Table~\ref{tab:sim-compare} shows the comparative results for the proposed PS$^2$F, DHPSF, and the Tetrapod PSF over three metrics - Mean Absolute Depth Error, RMS Depth Error, and multi-scale SSIM~\cite{wang2003multiscale}. We also perform simulations under varying levels of noise, to judge the robustness of the proposed PS$^2$F. We simulate vasculature scene renderings with Poisson noise, and also Gaussian read-out noise of levels $0.02$, $0.05$, and $0.1$ - which correspond to PSNRs of 34dB, 26dB, and 20dB. Fig. \ref{fig:sim-noise-analysis} shows that with increasing noise, the proposed PS$^2$F reconstruction only worsens slightly, showing the robustness of the proposed PSF and method to noise. \begin{figure}[h!] \centering \includegraphics[width=0.8\columnwidth]{RMSE_noise.pdf} \caption{Depth Estimation performance, for different noise levels (averaged over 5 scenes for each complexity level, shaded regions show variance across the 5 scenes). This shows the robustness of the proposed PSF and technique.} \label{fig:sim-noise-analysis} \end{figure} \subsection{Ablation study: Comparison with polarized stereo pair} The proposed PS$^2$F has been created using a novel polarizer-phase mask design and the usage of a polarization-based camera sensor. We perform an ablation study, comparing our proposed PS$^2$F (polarizer-phase mask design) with a PSF created using just polarizer halves in the Fourier (pupil) plane. This creates images in the $s$, $p$-polarized channels that correspond to different half apertures (taking inspiration from \cite{bando2008extracting}). With such a polarized stereo-vision effect, where each polarization channel sees a slightly different perspective, we attempt reconstruction of a 3D USAF skew target in simulation. The reconstructed intensity plots (depth color-coded) for the PS$^2$F and the Polarized-Stereo PSF are shown in Fig.~\ref{fig:ablation}. A polarizer-phase mask design gives better spatial resolution, and a $\sim2\times$ lower depth RMS error in reconstruction as compared to the polarizer-only design. This can possibly be attributed to the fact that the synthetic polarized-stereo PSF pair has a narrow baseline length, and also that the PSFs are no longer compact, leading to worse spatial resolution. \begin{figure}[t!] \centering \includegraphics[width=\columnwidth]{suppl_ablation_v1.pdf} \caption{\textbf{Comparing polarizer-phase mask design vs. polarizer-only design:} Simulated 3D USAF target reconstructions shown for the PS$^2$F case and for the polarized-stereo (polarizers-only) case. Showing depth color-coded reconstructed intensity plots. A polarizer-phase mask design gives a $\sim2\times$ lower depth RMS error in reconstruction as compared to the polarizer-only design, and poorer spatial resolution. Scale bars indicate $0.1$~mm.} \label{fig:ablation} \end{figure} \section{Supplementary: Experiments} \subsection{Pupil plane encoding using polarizers and phase mask} \begin{figure}[!t] \centering \includegraphics[width=\columnwidth]{fig10.pdf} \caption{\textbf{Polarizer-mask setup in the pupil plane.} (A) shows the 3D printed holder, in which two polarizer films (B) aligned to orthogonal polarizations (0 and 90 degree) will be placed, along with the fabricated DHPSF mask (C). (D) shows the entire setup kept in front of an LCD screen with polarized illumination. This demonstrates how such a setup allows for light from only one half of the mask to pass through. (E) shows an image of the polarizer-phase mask after keeping it in the pupil plane (imaged using a polarization camera).} \label{fig:polarizer-mask-setup} \end{figure} Figure~\ref{fig:polarizer-mask-setup} illustrates how the polarizer-phase mask encoding was achieved. \subsection{Phase mask fabrication} We used a two-photon photolithography 3D printer (Photonic Professional \textit{GT}, Nanoscribe GmbH~\cite{nanoscribe}) to print the phase mask structures using IP-Dip photoresist on a $700~\mu$m thick fused silica substrate. A $3$~mm diameter phase mask corresponding to the DHPSF phase mask profile was fabricated, with $2~\mu$m discretization in the $xy$-dimensions. For reliable fabrication, the phase profile was also quantized to 5 different levels. \subsection{Fluorescent sample preparation} The fluorescent bead sample was prepared by adding 10$\mu$m fluorescent beads solutions with varying concentrations in polydimethylsiloxane (PDMS) (Sylgard, Dow Corning; 10:1 elastomer:cross-linker weight ratio). The fluorescent cotton strand 3D sample was prepared by using a fluorescent highlighter on a tiny piece of a cotton ball, which consisted of strands with thicknesses ranging from $10$-$30~\mu$m. The fluorescent-highlighted sample was then placed in a PDMS solution. All samples were cured at room temperature for a minimum of 24 hours.\\ \subsection{Hyperparameter selection} In reconstruction, the optimization objective (Eq 5 in the main paper) has two hyperparameters - $\lambda_{TV}$ and $\lambda_{L1}$ - these control total variation (TV) and L1 regularization. The three broad types of experimental reconstructions shown in the submission (USAF target, fluorescent beads, fluorescent cotton strands) had different scene settings and properties, such as transmission vs. fluorescence illumination (different background signal statistics), extended vs. point-like scenes (different sparsity) and thus the hyperparameters were set accordingly to - \begin{itemize} \item USAF (planar) target: $(\lambda_{L1}, \lambda_{TV})=(0.02, 0.005)$ \item Fluorescent beads: $(\lambda_{L1}, \lambda_{TV})=(0.05, 0.00)$ \item Fluorescent strands: $(\lambda_{L1}, \lambda_{TV})=(0.002, 0.002)$ \end{itemize} For the same type of scene, high sensitivity to hyperparameter weights was not observed in our reconstructions.
\section{\label{sec:level1} Introduction} The Wentzel-Krammers-Brillouin theory has been one of the powerful techniques with which one can obtain analytical solutions of moderate accuracy to the Schrödinger equation \cite{CarlBender,DPark,Griffiths,Bender_Numerology_WKB,HBrian}. However, like other approximation methods, the WKB theory has its limited applicability only in specific space regions, normally far away from the classical turning points. It is well known that the WKB theory works well under the conditions of a rapidly-oscillating phase associated with a slowly-varying amplitude in wave mechanics. The problems that can be solved under such conditions are usually referred to be in the semi-classical limits of quantum physics \cite{,Berry_1972}. Let us first review the WKB theory. Consider the time independent Schrödinger equation in one dimension (the x coordinate) with the classical momentum $p(x) = \sqrt{2m[E-V(x)]}$, where $m$ is the particle mass, $E$ is the total energy, and $V(x)$ is the potential energy function, \begin{equation}\label{eq:1} \psi''(x) + \frac{p^{2}}{\hbar^{2}}\psi(x) = 0 \end{equation} The first order WKB approximation gives the fundamental solutions, \begin{equation} \psi(x) = \frac{1}{\sqrt{p}}e^{ \pm \frac{i}{\hbar}\int^{x}{pdt} } \end{equation} Because the Schrödinger equation is a second order linear differential equation, the complete approximated solution with a relative error of order $\hbar$ to the exact solution should be a linear combination of the two fundamental solutions, \begin{equation}\label{eq:3} \psi(x) = \frac{A}{\sqrt{p}}e^{ \frac{i}{\hbar}\int^{x}{pdt} } + \frac{B}{\sqrt{p}}e^{ -\frac{i}{\hbar}\int^{x}{pdt} } \end{equation} where A and B are two undetermined constants. One can see that the WKB approximation has a simple structure regardless of how complex the potential $V(x)$ is, as long as the phase integral can be carried out. However, the difficulty in extending the WKB theory emerges as soon as one proceeds to higher order approximations\cite{KembleArticle,Kemblebook,Hyouguchi_Letters}. For example, it is a tedious task to derive the connection formulas as the order increases, because the inherent singularity in the neighborhoods of the turning points (where the classical momentum is vanishing) can lead to stronger divergence for the higher order WKB series. Infact, going beyond the leading order of WKB theory without mathematical complications is very difficult, authough there exists techniques such as Borel summation of exact WKB analysis.\cite{Exact_WKB,Gergo_exact_WKB} Therefore, it is desirable to keep the simple solution form while improving the applicability range of the usual WKB theory. In this paper we propose an extended WKB approximation dubbed the alternating WKB (a-WKB) method by considering a heuristic ansatz of the wave function. The theoretical approach of the a-WKB method is presented in Sec. \uppercase\expandafter{\romannumeral2}. In Sec. \uppercase\expandafter{\romannumeral3}, the a-WKB method is applied to the eigenvalue problem of a harmonic oscillator and compared with the usual WKB method. In the a-WKB method, the extra correction terms due to the concept of reflection wave are shown that the contribution appear in only \textit{amplitude}, instead of in the \textit{phase} in Sec. \uppercase\expandafter{\romannumeral4} . The scattering problem of a particle from a repulsive inverse-square potential is investigated in Sec. \uppercase\expandafter{\romannumeral5}. We summarize our main findings and review previos works containing similar coupled differential equations to ours in Sec. \uppercase\expandafter{\romannumeral6}. \section{Theory of the alternating WKB (a-WKB) Approximation} We first consider a more general ansatz of the wave function than that of Eq. \eqref{eq:3} \begin{equation}\label{eq:4} \psi(x) = a(x)e^{iS} + b(x)e^{-iS} \end{equation} where $a(x)$ and $b(x)$ are the two complex amplitude variables which are in general not constants. Then the indefinite phase integral is defined as follows, \begin{equation} S(x) \equiv \frac{1}{\hbar}\int^{x}{pdt} \end{equation} We then substitute Eq. \eqref{eq:4} into Eq. \eqref{eq:1} and require the following two conditions, \begin{subequations} \begin{align} &\psi'(x) = i\frac{p}{\hbar}[ a(x)e^{iS} - b(x)e^{-iS}] \label{eq:6a} \\ &\psi''(x) = -\frac{p^{2}}{\hbar^{2}}[ a(x)e^{iS} + b(x)e^{-iS} ] \label{eq:6b} \end{align} \end{subequations} This procedure has a part of flavor in the method of variational parameters, but it is not exactly the same. The first condition implies that the variation of amplitudes $a(x)$ and $b(x)$ are sufficiently small compared with their respective phase terms. The second condition suggests a plane-wave like solution with the wave number being the classical momentum. Therefore, we obtain a set of coupled differential equations for the amplitude variables, \begin{equation}\label{eq:7} \left\{ \begin{array}{rcl} &a'(x) &= -\frac{p'}{2p}a(x) + \frac{p'}{2p}b(x) e^{-2iS} \\ \\ &b'(x) &= -\frac{p'}{2p}b(x) + \frac{p'}{2p}a(x) e^{2iS} \end{array} \right. \end{equation} We interpret these equations as follows. If a particle is moving forward (to the positve x direction from left to right), then the amplitudes $a(x)$ refers to the forward-going wave amplitude. However, when it confronts a finite potential, a reflection wave would be generated simultaneously as the wave proceeds into the interaction regions. Thus, the amplitude $b(x)$ refers to the backward-going wave amplitude. Such physics of incoming and reflected waves is governed by the coupling terms. One way to decouple the equations is to employ the method of averaging. Assuming the conditions of highly oscillating phases, the contributions of the inhomogeneous terms with the factors $e^{2iS}$ and $e^{-2iS}$ , respectively, would fade away in the background. This way we can obtain the first order WKB solutions, but the interference effects are totally ignored. Instead, we use the perturbation method to explore more about the mathematical structure of these equations. Let us rewrite the coupled differential equations in the matrix form by introducing a sufficiently small bookkeeping parameter $\epsilon$, \begin{equation} \frac{d}{dx} \begin{bmatrix} a \\ b \end{bmatrix} = -\frac{p'}{2p} \begin{bmatrix} 1 & -\epsilon e^{-2iS} \\ -\epsilon e^{2iS} & 1 \end{bmatrix} \begin{bmatrix} a \\ b \end{bmatrix} \end{equation} Consider now an expansion of the amplitudes, $a(x)$ and $b(x)$, up to the first order perturbation in $\epsilon$, \begin{equation} \frac{d}{dx} \begin{bmatrix} a_{0} + \epsilon a_{1} \\ \\ b_{0} + \epsilon b_{1} \end{bmatrix} = -\frac{p'}{2p} \begin{bmatrix} 1 & -\epsilon e^{-2iS} \\ -\epsilon e^{2iS} & 1 \end{bmatrix} \begin{bmatrix} a_{0} + \epsilon a_{1} \\ \\ b_{0} + \epsilon b_{1} \end{bmatrix} \end{equation} where $a_{0}$ and $b_0$ denote the unperturbed (zeroth order) terms, and $a_1$and $b_1$ represent the perturbed terms, respectively. Accordingly, we have the unperturbed equations, \begin{equation}\label{eq:10} \frac{d}{dx} \begin{bmatrix} a_{0} \\ b_{0} \end{bmatrix} = -\frac{p'}{2p} \begin{bmatrix} a_{0} \\ b_{0} \end{bmatrix} \end{equation} The solutions to Eq. \eqref{eq:10} are easily obtained, \begin{equation}\label{eq:11} \begin{bmatrix} a_{0} \\ b_{0} \end{bmatrix} \sim \begin{bmatrix} \frac{1}{\sqrt{p}} \\ \\ \frac{1}{\sqrt{p}} \end{bmatrix} \end{equation} We see that these are the usual first order WKB solutions. The first order perturbation equations ($O(\epsilon)$) are, \begin{equation}\label{eq:12} \frac{d}{dx} \begin{bmatrix} a_{1} \\ b_{1} \end{bmatrix} = -\frac{p'}{2p} \begin{bmatrix} a_{1} - b_{0}e^{-2iS} \\ \\ b_{1} - a_{0}e^{2iS} \end{bmatrix} \end{equation} By substituting Eq. \eqref{eq:11} into Eq. \eqref{eq:12}, we have \begin{equation}\label{eq:13} \frac{d}{dx} \begin{bmatrix} a_{1} \\ b_{1} \end{bmatrix} = -\frac{p'}{2p} \begin{bmatrix} a_{1} - \frac{1}{\sqrt{p}}e^{-2iS} \\ \\ b_{1} - \frac{1}{\sqrt{p}}e^{2iS} \end{bmatrix} \end{equation} Eq. \eqref{eq:13} can be solved by the Green’s function method in the region of $[0, x_{0}]$, namely, \begin{equation}\label{eq:14} \frac{d}{dx} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} + \frac{p'}{2p} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} = \begin{bmatrix} \delta(x-\xi) \\ \delta(x-\xi) \end{bmatrix} \end{equation} where $x_{0}$ is the nearest turning point in the neighborhood of the initial point $x = 0$, and the impulse delta function is applied right at the point of $x = \xi$. With the homogeneous boundary conditions, we have \begin{equation}\label{eq:15} \begin{bmatrix} G_{1}(0,\xi) \\ G_{2}(0,\xi) \end{bmatrix} = \begin{bmatrix} 0 \\ 0 \end{bmatrix} \end{equation} and the homogeneous solutions of Eq. \eqref{eq:14} are \begin{equation}\label{eq:16} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} = \begin{bmatrix} \frac{C_{1}}{\sqrt{p}} \\ \\ \frac{C_{2}}{\sqrt{p}} \end{bmatrix} \qquad \mbox{for}\qquad 0<x<\xi \end{equation} and \begin{equation} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} = \begin{bmatrix} \frac{C_{3}}{\sqrt{p}} \\ \\ \frac{C_{4}}{\sqrt{p}} \end{bmatrix} \qquad \mbox{for}\qquad \xi<x<x_{0} \end{equation} where $C_{1}$, $C_{2}$, $C_{3}$, and $C_{4}$ are the undetermined coefficients. Plugging the homogeneous conditions of Eq. \eqref{eq:15} into Eq. \eqref{eq:16}, we obtain $C_{1} = C_{2} = 0$. Next, the jump conditions are used such that \begin{equation} \begin{bmatrix} \frac{C_{3}}{\sqrt{p(\xi)}} \\ \\ \frac{C_{4}}{\sqrt{p(\xi)}} \end{bmatrix} = \begin{bmatrix} 1 \\ 1 \end{bmatrix} \qquad \mbox{for}\qquad \xi<x<x_{0} \end{equation} Hence we obtain the constants $C_{3}=C_{4}=\sqrt{p(\xi)}$. As a result, we have the Green’s functions associated with the Eq. \eqref{eq:13}, \begin{equation} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} = \begin{bmatrix} 0 \\ 0 \end{bmatrix} \qquad \mbox{for}\qquad 0<x<\xi \end{equation} and \begin{equation} \begin{bmatrix} G_{1}(x,\xi) \\ G_{2}(x,\xi) \end{bmatrix} = \begin{bmatrix} \frac{\sqrt{p(\xi)}}{\sqrt{p(x)}} \\ \\ \frac{\sqrt{p(\xi)}}{\sqrt{p(x)}} \end{bmatrix} \qquad \mbox{for}\qquad \xi<x<x_{0} \end{equation} With these, we can construct the function $a_{1}$ and $b_{1}$, \begin{eqnarray} \begin{bmatrix} a_{1}(x) \\ \\ b_{1}(x) \end{bmatrix} &= \begin{bmatrix} \int_{0}^{x_{0}}{G_{1}(x,\xi)\frac{ 1 }{\sqrt{p(\xi)}}\frac{p'(\xi)}{2p(\xi)}e^{-2iS}}d\xi \\\\ \int_{0}^{x_{0}}{G_{2}(x,\xi)\frac{1}{\sqrt{p(\xi)}}\frac{p'(\xi)}{2p(\xi)}e^{2iS}}d\xi \end{bmatrix} \nonumber\\ \nonumber\\ &= \begin{bmatrix} \frac{1}{\sqrt{p(x)}}\int_{0}^{x}{ \frac{p'(\xi)}{2p(\xi)}e^{-2iS(\xi)} }d\xi \\\\ \frac{1}{\sqrt{p(x)}}\int_{0}^{x}{ \frac{p'(\xi)}{2p(\xi)}e^{2iS(\xi)} }d\xi \end{bmatrix} \end{eqnarray} \\ Now we set $\epsilon = 1$, and obtain the first order perturbation for the amplitudes \begin{equation} \begin{bmatrix} a \\ b \end{bmatrix} = \begin{bmatrix} a_{0} + a_{1} \\ b_{0} + b_{1} \end{bmatrix} = \begin{bmatrix} \frac{ D_{1} }{\sqrt{p}} + \frac{1}{\sqrt{p(x)}}\int_{0}^{x}{ \frac{p'(\xi)}{2p(\xi)}e^{-2iS(\xi)} }d\xi \\ \\ \frac{ D_{2} }{\sqrt{p}} + \frac{1}{\sqrt{p(x)}}\int_{0}^{x}{ \frac{p'(\xi)}{2p(\xi)}e^{2iS(\xi)} }d\xi \end{bmatrix} \end{equation} According to perturbation theory, it is quite resonable to put the contants $D_{1}$ and $D_{2}$ at the unperturbed terms instead of putting infront of whole amplitude functions $a(x)$ and $b(x)$. Here we properly choose the constants $D_{1}$ and $D_{2}$ to be the factor $e^{\mp i\frac{\pi}{4}}$. Therefore, we obtain the first order perturbation solutions of the a-WKB theory with remained constant $D$, which can be determined by normalization, \begin{equation} \begin{aligned} \psi_{\text{aWKB} }(x) = D \left(\frac{ e^{-i\frac{\pi}{4}} }{\sqrt{p(x)}} + \frac{ 1 }{\sqrt{p(x)}}\int_{0}^{x}{ f(\xi) }d\xi\right) e^{\frac{i}{\hbar}\int_{x}^{x_{0}}{p}dt } \\ +D \left(\frac{ e^{i\frac{\pi}{4}} }{\sqrt{p(x)}} + \frac{1}{\sqrt{p(x)}}\int_{0}^{x}{ g(\xi) }d\xi\right) e^{-\frac{i}{\hbar}\int_{x}^{x_{0}}{p}dt } \end{aligned} \end{equation} The function $f(x)$ and $g(x)$ are defined as $\frac{p'(x)}{2p(x)}e^{-2iS(x)}$ and $\frac{p'(x)}{2p(x)}e^{2iS(x)} $. Notice that in the a-WKB theory, the correction terms appear only in the amplitude. We shall verify this statement in Sec. \uppercase\expandafter{\romannumeral4} by considering the potential well problem. \section{Harmonic Oscillator} \begin{figure}\label{Fig.1} \includegraphics[scale=0.35]{harmonic_oscillator.pdf \caption{\label{harmonic_oscillator} Harmonic oscillator in atomic units.} \end{figure} We apply the a-WKB method to the harmonic oscillator problem. We focus on the ground state because the WKB method usually gives worse approximation to the physical system at lower energy states, especially for the ground state energy. Consider a harmonic oscillator with $m = \omega = 1$, $E_{0} = 1/2$, and the turning points $x_{0} = 1$ (the atomic units) as shown in Fig.\ref{harmonic_oscillator}. Here $m$ is the mass, $\omega$ is the angular frequency, $E_{0}$ is the ground state energy, and $x_{0}$ represents the classical turning point. There are two distinguishing regions. The classically allowed (forbidden) region is associated with the $x$ domain where $E - V (x) > 0 \left( E - V (x) < 0 \right)$. In this case, the classically allowed region is referred to that of $[-1, 1]$. The a-WKB wave function in classical allowed region is, \begin{equation}\label{eq:24} \psi_{\text{aWKB}}(x) = \frac{D}{\sqrt{p(x)}} \left[ \left( e^{-i\frac{\pi}{4}} + \int_{0}^{x}{ f(\xi) }d\xi\right) e^{\frac{i}{\hbar}\int_{x}^{x_{0}}{p}dt } + c.c. \right] \end{equation} Here we use the abbreviation $c.c.$ which stands for complex conjugate. In the classically forbidden region ($|x| > 1$), as $|x|$ increases, the coupled differential equations, Eq. \eqref{eq:6a} and Eq. \eqref{eq:6b}, reduce to a single equation since the potential of harmonic oscillator grows monotonically. Therefore, the a-WKB solution reduces to the same wave function as the first order WKB approximation. One can see that the merit of choosing the constants $D_{1}$ and $D_{2}$ to be the factor $\mp e^{i\frac{\pi}{4}}$ can naturally satisfied the matched asymptotic expansion in the leading order at both side of classical allowed and classical forbidden region. The overall constant $D$ is then determined by the normalization condition. In Fig. \ref{ComparisonOnWholeRegion} we compare the a-WKB solution with the usual WKB approximation, as well as the exact wave function. We see a significant improvement on the wave function, thus demonstrating the utility of the a-WKB approximation. \begin{figure} \includegraphics[scale=0.35]{comparison_1_2_order_ho.pdf \caption{\label{ComparisonOnWholeRegion} Comparison of WKB(dotted line) and a-WKB(dashed line) approximation of ground state harmonic oscillator in the atomic unit, $m= \hbar= 1, E_{0}=0.5$, and exact solution(solid green line) } \end{figure} Because the a-WKB approximation contains the extra integration terms in the amplitude, it is interesting to study the well-known deficiency of the second order WKB approximation (see Fig. \ref{second_order_WKB}). We see that the second order WKB approximation does not even have the correct behavior around the turning point. This should be contrasted to the a-WKB approximation. The a-WKB approximation gives a better solution than the first order WKB approximation in the classical allowed region. As can be seen from Eq. \eqref{eq:24}, the a-WKB wave function is identical to the first order WKB wave function at the initial point $x = 0$. The perturbation due to the reflection wave starts to develop as the wave approaches to the turning point. The amplitudes in correction terms perform negative contribution, thus gradually reducing the divergent wave function and improving the solution. In contrast, the $2^{nd}$ order WKB wave function is a correction in the phase term so the (wrong) diverging behavior cannot be rectified. \begin{figure} \includegraphics[scale=0.35]{second_order_wkb.pdf} \caption{\label{second_order_WKB}Comparison of 1st(dotted line) and 2nd order(thinner solid line) WKB wave function in ground state harmonic oscillaor with atomic unit. Exact solution is represented by bolded dashed. } \end{figure} \begin{figure} \includegraphics[scale=0.35]{conditional_validation_of_second_order_wkb.pdf} \caption{\label{Order Comparison between 1st and 2nd WKB series}Comparison of $1^{ \text{st} }$ and $ 2^{\text{ nd } }$order WKB series in the classical region of ground state harmonic oscillator, turning point $x_{0} = 1 $. $S_{0}(x), S_{1}(x), S_{2}(x)$ are represented by solid line, dashed line and dotted line. } \end{figure} To see more closely the details, the WKB series up to the second order is written as \begin{equation} \psi(x) = e^{ \frac{1}{\hbar}S_{0}(x)+S_{1}(x)+ \hbar S_{2}(x) } \end{equation} where $S_{0}, S_{1}, S_{2}$ refer to the phase integral of $\int^{x}pdx'$, first and second order of WKB series, repectively. By definition, the series behaves asymptotically if the following relations are satisfied, \begin{alignat}{3} \frac{1}{\hbar} S_{0}(x) &>> S_{1}(x) \nonumber \\ S_{1}(x) &>> \hbar S_{2}(x) \label{eqn28}\\ \hbar S_{2}(x) &<< O(1) \nonumber \end{alignat} In Fig. \ref{Order Comparison between 1st and 2nd WKB series}, we compare the functions $S_{0}$, $S_{1}$ and $S_{2}$ in the classical region. We see clearly that the $2^{nd}$ of order WKB series does not behave asymptotically in the whole region. The function $S_{2}$ is actually larger than $S_1 (x)$ , and exhibits stronger divergence than the first order WKB approximation near the turning point. It is thus clear why the a-WKB theory has its supremacy over the usual WKB approximation. \section{Quantization condition of the a-WKB approximation} We found that the main contribution of correction terms in a-WKB wave function is on the amplitudes. Does it also improve the phase parts of the wave function? Infact, the quantization condition of the WKB approximation comes from the phase parts of exponent of wave function\cite{,Bender_Numerology_WKB}. Therefore we should derive the quantization condition associated to the a-WKB approximation to see if there has any extra correction. Consider a potential well with two simple turning points. The n-th order a-WKB wave function(n times reflections) has the form in classical allowed region \begin{equation} \psi_{\text{aWKB}}(z) \sim A(z)e^{\frac{i}{\hbar}\phi(z)} \equiv e^{ \frac{i}{\hbar}(S_{0}+\hbar S_{\text{a}}) } \end{equation} \begin{widetext} where $S_{a}(z)$ is defined as \begin{equation} e^{iS_{a}(z)} \equiv \frac{1}{\sqrt{p}} ( e^{-i\frac{\pi}{4}} + \int_{0}^{x}dt_{1} \frac{p'}{2p}e^{-2iS_{0}} \int_{0}^{t_{1}}dt_{2} \frac{p'}{2p}e^{2iS_{0}} \int_{0}^{t_{2}}dt_{3}...\int_{0}^{t_{n}}dt_{n-1}\frac{p'}{2p}e^{(-1)^{n}2iS_{0}} ) \end{equation} \end{widetext} the applitude $A(z)= e^{iS_{a}(z)}$ corresponds to reflection waves. $S_{0}(z) = \frac{1}{\hbar}\int{dz\sqrt{ E-V(z) }}$is now denoted as the phase integral. We raise the independent variable $x$ to be complex variable $z$. To represent a physical state, the a-WKB wave function should be single valued as the state evaluating along the contour $\gamma$ which encircles the two turning points. Therefore, we have to require(Keller has suggested the same condition satified by the $1^{\text{st}}$ order of WKB wave function\cite{Keller1958}) \begin{equation}\label{eq:29} \Delta \left[( S_{0}+\hbar S_{a} ) \right] = \oint_{\gamma}{ \frac{d}{dz}(S_{0}+\hbar S_{a}) }dz =2n\pi\hbar \end{equation} It is shown that the Eq. \eqref{eq:29} gives the Bohr-Sommerfield quantization condition with 1/2 phase shift. Therefore, the a-WKB wave function with all order of reflections contains no extra contribution to the quantization condition(see Appendix for the derivation). This verifies the statement that the correction of the a-WKB wave function is presented purely on \textit{amplitude}. \section{Scattering by Centrifugal Potential} Consider a particle scattering from a three dimensional centrifugal potential of the form of $l(l+1)/r^2$\cite{,Friedrich,BHolstein}. The particle beam is incident from infinity, with a total energy $E$ and the angular momentum $l$. The classical turning point is denoted as $r_{0}=\sqrt{(l(l+1)/E)}$. We shall focus on the classically allowed region which contains the essential information we need for the scattering problem. Following the same steps of the a-WKB theory in deriving Eqs. \eqref{eq:3} to \eqref{eq:13}, we have the Green’s function problem, now in a different spatial domain, $r \in [r_{0} ,\infty]$ \begin{equation} \vec{G}(r;\xi) = \begin{bmatrix} G_{1}(r;\xi) \\ G_{2}(r;\xi) \end{bmatrix} = \begin{bmatrix} \frac{C_{1}}{\sqrt{p(r)}} \\ \frac{C_{2}}{\sqrt{p(r)}} \end{bmatrix} \quad \mbox{for}\quad r_{0}<r<\xi \\ \\ \end{equation} \begin{equation} \vec{G}(r;\xi) = \begin{bmatrix} G_{1}(r;\xi) \\ G_{2}(r;\xi) \end{bmatrix} = \begin{bmatrix} \frac{C_{3}}{\sqrt{p(r)}} \\ \frac{C_{4}}{\sqrt{p(r)}} \end{bmatrix} \quad \mbox{for}\quad \xi<r<\infty \end{equation} By applying the homogeneous boundary condition at $r \to \infty$, we have \begin{equation} C_{3}=C_{4}=0 \end{equation} The constants $C_{1}$ and $C_{2}$ can then be determined by the jump condition \begin{equation} C_{1}=C_{2}=-\sqrt{p(\xi)} \end{equation} The associated Green’s functions are \begin{equation} \begin{aligned} \vec{G}(r;\xi) = \begin{bmatrix} \frac{ \sqrt{p(\xi)}}{\sqrt{p(r)}} \\ \\ \frac{ \sqrt{p(\xi)}}{\sqrt{p(r)}} \end{bmatrix} \end{aligned} \qquad \mbox{for}\qquad r_{0}<r<\xi \end{equation} The particular solution of Eq. \eqref{eq:13} corresponding to the repulsive potential is \begin{equation} \vec{A}_{1,p}(r) = \begin{bmatrix} a_{1,p} \\ b_{1,p} \end{bmatrix} = \int_{r_{0}}^{\infty}{ \hat{g}(r;\xi)\vec{F}(\xi) }d\xi \end{equation} where $\hat{g}$ stands for a 2 by 2 matrix, and $\vec{F}$ refers to the source term. \begin{equation} \hat{g}(r;\xi) \equiv \begin{bmatrix} G_{1}(r;\xi) & 0 \\ G_{2}(r;\xi) & 0 \end{bmatrix} \end{equation} \begin{equation} \vec{F}(\xi) = \begin{bmatrix} F_{1}(\xi) \\ F_{2}(\xi) \end{bmatrix} = \begin{bmatrix} \frac{-1}{ \sqrt{p(\xi)} } \frac{p'(\xi)}{2p(\xi)} e^{-2iS} \\ \\ \frac{-1}{ \sqrt{p(\xi)} } \frac{p'(\xi)}{2p(\xi)} e^{2iS} \end{bmatrix} \end{equation} Therefore, the particular solution is, \begin{equation} \vec{A_{1,p}}(r) = \begin{bmatrix} a_{1,p}(r) \\ b_{1,p}(r) \end{bmatrix} = \begin{bmatrix} \frac{-1}{ \sqrt{p} }\int_{r}^{\infty} {\frac{p'}{2p}e^{-2iS} }d\xi \\\\ \frac{-1}{ \sqrt{p} }\int_{r}^{\infty} {\frac{p'}{2p}e^{2iS} }d\xi \end{bmatrix} \end{equation} Now we set the parameter $\epsilon = 1$, and expand the perturbation series up to the first order, \begin{equation} \begin{aligned} \vec{A}(r) &\equiv \vec{A_{0}}(r) + \vec{A_{1}}(r) \\ \\ &= \begin{bmatrix} a_{0}(r) + a_{1}(r) \\ b_{0}(r) + b_{1}(r) \end{bmatrix} = \begin{bmatrix} \frac{D_{1}}{\sqrt{p}} + \frac{-1}{ \sqrt{p} }\int_{r}^{\infty} {\frac{p'}{2p}e^{-2iS} }d\xi \\ \frac{D_{2}}{\sqrt{p}} + \frac{-1}{ \sqrt{p} }\int_{r}^{\infty} {\frac{p'}{2p}e^{2iS} }d\xi \end{bmatrix} \end{aligned} \end{equation} where $D_{1}$ , $D_{2}$ are then chosen as the factor of $e^{\mp i\frac{\pi}{4}}$ . Finally, the a-WKB wave function is \begin{equation} \psi_{\text{aWKB}}(r) = \frac{D}{\sqrt{p(r)}} \left( e^{-i\frac{\pi}{4}} -\int_{r}^{\infty} {\frac{ p' }{ 2p }e^{-2iS } }d\xi \right) e^{ \frac{i}{\hbar} \int_{r}^{r_{0}}{ p dt } } + c.c. \end{equation} Here we consider the Langer modification for a regular wave function behavior near the origin in the radial Schrödinger equation. It amounts to replace the $l(l+1)$ factor in the potential function by the semi-classical $l(l+1/2)^2$ factor\cite{Berry_1972,Friedrich,Langer}. This modification turns the classical momentum $p(r)= \sqrt{2mE- \frac{l(l+1)\hbar^{2}}{r^{2}}}$ into the effective momentum $p^{eff} (r)= \sqrt{2mE-\frac{l(l+1/2)^{2}\hbar^{2}}{r^{2}}}$ , and the a-WKB wave function is written as, \begin{equation} \begin{aligned} \psi_{\text{aWKB}}(r) =& \frac{D}{\sqrt{p^{ \text{eff} }(r)}} \left( e^{-i\frac{\pi}{4}} -\int_{r}^{\infty} {\frac{ p^{ '\text{eff} } }{ 2p^{ \text{eff} } }e^{-2iS } }d\xi \right) e^{ \frac{i}{\hbar} \int_{r}^{r_{0}}{ p^{ \text{eff} } dt } } \\ \\ &+ c.c. \end{aligned} \end{equation} Setting $m = \hbar = l = 1$, and $E = 0.5$. In this case, the turning point is $x_{0} = 1.5$, and the effective potential is \begin{equation} p^{ \text{eff} }(r) = \sqrt{ ( 1-\frac{4/9}{ r^{2} } ) } \end{equation} In Fig. \ref{Centrifugal} we compare the a-WKB and the WKB wave functions, as well as the exact solution. We see that the a-WKB wave function performs better approximation than the usual WKB near the turning point. \begin{figure} \includegraphics[scale=0.35]{awkb_wkb_on_the_scattering.pdf} \caption{\label{Centrifugal} Wave function of WKB(dotted line) and a-WKB(dashed line) to the centrifugal model are compared with exact solution(solid line) in the region of $[1.5, \infty]$ with $m=\hbar=l=1$, and $E = 0.5$, where WKB wave function is represented by dotted line, aWKB wave function is represented by dashed line, and exact solution is solid line.} \end{figure} \section{Concluding Remarks and the Historical Review of Different Approaches of the WKB Theory} Historically, the coupled differential equation of Eq. \eqref{eq:7} or a similar one has been derived in several earlier works with different approaches. Each of them is very instructive to us for developing the a-WKB approximation. Therefore we intend to review these methods, briefly, in this section. Zwaan was the first one who introduced complex varialbles to explore the WKB theory . Following the Zwaan's treatment, Kemble, Fr\"{o}man and Fr\"{o}man provided a fine analysis\cite{KembleArticle,Kemblebook,Froman}. First, they give the approximate solution to the Schr\"{o}dinger equation in the similar form of Eq. \eqref{eq:4} \begin{equation}\label{eq:43} \psi(z) = a_{i}(z)f_{i}(z) + a_{v}(z)f_{v}(z) \end{equation} where $z$ is the complex variables, $f_{i}$ and $f_{v}$ are defined as \begin{equation} \left\{ \begin{array}{rcl} &f_{i}(z) \equiv \frac{1}{\sqrt{p}} e^{iS} \\ \\ &f_{v}(z) \equiv \frac{1}{\sqrt{p}} e^{-iS} \end{array} \right. \end{equation} Impose the condition \begin{equation} \psi'(z) = a_{i}f'_{i} + a_{v}f'_{v} \end{equation} This is physically equivalent to the short wave limit of our condition Eq. \eqref{eq:6a}, albeit using complex coordinate. On the second condition, Kemble, Fr\"{o}man and Fr\"{o}man found the governing equation of $f(z)$, which is different from our condition of Eq. \eqref{eq:6b}. \begin{equation}\label{eq:46} f'' + \left[(\frac{p}{\hbar})^{2} -Q \right]f = 0 \end{equation} where \begin{equation}\label{eq:47} Q = \frac{3}{4}(\frac{p'}{p})^{2} - \frac{p''}{2p} \end{equation} Having Eq. \eqref{eq:43} to Eq. \eqref{eq:47}, they obtained the coupled differential equation \begin{equation}\label{eq:48} \left\{ \begin{array}{rcl} &a_{i}'(z) = \frac{i\hbar}{2} \frac{Q}{p}[ a_{i} + a_{v}e^{-2iS} ] \\ \\ &a_{v}'(z) = -\frac{i\hbar}{2} \frac{Q}{p}[ a_{v} + a_{i}e^{2iS} ] \end{array} \right. \end{equation} The Eq. \eqref{eq:48} is similar to our coupled differential equation of Eq. \eqref{eq:7}, but it is not exactly the same because Eq. \eqref{eq:46} and Eq. \eqref{eq:6b} are different. On the other hand, it is shown that Bremmer's method can be used to derive the same coupled differential equation as Eq. \eqref{eq:7}\cite{Bremmer}. We follow the M. V. Berry's procedure\cite{Berry_1972} which is more straightforward. Start from the wave function \begin{equation}\label{eq:49} \psi(x) = \psi_{+} + \psi_{-} \equiv \frac{b_{+}(x)}{\sqrt{p(x)}}e^{ iS(x) } + \frac{b_{-}(x)}{\sqrt{p(x)}}e^{ -i{S(x)}} \end{equation} associated with the following two conditions \begin{equation}\label{eq:50} \psi'_{+}(x) = -\frac{p'}{2p}\psi_{+} + \frac{ip}{\hbar}\psi_{+} + \frac{p'}{2p}\psi_{-} \end{equation} and \begin{equation}\label{eq:51} \psi'_{-}(x) = -\frac{p'}{2p}\psi_{-} - \frac{ip}{\hbar}\psi_{-} + \frac{p'}{2p}\psi_{+} \end{equation} Combing from Eq. \eqref{eq:49} to Eq. \eqref{eq:51}, we have the coupled differential equations \begin{equation}\label{eq:52} b'_{\pm}(x) = \frac{p'}{2p}b_{\mp}(x)e^{\mp 2iS } \end{equation} Eq. \eqref{eq:52} is essentially equivalent to the coupled differential equation we derived in Eq. \eqref{eq:7}. Notice that the function $b_{-}(x)$ and $b_{+}(x)$ have the same effect to our amplitude functions in Eq. \eqref{eq:4} which describe reflections when the wave confronts a nonzero potential. Instead of considering the B.Cs from Berry ($b_{+}(\infty)=1$ and $b_{-}(\infty)=0$), we assume that neither $b_{+}(x)$ nor $b_{-}(x)$ would dominate the other one, and we give them the same weighting at the reference point, say $x=0$. \begin{equation}\label{eq:53} b_{+}(0) = b_{-}(0) = 1 \end{equation} Use the conditions of Eq. \eqref{eq:53} and integrate from $x=0$, we have \begin{equation} b_{\pm}(x) = 1 + \int_{0}^{x}{ \frac{p'}{2p}e^{ \mp 2iS }dt } \end{equation} Substitute into Eq. \eqref{eq:49} \begin{equation}\label{eq:55} \psi(x) \sim \frac{1}{\sqrt{p}} \left( 1 + \int_{0}^{x}{ \frac{p'}{2p}e^{-2iS}dt } \right)e^{ \frac{i}{\hbar}\int_{0}^{x} pdt } + c.c. \end{equation} Eq. \eqref{eq:55} corresponds to the a-WKB wave function in the pure classical allowed region. On the turning point problem, we use the connection formulas to match the wave function at classical allowed and classical forbidden region in the leading order. This essentially gives the a-WKB approximation modified by the constant $D$ from the classical forbidden side, which can be determind by normalization condition. \begin{equation} \psi(x) = \frac{D}{\sqrt{p}} \left( e^{-i\frac{\pi}{4}} + \int_{0}^{x}{ \frac{p'}{2p}e^{-2iS}dt } \right)e^{ \frac{i}{\hbar}\int_{0}^{x} pdt } + c.c. \end{equation} The conditions of Eq. \eqref{eq:50} and Eq. \eqref{eq:51} proposed by Berry come from the idea of dividing the potential function into a series of steps so that we can successfully express the related function with a constant in each step. Such method was also performed by the numerical approach so-called transfer matrix\cite{CFHuang}. In the transfer matrix method, the wave function takes the form of \begin{equation} \psi(x) = t(x)e^{\frac{i}{\hbar}p(x)x} + r(x)e^{-\frac{i}{\hbar}p(x)x} \end{equation} where $p(x)$ is the classical momentum as usual. For instance, we can write the wave function in $j$-th step \begin{equation} \psi(x) = t_{j}e^{\frac{i}{\hbar}p_{j}(x)x} + r_{j}e^{-\frac{i}{\hbar}p_{j}(x)x} \end{equation} By applying the continuous conditions on the $\psi(x)$ and $\psi'(x)$ at the joint between $j$-th, $j_{+1}$-th step and taking the first order of Taylor expansion, one can obtain \begin{equation}\label{eq:59} \frac{d}{dx} \begin{bmatrix} t(x) \\ r(x) \end{bmatrix} = \Gamma(x) \begin{bmatrix} t(x) \\ r(x) \end{bmatrix} \end{equation} with \begin{equation} \Gamma(x) \equiv \begin{bmatrix} -\frac{i}{\hbar}xp'(x) - \frac{p'}{2p} & \frac{p'}{2p}e^{-2ixp(x)/\hbar } \\ \frac{p'}{2p}e^{2ixp(x)/\hbar } & \frac{i}{\hbar}xp'(x) - \frac{p'}{2p} \end{bmatrix} \end{equation} As can be seen, the coupled differential equation of Eq. \eqref{eq:59} is identical to Eq. \eqref{eq:7} if we simply replace $t(x)$ and $r(x)$ by $a(x)e^{-\frac{i}{\hbar}xp}e^{iS}$ and $b(x)e^{\frac{i}{\hbar}xp}e^{-iS}$. However, the transfer matrix $\Gamma(x)$ is useful to calculate the wave function only with the uncoupled differential equation, namely, ignoring off diagonal terms. As one go further to solve the wave function with the transfer matrix containing off diagnal terms, the integrands of $\frac{p'}{2p}e^{-2ixp(x)/\hbar }$ and $\frac{p'}{2p}e^{2ixp(x)/\hbar }$ become mathematically cumbersome. Bremmer and Berry's coupled differential equation of Eq. \eqref{eq:52}, as well as ours of Eq. \eqref{eq:7} seems to be more intuitive because the complicated coupled terms in the transfer matrix $\Gamma(x)$ are now naturally formulated in the phase integral $e^{\pm 2iS}$ and amplitude function $\frac{p'}{2p}$. We have developed an alternating approximate scheme dubbed the a-WKB method by proposing an ansatz for the wave function together with two auxiliary conditions. This method is based on an extension of the usual WKB approximation with the extra perturbative correction terms in the amplitude. It is shown that a-WKB method is equivalent to the and Berry's treamnet by using the different conditions. We have demonstrated that the a-WKB method can be applied to the harmonic oscillator problem and a scattering problem. In particular, the a-WKB approximation serves as a proper way to avoid the more involved connection formulas as one proceeds to use higher order WKB approximations. In general, the a-WKB method performs much better than the first order WKB approximation around the turning points and the resulting analytical formulas are still quite manageable mathematically. \\ \begin{acknowledgments} This work was supported by the Ministry of Science and Technology of Taiwan through MOST 108-2221-E-002-002-MY3. We acknowledge the National Center for High-performance Computing (NCHC) for providing computing resources. \end{acknowledgments}
\section{Introduction} Two sample testing of equality of distributions, which is the homogeneity hypothesis, is fundamental in statistics and has a long history that dates back to \cite{kolmogorov1933,smirnov1948,cramer1928,von1928}. The literature on this topic can be categorized by the data type considered. Classical tests designed for low dimensional data include \cite{bickel1969, bickel1983, friedman1979, henze1988, schilling1986} among others. For recent developments that are applicable to data of arbitrary dimension, we refer to Energy Distance (ED) \citep{szekely2004} and Maximum Mean Discrepancy (MMD) \citep{sejdinovic2013nips}. To suit high dimensional regimes \citep{hall2005, aoshima2018survey}, extensions of Energy Distance and Maximum Mean Discrepancy were studied in \cite{zhu2019interpoint, chakraborty2019new, gao2021two}. Some other interesting developments of Energy distance for data residing in a metric space include \cite{lyons2013} and \cite{klebanov2005n}. In this paper, we focus on functional data that are random samples of functions on a real interval, e.g. $[0, 1] $ \citep{ramsay2004functional, hsing2015theoretical, doi:10.1146/annurev-statistics-041715-033624, davidian2004introduction}. Two sample inference for functional data is gaining more attention due to the explosion of data that can be represented as functions. A substantial literature has been devoted to comparing the mean and covariance functions between two groups of curves, see \cite{fan1998test, cuevas2004anova,10.2307/20441490, doi:10.1080/15598608.2010.10412005, doi:10.1198/jasa.2010.tm09239, horvath2012inference,zhang2014one, STAICU20151, pini2016interval, 10.1093/biomet/asw033, zhang2019new, guo2019new, yuan2020hypothesis, 10.1214/21-EJS1802}. However, the underlying distributions of two random functions can have the same mean and covariance function, but differ in other aspects. Testing homogeneity, which refers to a hypothesis testing procedure to determine the equality of the underlying distributions of any two random objects, is thus of particular interest and practical importance. The literature on testing the homogeneity for functional data is much smaller and restricted to the two-sample case based on either fully observed \citep{cabana2017permutation, benko2009common, wynne2020kernel,krzysko2021two} or intensely measured functional data \citep{hall2007two, jiang2019asymptotics}. For intensively measured functional data, a presmoothing step is often adopted to each individual curve in order to construct a smooth curve before carrying out subsequent analysis, which may reduce mean square error \citep{10.2307/24310140} or, with luck, remove the noise, a.k.a. measurement error, in the observed discrete data. This results in a two-stage procedure, smoothing first and then testing homogeneity based on the presmoothed curves. For example, \cite{hall2007two, jiang2019asymptotics} adopted such an approach. In addition, the energy distance in \citep{szekely2004} could be extended to the space of $L^2$ functions to characterize the distribution differences of random functions, provided the functional data are fully observed without errors \citep{klebanov2005n}. Specifically, the energy distance between two random functions $X$ and $Y$ is defined as \begin{align} \label{eq:ed} \text{ED}(X,Y) = 2E[\| X-Y \|_{L^2}] - E[\| X-X' \|_{L^2}] - E[\| Y-Y' \|_{L^2}], \end{align} where $X',Y'$ are i.i.d copies of $X,Y$ respectively and $ \| X-Y \|_{L^2} = (\int_{0}^1 (X(t) - Y(t))^2 dt)^{1/2} $. According to the results of \cite{lyons2013} and \cite{klebanov2005n}, $ \text{ED}(X,Y) $ fully characterizes the distributions of $X$ and $Y$ in the sense that $ \text{ED}(X,Y) \geq 0 $ and $ \text{ED}(X,Y)=0 \Leftrightarrow X =^d Y $, where we use $X=^d Y$ to indicate that $X,Y$ are identically distributed. Given two samples of functional data $\{ X_i \}_{i=1}^n$ and $\{Y_i\}_{i=n+1}^{n+m},$ which are intensively measured at some discrete time points, the reconstructed functions, denoted as $\{ \widehat{X}_i \}_{i=1}^n$ and $\{ \widehat{Y}_i\}_{i=n+1}^{n+m}$, can be obtained using the aforementioned presmoothing procedure. Then $\text{ED}(X,Y)$ can be estimated by the $U$-type statistic \begin{multline} \label{eq:2step} \text{ED}_n (X,Y) = \frac{2}{mn} \sum\limits_{i_1=1}^{n} \sum\limits_{i_2=n+1}^{n+m} \| \widehat{X}_{i_1}-\widehat{Y}_{i_2} \|_{L^2} \\ - \frac{2}{n(n-1)} \sum\limits_{1 \leq i_1 < i_2 \leq n} \| \widehat{X}_{i_1}-\widehat{X}_{i_2} \|_{L^2} - \frac{2}{m(m-1)} \sum\limits_{n+1 \leq i_1 < i_2 \leq n+ m} \| \widehat{Y}_{i_1}-\widehat{Y}_{i_2} \|_{L^2}. \end{multline} While the above presmoothing procedure to reconstruct the original curves may be promising for intensely measured functional data, it has not yet been utilized to our knowledge, perhaps due to the technical and practical challenges to implement this approach as the level of intensity in the measurement schedule and the proper amount of smoothing are both critical. First, the distance between the reconstructed functional data and its target (the true curve) needs to be tracked and reflected in the subsequent calculations. Second, such a distance would depend on the intensity of the measurements and the bandwidth used in the presmoothing stage. Neither is easy to nail down in practice. Furthermore, in real world applications, such as in longitudinal studies, each subject often may only have a few measurements, leading to sparse functional data \citep{yao2005functional}. Here presmoothing individual data no longer works and one must borrow information from all subjects to reconstruct the trajectory of an individual subject. The PACE approach in \citet{yao2005functional} offers such an imputation method, yet it does not lead to consistent estimates of the true curve for sparsely observed functional data as there are not enough data available for each individual. Consequently, the quantities $E[\| X-Y \|_{L^2}], E[\| X-X' \|_{L^2}]\ \text{and} \ E[\| Y-Y' \|_{L^2}]$ in equation \eqref{eq:ed} are not consistently estimable as these expectations are outside the corresponding $L^2$ norms. \cite{10.2307/24773028} reduced the problem to testing the homogeneity of the scores of the two processes by assuming that the random functions are finite dimensional. Such an approach would not be consistent either as the scores still cannot be consistently estimated for sparse functional data. To our knowledge, there exists no consistent test of homogeneity for sparse functional data. In fact, it is not feasible to test full homogeneity based on sparsely observed functional data as there are simply not enough data for such an ambitious goal. This seems disappointing, since although it has long been recognized that sparse functional data are much more challenging to handle than intensively measured functional data, much progress has been made to resolve this challenge. For instance, both the mean and covariance function can be estimated consistently at a certain rate \citep{yao2005functional, li2010uniform, zhang2016sparse} for sparsely observed functional data. Moreover, the regression coefficient function in a functional linear model can also be estimated consistently with rates \citep{yao2005regression}. This motivated us to explore a less stringent concept of homogeneity that can be tested consistently for sparse functional data. In this paper we provide the answer by proposing a test of marginal homogeneity for two independent samples of functional data. For ease of presentation we assume that the random functions are defined on the unit interval $[0, 1]$. \begin{definition} Two random functions $X$ and $Y$ defined on [0, 1] are marginal homogeneous if $$\ \; X(t) =^d Y(t) \text{ for almost all } t \in [0,1] .$$ \end{definition} From this definition we can see that, unlike testing homogeneity that involves testing the entire distribution of functional data, testing marginal homogeneity only involves simultaneously testing the marginal distributions at all time points. This is a much more manageable task that works for all sampling designs, be it intensively or sparsely observed functional data, and it is often adequate in many applications. Testing marginal homogeneity is not new in the literature and has been investigated by \cite{zhu2019interpoint, chakraborty2019new} for high-dimensional data. They show that the marginal tests can be more powerful than their joint counterparts under the high dimensional regime. In a larger context, the idea of aggregating marginal information originates from \cite{zhu2020distance}, where they consider a related problem of testing the independence between two high-dimensional random vectors. For real applications of testing marginal homogeneity, taking the analysis of biomarkers over time in clinical research as an example, comparing differences between marginal distributions of the treatment and control groups may be sufficient to establish the treatment effect. To contrast stocks in two different sectors, the differences between marginal distributions might be more important than the differences between joint distributions. In addition, differences between marginal distributions can be seen as the main effect of differences between distributions. Thus, it makes good sense to test marginal homogeneity, especially in situations where joint distribution testing is not feasible or inefficient. Let $\lambda$ be the Lebesgue measure on $\mathbb{R}$. The focus of this paper is to test \begin{align} \label{hypotheses} \begin{array}{c} H_0: X(t) =^d Y(t) \text{ for almost all } t \in [0,1], \\ \text{versus} \\ H_A : \text{there exists a set } \mathbb{T} \subseteq [0,1] \text{ such that } \lambda(\mathbb{T}) >0 \text{ and } X(t) \neq^d Y(t) \text{ if } t \in \mathbb{T}. \end{array} \end{align} This can be accomplished through the marginal energy distance (MED) defined as: \begin{align} \label{MED} \text{MED}(X,Y)=\int 2E \left[ \left| X(t) - Y(t) \right| \right] - E \left[ \left| X(t) - X'(t) \right| \right] - E \left[ \left| Y(t) - Y'(t) \right| \right]dt , \end{align} where $X'$ and $ Y'$ are independent copies of $X$ and $Y$ respectively. Indeed, MED is a metric for marginal distributions in the sense that $ \text{MED}(X,Y) \geq 0 $ and $\text{MED}(X,Y)=0 \Leftrightarrow X(t) = ^d Y(t)$ for almost all $t \in [0,1]$. A key feature of our approach is that it can consistently estimate $\text{MED}(X,Y)$ for all types of sampling plans. Moreover, $ E \left[ \left| X(t) - Y(t) \right| \right] $, $ E \left[ \left| X(t) - X'(t) \right| \right] $ and $ E \left[ \left| Y(t) - Y'(t) \right| \right] $ can all be reconstructed consistently for both intensively and sparsely observed functional data. Such a unified procedure for all kinds of sampling schemes may be more practical as the separation between intensively and sparsely observed functional data is usually unclear in practical applications. Moreover, it could happen that while some of the subjects are intensively observed, others are sparsely observed. In the extremely sparse case, our method can still work if each subject only has one measurement. Measurement errors (or noise) are common for functional data, so it is important to accommodate them. If noise is left unattended, there will be bias in the estimates of MED as the observed distributions are no longer the true distributions of $X$ and $Y$. One might hope that the measurement errors can be averaged out during the estimation of the function $E[ | X(t) - Y(t) |]$ in $\text{MED}(X,Y)$ in analogy to estimating the mean function $\mu(t) = E[X(t)]$ or covariance function $C(s,t) = E[(X(t) - \mu(t))(X(s) -\mu(s))] $ \citep{yao2005functional}. However, this is not the case. To see why, let $e_1$, $e_1'$, $e_2$ and $e_2'$ be independent white noise. When estimating mean or covariance function at any fixed time $t, s \in [0,1] $, it holds that $\mu(t) = E[X(t) + e_1] $ and $C(s,t) = E [(X(t) - \mu(t) + e_1) ( X(s) - \mu(s) + e_1' )] $ for $t \neq s$. But $ E[| X(t) - Y(s) + e_1 - e_2 |] \neq E[ | X(t) - Y(s) |] $. Likewise, we can see that the energy distance ED in \eqref{eq:ed} would have the same challenge to handle measurement errors unless these errors were removed in a presmoothing step before carrying out the test. So the challenges with measurement errors is not triggered by the use of the $L^1$ norm in MED. The $L^2$ norm in ED will face the same challenge. For intensely measured functional data, a presmoothing step is often used to handle measurement errors in the observed data in the hope that smoothing will remove the error. However, this is a delicate issue, as it is difficult to know the amount of smoothing needed in order for the subsequent analysis to retain the same convergence rate as if the true functional data were fully observed without errors. For instance, \cite{10.1214/009053606000001505} study the effects of smoothing to obtained reconstructed curves and show that in order to retain the same convergence rate of mean estimation for fully observed functional data the number of measurements per subject that generates the curves must be of higher order than the number of independent subjects. This requires functional data that are intensively sampled well beyond ultra dense (or dense) functional data that have been studied in the literature \citep{zhang2016sparse}. In this paper, we propose a new way of handling measurement errors so that the $\text{MED}$-based testing procedure is still consistent in the presence of measurement errors. The key idea is to show that when the measurement errors $e_1$ and $ e_2$ of $X$ and $Y$, respectively, are identically distributed, i.e, $e_1 = ^d e_2$, the $\text{MED}(X,Y)$-based approach can still be applied to the contaminated data with consistency guaranteed under mild assumptions (cf. Corollary \ref{cor:powerconta}). When $e_1 \neq^d e_2$, we propose an error-augmentation approach, which can be applied jointly with our unified estimation procedure. The rest of the paper is organized as follows. Section \ref{sec:3} contains the main methodology and supporting theory about testing marginal homogeneity. Numerical studies are presented in Section \ref{sec:sim} The conclusion is in Section \ref{sec:con} All technical details are postponed to Section \ref{sec:tec} \begin{comment} \section{Energy Distance} \label{sec:2} Before introducing statistics for functional data, we first review some Energy distance based methods, which can be applied to data of different types. For now, we have been measuring the discrepancies between distributions at the population level. In fact, the Energy distance type metric can be estimated conveniently using $U$-statistics, i.e., given two samples $V_1, V_2, \cdots, V_n \sim^{i.i.d} V$ and $W_{n+1}, W_{n+2}, \cdots, W_{n+m} \sim^{i.i.d} W$, an unbiased estimator for $\text{ED}^d(V,W)$ is given by \begin{multline*} \text{ED}_n^{d} (V,W) = \frac{2}{mn} \sum\limits_{i_1=1}^{n} \sum\limits_{i_2=n+1}^{n+m} d(V_{i_1}, W_{i_2}) \\ - \frac{2}{n(n-1)} \sum\limits_{1 \leq i_1 < i_2 \leq n} d(V_{i_1}, V_{i_2}) - \frac{2}{m(m-1)} \sum\limits_{n+1 \leq i_1 < i_2 \leq n+ m} d(W_{i_1}, W_{i_2}). \end{multline*} Estimators for $\text{ED}^{2}(U,V)$, $\text{ED}^{k}(U,V)$ can be constructed similarly and are denoted as $\text{ED}_n^{2}(U,V)$, $\text{ED}^k_n(U,V)$, where the subscript $n$ is used to indicate the sample version. To conduct hypothesis testing, the $\text{ED}$-based statistics are often jointly used with the permutation tests due to its implementation convenience and accurate size, see \cite{lehmann2006testing} for more details on permutation related results. In theory, as $n \rightarrow \infty$, the sample estimate $\text{ED}^d_n(Z,W)$ will converge to $\text{ED}^d(Z,W)$ in probability. \begin{equation} \label{eq:l1} \text{ED}^{1}(U,V) = 2E[\|U - V\|_1] - E[ \|U - U'\|_1] - E[\| V - V'\|_1]. \end{equation} They show $\text{ED}^{1}(U,V)$ can detect the differences between marginal distributions under the high dimensional regime. Thus, $\text{ED}^{1}$-based test is advantageous over $\text{ED}^{k}$ (with $k$ being $L^2$-distance, Gaussian or Laplacian kernel), which can only detect the marginal mean and variance discrepancies between high dimensional distributions. \end{comment} \section{Testing Marginal Homogeneity} \label{sec:3} \iffals \begin{align*} \begin{array}{c} H_0: X(t) =^d Y(t) \text{ for all } t \in [0,1], \\ \text{versus} \\ H_A : X(t) \neq^d Y(t) \text{ for some } t \in [0,1]. \end{array} \end{align*} \f We first consider the case where there are no measurement errors and postpone the discussion of measurement errors to the end of this section. Let $\{ X_i \}_{i=1}^n$ and $\{Y_i\}_{i=n+1}^{n+m}$ be i.i.d copies of $X$ and $Y$, respectively. In practice, the functions are only observed at some discrete points, i.e., \begin{align*} \begin{array}{ll} x_{ij} = X_i(T_{ij}), & \text{ if } i = 1,2, \dots, n, \ j = 1,2, \dots, N_i , \\ y_{ij} = Y_{i}(T_{ij}), & \text{ if } i = n+1, \dots,n+ m, \ j = 1,2, \dots, N_i. \end{array} \end{align*} This sampling plan allows the two samples to be measured at different schedules and additionally each subject within the sample has its own measurement schedule. This is a realistic assumption but the consequence is that two-dimensional smoothers will be needed to estimate the targets. Fortunately, the convergence rate of our estimator attains the same convergence rate as that of a one-dimensional smoothing method. This intriguing phenomenon will be explained later. For notational convenience, denote $ Z_i = X_i, \text{ if } i = 1,2, \dots, n$ and $ Z_i = Y_i, \text{ if } i =n+1, \dots, n+m $. Let $\mathbf{Z} = (\mathbf{z}_1 ; \dots ; \mathbf{z}_{n+m} )$ be the combined observations, where for $i=1,2, \dots, n+m$, $\mathbf{z}_i $ is a vector of length $N_i$, \begin{align*} \mathbf{z}_i= (z_{i1}, z_{i2}, \dots, z_{i,N_i})^T = \left\lbrace \begin{array}{ll} (x_{i1}, x_{i2}, \dots , x_{i,N_i})^T, & \text{ if } 1 \leq i \leq n, \\ (y_{i1}, y_{i2}, \dots , y_{i,N_i})^T, & \text{ if } n+ 1 \leq i \leq n +m. \end{array} \right. \end{align*} The observations corresponding to $X$ and $Y$ are defined as $\mathbf{X} = (\mathbf{z}_1; \dots ; \mathbf{z}_n)$ and $ \mathbf{Y} = (\mathbf{z}_{n+1}; \dots ; \mathbf{z}_{n+m})$ respectively. \begin{comment} To test if $X =^d Y$, a straightforward approach is by first reconstructing each trajectory $\widehat{Z}_i$ based on the discrete observations $\mathbf{z}_i$ and compute the sample version of $\text{ED}(X,Y)$ via $U$-statistic, \begin{multline} \label{eq:est} \text{ED}_n(\mathbf{Z}) = \frac{2}{mn} \sum_{i_1=1}^n\sum_{i_2=n+1}^{n+m} \| \widehat{Z}_{i_1} - \widehat{Z}_{i_2} \|_{L^2} \\ - \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2\leq n} \| \widehat{Z}_{i_1} - \widehat{Z}_{i_2} \|_{L^2} - \frac{2}{m(m-1)} \sum_{n+1 \leq i_1 < i_2\leq n+m} \| \widehat{Z}_{i_1} - \widehat{Z}_{i_2} \|_{L^2}, \end{multline} The critical value of $ \text{ED}_n(\mathbf{Z}) $ can be conveniently determined by permutations. Details about permutation test are postponed later. Thus, we can reject the null hypothesis and conclude that $X \neq^d Y$ if $ \text{MED}_n(\mathbf{Z}) $ is larger than this threshold. \end{comment} \begin{comment} To quantify the difference between marginal distributions, define marginal energy distance (MED) as \begin{align*} \text{MED}(X,Y) = \int 2E \left[ \left| X(t) - Y(t) \right| \right] - E \left[ \left| X(t) - X'(t) \right| \right] - E \left[ \left| Y(t) - Y'(t) \right| \right]dt \end{align*} where $\|X - Y\|_{L^1} = \int_{0}^{1} |X(t) - Y(t)| dt$. MED is a metric between marginal distributions, i.e., $ \text{MED}(X,Y) \geq 0 $ and $ \text{MED}(X,Y) =0$ iff $ X(t) =^d Y(t) \text{ for all } t \in [0,1]$. \end{comment} To estimate $\text{MED}(X,Y)$ in (\ref{MED}), note that we actually have no observations for the one-dimensional functions $ E \left[ \left| X(t) - Y(t) \right| \right] $, $E \left[ \left| X(t) - X'(t) \right| \right]$ and $ E \left[ \left| Y(t) - Y'(t) \right| \right] $, due to the longitudinal design where $X$ and $Y$ are observed at different time points. Thus, the sampling schedule for $X$ and $Y$ are not synchronized. A consequence of such asynchronized functional/longitudinal data is that a one-dimensional smoothing method that has typically been employed to estimate a one-dimensional target function, e.g. $ E \left[ \left| X(t) - Y(t) \right| \right], $ does not work here. However, a workaround is to estimate the following two-dimensional functions first: \begin{align*} &G_{1}(t_1,t_2) := E\left[ \left| X(t_1) - Y(t_2) \right| \right], \\ &G_{2} (t_1,t_2):= E \left[ \left| X(t_1) - X'(t_2) \right| \right], \\ &G_{3}(t_1,t_2):= E \left[ \left| Y(t_1) - Y'(t_2) \right| \right], \end{align*} then set $t_1=t_2=t $ in all three estimators. Since $G_1, G_2, G_3$ can all be recovered by some local linear smoother, $\text{MED}(X,Y)$ admits consistent estimates for both intensively and sparsely observed functional data. For instance, $G_{1}(t_1,t_2)$ can be estimated by $\widehat{G}_{1}(t_1,t_2) = \hat{\beta}_0$, where \begin{multline} \label{eq:xy} (\hat{\beta}_0, \hat{\beta}_1, \hat{\beta}_2) = \underset{\beta_0, \beta_1, \beta_2}{\text{argmin}} \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+ m} \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} K_{h_x}(T_{i_1 j_1} - t_1) \times \\ K_{h_y}(T_{i_2 j_2}- t_2) \left[\left| z_{i_1 j_1} - z_{i_2 j_2} \right| - \beta_0 - \beta_1 (T_{i_1j_1}-t_1) - \beta_2 (T_{i_2j_2} - t_2) \right]^2, \end{multline} and $G_2(t_1, t_2)$ can be estimated by $\widehat{G}_{2}(t_1, t_2) = \hat{\alpha}_0$, where \begin{multline} \label{eq:xx} (\hat{\alpha}_0, \hat{\alpha}_1, \hat{\alpha}_2) = \underset{\alpha_0, \alpha_1, \alpha_2}{\text{argmin}} \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} K_{h_x}(T_{i_1j_1}- t_1) K_{h_x}(T_{i_2j_2}- t_2) \\ [ \left| z_{i_1j_1} - z_{i_2 j_2} \right| - \alpha_0 - \alpha_1 (T_{i_1j_1}-t_1) - \alpha_2 (T_{i_2 j_2} - t_2) ]^2. \end{multline} $G_{3}(t_1,t_2)$ can be estimated similarly as $G_{2}(t_1,t_2)$ by an estimator $\widehat{G}_{3}(t_1,t_2)$. In the above, $N_{i1}$ and $N_{i2}$ should be understood as the respective length of the vector $\mathbf{z}_{i1}$ and $\mathbf{z}_{i2}$ and $K_{h}(\cdot) = K(\cdot/h)/h$ is a one-dimensional kernel with bandwidth $h$. The sample estimate of $ \text{MED}(X,Y) $ can be constructed as \begin{align} \label{eq:MED} \text{MED}_n(\mathbf{Z}) := \int_{0}^1 2 \widehat{G}_{1}(t,t) - \widehat{G}_{2}(t, t) - \widehat{G}_{3}(t, t) dt. \end{align} For hypothesis testing, the critical value or $p$-value can be determined by permutations \citep{lehmann2006testing}. To be more specific, let $\pi : \{ 1,2, \dots, n+m \} \rightarrow \{ 1,2, \dots, n+m \}$ be a permutation. There are $(n+m)!$ number of permutations in total and we denote the set of permutations as $\mathbb{P}_{n+m} = \{ \pi_l : l=1,2, \dots, (n+m)! \}$. For $l=1,2, \dots, (n+m)!$, define the permutation of $\pi_l$ on $\mathbf{Z}$ as: \begin{align} \pi_l \cdot \mathbf{Z} = (\mathbf{z}_{\pi_l(1) } ; \mathbf{z}_{\pi_l(2) }; \dots ; \mathbf{z}_{\pi_l(n+m)} ). \end{align} Write the statistic that is based on the permuted sample $\pi_l \cdot \mathbf{Z} $ as $ \text{MED}_n(\pi_l \cdot \mathbf{Z}) $ and let $ \Pi_1, \dots, \Pi_{S-1} $ be i.i.d and uniformly sampled from $ \mathbb{P}_{n+m} $, we define the permutation based $p$-value as \begin{align*} \widehat{p} = \frac{1}{S} \left\{ 1 + \sum\limits_{l=1}^{S-1} \mathbb{I}_{ \left\lbrace \text{MED}_n( \Pi_l \cdot \mathbf{Z}) \geq \text{MED}_n( \mathbf{Z}) \right\rbrace } \right\}. \end{align*} Then, the level-$\alpha$ permutation test w.r.t. $\text{MED}_n( \mathbf{Z})$ can be defined as: $$ \text{Reject } H_0, \text{ if } \widehat{p} \leq \alpha. $$ \subsection{Convergence Theory} In this subsection, we show that $\text{MED}_{n}(\mathbf{Z})$ is a consistent estimator and develop its convergence rate. \begin{assumption} \label{ass:1} \begin{itemize} \item[A.1] The kernel function $K(\cdot) \geq 0$ is symmetric, Lipschitz continuous, supported on $[-1, 1]$ and satisfies \begin{align*} \int K(u)du = 1, \; \int_{0}^1 u^2 K(u) du < \infty \text{ and } \int_{0}^1 K(u)^2 du < \infty. \end{align*} \item[A.2] Let $\{ T_{ij} : 1 \leq i \leq n, 1 \leq j \leq N_i \} \sim^{i.i.d} T_x$, $\{ T_{ij} : n+ 1 \leq i \leq n +m, 1 \leq j \leq N_i \} \sim^{i.i.d} T_y$ and denote the density functions of $T_x, T_y$ by $g_{x}$, $g_y$ respectively. There exists constants $c$ and $ C$ such that $0 < c \leq g_{x} (s), g_{y} (t) \leq C < \infty$ for any $s,t \in [0, 1]$. \item[A.3] $\{ X_{i_1}, Y_{i_2}, T_{ij} : 1 \leq i_1 \leq n, n+1 \leq i_2 \leq n+m, 1 \leq i \leq n+m, 1 \leq j \leq N_i \}$ are mutually independent. \item[A.4] The second order partial derivatives of $G_{1}, G_{2}, G_{3}$ are bounded on $[0,1]$. \item[A.5] $ \sup_{t} E|X(t)|^2 < \infty $ and $\sup_{t} E|Y(t)|^2 < \infty $. \end{itemize} \end{assumption} \begin{remark} \label{eq:rmk} Conditions A.1 - A.3 and A.5 are fairly standard and also used in \cite{li2010uniform}. Condition A.4 may seem a bit problematic at first, as the absolute value function $| \cdot | $ is not differentiable at 0. However, its expectation can easily be differentiable. For instance, if the density functions of $X(t)$, $Y(t)$ are $f_{x}(\cdot|t)$, $f_{y}(\cdot|t)$ respectively, then we have $G_{1}(s,t) = \int\int |u-v| f_{x}(u|s)f_{y}(v|t) dudv$. Assuming the conditions of the Leibniz integral rule, we can interchange the partial derivatives and integration, i.e., \begin{align*} \frac{\partial^2 }{\partial s \partial t} G_{1}(s,t) = \int\int |u-v| \frac{\partial }{\partial s} f_{x}(u|s) \frac{\partial }{\partial t} f_{y}(v|t) dudv. \end{align*} Thus, the partial derivatives of $G_{1}(s,t)$ are bounded if the second order partial derivatives of $ f_{x}(u|s) $, $ f_{y}(v|t) $ w.r.t. $s,t$ exist for all $u,v$ and \begin{align}\label{eq:partial} \sup_{u} \left| \frac{\partial^2 }{ \partial s \partial s } f_{x}(u|s) \right| < \infty \text{ and } \sup_{v} \left| \frac{\partial^2 }{ \partial t \partial t } f_{y}(v|t) \right| < \infty. \end{align} A more specific example is when $X(t)$ is Gaussian and $Y(t)$ is a mixture of Gaussians with density functions \begin{align*} f_{x}(u|t) & = \frac{1}{\sigma_{1}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u - \mu_{1}(t)}{\sigma_{1}(t) } \right)^2 }, \\ f_{y}(u|t) & = \frac{1}{2} \frac{1}{\sigma_{2}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u - \mu_{2}(t)}{\sigma_{2}(t) } \right)^2 } + \frac{1}{2} \frac{1}{\sigma_{2}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u + \mu_{2}(t)}{\sigma_{2}(t) } \right)^2 }. \end{align*} Then A.4 holds if we assume $\sigma_{1}(s), \sigma_{2}(t) $ are bounded from below by a positive constant, $ \mu_1(s), \mu_2(t), \sigma_{1}(s), \sigma_{2}(t)$ are bounded and have bounded second order derivatives. Similar conclusions can be drawn for $G_2$ and $G_3$. Therefore, Condition A.4 is not restrictive as it is customary to assume that the mean and covariance functions for functional data are differentiable. \begin{comment} consider the example that $X(t)$ is Gaussian and $Y(t)$ is mixture of Gaussians with density functions \begin{align*} f_{x}(u|t) & = \frac{1}{\sigma_{1}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u - \mu_{1}(t)}{\sigma_{1}(t) } \right)^2 }, \\ f_{y}(u|t) & = \frac{1}{2} \frac{1}{\sigma_{2}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u - \mu_{2}(t)}{\sigma_{2}(t) } \right)^2 } + \frac{1}{2} \frac{1}{\sigma_{2}(t) \sqrt{2\pi}} e^{ - \frac{1}{2} \left( \frac{u + \mu_{2}(t)}{\sigma_{2}(t) } \right)^2 }. \end{align*} Then, $G_{Z}(s,t) = \int\int |u-v| f_{x}(u|s)f_{y}(v|t) dudv$ is a smooth function in $s,t$ provided that $ \mu_1(s), \mu_2(t), \sigma_{1}(s), \sigma_{2}(t) $ are smooth functions and $\sigma_{1}(s)$, $\sigma_{2}(t)$ are bounded below by some constant greater than 0. Some straight-forward calculations imply that A.4 holds if we further assume $ \mu_1(s), \mu_2(t), \sigma_{1}(s), \sigma_{2}(t) $ have bounded first order derivatives. Similar conclusions can be drawn on $G_X$ and $G_Y$. In general, Condition A.4 is not restrictive as it is customary to assume that mean and covariance functions for functional data are differentiable. \end{comment} \end{remark} The next assumption specifies the relationship of the number of observations per subject and the decay rate of the bandwidth parameters $h_x, h_y$. \begin{assumption} \label{ass:2new} Suppose $h_x:=h_x(n), h_y:=h_y(m) \rightarrow 0$ and \begin{align*} & \log \left( \frac{n}{ \sum_{i=1}^n N_i^{-1}/n} \right) \frac{\max_{1 \leq i \leq n} N_i^{-1}}{h_x} \frac{\max_{1 \leq i \leq n} N_i^{-1}}{ \sum_{i=1}^n N_i^{-1}/n} \frac{1}{ n} \rightarrow 0, \\ & \log \left( \frac{m}{ \sum_{i=n+1}^{n+m} N_i^{-1}/m} \right) \frac{\max_{n+1 \leq i \leq n+m} N_i^{-1}}{h_y} \frac{\max_{n+1 \leq i \leq n+m} N_i^{-1}}{ \sum_{i=n+1}^{n+m} N_i^{-1}/m} \frac{1}{m} \rightarrow 0. \end{align*} \end{assumption} The following theorem states that we can consistently estimate $\text{MED}(X,Y)$ with sparse observations. \begin{theorem} \label{thm:1} Under Assumptions \ref{ass:1} and \ref{ass:2new}, \begin{align*} \left| \emph{\text{MED}}_n(\mathbf{Z}) - \emph{\text{MED}}(X,Y) \right| = O_p\left(h_x^2+ \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_i } + h_y^2 + \sqrt{ \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_i } \right), \end{align*} where $\{\phi_i: i=1,2, \dots, n+m\}$ are defined as \begin{align*} \phi_i = \left\lbrace \begin{array}{ll} \frac{N_ih_x + N_i(N_i-1)h_x^2}{ N_i^2 h_x^2}, & 1\leq i \leq n , \\ \frac{N_i h_y + N_i(N_i-1)h_y^2}{N_i^2 h_y^2}, & n+1 \leq i \leq n+m . \end{array} \right. \end{align*} \end{theorem} \begin{remark} For any two-dimensional function $F \in L^2([0,1]^2)$, define the $L^2$-norm as $\|F\|_2 := (\int_{t_1}\int_{t_2} [F(t_1, t_2)]^2 dt_1 dt_2 )^{1/2}$. The above theorem is a consequence of \begin{align*} \left\| G_I(t_1, t_2) - \widehat{G}_{I}(t_1, t_1) \right\|_2 & = O_p\left(h_x^2+ \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_i } + h_y^2 + \sqrt{ \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_i } \right), \end{align*} where $I=1,2,3$. Compared with the mean function $\mu(t) = E[X(t)]$ and the covariance function $ C_X(s,t) = E[(X(s) - \mu(s))(X(t) - \mu(t))] $, $ G_1, G_2, G_3$ are functions involving two independent stochastic processes. An intriguing phenomenon is that even though $ G_1, G_2, G_3 $ are two-dimensional functions, the convergence rate of their linear smooth estimates is the same as for a one-dimensional function, such as the mean function. This is because the expectation $G_1(s,t) = E[|X(s) - Y(t)|]$ involves two independent stochastic processes and $n \times m$ pairs $\{ (X_{i_1}, Y_{i_2}):i_1=1,\dots, n, i_2=n+1, \dots,n+m \}$ are used in the linear smoother, leading to a faster convergence rate. This distinguishes this situation from the standard estimation of a bivariate function. For instance, if the goal is to estimate $E[|X(s) - X(t)|]$, the convergence rate would be slower and would be the same as that for a two-dimensional smoother. We further point out that even though a two dimensional smoothing method is used to estimate $G_I (s, t)$, we only need to evaluate its values at the diagonal where $s=t$. Therefore, the computational effort is manageable. \end{remark} Given two sequences of positive real numbers $a_n$ and $b_n$, we say that $a_n$ and $b_n$ are of the same order as $n \rightarrow \infty$ (denoted as $a_n \asymp b_n$) if there exists constants $0 < c_1 < c_2 < \infty$ such that $ c_1 \leq \lim_{n \rightarrow \infty} a_n/b_n \leq c_2 $ and $ c_1 \leq \lim_{n \rightarrow \infty} b_n/a_n \leq c_2 $. The convergence rates of $\text{MED}_n(\mathbf{Z})$ for different sampling plans are provided in the following corollary. \begin{corollary} \label{cor:1} Under \ref{ass:1} and \ref{ass:2new}, and further assume $m(n) \asymp n$. \begin{itemize} \item[(i)] When $N_i \asymp C$ for all $i=1,2, \dots, n+m$, where $0 < C <\infty$ is a constant, and $h_x \asymp h_y \asymp n^{-1/5}$, we have \begin{align*} \left| \emph{\text{MED}}_n(\mathbf{Z}) - \emph{\text{MED}}(X,Y) \right| = O_p \left( \frac{1}{n^{2/5}} \right). \end{align*} \item[(ii)] When $N_i \asymp n^{1/4}$ for all $i=1,2, \dots, n+m$ and $h_x \asymp h_y \asymp n^{-1/4}$, we have \begin{align*} \left| \emph{\text{MED}}_n(\mathbf{Z}) - \emph{\text{MED}}(X,Y) \right| = O_p \left( \frac{1}{\sqrt{n}} \right). \end{align*} \end{itemize} \end{corollary} \subsection{Validity of the Permutation Test and Power Analysis} We now justify the permutation based test for sparsely observed functional data. Under the null hypothesis and the mild assumption that $\{N_i\}$ are i.i.d across subjects, the size of the test can be guaranteed by the fact that the distribution of the sample is invariant under permutation. For a rigorous argument, see Theorem 15.2.1 in \cite{lehmann2006testing}. Thus, the permutation test based on the test statistic (\ref{eq:MED}) produces a legitimate size of the test. The power analysis is much more challenging and will be presented below. Let $\pi \in \mathbb{P}_{n+m}$ be a fixed permutation and $\widehat{G}_{\pi, I}(t_1, t_2), I=1,2,3$ be the estimated functions from algorithms \eqref{eq:xy} and \eqref{eq:xx} using permuted samples $\pi \cdot \mathbf{Z}$. The conditions on the decay rate of bandwidth parameters $h_x, h_y$ that ensure the convergence of $\widehat{G}_{\pi, I}$ for any fixed permutation $\pi$ are summarized below. \begin{assumption} \label{ass:power} Suppose $h_x:=h_x(n), h_y:=h_y(m) \rightarrow 0$ and \begin{align*} & \sup\limits_{\pi \in \mathbb{P}_{n+m}} \log \left( \frac{n^2}{ \sum_{i=1}^n N_{\pi(i)}^{-1}} \right) \frac{\max\limits_{1 \leq i \leq n} N_{\pi(i)}^{-1}}{\min\{ h_x, h_y \} } \frac{\max\limits_{1 \leq i \leq n} N_{\pi(i)}^{-1}}{ \sum_{i=1}^n N_{\pi(i)}^{-1}/n} \frac{1}{ n} \rightarrow 0, \\ &\sup\limits_{\pi \in \mathbb{P}_{n+m}} \log \left( \frac{m^2}{ \sum_{i=n+1}^{n+m} N_{\pi (i)}^{-1}} \right) \frac{\max\limits_{n+1 \leq i \leq n+m} N_{\pi(i)}^{-1}}{\min\{h_x, h_y\} } \frac{\max\limits_{n+1 \leq i \leq n+m} N_{\pi(i)}^{-1}}{ \sum_{i=n+1}^{n+m} N_{\pi(i)}^{-1}/m} \frac{1}{m} \rightarrow 0. \end{align*} \end{assumption} Let $\Pi$ be a random permutation uniformly sampled from $\mathbb{P}_{n+m}$. If the sample is randomly shuffled, it holds that $Z_{\Pi(i)} =^d Z_{\Pi(j)}$ and $\text{MED}(Z_{\Pi(i)} , Z_{\Pi(j)} ) = 0$ for any $i,j=1,2, \dots, n+m$. For the sample estimate $\text{MED}_n(\Pi \cdot \mathbf{Z})$ based on the permuted sparse observations, we show that $\text{MED}_n(\Pi \cdot \mathbf{Z})$ converges to 0 in probability. \begin{theorem} \label{thm:per} Under Assumptions \ref{ass:1} and \ref{ass:power}, \begin{align*} \left|\emph{\text{MED}}_n( \Pi \cdot \mathbf Z) \right| =O_p \left( \sup\limits_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_{\pi(i)} } + \sup\limits_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_{\pi(i)} } \right), \end{align*} where $ \Pi \sim \text{Uniform}(\mathbb P_{n+m})$ and is independent of the data. \end{theorem} For the original data that have not been permuted, $ \text{MED}_n(\mathbf{Z}) \rightarrow^p \text{MED}(X , Y )$, which is strictly positive under the alternative hypothesis. On the other hand, we know from Theorem \ref{thm:per} that the permuted statistics converges to 0 in probability. This suggests that, under mild assumptions, the probability of rejecting the null approaches 1 as $n,m \rightarrow \infty$. We make this idea rigorous in the following theorem. \begin{theorem} \label{thm:power} Under Assumption \ref{ass:1} and \ref{ass:power}, for any fixed $S > 1/ \alpha$ we have \begin{align*} P_{H_A} \left( \widehat{p} \leq \alpha \right) \rightarrow 1. \end{align*} \end{theorem} \begin{remark} Since Assumption 3 implies Assumption 2, Theorem 1 holds under the assumption of Theorem \ref{thm:power} and it facilitates the proof of Theorem 3. \end{remark} \subsection{Handling of Measurement Errors} \label{sec:correct} With measurement errors present, the actual observed data are: \begin{align*} \widetilde{x}_{ij} &= X_i(T_{ij})+ e_{ij}, \; i=1,2, \dots, n, \; j = 1,2, \dots, N_i , \\ \widetilde{y}_{ij} &= Y_{i}(T_{ij})+e_{ij}, \; i = n+ 1, \dots, n+ m, \; j = 1,2, \dots, N_i. \end{align*} where $ \{ e_{ij} : i=1,2 , \dots, n, j=1,2, \dots, N_i \} \sim^{i.i.d} e_1 $, $ \{e_{ij} : i=n+1, \dots, n+m, j=1,2, \dots, N_i \} \sim^{i.i.d} e_2 $, and $e_1, e_2$ are mean 0 independent univariate random variables. Denote the combined noisy observations by $ \widetilde{\mathbf{Z}} = (\widetilde{\mathbf{z}}_1 ; \dots ; \widetilde{\mathbf{z}}_{n+m}) $, where \begin{align*} \widetilde{\mathbf{z}}_i= (\widetilde{z}_{i1}, \widetilde{z}_{i2}, \dots, \widetilde{z}_{i,N_i})^T = \left\lbrace \begin{array}{ll} (\widetilde{x}_{i1}, \widetilde{x}_{i2}, \dots , \widetilde{x}_{i,N_i})^T, & \text{ if } 1 \leq i \leq n, \\ (\widetilde{y}_{i1}, \widetilde{y}_{i2}, \dots , \widetilde{y}_{i,N_i})^T, & \text{ if } n+ 1 \leq i \leq n +m. \end{array} \right. \end{align*} The local linear smoothers described in Equations \eqref{eq:xy} and \eqref{eq:xx} are then applied with the input data $\{ z_{i_1j_1} \}$ and $\{ z_{i_2j_2} \}$ replaced, respectively, by $\{ \widetilde{z}_{i_1j_1} \}$ and $\{ \widetilde{z}_{i_2j_2} \}$. The resulting outputs are denoted as $ \widehat{H}_{1}, \widehat{H}_{2}, \widehat{H}_{3}$, leading to the estimator \begin{align} \label{eq:medZtilde} \text{MED}_n(\widetilde{\mathbf{Z}}) = \int 2 \widehat{H}_{1}(t,t) - \widehat{H}_{2}(t,t) - \widehat{H}_{3}(t,t) dt. \end{align} Correspondingly, the proposed test with contaminated data $\widetilde{\mathbf{Z}}$ is: $$ \text{Reject } H_0, \text{ if } \widetilde{p} \leq \alpha. $$ where $ \widetilde{p} = \frac{1}{S} \big\{ 1 + \sum_{l=1}^{S-1} \mathbb{I}_{ \{ \text{MED}_n( \Pi_l \cdot \widetilde{\mathbf{Z}}) \geq \text{MED}_n( \widetilde{\mathbf{Z}}) \} } \big\}. $ To study the convergence of $\text{MED}_n(\widetilde{\mathbf{Z}})$, define the two-dimensional functions $ H_{1}(s,t) $, $H_{2}(s,t) $ and $ H_{3}(s,t) $ as \begin{align} \label{tildeG} \begin{split} & H_{1}(s,t) = E[ |X(s) + e_{1} - Y(t) - e_2 | ] , \\ & H_{2}(s,t) = E[ | X(s) + e_1 - X'(t) - e_{1}' | ], \\ & H_{3}(s,t) = E[ |Y(s) + e_2 - Y'(t) - e_{2}' | ], \end{split} \end{align} where $e_{1}'$ and $e_{2}'$ are independent and identical copies of $e_1$ and $e_2$ respectively. The target of $\text{MED}_n(\widetilde{\mathbf{Z}})$ is shown to be \begin{align} \widetilde{\text{MED}}(X,Y) = \int 2 H_{1}(t,t) - H_{2}(t,t) - H_{3}(t,t) dt, \end{align} \begin{remark} An unpleasant fact is that $\widetilde{\emph{\text{MED}}}(X, Y)$ involves errors, which cannot be easily removed due to the presence of the absolute error function in (\ref{tildeG}). The handling of measurement errors in both method and theory is thus very different here from conventional approaches for functional data, where one does not deal with the absolute function. The energy distance with $L^2$-norm in Equation \eqref{eq:2step} also has this issue. Thus, measurement errors would also be a challenge for the full homogeneity test even if we can approximate the $L^2$ norm in the energy distance well. \end{remark} To show the approximation error of $\text{MED}_n(\widetilde{\mathbf{Z}})$ we need the following assumptions. \begin{assumption} \label{ass:2} \begin{itemize} \item[D.1] $ E[e_1^2] < \infty $ and $E[e_2^2] < \infty$. \item[D.2] $\{ X_{i_1}, Y_{i_2}, T_{ij}, e_{ij} : 1 \leq i_1 \leq n, n+1 \leq i_2 \leq n+m, 1 \leq i \leq n+m, 1 \leq j \leq N_i \}$ are mutually independent. \item[D.3] \label{ass:a4} The second order partial derivatives of $H_{1}, H_{2}, H_{3}$ are bounded on $[0,1]$. \end{itemize} \end{assumption} \begin{remark} Using the notations $f_x(\cdot|t)$ and $f_y(\cdot|t)$ in Remark \ref{eq:rmk}, let the density functions of $e_1$ and $e_2$ be $\eta_{1}( \cdot )$ and $\eta_{2}(\cdot )$ respectively. Under the conditions of the Leibniz integral rule, \begin{align*} \frac{\partial^2}{\partial s \partial t} H_{1}(s,t) = \int |u-v + a-b| \frac{\partial}{\partial s} f_{x}(u|s) \frac{\partial}{ \partial t} f_{y}(v|t) \eta_{1}(a) \eta_{2}(b)dudvdadb, \end{align*} which admits bounded second-order partial derivatives if \eqref{eq:partial} holds. Similar conclusions can be drawn for $ H_2 $ and $H_3$. Therefore, Assumption D.3 is mild. \end{remark} \begin{corollary} \label{cor:2} Under Assumptions \ref{ass:1}, \ref{ass:2new} and \ref{ass:2}, we have \begin{align*} \left| \emph{\text{MED}}_n(\widetilde{\mathbf{Z}}) - \widetilde{\emph{\text{MED}}}(X,Y) \right| = O_p\left(h_x^2+ \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_i } + h_y^2 + \sqrt{ \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_i } \right). \end{align*} \end{corollary} By the property of energy distance, it holds that $ \widetilde{\text{MED}}(X, Y) \geq 0 $ and $ \widetilde{\text{MED}} (X, Y ) = 0 \Leftrightarrow X(t) + e_1 =^d Y(t) + e_2$ for almost all $t \in [0,1]$. Then, we show that under the following assumptions, the condition that $ X(t) + e_1 =^d Y(t) + e_2 $ would imply the homogeneity of $X(t)$ and $Y(t)$. \begin{assumption} \label{ass:3} Suppose that for any $t \in [0,1]$ \begin{itemize} \item[E.1] $X(t)$, $Y(t)$ are continuous random variables with density functions $f_{x}(\cdot|t)$, $f_{y}(\cdot|t)$ respectively. \item[E.2] $e_1 $, $e_2$ are i.i.d continuous random variables with characteristic function $\phi(\cdot)$ and the real zeros of $\phi(\cdot)$ has Lebesgue measure 0. \item[E.3] $ \{ X(t), Y(t), e_1, e_2 \} $ are mutually independent. \end{itemize} \end{assumption} For common distributions, such as Gaussian and Cauchy, their characteristic functions are of exponential form and have no real zeros. Some other random variables, such as Exponential, Chi-square and Gamma, have characteristic functions of the form $\psi(t) = (1-it \theta)^{-k}$ with only a finite number of real zeros. Since it is common to assume Gaussian measurement errors, the restriction on the real zeros of the characteristic function in Assumption \ref{ass:3} (E.2) is very mild. \begin{theorem} \label{thm:2} Under Assumption \ref{ass:3}, for any $t \in [0,1]$, \begin{align*} & X(t) + e_1 = ^d Y(t) + e_2 \Leftrightarrow X(t) =^d Y(t). \end{align*} \end{theorem} Based on the above theorem, we have the important property that $ \widetilde{\text{MED}} (X,Y) = 0 \Leftrightarrow X(t) =^d Y(t) $ for almost all $t \in [0,1]$. As discussed before, $\widetilde{\text{MED}}(X,Y)$ can be consistently estimated by $\text{MED}_n(\widetilde{\mathbf{Z}}) $ and the test can be conducted via permutations. Consequently, the data contaminated with measurement errors can still be used to test marginal homogeneity as long as $e_1 = ^d e_2$. We make this statement rigorous below. \begin{corollary} \label{cor:powerconta} Under Assumptions \ref{ass:1}, \ref{ass:power}, \ref{ass:2}, \ref{ass:3}. For any fixed $S > 1/ \alpha$, \begin{align*} P_{H_A} \left( \widetilde{p} \leq \alpha \right) \rightarrow 1. \end{align*} \end{corollary} The circumstance of Identically distributed errors among the two samples is a strong assumption that nevertheless can be satisfied in many real situations, for example, when the curves $\{X_i\}$ and $\{ Y_j \}$ are measured by the same instrument. The PBC data in Section \ref{sec:realdata} underscore this phenomenon. When $ e_1 \neq ^d e_2 $, not all is lost and we show that some workarounds exist. In particular, we propose an error-augmentation method that raises the noise of one sample to the same level as that of the other sample. For instance, suppose that $e_1 \sim N(0, \sigma_{1}^2)$ and $ e_2 \sim N(0, \sigma_{2}^2)$. Then the variances $\sigma_{1}^2$ and $\sigma_{2}^2$ can be estimated consistently using the R package ``fdapace" \citep{fdapace} under both intensive and sparse designs with estimates $\widehat{\sigma}_{1}^2 $ and $\widehat{\sigma}_{2}^2$ respectively. \cite{yao2005functional} showed that \begin{align*} \left| \widehat{\sigma}_1 - \sigma_1 \right| = O_p\left( \frac{1}{\sqrt{n}} \left( \frac{1}{h_{G_x}^2} + \frac{1}{h_{V_x}} \right) \right), \end{align*} where $ h_{G_x}$ and $ h_{V_x} $ are the bandwidth parameters for estimating the covariance function $\text{cov}(X(s), X(t))$ and the diagonal function $ \text{cov}(X(t), X(t)) + \sigma_1^2 $ respectively. A different estimator for $\widehat{\sigma}_1$ that has a better convergence rate is provided in \cite{linandwang2022}. An analogous result holds for $ \widehat{\sigma}_2 $. Then, by adding additional Gaussian white noise, we obtain the error-augmented data $\{ \breve{x}_{ij} \}$, $\{ \breve{y}_{ij}\}$ as follows, \begin{align*} \left\lbrace \begin{array}{ll} \left. \begin{array}{ll} \breve{x}_{ij} = \widetilde{x}_{ij} + \epsilon_{ij}, & \text{ for } i=1,2, \dots, n, j=1, \dots, N_i, \\ \breve{y}_{ij} = \widetilde{y}_{ij}, & \text{ for } i=n+1, \dots, n+m, j=1, \dots, N_i, \end{array} \right\rbrace & \text{ if } \widehat{\sigma}_{1}^2 < \widehat{\sigma}_{2}^2, \\ [\bigskipamount] \left. \begin{array}{ll} \breve{x}_{ij} = \widetilde{x}_{ij}, & \text{ for } i=1,2, \dots, n, j=1, \dots, N_i , \\ \breve{y}_{ij} = \widetilde{y}_{ij}+ \epsilon_{ij}, & \text{ for } i=n+1, \dots, n+m, j=1, \dots, N_i, \end{array} \right\rbrace & \text{ if } \widehat{\sigma}_{1}^2 > \widehat{\sigma}_{2}^2, \end{array} \right. \end{align*} where $\{ \epsilon_{ij} \} \sim^{i.i.d} N(0, |\widehat{\sigma}_{2}^2 - \widehat{\sigma}_{1}^2| )$. With $\breve{\mathbf{Z}}$ being the combined error-augmented data, the proposed test is: $$ \text{Reject } H_0 \text{ if } \breve{p} \leq \alpha, $$ where $ \breve{p} = \frac{1}{S} \big\{ 1 + \sum_{l=1}^{S-1} \mathbb{I}_{ \{ \text{MED}_n( \Pi_l \cdot \breve{\mathbf{Z}}) \geq \text{MED}_n( \breve{\mathbf{Z}}) \} } \big\}$. The normal error assumption is common in practice. The variance augmentation approach also works for any parametric family of distributions that is close under convolution (i.e., the sum of two independent distributions from this family is also a member of the family) and that has the property that the first two moments of a distribution determine the distribution. \begin{comment} Here, we discuss the situation when observations are contaminated by measurement errors. Given two random variables $W_1$ and $W_2$, we use the notation $W_1 \perp W_2$ to indicate that $W_1$ is independent of $W_2$. Consider the following model \begin{align*} \widetilde{X}_{i}(t) &= X_i(t) + \epsilon_{i}(t), \; i=1,2, \cdots, n, \\ \widetilde{Y}_{i}(t) &= Y_{i}(t)+\epsilon_{i}(t), \; i = n+ 1,2, \cdots, n+ m, \end{align*} where $ \{ \epsilon_{i} : i=1,2 , \cdots, n \} \sim^{i.i.d} \epsilon_x $, $ \{\epsilon_{i} : i=n+1, \cdots, n+m \} \sim^{i.i.d} \epsilon_y $ and $\epsilon_x$, $\epsilon_y$ are stationary random processes such that $\epsilon_x (t_1) \perp \epsilon_x(t_2)$, $\epsilon_y(t_1) \perp \epsilon_y(t_2)$ if $t_1 \neq t_2$. Also, the contaminated curves $ \{ \widetilde{X}_i, \widetilde{Y}_j : i =1,2, \cdots,n, j=n+1, \cdots, n+m \}$ are only observed at discrete time points, \begin{align*} \widetilde{x}_{ij} &= X_i(T_{ij})+ \epsilon_{i}(T_{ij}), \; i=1,2, \cdots, n, \; j = 1,2, \cdots, N_i , \\ \widetilde{y}_{ij} &= Y_{i}(T_{ij})+\epsilon_{i}(T_{ij}), \; i = n+ 1,2, \cdots, n+ m, \; j = 1,2, \cdots, N_i. \end{align*} Similarly, the combined noisy observations are denoted as $ \widetilde{\mathbf{Z}} = (\widetilde{\mathbf{z}}_1 ; \cdots ; \widetilde{\mathbf{z}}_{n+m}) $, where \begin{align*} \widetilde{\mathbf{z}}_i= (\widetilde{z}_{i1}, \widetilde{z}_{i2}, \cdots, \widetilde{z}_{i,N_i})^T = \left\lbrace \begin{array}{ll} (\widetilde{x}_{i1}, \widetilde{x}_{i2}, \cdots , \widetilde{x}_{i,N_i})^T, & \text{ if } 1 \leq i \leq n, \\ (\widetilde{y}_{i1}, \widetilde{y}_{i2}, \cdots , \widetilde{y}_{i,N_i})^T, & \text{ if } n+ 1 \leq i \leq n +m. \end{array} \right. \end{align*} The local linear smoother described in Equation \eqref{eq:xy} and \eqref{eq:xx} are then applied with input data $\{ z_{i_1j_1} \}$ and $\{ z_{i_2j_2} \}$ replaced, respectively, by $\{ \widetilde{z}_{i_1j_1} \}$ and $\{ \widetilde{z}_{i_2j_2} \}$. The resulting outputs are denoted as $ \widehat{G}_{\widetilde{\mathbf{Z}}}, \widehat{G}_{\widetilde{\mathbf{X}}}, \widehat{G}_{\widetilde{\mathbf{Y}}}$, which lead to the estimator $$ \text{MED}_n(\widetilde{\mathbf{Z}}) = \int 2 \widehat{G}_{\widetilde{\mathbf{Z}}}(t,t) - \widehat{G}_{\widetilde{\mathbf{X}}}(t,t) - \widehat{G}_{\widetilde{\mathbf{Y}}}(t,t) dt. $$ Let $ G_{\widetilde{Z}}(s,t) = E[ |\widetilde{X}(s) - \widetilde{Y}(t)| ], G_{\widetilde{X}}(s,t) = E[ | \widetilde{X}(s) - \widetilde{X}'(t) | ], G_{\widetilde{Y}}(s,t) = E[ |\widetilde{Y}(s) - \widetilde{Y}'(t) | ]$, and suppose that the first order partial derivatives of $G_{\widetilde{Z}}, G_{\widetilde{X}}, G_{\widetilde{Y}}$ are bounded on $[0,1]$, then it follows from Theorem \ref{thm:1} that $\text{MED}_n(\widetilde{\mathbf{Z}}) $ is a consistent estimator of \begin{align*} \text{MED}(\widetilde{X},\widetilde{Y}) = \int 2 G_{\widetilde{Z}}(t,t) - G_{\widetilde{X}}(t,t) - G_{\widetilde{Y}}(t,t) dt \end{align*} and it holds that $ \text{MED}(\widetilde{X},\widetilde{Y}) \geq 0 $ and $ \text{MED} (\widetilde{X},\widetilde{Y}) = 0 \Leftrightarrow X(t) + \epsilon_x(t) =^d Y(t) + \epsilon_y(t)$ for all $t \in [0,1]$. \end{comment} \begin{comment} Let the density functions of $\epsilon_x(t), \epsilon_y(t)$ be $f_{1}(\cdot |t)$ and $f_{2}(\cdot | t)$ respectively. Continue with Remark \eqref{eq:rmk}, \begin{align*} G_{\widetilde{Z}}(s,t) = \int |a-b+u-v|f_{x}(a|s) f_{y}(b|t)f_{1}(u|s)f_{2}(v|t)dadbdudv, \end{align*} which is smooth as long as $f_{x}(a|s)$, $f_{y}(b|t)$, $f_{1}(u|s)$ and $f_{2}(v|t)$ are smooth in terms of $s,t$. Similar conclusions can be drawn for $ G_{\widetilde{X}} $ and $G_{\widetilde{Y}}$. \end{comment} \begin{comment} When the measurement errors of $X$, $Y$ are identically distributed, i.e., $\epsilon_x = ^d \epsilon_y$, it can be shown that, under mild conditions, $X(t) + \epsilon_x(t) =^d Y(t) + \epsilon_y(t) \Leftrightarrow X(t) =^d Y(t)$, which hints that the data contaminated with measurement errors can still be used to test marginal homogeneity as if there is none measurement error provided that the measurement errors of $X$ and $Y$ are identically distributed. The circumstance of identically distributed error, although strong, can correspond to many real situations. For example, it happens when the curves $\{X_i\}$, $\{ Y_j \}$ are measured by the same machine. \begin{assumption} \label{ass:2} Suppose that for any $t \in [0,1]$ \begin{itemize} \item[B.1] $X(t)$, $Y(t)$ are continuous random variables with density functions $f_{x}(\cdot|t)$, $f_{y}(\cdot|t)$ respectively. \item[B.2] $\epsilon_x(t)$, $\epsilon_y(t)$ are i.i.d continuous random variables with characteristic function $\phi(\cdot | t)$ and the real zeros of $\phi(\cdot|t)$ has Lebesgue measure 0. \item[B.3] $ \{ X(t), Y(t), \epsilon_x(t), \epsilon_y(t) \} $ are mutually independent. \end{itemize} \end{assumption} \end{comment} \begin{comment} For commonly seen random variables such as Gaussian and Cauchy, their characteristic functions are of exponential form and has no real zeros. Some other random variables, such as Exponential, Chi-square and Gamma, have characteristic functions of the form $\psi(t) = (1-it \theta)^{-k}$ with only finite number of real zeros. Since it is common to assume Gaussian measurement errors, the restriction on the real zeros of the characteristic function in assumption B.2 is very mild. \begin{theorem} \label{thm:2} Under Assumption \ref{ass:2}, \begin{align*} & X(t) + \epsilon_x(t) = ^d Y(t) + \epsilon_y(t) \Leftrightarrow X(t) =^d Y(t). \end{align*} \end{theorem} Based on the above theorem, we have the property that $ \text{MED}(\widetilde{X},\widetilde{Y}) = 0 \Leftrightarrow X(t) =^d Y(t) $ for all $t \in [0,1]$. As discussed before, $\text{MED}(\widetilde{X},\widetilde{Y})$ can be consistently estimated by $\text{MED}_n(\widetilde{\mathbf{Z}}) $ and the test can be conducted via permutations. \end{comment} \begin{comment} \JL{This paragraph needs to be checked/rewritten to be consistent with the equal errors part. Also $\perp$ is used here, so need to define it.} When $ \epsilon_x \neq ^d \epsilon_y $, we propose an error-augmentation method that raises the noise of one sample to the same level as the other sample. For each $t$, suppose that $\epsilon_x(t)$, $\epsilon_y (t) $ are, respectively, independent copies of $ \epsilon_x $ and $ \epsilon_y. $ That is, for all $t \in [0,1]$, $\epsilon_x(t) \sim N(0, \sigma_{1}^2)$, $ \epsilon_y(t) \sim N(0, \sigma_{2}^2) $ and $\epsilon_x (t_1) \perp \epsilon_x(t_2)$, $\epsilon_y(t_1) \perp \epsilon_y(t_2)$ if $t_1 \neq t_2$. The variances $\sigma_{1}^2$ and $\sigma_{2}^2$ can be estimated consistently using the R package ``fdapace" \citep{fdapace} under both the dense and sparse design. In addition, let the estimated $\sigma^2_{1}$, $\sigma_{2}^2$ be denoted as $ \widehat{\sigma}_{1}^2 $, $\widehat{\sigma}_{2}^2$ respectively, \cite{yao2005functional} show that \begin{align*} \left| \widehat{\sigma}_1 - \sigma_1 \right| = O_p\left( \frac{1}{\sqrt{n}} \left( \frac{1}{h_{G_x}^2} + \frac{1}{h_{V_x}} \right) \right), \end{align*} where $ h_{G_x}$ and $ h_{V_x} $ are the bandwidth parameters for estimating the covariance function $\text{cov}(X(s), X(t))$ and the diagonal function $ \text{cov}(X(t), X(t)) + \sigma_1^2 $ respectively. Similar result holds for $ \widehat{\sigma}_2 $. Then, by adding additional Gaussian white-noise, we can obtain the error-augmented data $\{ \breve{x}_{ij} \}$, $\{ \breve{y}_{ij}\}$ as \begin{align*} \left\lbrace \begin{array}{ll} \left. \begin{array}{ll} \breve{x}_{ij} = \widetilde{x}_{ij} + e_{ij}, & \text{ for } i=1,2, \cdots, n, j=1, \cdots, N_i, \\ \breve{y}_{ij} = \widetilde{y}_{ij}, & \text{ for } i=n+1, \cdots, n+m, j=1, \cdots, N_i, \end{array} \right\rbrace & \text{ if } \widehat{\sigma}_{1}^2 < \widehat{\sigma}_{2}^2, \\ [\bigskipamount] \left. \begin{array}{ll} \breve{x}_{ij} = \widetilde{x}_{ij}, & \text{ for } i=1,2, \cdots, n, j=1, \cdots, N_i , \\ \breve{y}_{ij} = \widetilde{y}_{ij}+ e_{ij}, & \text{ for } i=n+1, \cdots, n+m, j=1, \cdots, N_i, \end{array} \right\rbrace & \text{ if } \widehat{\sigma}_{1}^2 > \widehat{\sigma}_{2}^2, \end{array} \right. \end{align*} where $\{ e_{ij} \} \sim^{i.i.d} N(0, |\widehat{\sigma}_{2}^2 - \widehat{\sigma}_{1}^2| )$. Similarly, let $\breve{\mathbf{Z}}$ be the combined error-augmented data, we can conduct the test as $$ \text{Reject } H_0 \text{ if } \text{MED}_n( \breve{\mathbf{Z}}) > Q_{\breve{R},1-\alpha}. $$ where let $ \breve{\pi}_1, \cdots, \breve{\pi}_S $ be i.i.d and uniformly sampled from $ \mathbb{P}_{n+m} $ and \begin{align*} \breve{R}(t):=\frac{1}{S} \left(\mathbb{I}_{ \left\lbrace \text{MED}_n( \breve{\mathbf{Z}}) \leq t \right\rbrace } + \sum\limits_{i=1}^{S-1} \mathbb{I}_{ \left\lbrace \text{MED}_n( \breve{\pi}_l \cdot \breve{\mathbf{Z}}) \leq t \right\rbrace } \right). \end{align*} \end{comment} \section{Numerical Studies} \label{sec:sim} In this section, we examine the proposed testing procedure for both synthetic and real data sets. \subsection{Performance on simulated data} For simulations, we set $\alpha = 0.05$ and perform 500 Monte Carlo replications with 200 permutations for each test. The following example is used to examine the size of our test. \begin{example} \label{exp:1} The stochastic processes $\{ X_i \}_{i=1}^n$ are i.i.d copies of $X$ and $\{ Y_{i} \}_{i=n+1}^{n+m}$ are i.i.d copies of $Y$, where for $t\in [0, 1]$, \begin{align*} & X(t) = \xi_1 \left( -\cos(2\pi t) \right) + \xi_2 \left( \sin(2 \pi t) \right) , \\ & Y(t) = \varsigma_1 \left( -\cos(2\pi t) \right) + \varsigma_2 \left( \sin(2 \pi t) \right), \end{align*} and $ \xi_1, \xi_2, \varsigma_1, \varsigma_2, \sim^{i.i.d} N(0,1) $. These curves are observed at discrete time points \begin{align*} & \widetilde{x}_{ij} = X_i(T_{ij}) + e_{ij}, i = 1,2, \dots, n, j=1,2, \dots, N_i, \\ & \widetilde{y}_{ij} = Y_i(T_{ij}) + e_{ij}, i = n+1, \dots, n+m, j=1,2, \dots, N_i, \end{align*} where $ \{ T_{ij} : i = 1, 2, \dots, n+m, j = 1,2, \dots, N_i \} \sim^{i.i.d} \text{Uniform}[0,1] $ and the measurement errors are Gaussian \begin{align*} & \{ e_{ij} : i = 1,2, \dots, n, j=1,2, \dots, N_i \} \sim^{i.i.d} N(0, \sigma_{1}^2), \\ & \{ e_{ij} : i=n+1, \dots, n+m, j=1,2, \dots, N_i \} \sim^{i.i.d} N(0, \sigma_{2}^2). \end{align*} \end{example} \begin{table} \centering \begin{tabular}{cccccc} \hline $(n,m)$ & $\sigma_{1}$ & $\sigma_{2}$ & $\text{MED}$ & FPCA \\ \hline \multirow{3}{*}{$(100, 70)$} & 0 & 0 & 0.05 & 0 \\ & 0.2 & 0.2 & 0.04 & 0.004 \\ & 0.05 & 0.25 & 0.058 & 0.006 \\ \hline \multirow{3}{*}{$(150, 130)$} & 0 & 0 & 0.054 & 0.002 \\ & 0.2 & 0.2 & 0.046 & 0.002 \\ & 0.05 & 0.25 & 0.056 & 0 \\ \hline \end{tabular} \caption{Comparison of Type I errors under the sparse design.} \label{tab:size} \end{table} The quantities $\{N_i\}$ are used to control the sparsity level. Under sparse designs, $N_i$ are uniformly selected from $2\sim10$ for all $i=1,2, \dots, n+m$, and $ \sigma_{1}^2 $ and $\sigma^2_{2}$ account for the magnitude of measurement errors. If $ \sigma_{1}^2 = \sigma^2_{2} =0 $, this corresponds to the case that there is no measurement error. If $ \sigma_{1}^2 \neq \sigma^2_{2} $, the error augmentation method described in Section \ref{sec:correct} is used. When applying the $\text{MED}$-based permutation test, the bandwidth parameters are selected as $h_x=h_y=0.2$. Table \ref{tab:size} contains the size comparison results under this sparse design. As a baseline method for comparison, the FPCA approach is also included, where we first impute the principal scores and then apply the energy distance on the imputed scores. To be more specific, the first step is to reconstruct the principal scores using the R package ``fdapace" \citep{fdapace}. Then, a two sample tests for multivariate data is applied on the recovered scores. For this, we choose the energy distance based procedure and conduct the hypothesis testing via the R package ``energy" \citep{eneryRpackage}. The FPCA approach has two drawbacks. First, the scores can not be estimated consistently; second, the infinite dimensional vector of scores has to be truncated for computational purposes, which causes information loss. When the Gaussian errors have different variances, the same error-augmentation method is applied to the FPCA approach. From Table \ref{tab:size}, we see that the $\text{MED}$-based methods have satisfactory size, while the FPCA based approach is undersized, likely due to the inherent inaccuracy in the imputed scores. \begin{example} \label{exp:size} The stochastic processes $\{ X_i \}_{i=1}^n$ are i.i.d copies of $X$, $\{ Y_{i} \}_{i=n+1}^{n+m}$ are i.i.d copies of $Y$, where for $t\in [0,1]$, \begin{align*} & X(t) = \xi_1 \left( -\cos(2\pi t) \right) + \xi_2 \left( \sin(2 \pi t) \right) , \\ & Y(t) = \varsigma_1 t ^2 + \varsigma_2 \left( \sqrt{1- t^4} \right), \end{align*} $ \xi_1, \xi_2 \sim^{i.i.d} N(0,1), $ and $\varsigma_1, \varsigma_2$ are independently sampled from the following mixture of Gaussian distributions, \begin{align*} & P(\varsigma \leq a) = \frac{1}{2} P(N(\mu_{\varsigma},\sigma^2_{\varsigma}) \leq a) + \frac{1}{2 } P( N(-\mu_{\varsigma},\sigma^2_{\varsigma}) \leq a ) \text{ for any } a \in \mathbb{R}. \end{align*} The curves are observed at discrete time points \begin{align*} & \widetilde{x}_{ij} = X_i(T_{ij}) + e_{ij}, i = 1,2, \dots, n, j=1,2, \dots, N_i, \\ & \widetilde{y}_{ij} = Y_i(T_{ij}) + e_{ij}, i = n+1, \dots, n+m, j=1,2, \dots, N_i, \end{align*} where $ \{ T_{ij} : i = 1, 2, \dots, n+m, j = 1,2, \dots, N_i \} \sim^{i.i.d} \text{Uniform} \ [0,1] $ and the measurement errors are Gaussian, \begin{align*} & \{ e_{ij} : i = 1,2, \dots, n, j=1,2, \dots, N_i \} \sim^{i.i.d} N(0, \sigma_{1}^2), \\ & \{ e_{ij} : i=n+1, \dots, n+m, j=1,2, \dots, N_i \} \sim^{i.i.d} N(0, \sigma_{2}^2). \end{align*} \end{example} In the above example, $X(s)$ and $Y(t)$ both have zero mean functions. The variances of $\varsigma_1$ and $\varsigma_2$ are the same and equal to $ \mu_{\varsigma}^2 + \sigma^2_{\varsigma} $. By selecting $\mu_{\varsigma}$ and $\sigma^2_{\varsigma}$ such that $ \mu_{\varsigma}^2 + \sigma^2_{\varsigma} =1 $, we have ${\rm var}(X(t)) = {\rm var}(Y(t))$. Under this scenario, $X(s)$ and $Y(t)$ have the same marginal mean and variance, but different marginal distributions. In this example, we set $\mu_{\varsigma} = 0.98$ and $\sigma_{\varsigma} = 0.199$. The power comparison results are provided in Table \ref{tab:power}. The FPCA approach is not included for power analysis as we have already shown in Example \ref{exp:1} that it is undersized. Instead, we compare the power of $\text{MED}$-based tests under the sparse design with the result from dense regular data, where the sparse data are uniformly sampled. The same error-augmentation approach is applied for $\text{MED}$ with dense regular data when $ \sigma_1 \neq \sigma_2 $. From Table \ref{tab:power}, we can observe a moderate power drop from dense to sparse data. In addition, as the sample size increases, the power grows above 0.94 when $\sigma_1=\sigma_2$ and almost doubles when $\sigma_1 \neq \sigma_2$ under the sparse design. \begin{table} \centering \begin{tabular}{ccccc} \hline $(n,m)$ & $\sigma_{1}$ & $\sigma_{2}$ & $\text{MED}$ (dense) & $\text{MED}$ (sparse) \\ \hline \multirow{3}{*}{$(100, 70)$} & 0 & 0 & 0.998 & 0.806 \\ & 0.2 & 0.2 & 0.866 & 0.53 \\ & 0.05 & 0.25 & 0.76 & 0.412 \\ \hline \multirow{3}{*}{$(150, 130)$} & 0 & 0 & 1 & 0.998 \\ & 0.2 & 0.2 & 1 & 0.946 \\ & 0.05 & 0.25 & 0.998 & 0.786 \\ \hline \end{tabular} \caption{Power comparison.} \label{tab:power} \end{table} \subsection{Applications to real data} \label{sec:realdata} In this subsection, we apply the proposed $\text{MED}$-based tests to two real data sets. \\ \noindent \textbf{PBC Data}: \ The first data set is the primary biliary cirrhosis (PBC) data from Mayo Clinic \citep{fleming2011counting}. This data set is from a clinical trial studying primary biliary cirrhosis of the liver. There were 312 patients assigned to either the treatment or control group. The drug D-penicillamine is given to the treatment group. Here, we are interested in testing the equality of the marginal distributions of Prothrombin time, which is a blood test that measures how long it takes blood to clot. The trajectories of Prothrombin time for different subjects are plotted in Figure \ref{fig:straw}, there are on average 6 measurements per subjects. For our tests, the bandwidth is set to be 2. Here, the equal distribution assumption for errors seems to work (the estimated variances for treatment and control group are 0.96 and 1.009 respectively). By using 200 permutations, the $p$-value of the $\text{MED}$-based test is 0.54, which means that there is not enough evidence to conclude that the maginal distributions of Prothrombin time are different between the two groups. This conclusion matches with existing knowledge that D-penicillamine is ineffective to treat primary biliary cirrhosis of liver. \\ \noindent \textbf{Strawberry Data}: \ In the food industry, there is a continuing interest in distinguishing the pure fruit purees from the adulterated ones \citep{straw}. One practical way to detect adulteration is by looking at the spectra of the fruit purees. Here, we are interesting in testing the marginal distribution between the spectra of strawberry purees (authentic samples) and non-strawberry purees (adulterated strawberries and other fruits). The strawberry data can be downloaded from the UCR Time Series Classification Archive \citep{UCRArchive2018} (\url{https://www.cs.ucr.edu/~eamonn/time_series_data_2018/}). The single-beam spectra of the purees were normalized to back-ground spectra of water and then transformed into absorbance units. The spectral range was truncated to 899-1802 $\text{cm}^{-1}$ (235 data points). The two samples of spectra are plotted in Figure \ref{fig:straw} and more information about this data set can be found at \cite{straw}. The estimated variances of the measurement errors are 0.000279 and 0.00031 for the two samples, which indicates that there are practically no measurement errors. To check the performance of our method, we analyze the data using all 235 measurements as well as a sparse subsamples that contain $2$ to $10$ observations per subject. The R package ``energy" is applied for the complete data. Both tests are conducted with 200 permutations and have $p$-value 0.005. Thus, we have strong evidence to conclude that the marginal distributions between the spectra of strawberry and non-strawberry purees are significantly different and our test produced similar results regardless of the sampling plan. \begin{comment} \begin{figure} \begin{center} \includegraphics[scale=0.3]{protho.pdf} \includegraphics[scale=0.3]{Strawberry.pdf} \end{center} \caption{Trajectories of the strawberry data. Blue curves are for strawberry puree and the red curves are for the adulterated strawberries and other fruits. \label{fig:straw} \end{figure} \end{comment} \begin{figure} \centering \begin{subfigure}{.5\textwidth} \centering \caption{PBC} \label{fig:sub1} \includegraphics[scale=0.5]{protho.pdf} \end{subfigure}% \begin{subfigure}{.5\textwidth} \centering \caption{Strawberry} \label{fig:sub2} \includegraphics[scale=0.5]{Strawberry.pdf} \end{subfigure} \caption{Trajectories of real data.} \label{fig:straw} \end{figure} \section{Conclusion} \label{sec:con} The literature on testing homogeneity for functional data is scarce probably because most approaches rely on intense measurement schedules and the hope that measurement errors can be addressed by presmoothing the data. Since reconstruction of noise-free functional data is not feasible for sparsely observed functional data, a test of homogeneity is infeasible. In this work, we show what is feasible for sparse functional data, a.k.a. longitudinal data, and propose a test of marginal homogeneity that adapts to the sampling plan and provides the corresponding convergence rate. Our test is based on Energy distance with a focus on testing the marginal homogeneity. To the best of our knowledge, this is the only nonparametric test with theoretical guarantees under sparse designs, which are ubiquitous. There are several twists in our approach, including the handling of asynchronized longitudinal data and the unconventional way that measurement errors affect the method and theory. The asynchronization of the data can be overcome completely as we demonstrated in Section 2.1, but the handling of measurement errors requires some compromise when the distributions of the measurement errors are different for the two samples. This is the price one pays for lack of data, and is not due to the use of the $L_1$ norm associated with testing the marginal homogeneity, as an $L^2$ norm for testing full homogeneity would also face the same challenge with measurement errors unless a presmoothing step has been employed to eliminate the measurement errors. As we mentioned in Section 1 this would require a super intensive sampling plan well beyond the usual requirement for dense or ultra dense functional data \citep{zhang2016sparse}. While the new approach may involve error-augmentation, numerical results show that the efficiency loss is minimal. Moreover, such an augmentation strategy is not uncommon. For instance, an error augmentation method has also been adopted in the SIMEX approach \citep{doi:10.1080/01621459.1994.10476871} to deal with measurement errors for vector data. While testing marginal homogeneity has its own merits and advantages over a full-fledged test of homogeneity, our intention is not to particularly endorse it. Rather, we point out what is feasible and infeasible for sparsely or intensively measured functional data and develop theoretical support for the proposed test. To the best of our knowledge, we are the first to provide the convergence rate for the permuted statistics for sparse functional data. This proof and the proof of consistency for the proposed permutation test is non-conventional and different from the the multivariate/high-dimensional case. \iffals \JL{I'll rewrite this section by not advocating the marginal approach, rather that it's the reality of what can be done for sparse data.} We propose a two sample testing procedure based on Energy distance. Our test is useful for both dense and sparse functional data. To the best of our knowledge, this is the only nonparametric test with theoretical guarantees under the sparse design. Furthermore, we proposed a method to handle measurement errors in the data. While our approach involves error-augmentation when the error distributions between the two samples are different, such an augmentation strategy is not uncommon. For instance, an error augmentation method is also adopted in the SIMEX approach \citep{doi:10.1080/01621459.1994.10476871}. \f \begin{comment} \section{Technical Details} \label{sec:tec} \subsection{Proof of Theorem \ref{thm:1}} Here, we show that \begin{align*} & \| \widehat{G}_{1} - G_{1} \|_2 = O_p\left(h_x^2+ \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_i } + h_y^2 + \sqrt{ \frac{1}{n^2} \sum_{i=n+1}^{n+m} \phi_i } \right) , \\ & \| \widehat{G}_{2} - G_{2} \|_2 = O_p \left(h_x^2+ \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_i } \right) \end{align*} and $\| \widehat{G}_{3} - G_{3} \|_2$ can be bounded similarly. For $p_1, p_2=0,1,2$, set \begin{multline*} \mathcal{T}_{i_1 i_2}^{p_1p_2}(t_1, t_2) = \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} K_{h_{i_1}}\left(T_{i_1 j_1}- t_1 \right) K_{i_2} \left(T_{i_2 j_2}- t_2 \right) \\ \left( \frac{T_{i_1j_1}-t_1}{h_{i_1}} \right)^{p_1} \left( \frac{T_{i_2j_2} - t_2}{h_{i_2}} \right)^{p_2}, \end{multline*} where $ h_{i} = h_x, \text{ if } 1\leq i \leq n, $; otherwise $h_i = h_y$. The weighted raw data are denoted as \begin{multline*} \mathcal{Z}_{i_1 i_2}^{p_1p_2}(t_1, t_2) = \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} K_{h_{i_1}} \left(T_{i_1 j_1}- t_1\right) K_{h_{i_2}} \left( T_{i_2 j_2}- t_2 \right) \\ \left( \frac{T_{i_1j_1}-t_1}{h_{i_1}} \right)^{p_1} \left( \frac{T_{i_2j_2} - t_2}{h_{i_2}} \right)^{p_2} \left| z_{i_1j_1} - z_{i_2j_2} \right|. \end{multline*} If there is no confusion, we will only write $\mathcal{T}_{i_1 i_2}$, $\mathcal{Z}_{i_1i_2} $ instead of $\mathcal{T}_{i_1 i_2}^{p_1p_2}(t_1, t_2)$, $ \mathcal{Z}_{i_1 i_2}^{p_1p_2}(t_1, t_2) $ for simplicity. Both \eqref{eq:xy} and \eqref{eq:xx} admit closed form solutions and some algebra show that for $I =1,2$ \begin{align*} \widehat{G}_{I}(t_1,t_2) & = \frac{ W_{I,1}(t_1,t_2) V_{I}^{0,0}(t_1,t_2) - W_{I,2}(t_1,t_2) V_{I}^{1,0}(t_1,t_2) + W_{I,3} (t_1,t_2) V_{I}^{0,1}(t_1,t_2) }{W_{I,1}(t_1,t_2) U_{I}^{0,0}(t_1,t_2) - W_{I,2}(t_1,t_2) U_{I}^{1,0}(t_1,t_2) + W_{I,3} (t_1,t_2) U_{I}^{0,1}(t_1,t_2)}. \end{align*} Here, $\{W_{I, J} : I=1,2, J=1,2,3 \}$ are two-dimensional functions defined as \begin{align*} W_{I,1} (t_1, t_2)& = (U_{I}^{2,0} (t_1, t_2)U_{I}^{0,2}(t_1, t_2) - (U_{I}^{1,1} (t_1, t_2))^2, \\ W_{I,2} (t_1, t_2) & = (U_{I}^{1,0} (t_1, t_2)U_{I}^{0,2}(t_1, t_2) - U_{I}^{0,1} (t_1, t_2) U_{I}^{1,1} (t_1, t_2), \\ W_{I,3} (t_1, t_2) & = (U_{I}^{1,0} (t_1, t_2)U_{I}^{1,1}(t_1, t_2) - U_{I} (t_1, t_2;0,1) U_{I}^{2,0} (t_1, t_2), \end{align*} where, for $p_1,p_2=0,1,2$, $\{U_I, V_I: I=1,2\}$ have the following expressions \begin{align*} & U_{1}^{p_1p_2}(t_1,t_2) = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+m} \mathcal{T}_{i_1 i_2}^{p_1p_2}(t_1, t_2), \\ & U_{2}^{p_1p_2}(t_1, t_2) = \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \mathcal{T}_{i_1 i_2}^{p_1p_2}(t_1, t_2), \\ & V_{1}^{p_1p_2}(t_1,t_2) = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+m} \mathcal{Z}_{i_1 i_2}^{p_1 p_2}(t_1, t_2), \\ & V_{2}^{p_1p_2}(t_1, t_2) = \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \mathcal{Z}_{i_1 i_2}^{p_1 p_2}(t_1, t_2). \end{align*} By some straight-forward calculations, for $I =1,2$ \begin{multline} \label{eq:denome} \widehat{G}_{I}(t_1,t_2) - G_{I}(t_1,t_2) \\ = \frac{ W_{I,1}(t_1,t_2) F_{I}^{0,0}(t_1,t_2) - W_{I,2}(t_1,t_2) F_{I}^{1,0}(t_1,t_2) + W_{I,3} (t_1,t_2) F_{I}^{0,1}(t_1,t_2)}{W_{I,1}(t_1,t_2) U_{I}^{0,0}(t_1,t_2) - W_{I,2}(t_1,t_2) U_{I}^{1,0}(t_1,t_2) + W_{I,3} (t_1,t_2) U_{I}^{0,1}(t_1,t_2)}, \end{multline} where for $p_1,p_2=0,1,2$, \begin{multline*} F_{1}^{p_1p_2}(t_1,t_2) = V_{1}^{p_1p_2}(t_1,t_2) -G_{1}(t_1,t_2)U_{1}^{p_1p_2}(t_1,t_2) \\ - h_x \frac{\partial G_{1} (t_1, t_2)}{\partial t_1} U_{1}^{p_1+1, p_2}(t_1,t_2) - h_y\frac{\partial G_{1} (t_1, t_2)}{\partial t_2 } U_{1}^{p_1, p_2+1}(t_1,t_2), \end{multline*} and \begin{multline*} F_{2}^{p_1p_2}(t_1,t_2) = V_{2}^{p_1p_2}(t_1,t_2) -G_{2}(t_1,t_2)U_{2}^{p_1p_2}(t_1,t_2) \\ - h_x \frac{\partial G_{2} (t_1, t_2)}{\partial t_1} U_{2}^{p_1+1,p_2}(t_1,t_2) - h_x\frac{\partial G_{2}(t_1,t_2) }{\partial t_2} U_{2}^{p_1, p_2+1}(t_1,t_2). \end{multline*} Lemma \ref{lem:power2} entails that the denominator in \eqref{eq:denome} is bounded away from 0 with high probability. Thus, for $I=1,2$, we have \begin{align*} \| \widehat{G}_{I} - G_{I} \|_{2} = O_p \left( \| F_I^{0,0} \|_{2} + \| F_I^{1,0} \|_{2} + \| F_I^{0,1} \|_{2} \right). \end{align*} It can be easily seen that $E \left[ \left. \left| z_{i_1j_1} - z_{i_2j_2} \right| \right| T_{i_1j_1}, T_{i_2j_2} \right] = G_{1}(T_{i_1j_1}, T_{i_2, j_2})$ if $1 \leq i_1 \leq n $, $n+1 \leq i_2 \leq n+m$ and $E \left[ \left. \left| z_{i_1j_1} - z_{i_2j_2} \right| \right| T_{i_1j_1}, T_{i_2j_2} \right] = G_{2}(T_{i_1j_1}, T_{i_2, j_2})$ if $1 \leq i_1 < i_2 \leq n $. By Taylor expansion \begin{align*} \| F_{I}^{p_1p_2}\|_2 = O(\| L_{I}^{p_1p_2} \|_2 + h_x^2 + h_y^2 + h_x h_y), \end{align*} where \begin{align*} L_{1}^{p_1p_2}(t_1,t_2) & = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+ m} \mathcal{Z}_{i_1i_2}^{p_1p_2} - E\big[\mathcal{Z}_{i_1i_2}^{p_1p_2} \big| \{T_{i_1j_1}\}_{j_1=1}^{N_{i_1}}, \{T_{i_2, j_2} \}_{j_2=1}^{N_{i_2}} \big], \\ L_{2}^{p_1p_1}(t_1,t_2) & = \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \mathcal{Z}_{i_1i_2}^{p_1p_2} - E \big[ \mathcal{Z}_{i_1i_2}^{p_1p_2} \big| \{T_{i_1j_1}\}_{j_1=1}^{N_{i_1}}, \{T_{i_2, j_2} \}_{j_2=1}^{N_{i_2}}\big]. \end{align*} \begin{lemma}\label{lem:1} Under Assumption \ref{ass:1} and \ref{ass:2new}, for $I=1,2$ and $p_1, p_2=0,1,2$, it holds that \begin{align*} E\left[ \| L_I^{p_1p_2} \|_2^2 \right] \lesssim \frac{1}{n^2} \sum_{i=1}^n \phi_i + \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_i. \end{align*} \end{lemma} \begin{proof} By setting \begin{multline*} w_{i_1i_2, j_1j_2}^{p_1p_2}(t_1, t_2) = K_{h_{i_1}}(T_{i_1 j_1}- t_1) K_{h_{i_2}}(T_{i_2 j_2}- t_2) \left( \frac{T_{i_1j_1}-t_1}{h_{i_1}} \right)^{p_1} \left( \frac{T_{i_2j_2} - t_2}{h_{i_2}} \right)^{p_2} \\ \left\lbrace \left| z_{i_1j_1} - z_{i_2j_2} \right| - E \left[ \left. \left| z_{i_1j_1} - z_{i_2j_2} \right| \right| T_{i_1j_1}, T_{i_2j_2} \right] \right\rbrace, \end{multline*} we can write \begin{align*} \overline{\mathcal{Z}}_{i_1i_2}^{p_1p_2} & = \mathcal{Z}_{i_1i_2}^{p_1p_2} - E \big[ \mathcal{Z}_{i_1i_2}^{p_1p_2} \big| \{T_{i_1j_1}\}_{j_1=1}^{N_{i_1}}, \{T_{i_2, j_2} \}_{j_2=1}^{N_{i_2}}\big] \\ &= \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} w_{i_1i_2, j_1j_2}^{p_1p_2}(t_1, t_2). \end{align*} For any $p_1,p_2=0,1,2$, \begin{align*} & E\left[ w_{i_1i_2, j_1j_2}^{p_1p_2}(t_1, t_2) \right] = E \left[ E\left[ \left. w_{i_1i_2}(t_1, t_2) \right| T_{i_1j_1}, T_{i_2 j_2} \right] \right] = 0, \end{align*} which implies that $E\big[\overline{\mathcal{Z}}_{i_1i_2}^{p_1p_2}\big] = 0$ and $ E\big[ \overline{\mathcal{Z}}_{i_1i_2}^{p_1p_2} \overline{\mathcal{Z}}_{i_1'i_2'}^{p_1p_2} \big] \neq 0, $ only if $ \{i_1, i_2\} \cap \{ i_1', i_2' \} \neq \emptyset$. If $1 \leq i_1, i_1' \leq n$, $n+1 \leq i_2, i_2' \leq n+m$, $i_1=i_1'$, we have $h_{i_1} = h_{i_1'}$, $h_{i_2} = h_{i_2'}$ and \begin{multline*} h_{i_1}^2h_{i_2}^2 \left| \int \int E\left[ w_{i_1i_2, j_1j_2}(t_1, t_2)w_{i_1'i_2', j_1'j_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \\ \leq C \times \left\{ \begin{array}{ll} h_{i_1} h_{i_2}^2, & j_1=j_1', i_2 \neq i_2'; \\ h_{i_1} h_{i_2}^2, & j_1=j_1', i_2 = i_2', j_2 \neq j_2'; \\ h_{i_1} h_{i_2}, & j_1=j_1', i_2 = i_2', j_2 = j_2'; \\ h_{i_1}^2 h_{i_2}^2, & j_1 \neq j_1', i_2 \neq i_2'; \\ h_{i_1}^2 h_{i_2}^2, & j_1\neq j_1', i_2 = i_2', j_2 \neq j_2'; \\ h_{i_1}^2 h_{i_2}, & j_1\neq j_1', i_2 = i_2', j_2 = j_2'. \\ \end{array} \right. , \end{multline*} where $C$ is a constant that depends on $T_x, T_y, K, X, Y$. Using the above inequality, if $ i_1= i_1' $ and $ i_2=i_2' $, we have \begin{align} \label{eq:add1} \begin{split} & \left| \int \int E\left[ \overline{\mathcal{Z}}_{i_1i_2}(t_1, t_2)\overline{\mathcal{Z}}_{i_1'i_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \\ \leq & \frac{1}{N_{i_1}} \frac{1}{N_{i_2}} \sum_{j_1=1}^{N_{i_1}} \sum_{j_2=1}^{N_{i_2}} \frac{1}{N_{i_1'}} \frac{1}{N_{i_2'}} \sum_{j_1'=1}^{N_{i_1'}} \sum_{j_2'=1}^{N_{i_2'}} \left| \int \int E\left[ w_{i_1i_2, j_1j_2}(t_1, t_2)w_{i_1'i_2', j_1'j_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \\ \leq & C \left\{ \frac{N_{i_1}h_{i_1} + N_{i_1}(N_{i_1}-1) h_{i_1}^2}{N_{i_1}^2h_{i_1}^2} \frac{N_{i_2}h_{i_2} + N_{i_2}(N_{i_2}-1)h_{i_2}^2 }{N_{i_2}^2h_{i_2}^2} \right\} \\ = & C \phi_{i_1} \phi_{i_2}. \end{split} \end{align} If $i_1=i_1'$, $i_2\neq i_2'$, we have \begin{multline} \label{eq:add2} \left| \int \int E\left[ \overline{\mathcal{Z}}_{i_1i_2}(t_1, t_2)\overline{\mathcal{Z}}_{i_1'i_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \\ \leq C \left\{ \frac{N_{i_1}h_1 + N_{i_1}(N_{i_1}-1) h_1^2}{N_{i_1}^2h_1^2} \right\} \frac{N_{i_2}N_{i_2'}h_2^2 }{N_{i_2}N_{i_2'}h_2^2} = C \phi_{i_1}. \end{multline} Similarly, for the case $i_1 \neq i_1'$ and $i_2=i_2'$, it holds that \begin{align} \label{eq:add3} \left| \int \int E\left[ \overline{\mathcal{Z}}_{i_1i_2}(t_1, t_2)\overline{\mathcal{Z}}_{i_1'i_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \leq C \phi_{i_2}. \end{align} Then, it follows from the above inequalities that \begin{align*} &E\left[ \int_{0}^{1} \int_{0}^{1} L_{1}(t_1,t_2)^2 dt_1 dt_2 \right] \\ \lesssim & \frac{1}{n^2m^2} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+ m} \sum_{1 \leq i_1' \leq n} \sum_{n+1 \leq i_2' \leq n+ m} \left| \int \int E\left[ \overline{\mathcal{Z}}_{i_1i_2}(t_1, t_2)\overline{\mathcal{Z}}_{i_1'i_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \\ \lesssim & \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{i_1} + \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{i_2} + \frac{1}{n^2m^2} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{i_1} \phi_{i_2} \\ \lesssim & \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{i_1} + \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{i_2}. \end{align*} The bound for $ E[ \int_{0}^{1} \int_{0}^{1} L_{2}(t_1,t_2)^2 dt_1dt_2 ]$ can be derived using the same tactics. \end{proof} \subsection{Proof of Theorem \ref{thm:per}} \label{sec:5.2} For any permutation $\pi$, let $\widehat{G}_{\pi,I}$ be the estimated function based on the permuted sample \begin{align*} \pi \cdot \mathbf{Z} = (\mathbf{z}_{\pi(1) } ; \mathbf{z}_{\pi(2) }; \dots ; \mathbf{z}_{\pi(n+m)} ). \end{align*} Correspondingly, the explicit form of $\widehat{G}_{\pi, I}$ depends on the following quantities \begin{align*} & U_{\pi,1}^{p_1p_2}(t_1,t_2) = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+m} \mathcal{T}_{\pi(i_1) \pi(i_2)}, \\ & U_{\pi,2}^{p_1p_2}(t_1, t_2) = \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \mathcal{T}_{\pi(i_1) \pi(i_2)}, \\ & V_{\pi, 1}^{p_1p_2}(t_1,t_2) = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+m} \mathcal{Z}_{\pi(i_1) \pi(i_2)}, \\ & V_{\pi, 2}^{p_1p_2}(t_1, t_2) = \frac{2}{n(n-1)} \sum_{1 \leq i_1 < i_2 \leq n} \mathcal{Z}_{\pi(i_1) \pi(i_2)}. \end{align*} For $I=1,2, J=1,2,3$, let $W_{\pi, I, J}^{p_1p_2}$ be defined similarly with $W_{I, J}^{p_1p_2}$ with $U_I, V_I$ replaced by $ U_{\pi, I}, V_{\pi, I}$ respectively. Then it can be shown that $ \widehat{G}_{\pi, I}(t_1,t_2)$ admits a similar form as $ \widehat{G}_{ I}(t_1,t_2)$ with $W_{I}, U_I, V_I$ replaced by $W_{\pi, I}$, $U_{\pi, I}, V_{\pi, I}$. The following lemma shows that $U_{\pi, I}, V_{\pi, I}$ as well as $U_{\Pi, I}, V_{\Pi, I}$ converge to their mean functions, where $\Pi$ is a random permutation sampled uniformly from $\mathbb{P}_{n+m}$. \begin{lemma} \label{lemma:power} Under the assumptions of Theorem \ref{thm:per}, for any $p_1, p_2=0,1,2$, we have \begin{itemize} \item[(i)] for any fixed permutation $\pi \in \mathbb{P}_{n+m}$, \begin{multline*} \int \int E \left( V_{\pi,I}^{p_1p_2}(t_1,t_2) - E\left[ V_{\pi,I}^{p_1p_2}(t_1,t_2) \right] \right)^2 dt_1 dt_2 \\ \leq C \left\{ \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{\pi(i_1)} + \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{\pi(i_2)} \right\}, \end{multline*} where $C$ is a constant that depends on $T_x, T_y, K, X, Y$. Moreover, $\int \int E ( U_{\pi,I}^{p_1p_2}(t_1,t_2) - E[ U_{\pi,I}^{p_1p_2}(t_1,t_2) ] )^2 dt_1 dt_2 $ satisfies the same bound. \item[(ii)] For any random permutation $\Pi$ sampled from $\mathbb{P}_{n+m}$ uniformly, \begin{multline*} \int \int E \left( V_{\Pi,I}^{p_1p_2}(t_1,t_2) - E\left[ V_{\Pi,I}^{p_1p_2}(t_1,t_2) \right] \right)^2 dt_1 dt_2 \\ \leq C \left\{ \sup_{\pi \in \mathbb{P}_{n+m}} \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{\pi(i_1)} + \sup_{\pi \in \mathbb{P}_{n+m}} \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{\pi(i_2)} \right\}, \end{multline*} where $C$ is a constant that depends on $T_x, T_y, K, X, Y$. Moreover, $\int \int E ( U_{\Pi,I}^{p_1p_2}(t_1,t_2) - E[ U_{\Pi,I}^{p_1p_2}(t_1,t_2) ] )^2 dt_1 dt_2 $ can attain the same rate as above. \end{itemize} \end{lemma} \begin{proof} (i) Here, we only show the result for $V_{\pi, 1}$, as the proof for $V_{\pi, 2}$ is similar. Set $ \widetilde{V}_{\pi, 1} = \frac{1}{nm} \sum_{1 \leq i_1 \leq n} \sum_{n+1 \leq i_2 \leq n+m} \widetilde{\mathcal{Z}}_{\pi(i_1) \pi(i_2)} , $ where \begin{align*} \widetilde{\mathcal{Z}}_{\pi(i_1) \pi(i_2)} = \mathcal{Z}_{\pi(i_1) \pi(i_2)} - E[\mathcal{Z}_{\pi(i_1) \pi(i_2)}]. \end{align*} Since $E[\widetilde{\mathcal{Z}}_{\pi(i_1) \pi(i_2)} \widetilde{\mathcal{Z}}_{\pi(i_1') \pi(i_2')} ] \neq 0$ only if $ \{ \pi(i_1), \pi(i_2) \} \cap \{ \pi(i_1'), \pi(i_2') \} \neq \emptyset $, by similar arguments with Equations \eqref{eq:add1}, \eqref{eq:add2} and \eqref{eq:add3}, we can show that \begin{align*} \left| \int \int E\left[ \widetilde{\mathcal{Z}}_{i_1i_2}(t_1, t_2)\widetilde{\mathcal{Z}}_{i_1'i_2'}(t_1, t_2) \right] dt_1 dt_2 \right| \leq C \left\{ \begin{array}{cc} \phi_{\pi(i_1)} \phi_{\pi(i_2)} & \text{ if } i_1 = i_1', i_2 = i_2', \\ \phi_{\pi(i_1)} & \text{ if } i_1 = i_1', i_2 \neq i_2', \\ \phi_{\pi(i_2)} & \text{ if } i_1 \neq i_1', i_2 = i_2', \end{array} \right. \end{align*} which implies \begin{multline*} \int \int E \left( V_{\pi,I}^{p_1p_2}(t_1,t_2) - E\left[ V_{\pi,I}^{p_1p_2}(t_1,t_2) \right] \right)^2 dt_1 dt_2 \\ \leq 2 C \left\{ \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{\pi(i_1)} + \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{\pi(i_2)} \right\}. \end{multline*} The bound on $\int \int E ( U_{\pi,I}^{p_1p_2}(t_1,t_2) - E[ U_{\pi,I}^{p_1p_2}(t_1,t_2) ] )^2 dt_1 dt_2 $ can be shown similarly. (ii) The result follows from the inequality \begin{multline*} \int \int E \left( V_{\pi,I}^{p_1p_2}(t_1,t_2) - E\left[ V_{\pi,I}^{p_1p_2}(t_1,t_2) \right] \right)^2 dt_1 dt_2 \\ \leq 2 C \left\{ \sup_{\pi \in \mathbb{P}_{n+m}} \frac{1}{n^2} \sum_{1 \leq i_1 \leq n} \phi_{\pi(i_1)} + \sup_{\pi \in \mathbb{P}_{n+m}} \frac{1}{m^2} \sum_{n+1 \leq i_2 \leq n+ m} \phi_{\pi(i_2)} \right\} \end{multline*} and the fact that $C$ is a constant independent of $\pi$. \end{proof} \begin{lemma} \label{lem:power2} Under the assumptions of Theorem \ref{thm:per}, for any $p_1, p_2=0,1,2$, we have \begin{itemize} \item[\emph{(i)}] for any fixed permutation $\pi \in \mathbb{P}_{n+m}$, \begin{align*} \emph{\text{sup}}\limits_{t_1,t_2} \left| U_{\pi,I}^{p_1p_2}(t_1,t_2) - E\left[ U_{\pi,I}^{p_1p_2}(t_1,t_2) \right] \right| = o_p(1). \end{align*} \item[\emph{(ii)}] For any random permutation $\Pi$ sampled from $\mathbb{P}_{n+m}$ uniformly, \begin{align*} \emph{\text{sup}}\limits_{t_1,t_2} \left| U_{\Pi,I}^{p_1p_2}(t_1,t_2) - E\left[ U_{\Pi,I}^{p_1p_2}(t_1,t_2) \right] \right| = o_p(1). \end{align*} \end{itemize} \end{lemma} \begin{proof} (i) We can write $ \mathcal{T}_{i_1i_2}^{p_1p_2} = \mathcal{S}_{i_1}^{p_1} \mathcal{S}_{i_2}^{p_2} $ and \begin{align*} U_{\pi,1}^{p_1p_2}(t_1,t_2) & = R_{\pi, 1}^{p_1}(t_1) R_{\pi, 2}^{p_2}(t_2), \\ U_{\pi,2}^{p_1p_2}(t_1,t_2) & = R_{\pi, 1}^{p_1}(t_1) R_{\pi, 1}^{p_1}(t_1) - \frac{1}{n^2} \sum_{i_1=1}^n (\mathcal{S}_{\pi(i_1)}^{p_1}(t_1))^2, \end{align*} where $ R_{\pi, 1}^{p_1}(t_1) & = \frac{1}{n} \sum_{i_1=1}^n \mathcal{S}_{\pi(i_1)}^{p_1}(t_1) \text{ and } R_{\pi, 2}^{p_2}(t_2) = \frac{1}{m} \sum_{i_2=n+1}^{n+m} \mathcal{S}_{\pi(i_2)}^{p_2}(t_2)$ and \begin{align*} \mathcal{S}_{i_1}^{p_1}(t_1) & = \frac{1}{N_{i_1}} \sum_{j_1=1}^{N_{i_1}} K_{h_{i_1}}(T_{i_1 j_1}- t_1) \left( \frac{T_{i_1j_1}-t_1}{h_{i_1}} \right)^{p_1}, \\ \mathcal{S}_{i_2}^{p_2}(t_2) & = \frac{1}{N_{i_2}} \sum_{j_2=1}^{N_{i_2}} K_{h_{i_2}}(T_{i_2 j_2}- t_2) \left( \frac{T_{i_2j_2} - t_2}{h_{i_2}} \right)^{p_2} . \end{align*} It is sufficient to show that $\text{sup}_{t_1} \left| R_{\pi,I}(t_1) - E\left[ R_{\pi,I}(t_1) \right] \right| = o_p(1)$. Here, we only provide details for the case $I=1$ and the rest can be shown similarly. The superscrips $p_1, p_2$ will be dropped if there is no confusion. Set \begin{align*} \widetilde{R}_{\pi, 1} = \frac{1}{n} \sum_{1 \leq i_1 \leq n} \widetilde{\mathcal{S}}_{\pi(i_1)} , \end{align*} where $ \widetilde{\mathcal{S}}_{\pi(i_1)} = \mathcal{S}_{\pi(i_1)} - E[\mathcal{S}_{\pi(i_1)}]$, and \begin{align*} a_n = \left\{ \log\left( \frac{n}{\sum_{i=1}^n N_{\pi(i)}^{-1}/n} \right) \frac{\sum_{i=1}^n N_{\pi(i)}^{-1}/n }{\min \{ h_x, h_y \} } \frac{1}{n} \right\}^{1/2} , \ \ b_n = \frac{1}{(n^{-2}\sum_{i=1}^n N_{\pi(i)}^{-1})^2} . \end{align*} Let $\chi (b_n) \subset [0,1]$ be the set of equally spaced grid points of size $b_n$, then \begin{align*} \sup\limits_{t_1} \big| \widetilde{R}_{\pi, 1}(t_1) \big| & \lesssim \sup\limits_{t_1 \in \chi(b_n)} | \widetilde{R}_{\pi, 1}(t_1) | + \sup_{|t_1 -t_1'|< 1/b_n } | \widetilde{R}_{\pi, 1}(t_1) - \widetilde{R}_{\pi, 1}(t_1') |. \end{align*} To bound the second term, note that \begin{multline*} | \widetilde{R}_{\pi, 1}(t_1) - \widetilde{R}_{\pi, 1}(t_1') | \leq \\ \frac{1}{n}\sum_{1 \leq i_1 \leq n} \Big\{ \left| \mathcal{S}_{\pi(i_1)}(t_1) - \mathcal{S}_{\pi(i_1)}(t_1') \right| + \left| E\big[\mathcal{S}_{\pi(i_1)}(t_1)\big] - E\big[\mathcal{S}_{\pi(i_1)}(t_1')\big] \right| \Big\}. \end{multline*} For any $t_1, t_1'$ such that $|t_1-t_1'| < 1/b_n$, \begin{align} \label{eq:hh1} \begin{split} & \sup_{|t_1 -t_1'|< 1/b_n } \frac{1}{n}\sum_{1 \leq i_1 \leq n} \left| \mathcal{S}_{\pi(i_1)}(t_1) - \mathcal{S}_{\pi(i_1)}(t_1') \right| \\ \leq & C_K \sup_{|t_1 -t_1'|< 1/b_n } \frac{1}{n}\sum_{1 \leq i_1 \leq n} \frac{1}{N_{\pi(i_1)}} \sum_{j_1=1}^{N_{\pi(i_1)}} \bigg\{ \left| K_{h_{\pi(i_1)}}(T_{\pi(i_1) j_1}- t_1) - K_{h_{\pi(i_1)}}(T_{\pi(i_1) j_1}- t_1') \right| \\ & \hspace{3cm} + \left| \left( \frac{T_{\pi(i_1)j_1}-t_1}{h_{\pi(i_1)}} \right)^{p_1} - \left( \frac{T_{\pi(i_1)j_1}-t_1'}{h_{\pi(i_1)}} \right)^{p_1} \right| \bigg\} \\ \leq & C_K' \frac{1}{b_n} \frac{1}{\min \{h_x^2, h_y^2\} } \rightarrow 0, \end{split} \end{align} where $C_K, C_K'$ are constants that depend on $K$. Similarly, we can show that $$ \sup_{|t_1 -t_1'|< 1/b_n} \frac{1}{n}\sum_{1 \leq i_1 \leq n} \left| E[\mathcal{S}_{\pi(i_1)}(t_1)] - E[\mathcal{S}_{\pi(i_1)}(t_1')] \right| \leq C_K' \frac{1}{b_n} \frac{1}{\min \{h_x^2, h_y^2\} } \rightarrow 0. $$ By similarly arguments as in the proof of Lemma 2 \cite{zhang2016sparse}, we can show that if $M$ is large enough \begin{align} \label{eq:hh2} P \left( \sup\limits_{t_1 \in \chi(b_n)} \left| \widetilde{R}_{\pi, 1}(t_1) \right| > M a_n \right) \leq 2 \left( \frac{\max_{1 \leq i \leq n+m} N_{i}^{-1}}{n} \right)^{M-C_{K, T} - 2} \rightarrow 0, \end{align} where $C_{K,T}$ is a constant that depends on $K$ and $T_x, T_y$. (ii) The result follows from the fact that the upper bounds in \eqref{eq:hh1} and \eqref{eq:hh2} hold uniformly for all permutations in $\mathbb{P}_{n+m}$. \end{proof} Next, we resume the proof of Theorem \ref{thm:per}. For any $1\leq i_1 \neq i_2 \leq n+m$, we set \begin{align*} \widetilde{T}(t_1, t_2) = E \left[ \mathcal{T}_{\Pi(i_1), \Pi (i_2)} \right] \text{ and } \widetilde{G}(t_1, t_2) = E\left[ \mathcal{Z}_{\Pi(i_1), \Pi(i_2)} \right]. \end{align*} By symmetry of the kernel $K$, $E[U_{\Pi, I}^{p_1p_2}] = 0$ if $p_1=1$ or $p_2=1$. Consequently, by Lemma \ref{lemma:power} and \ref{lem:power2}, if $p_1=1 \text{ or }p_2=1$, \begin{align*} \left\| U_{\Pi, I}^{p_1p_2} (t_1, t_2) \right\|_2 = o_p(1) \text{ and } \sup_{t_{1}, t_{2}} \left| U_{\Pi, I}^{p_1p_2} (t_1, t_2) \right| = o_p(1) . \end{align*} Otherwise, $\sup_{t_{1}, t_{2}} \left| U_{\Pi, I}^{p_1p_2} (t_1, t_2) \right|$ would converge to a positive constant in probability. This also implies that the denominator of $\widehat{G}_{\Pi, I}(t_1, t_2)$ is bounded away from 0 with high probability. Thus, \begin{align*} & \int \int \left| \widehat{G}_{\Pi, I}(t_1, t_2) - \frac{\widetilde{G}(t_1, t_2)}{\widetilde{T}(t_1, t_2)} \right| dt_1 dt_2 \\ = & O_p \bigg( \int \int | U_{\Pi, I}^{0,0}(t_1, t_2) - \widetilde{T}(t_1, t_2) | dt_1 dt_2 + \int \int \left| V_{\Pi, I}^{0,0}(t_1, t_2) - \widetilde{G}(t_1, t_2) \right| dt_1 dt_2 \\ & + \int \int | U_{\Pi, I}^{1,0}(t_1, t_2) V_{\Pi, I}^{1,0}(t_1, t_2) | dt_1 dt_2 + \int \int | U_{\Pi, I}^{0,1}(t_1, t_2) V_{\Pi, I}^{1,0}(t_1, t_2) | dt_1 dt_2 \\ & + \int \int | U_{\Pi, I}^{1,0}(t_1, t_2) V_{\Pi, I}^{0,1}(t_1, t_2) | dt_1 dt_2 + \int \int | U_{\Pi, I}^{0,1}(t_1, t_2) V_{\Pi, I}^{0,1}(t_1, t_2) | dt_1 dt_2 \bigg) \\ = & O_p \left( \left\| U_{\Pi, I}^{0,0} - \widetilde{T} \right\|_2 + \left\| V_{\Pi, I}^{0,0} - \widetilde{G} \right\|_2 + \left\| U_{\Pi, I}^{1,0} \right\|_2 + \left\| U_{\Pi, I}^{0,1} \right\|_2 \right) \\ = & O_p \left( \sup_{\pi \in \mathbb{P}_{n+m}} \sqrt{\frac{1}{n^2} \sum_{i_1=1}^n \phi_{\pi(i_1)}} + \sup_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{m^2} \sum_{i_2=n+1}^{n+m} \phi_{\pi(i_2)}} \right), \end{align*} which entails that \begin{multline*} \left|\text{MED}(\Pi \cdot \mathbf{Z}) \right| \leq \int_{0}^1 2 \left| \widetilde{G}_{\Pi, 1} (t, t) - \frac{\widetilde{G}(t, t)}{\widetilde{T}(t, t)} \right| \\ + \left| \widetilde{G}_{\Pi, 2} (t,t) - \frac{\widetilde{G}(t, t)}{\widetilde{T}(t, t)} \right| + \left| \widetilde{G}_{\Pi, 3}(t,t)-\frac{\widetilde{G}(t, t)}{\widetilde{T}(t, t)} \right| dt \\ = O_p \left( \sup_{\pi \in \mathbb{P}_{n+m}} \sqrt{\frac{1}{n^2} \sum_{ i_1 =1}^n \phi_{\pi(i_1)}} +\sup_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{m^2} \sum_{i_2=n+1}^{n+m} \phi_{\pi(i_2)}} \right) . \end{multline*} \subsection{Proof of Theorem \ref{thm:power}} Under $H_A$, set $c = \text{MED}(X, Y) \geq 0$. By Theorem \ref{thm:1} and \ref{thm:per}, \begin{align*} P_{H_A} \left( \widehat{p} \leq \alpha \right) & \geq P_{H_A} \left( |\text{MED}_n( \mathbf{Z})-c| < c/2, |\text{MED}_n( \Pi_l \cdot \mathbf{Z})| < c/2, l = 1, \dots, S-1 \right) \\ & \geq 1 - \left\{ P_{H_A} \left( |\text{MED}_n( \mathbf{Z})-c| \geq c/2\right) + (S-1) P_{H_A}\left( |\text{MED}_n( \Pi_l \cdot \mathbf{Z})| \geq c/2 \right) \right\} \\ & \rightarrow 1. \end{align*} \subsection{Proof of Corollary \ref{cor:2}} Under Assumptions \ref{ass:1}, \ref{ass:2new} and \ref{ass:2}, the result can be shown by adopting similar arguments as in the proof of Theorem \ref{thm:1} by replacing $z_{i_1j_1}, z_{i_2j_2}$, $G_{1}$, $G_{2}$, $\widehat{G}_{1}$, $\widehat{G}_{2}$ with $\widetilde{z}_{i_1j_1}, \widetilde{z}_{i_2j_2}$, $H_{1}$, $H_{2}$, $ \widehat{H}_{1} $, $\widehat{H}_{2} $ respectively. \subsection{Proof of Theorem \ref{thm:2}} \begin{proof} Denote the density functions of $ X(t) + e_1$, $Y(t) + e_2 $, $X(t)$, $Y(t)$, $e_1$, $e_2$ as $\widetilde{f}_{x}(\cdot|t)$, $\widetilde{f}_y(\cdot|t)$, $f_{x}(\cdot|t)$, $f_y(\cdot|t)$, $\eta_1(\cdot)$ and $\eta_2( \cdot)$ respectively. Under Assumption \ref{ass:3}, we have $\eta_1 = \eta_2$. Suppose $ X(t) + e_1 = ^d Y(t) + e_2 $, then for all $a \in \mathbb{R}$ \begin{align*} \int_{-\infty}^{\infty} f_{x}(u|t) \eta_{1}(a-u) du = \widetilde{f}_{x}(a|t) = \widetilde{f}_y(a|t) = \int_{- \infty}^{\infty} f_y(u|t) \eta_{2}(a-u) du, \end{align*} which entails that $f_{x}(u|t) - f_{y}(u|t)$ is orthogonal to $l_{a}(u) = \eta_{1}( a - u )$ for any $a \in \mathbb{R}$. By Wiener's Tauberian theorem, the span of $\{ l_{a}(u) = \eta_{1}( a - u ) \ | \ a \in \mathbb{R} \}$ forms a dense subset of $L^2(\mathbb{R})$. Thus, we can conclude that $ f_{x}(u|t) - f_{y}(u|t)=0 $. \end{proof} \subsection{Proof of Corollary \ref{cor:powerconta}} \begin{proof} Under Assumptions \ref{ass:1}, \ref{ass:power} and \ref{ass:2}, we can adopt similar arguments in Section \ref{sec:5.2} and show that \begin{align*} \left|\text{MED}_n( \Pi \cdot \widetilde{\mathbf Z}) \right| =O_p \left( \sup\limits_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{n^2} \sum_{i=1}^{n} \phi_{\pi(i)} } + \sup\limits_{\pi \in \mathbb{P}_{n+m}} \sqrt{ \frac{1}{m^2} \sum_{i=n+1}^{n+m} \phi_{\pi(i)} } \right). \end{align*} By Corollary \ref{cor:2} and Theorem \ref{thm:2}, the result follows similarly from Theorem \ref{thm:power}. \end{proof} \end{comment} \bibliographystyle{rss}
\section{Introduction} In this paper, we consider slow-fast systems of the form \begin{equation}\eqlab{uvfast} \begin{aligned} \dot u &=\epsilon U(u,v,\epsilon),\\ \dot v&=V(u,v,\epsilon), \end{aligned} \end{equation} with $U$ and $V$ sufficiently smooth and $0<\epsilon\ll 1$. Systems of this form occur in many different applications, including neuroscience \cite{amir2002a,izhi,rinzel1987a}, biology and chemical reaction networks \cite{epstein1996a,goryachev1997a} and many other areas, see \cite{desroches2012a,kuehn2015} for further references. Geometric Singular Perturbation Theory \cite{jones_1995} is a collection of methods, based upon the ground-breaking work of Fenichel \cite{fen1,fen2,fen3}, that can be used to study systems of the form \eqref{uvfast} for $0<\epsilon\ll 1$. The point of departure for this theory, is the critical manifold \begin{align*} S= \{(u,v)\vert V(u,v,0)=0\}, \end{align*} which is a set of equilibria for the associated layer problem: \begin{equation}\eqlab{layeruv} \begin{aligned} \dot u &=0,\\ \dot v &=V(u,v,0), \end{aligned} \end{equation} obtained by setting $\epsilon=0$ in \eqref{uvfast}. $S$ is said to be normally hyperbolic if all eigenvalues of $D_vV(u,v)$ for all $(u,v)\in S$ have nonzero real part. In particular, it is attracting (repelling) if the real part of these eigenvalues is negative (positive, respectively). Fenichel's theory, see \cite{fen1,fen2,fen3,jones_1995}, then says that compact submanifolds $S_0$ of $S$ perturb to diffeomorphic locally invariant manifolds $S_\epsilon$ for all $0<\epsilon\ll 1$. The reduced flow on $S_\epsilon$ is to leading order given by the reduced problem: \begin{align*} u' &= U(u,v,0),\\ 0&=V(u,v,0), \end{align*} obtained by writing \eqref{uvfast} in terms of the slow time $\tau = t\epsilon^{-1}$ with $()'=\frac{d}{d\tau}$ and subsequently letting $\epsilon\rightarrow 0$. Moreover, stable and unstable manifolds of $S_0 also perturb to $W^s(S_\epsilon)$ and $W^u(S_\epsilon)$, each having invariant foliations by fibers; for full details see e.g. \cite{jones_1995}. Following work by Dumortier and Roussarie on the blowup method \cite{dumortier_1996}, there was in the early parts of the 2000s an effort \cite{krupa_extending_2001,krupa2001a,krupa2008a,szmolyan_canards_2001,wechselberger_existence_2005} to extend the geometric theory of Fenichel to points where normal hyperbolicity breaks down. The simplest type of breakdown is perhaps observed in $\mathbb R^2$ and $\mathbb R^3$ with $S$ having folds, that divide the critical manifold into attracting and repelling subsets. In these cases, canard solutions are solutions of \eqref{uvfast} for $0<\epsilon\ll 1$ that -- counter-intuitively -- follow the attracting and repelling branches of the critical manifold by passing close to the fold. Canards are well described in $\mathbb R^2$ and $\mathbb R^3$ \cite{dumortier_1996,Benoit81,beno2001a,krupa2001a,szmolyan_canards_2001,krupa2008a,vonew} and play an important role in applications and in the global dynamics of systems of the form \eqref{uvfast}, see e.g. \cite{desroches2012a,mujica2017a,Vo}. In $\mathbb R^2$, canards of folded critical manifolds require an unfolding parameter \cite{krupa2001a} and here canard orbits may be limit cycles. In fact, the reference \cite{krupa2001a} proves that there is a family of periodic orbits -- under some non-degeneracy conditions -- that include small (i.e. of size $o(1)$) Hopf cycles and canard cycles of size $\mathcal O(1)$. Under some additional global properties, such family may be extended further to include canards with ``head'', see \cite{krupa2001a}, and eventually relaxation oscillations, as in the van der Pol system \cite{pol1920a,krupa2001a}. This situation is also known as the canard explosion \cite{bronsBarEli}, due to the fact that the canard limit cycles of different amplitude differ in parameter values by an order of $\mathcal O(e^{-c/\epsilon})$ for all $0<\epsilon\ll 1$, see full details in \cite{krupa2001a}. In $\mathbb R^3$, on the other hand, with $u\in \mathbb R^2$ and $v\in \mathbb R$ in \eqref{uvfast}, canards of folded critical manifolds are generic, without parameters. They appear persistently at so-called folded singularities, which are singular points on the fold of a ``desingularized'' reduced problem, see \cite{beno2001a,szmolyan_canards_2001}. The folded singularities come in different generic versions: folded node, folded saddle and folded focus depending on the type of singularity, with only the former two producing canard solutions. The folded node is of particular interest due to its connection to mixed-mode oscillations. In summary, the folded node gives rise to a weak canard (under a nonresonance condition) and a strong canard -- essentially due to the weak and the strong directions of the linearization of the node -- and close to folded singularity, it has been shown, using blowup \cite{szmolyan_canards_2001}, that the tangent space of the attracting slow manifold twists a finite number of times along the weak canard. This implies, due to the contractivity towards the weak canard on the attracting side of the critical manifold, that an open set of points twist upon passage through the folded node. Upon composition with a global return mapping, this provides a simple mechanism for producing attracting limit cycles, see \cite{brons-krupa-wechselberger2006:mixed-mode-oscil}, that are of mixed-mode type, see also \cite{desroches2012a}. The folded saddle-node is a bifurcation of the folded singularity. It comes in different types I and II, but the unfolding of type II -- at the level of the reduced problem -- produces to a transcritical bifurcation of a true singularity and a folded one. In this paper, we will only focus on the type II and we will therefore continue to refer to this case simply as the folded saddle-node. This bifurcation is known to give rise to a Hopf bifurcation \cite{krupa2008a}. It has been called a singular Hopf bifurcation \cite{guckenheimer2008a} due to the fact that the linearization (upon blowup) has eigenvalues of the form $\sim \pm i \omega,\sim \epsilon\lambda$, $\omega,\lambda\ne 0$ as $\epsilon\rightarrow 0$ at the Hopf; it is therefore a zero-Hopf bifurcation \cite{baldom2008a,baldom2019a} for $\epsilon=0$. The interest in the folded saddle-node comes from the fact that it marks the onset/termination of mixed-mode oscillations through the folded node. However, in the author's opinion, the details of this onset/termination and the connection of mixed-mode oscillations with the Hopf cycles is still not fully understood. \cite[Theorem 4.2]{brons-krupa-wechselberger2006:mixed-mode-oscil} relates to the connection problem, but only indirectly. Specifically, the bifurcation described in this theorem -- where a return mechanism transverses to the strong canard -- does not relate to the Hopf bifurcation. The reference \cite{mujica2017a} is another interesting study, based upon detailed numerical computations. Here the Hopf cycles are continued using the software package AUTO and it is demonstrated (for a fixed small value of $\epsilon>0$) that these cycles are of relaxation type without mixed-modes. In particular, the periodic orbits of mixed-mode type in the model system of \cite{mujica2017a} form isolas that are disconnected from the branch of Hopf cycles. At the same time, \cite{guckenheimer2008a} studies a normal form for the folded saddle-node, computing the Lyapunov coefficient and demonstrating additional bifurcations (periodic doubling and torus) using numerical computations along the branch of period orbits that appear from the Hopf. \cite{kukmgp} describes the onset of mixed-modes in a cusped saddle-node in a system with symmetry. It has also been speculated, following work on the Koper model \cite{koper1991a,koper1992a}, that the folded saddle-node may be associated with homoclinics and Shilnikov bifurcations. This was demonstrated for the Koper model in \cite{guckenheimer2015a} using sophisticated numerical methods, among other things. At the same time, it is by now known \cite{baldom2019a,broer1984a} that generic unfoldings of the zero-Hopf bifurcation produce Shilnikov bifurcations. In future work, the present author hopes to pursue these bifurcations rigorously in the context of the folded saddle-node. In preparation, we will in this paper extend the results of \cite{krupa2001a} on the family of periodic canard orbits in $\mathbb R^2$ to the $\mathbb R^3$-context. Whereas the family of canard cycles have ``explosive growth'' in the planar context, the growth rate is regular in $\mathbb R^3$, with canard cycles of different amplitude corresponding (in general) to parameter values that differ by an $\mathcal O(1)$-amount. For this reason, we have chosen to call the phenomena we describe as the ``dud canard'' instead of the canard explosion. \subsection{Setting} We consider the following normal form for the folded node/folded saddle-node\cite{krupa2008a}: \begin{equation}\eqlab{fast} \begin{aligned} \dot x &=\epsilon (y-(\mu+1)z+F(x,y,z,\epsilon,\mu)),\\ \dot y &=\epsilon (\frac12 \mu+ G(x,y,z,\epsilon,\mu)),\\ \dot z &= x+z^2 + z H(x,y,z,\epsilon,\mu), \end{aligned} \end{equation} in the regime $\epsilon>0$, $\epsilon\sim 0$ and $\mu\sim 0$, where $F$, $G$ and $H$ are higher order in the following sense: \begin{align*} F(x,y,z,\epsilon,\mu)=\mathcal O(x,\epsilon,(\vert y\vert+\vert z\vert)^2),\quad G(x,y,z,\epsilon,\mu)=\mathcal O(x,y,z,\epsilon), \end{align*} and \begin{align*} H(x,y,z,\epsilon,\mu)=\mathcal O(xz,xy,z^2,\epsilon). \end{align*} (In comparison with \cite{krupa2008a} we have $z H$ instead of $H$. This plays little role, but the latter can be brought into the former by a transformation of $x$.) $\mu$ will be our bifurcation parameter. The system is normalized in such a way that the layer problem has a critical manifold $S$ of the form $x=m(y,z):=-z^2(1+\mathcal O(y,z))$ with a fold along the line defined by $(0,y,0)$. Locally, the manifold $S$ is normally attracting along $z<0$ and normally repelling along $z>0$. Upon desingularization, the reduced problem takes the following form: \begin{equation}\eqlab{reduced} \begin{aligned} y' &=-z(1+L(y,z,\mu)) (\mu+2G(m(y,z),y,z,0,\mu)),\\ z' &=y-(\mu+1)z+F(m(y,z),y,z,0,\mu), \end{aligned} \end{equation} with $x=m(y,z)$ and $L$ smooth satisfying $L(0,0,\mu)=0$. Consequently, we have a folded singularity \cite{szmolyan_canards_2001} at $(x,y,z)=(0,0,0)$. In fact, $(y,z)=(0,0)$ is partially hyperbolic for \eqref{reduced} with $\mu=0$ and a center manifold reduction shows that \eqref{reduced} undergoes a transcritical bifurcation for $\mu=0$ if \begin{align} \lambda:=\partial_y G(\textbf 0)+\partial_z G(\textbf 0)\ne 0.\eqlab{assumption} \end{align} It is this bifurcation that is known as a folded saddle-node (of type II \cite{krupa2008a}) the slow-fast system \eqref{fast}. We illustrate the bifurcation in \figref{fig:fsn} in terms of the slow time; in comparison with \eqref{reduced} the directions on the repelling sheet are therefore reversed \cite{szmolyan_canards_2001}. Here \figref{fig:fsn} (a) shows $\lambda<0$ whereas \figref{fig:fsn} (b) shows $\lambda>0$. \begin{figure}[h!] \begin{center} \subfigure[$\lambda<0$]{\includegraphics[width=.93\textwidth]{./foldedC123.pdf}} \subfigure[$\lambda>0$]{\includegraphics[width=.93\textwidth]{./foldedC456.pdf}} \end{center} \caption{Illustration of the folded saddle-node of type II. It is a transcritical bifurcation of the desingularized reduced problem \eqref{reduced}. The dynamics illustrated here is in terms of the slow time. Consequently, in comparison with \eqref{reduced}, we have reversed the directions on the repelling sheet. The sign of $\lambda$ (see \eqref{assumption}) determines on what side of the two sheets the saddle point $q$ appears. For $\lambda<0$, shown in (a), $q$ lies on the repelling sheet $C_r$, appearing for $\mu\gtrsim 0$. For $\lambda<0$, shown in (b), $q$ lies on the the attracting sheet $C_a$, appearing for $\mu\lesssim 0$. In both cases, the folded singularity $p$ is a folded node on one side of $\mu$ and folded saddle on the other side. The strong singular canard $\gamma_0(\mu)$ divides $C_a$ into separate sets. In particular, for $\mu\ge 0$ and $\lambda<0$, the dark shaded region in (a) is a funnel region. Here all points pass through $p$ upon following the reduced flow. } \figlab{fig:fsn} \end{figure} On the other hand, for each $\mu\sim 0$, there exists a strong stable manifold $\gamma_0(\mu)$ of $(y,z)=0$ for \eqref{reduced}, known as the strong singular canard in the $(x,y,z)$-space. It is well-known \cite{szmolyan_canards_2001} that $\gamma_0(\mu)$ persists as a (maximal) canard $\gamma_{\epsilon}(\mu)$ connecting fixed copies of Fenichel slow manifolds $S_{a,\epsilon}$ and $S_{r,\epsilon}$ as perturbations of (appropriate) compact subsets $S_{a,0}$ and $S_{r,0}$ of $S_a:=S\cap \{z<0\}$ respectively $S_r:=S\cap \{z>0\}$ for all $0<\epsilon\ll 1$. \subsection{Main result} In this paper, we are interested in canard cycles, i.e. periodic orbits that follow $\gamma_0(\mu)$ on $S$. In particular, under the assumption \eqref{assumption} and analyticity of $F$, $G$ and $H$, we prove the existence of a family of periodic orbits: \begin{thm}\thmlab{main} Suppose \eqref{assumption} and that $F$, $G$ and $H$ are analytic functions in the phase space variables $x$, $y$, $z$ and smooth in $\epsilon$ and $\mu$. Then there exists an $h_1>0$ such that for all $0<\epsilon\ll 1$ the following holds: There is a Hopf bifucation of \eqref{fast} at $\mu=\mu_H(\sqrt \epsilon)$, $\mu_H(0)=0$, and a family of periodic orbits $\Gamma_{h,\epsilon}$, with $h\in (0,h_1]$ measuring the amplitude (to be defined later), along $\mu=\overline \mu(h,\epsilon)$, where $\overline \mu$ is continuous and satisfies $\overline \mu(0,\epsilon)=\mu_H(\sqrt \epsilon)$. For each $h\in (0,h_1]$ fixed, $\Gamma_{h,\epsilon}$ converges as $\epsilon\rightarrow 0$ in Haussdorf distance to a singular canard cycle, consisting of a segment of $\gamma_0(\overline \mu(h,0))$ across the fold and a fast jump. The limit is uniform on compact subsets of $(0,h_1]$. \end{thm} To prove \thmref{main}, we follow \cite{szmolyan_canards_2001}, and apply the following blowup transformation to the extended system (\eqref{fast},$\dot \epsilon=0$) of the folded singularity: \begin{align*} (r,(\bar x,\bar y,\bar z,\bar \epsilon))\mapsto \begin{cases} x &=r^2 \bar x,\\ y&=r\bar y,\\ z &=r\bar z,\\ \epsilon &=r^2\bar \epsilon. \end{cases} \end{align*} for $r\ge 0$, $(\bar x,\bar y,\bar z,\bar \epsilon)\in S^3$ and $\mu\sim 0$. In constrast to \cite{krupa2008a}, we will not blowup $\mu = 0$ (at this stage, at least). We use two separate charts $\bar x=-1$ and $\bar \epsilon=1$ with chart-specific coordinates $ (\epsilon_1,r_1,y_1,z_1)$ and $ (r_2,x_2,y_2,z_2)$ defined by : \begin{align} (\epsilon_1,r_1,y_1,z_1) \mapsto \begin{cases} x &=-r_1^2,\\ y&=r_1y_1,\\ z &=r_1 z_1,\\ \epsilon &=r_1^2 \epsilon_1, \end{cases}\eqlab{barxneg1}\\ (r_2,x_2,y_2,z_2) \mapsto \begin{cases} x &=r_2^2x_2,\\ y&=r_2y_2,\\ z &=r_2 z_2,\\ \epsilon &=r_2^2, \end{cases} \eqlab{bareps1} \end{align} respectively. The charts overlap on $\bar x<0$ and here the change of coordinates are given by the expressions: \begin{equation}\eqlab{cc12} \begin{aligned} r_2 & = r_1 \sqrt{\epsilon_1},\\ x_2 &= -\epsilon_1^{-1},\\ y_2 &= y_1/\sqrt{\epsilon_1},\\ z_2 &= z_1/\sqrt{\epsilon_1}, \end{aligned} \end{equation} for $\epsilon_1>0$. In \cite{guckenheimer2008a,krupa2008a} the authors also describe the Hopf bifurcation in the $\bar \epsilon=1$-chart. The new contribution of \thmref{main} is that we describe the family of periodic orbits bifurcating from the Hopf in a full (i.e. $\epsilon$-independent) neighborhood of the folded saddle-node for all $0<\epsilon\ll 1$. This family includes periodic orbits with amplitude $h=\mathcal O(1)$ following the strong canard $\gamma_0(\mu)$ for $\mu\sim 0$ and all $0<\epsilon\ll 1$. Our approach is similar to the approach for the analysis of the canard explosion in the planar case, see \cite{krupa2001a}. The paper \cite{krupa2001a} also uses a Melnikov approach to extend the Hopf cycles in the associated scaling chart and then subsequently extend these to canard cycles by working in directional charts. The latter connection problem is already complicated in \cite{krupa2001a}. In the present paper, we feel that our proof in $\mathbb R^3$ is relatively simple. It basically extends the classical way of obtaining canard cycles in $\mathbb R^2$, by flowing points forward and backward along the attracting and repelling sheets and then extending this close to the folded saddle-node through blowup (using the chart $\bar x=-1$). Shilnikov variables \cite{deng1988a} and normal forms \cite{Guckenheimer97,ilyashenko1991a} are used to study the necessary transition maps. This approach could also be used as an alternative to solving the connection problem in $\mathbb R^2$, although we have not attempted to do so. In contrast, the problem of connecting the Hopf cycles with the ones obtained by the Melnikov analysis is more complicated here in $\mathbb R^3$ than in the $\mathbb R^2$-context of \cite{krupa2001a}. This is also related to our assumption on analyticity of $F$, $G$ and $H$ in \thmref{main}, which may seem unusual for results in this direction. To explain the difficulty, we recall from \cite{krupa2008a} that there is a critical manifold in the $\bar \epsilon=1$-chart for $\mu=r_2=0$ given by $x_2=y_2^2,z_2=y_2$, $y_2\in \mathbb R$. The linearization has imaginary eigenvalues at $y_2=0$ and the reduced problem has a hyperbolic equilibrium precisely at this point; this is what produces the Hopf bifurcation for all $0<r_2\ll 1$. However, it is nontrivial to study the Hopf cycles in a fixed (small) neighborhood of the Hopf since the eigenvalues are of the form $\pm (1+\mathcal O(r_2)) i, r_2 (\lambda+\mathcal O(r_2))$, i.e. a zero-Hopf bifurcation occurs at $\mu=r_2=0$. We find that there are two ways to perform the analysis of the Hopf cycles: (i) Perfom a center manifold reduction and apply the Hopf bifurcation theorem there. Or: (ii) Straighten out the (strong) unstable/stable manifold ($\lambda\gtrless 0$, respectively), introduce polar coordinates in the transverse direction and apply Melnikov-like methods to construct the periodic orbits as fixed-points of a return map. However, both approaches are not uniform in $\epsilon\rightarrow 0$ in the smooth setting, due to the fact that the eigenvalue with the nonzero real part $\sim \epsilon\lambda$, providing the necessary hyperbolicity, goes to zero as $\epsilon\rightarrow 0$. Nevertheless, in the analytic case, we are able to extend the proof of the unstable/stable manifold to a fixed neighborhood (by following \cite[Section 3]{de2020a}), and this allows us to apply approach (ii). In the author's opinion, there is no clear way to extend the proof of the unstable/stable in the smooth setting since this rests on the exponential estimates (that are nonuniform in the present context). \subsection{Overview} In the remainder of the paper, we work to prove \thmref{main}. First in \secref{small}, we describe the small periodic orbits, extending all the way down to the Hopf bifurcation. Subsequently, in \secref{inter}, we describe the ``intermediate orbits'' that connect small periodic orbits with the canard orbits. In \secref{completing}, we complete the proof of the theorem. Finally, in \secref{discuss} we conclude the paper through a discussion of the results and potential future work. \section{Existence of small periodic orbits}\seclab{small} In the scaling chart $\bar \epsilon=1$, applying \eqref{bareps1} to (\eqref{fast},$\dot \epsilon=0$) gives \begin{equation}\eqlab{eqn2} \begin{aligned} \dot x_2 &= y_2-(\mu+1)z_2 + r_2 F_2(x_2,y_2,z_2,r_2,\mu),\\ \dot y_2 &= \frac12 \mu +r_2 a_1 y_2 + r_2 a_2 z_2 + r_2^2 \overline G_2(x_2,y_2,z_2,r_2,\mu),\\ \dot z_2 &=x_2 + z_2^2+r_2 z_2 H_2(x_2,y_2,z_2,r_2,\mu), \end{aligned} \end{equation} and $\dot r_2=0$. Here the desingularization has been achieved by dividing the right hand side by $r_2$. As in \cite{krupa2008a}, we have put $a_1=\partial_x G(\textnormal{\textbf{0}})$, $a_2=\partial_z G(\textnormal{\textbf{0}})$. Recall then that \begin{align*} \lambda:=a_1+a_2\ne 0, \end{align*} by assumption \eqref{assumption}. For $\mu\sim 0$, $r_2\sim 0$ this system is slow-fast with $x_2$ and $z_2$ being fast and $y_2$ slow. In particular, $\mu=r_2=0$ gives an associated layer problem \begin{equation}\eqlab{layer2} \begin{aligned} \dot x_2 &= y_2-z_2,\\ \dot y_2 &=0,\\ \dot z_2 &=x_2+z_2^2. \end{aligned} \end{equation} \begin{lemma}\lemmalab{C2} The set $C_2$ defined by $x_2=y_2^2,z_2=y_2$, $y_2\in \mathbb R$ is a critical manifold of \eqref{layer2}. It is normally attracting for $y_2<0$ and repelling for $y_2>0$. On the other hand, $y_2=0$ is degenerate for $C_2$ with the linearization around $(0,0,0)$ having two imaginary eigenvalues $\pm i$. In particular, \eqref{layer2} is time-reversible within $y_2=0$ and the orbit \begin{equation}\eqlab{gamma20} \begin{aligned} \gamma_{2,0}(0): \begin{cases} x_2(t_2)& = -\frac14 t_2^2+\frac12,\\ z_2(t_2) &=\frac12 t_2, \end{cases} \end{aligned} \end{equation} is a separatix in the $(x_2,z_2)$-plane separating closed periodic orbits $\phi_h(t)=(x_{2h}(t),z_{2h})$, $x_{2h}(0)=-h,z_{2h}(0)=0$ with period $T_0(h)>0$ for any $h>0$, from unbounded orbits. \end{lemma} \begin{proof} The statement regarding the stability of $C_2$ follows from simple calculations. The analysis for $y_2=0$ is also straightforward, see \cite{krupa2001a}. \end{proof} We illustrate the dynamics in the $(x_2,z_2)$-plane in \figref{fig:x2z2}. \begin{figure}[h!] \begin{center} {\includegraphics[width=.53\textwidth]{./x2z2ppyEq0.pdf}} \end{center} \caption{The layer problem \eqref{layer2} in the $(x_2,z_2)$-plane. The canard orbit $\gamma_0$ (orange) separates bounded periodic orbits $\phi_h$ (green) from unbounded orbits. } \figlab{fig:x2z2} \end{figure} The solution \eqref{gamma20} for $\mu=0$ belongs to a $\mu$-family of solutions \begin{equation}\eqlab{gamma2} \begin{aligned} \gamma_{2,0}(\mu): \begin{cases} x_2(t_2)& = -\frac14 t_2^2+\frac12,\\ y_2(t_2) &=\frac{\mu}{2}t_2 ,\\ z_2(t_2) &=\frac12 t_2, \end{cases} \end{aligned} \end{equation} of \eqref{eqn2} for $\mu=r_2=0$. It corresponds to the blowup of the strong canard $\gamma_0(\mu)$, see \cite{szmolyan_canards_2001}. To describe the reduced problem on $C_2$, we consider the scaling \begin{align} \mu = r_2 \mu_2.\eqlab{mu2scaling} \end{align} Then with $\mu_2$ fixed, $r_2=0$ implies $\mu=0$. \cite{krupa2008a} also uses this scaling, but it is strictly speaking not necessary for the present analysis. In fact, it will be crucial to our approach not to scale $\mu$ for the description of the intermediate periodic orbits that we study later in the paper. With \eqref{mu2scaling}, we obtain the following reduced problem on $C_2$: \begin{align*} y_2' &= \frac12 \mu_2+\lambda y_2, \end{align*} in terms of a slow time. For $\lambda\ne 0$, we have a hyperbolic equilibrium at \begin{align}\eqlab{y2eq} y_2= -\frac{\mu_2}{2\lambda}. \end{align} Consequently, for $\mu_2=0$ this equilibrium lies at $y_2=0$, corresponding to the degenerate point of $C_2$. This gives rise to the (singular) Hopf bifurcation \cite{guckenheimer2008a}. As advertised in the introduction, we study the Hopf bifurcation and periodic orbits of \eqref{eqn2} by first proving existence of a slow manifold $Z_{2,r_2}$ (following \cite[Section 3]{de2020a}), analytic in the space variables and smooth in $\epsilon$ and $\mu_2$, as a perturbation of $C_2$ in a fixed neighborhood of $y_2=0$. Then upon straightening out this one-dimensional manifold we ensure that polar coordinates in the normal directions are well-defined. This allows us to set up a return map, defined in a full neighborhood, which we describe using Melnikov theory. \subsection{Existence of an analytic slow manifold} It will be convinient to write $u:=(x_2,z_2)\in \mathbb R^2$. Moreover, let $\Omega(\nu)$ for $\nu>0$ denote the space of real-analytic functions $m:[-\nu,\nu]\rightarrow \mathbb R^2$. Then the following holds. \begin{proposition}\proplab{slowmanifold} Fix $k\in \mathbb N$. Then there are constants $\delta>0$ and $\nu>0$, both sufficiently small, such that the following holds: There exists a locally invariant one-dimensional manifold $Z_{2,r_2}$ of \eqref{eqn2} of the graph form: \begin{align*} u = m(y_2,r_2,\mu_2), \end{align*} where $(r_2,\mu_2)\mapsto m(\cdot,r_2,\mu_2)\in \Omega(\nu)$ is $C^k$ for $0\le r_2\ll 1$, $\mu_2\in [-\delta,\delta]$, satisfying $m(y_2,0,\mu_2)=(y_2,y_2^2)$ such that $Z_{2,0}$ is a submanifold of $C_2$. \end{proposition} \begin{proof} It is elementary to transform the system into the following form: \begin{equation}\eqlab{tildeuv} \begin{aligned} \dot{\tilde u} &= A(v,\mu_2)\tilde u + r_2 R_2(\tilde u,v,r_2,\mu_2),\\ \dot v &=r_2 \left(v + r_2 P_2(\tilde u,v,r_2,\mu_2)\right), \end{aligned} \end{equation} where $v=y_2+\frac{1}{2\lambda}\mu_2$ and \begin{align*} A(v,\mu_2) = \lambda^{-1}\begin{pmatrix} 0 & -1\\ 1 & 2v-\frac{1}{\lambda}\mu_2 \end{pmatrix}. \end{align*} Henceforth we drop the tilde on $u$. The eigenvalues of $A(v,\mu_2)$ are $\pm \lambda^{-1} i+\mathcal O(v,\mu_2)$. Consequently, for all $\mu_2$ sufficiently small, we have that \begin{align*} (q I-A(0,\mu_2))^{-1} \end{align*} exists for all $q\in \mathbb R$. \begin{lemma}\lemmalab{qIAinv} Fix $\mu_2 \in [-\delta,\delta]$ with $\delta>0$ sufficiently small. Then there exists a constant $C>0$ such that \begin{align*} \Vert (qI-A(0,\mu_2))^{-1}\Vert \le \frac{C}{\vert q\vert +1}, \end{align*} for all $q\in \mathbb R$ and all $\mu_2 \in [-\delta,\delta]$. \end{lemma} \begin{proof} Clearly, \begin{align*} \Vert (qI-A(0,\mu_2))^{-1}\Vert \le \tilde C \vert \chi(\mu_2) - q\vert^{-1}, \end{align*} where $\chi(\mu_2)$ and $\overline \chi(\mu_2)$ are the imaginary eigenvalues of $A(0,\mu_2)$. We then use that $\chi(\mu_2)$ is close to imaginary axes to obtain the result. \end{proof} Henceforth we suppress the dependency of $\mu_2$. We then write the system as a first order system in the following form: \begin{equation}\eqlab{eqnsm} \begin{aligned} r_2 v \frac{du}{dv} -A(0) u&= \left(A(v)-A(0)\right) u+ r_2 R_2(u,v,r_2)\\ &- r_2^2 P_2(u,v,r_2)\frac{du}{dv}. \end{aligned} \end{equation} We will solve this equation by using a fixed-point argument. For this purpose define the following Banach norm \begin{align*} \Vert \sum_{k=0}^\infty h_k v^k \Vert:=\sum_{k=0}^\infty \vert h_k\vert \nu^k. \end{align*} on the space $\Omega(\nu)$ of real analytic functions $h(v)=\sum_{k=0}^\infty h_k v^k\in \mathbb R^n$ with $\Vert h\Vert<\infty$ defined on $\vert v\vert \le \nu$ with $\nu>0$. Now, consider first \begin{align} r_2v\frac{du}{dv} - A(0) u =F(v),\eqlab{lineqn} \end{align} with $F:=\sum_{k=0}^{\infty}F_k (\cdot)^k\in \Omega(\nu)$. Then we have the following: \begin{lemma} Let $0\le r_2\ll 1$ and define the linear operator $T_k:\mathbb R^2\rightarrow \mathbb R^2$ by \begin{align*} T_k(F_k) = (r_2 k - A(0))^{-1} F_k,\quad F_k\in \mathbb R^2. \end{align*} Then \begin{align} \Vert T_k\Vert \le \frac{C}{r_2k+1},\eqlab{Tkmap} \end{align} with $C>0$ from \lemmaref{qIAinv} and \begin{align} u(v) =T(F)(v):= \sum_{k=0}^{\infty}T_k(F_k) v^k,\eqlab{Tdef} \end{align} is the unique real-analytic solution of \eqref{lineqn} in $\Omega(\nu)$. The operator $T:\Omega(\nu)\rightarrow \Omega(\nu)$ defined by \eqref{Tdef} is continous: \begin{align} \Vert T(F) \Vert \le C\Vert F\Vert,\eqlab{Tcont} \end{align} for all $F\in \Omega(\nu)$. \end{lemma} \begin{proof} Straightforward. \eqref{Tkmap} follows from \lemmaref{qIAinv} which gives \eqref{Tcont} upon using \eqref{Tdef}. \end{proof} The following is also important. \begin{lemma}\lemmalab{estTfp} Consider $F\in \Omega(\nu)$. Then \begin{align*} \Vert T(r_2 F')\Vert \le C\nu^{-1} \Vert F\Vert, \end{align*} for all $0<r_2\ll 1$. \end{lemma} \begin{proof} If $F(v)=\sum_{k=0}^\infty F_k v^k$ then $F'(v)=\sum_{k=0} (k+1) F_{k+1} v^k$. Consequently, \begin{align*} T(r_2 F')(v) = \sum_{k=0}^\infty T_k (r_2(k+1) F_{k+1}) v^k, \end{align*} and by \lemmaref{qIAinv}: \begin{align*} \Vert T(r_2 F') \Vert &\le \sum_{k=0}^\infty \frac{Cr_2 (k + 1)}{r_2 k+1} \vert F_{k+1}\vert \nu^k \le C\sum_{k=0}^\infty \vert F_{k+1}\vert \nu^k\le C\nu^{-1} \sum_{k=0}^\infty \vert F_{k}\vert \nu^k. \end{align*} \end{proof} To solve \eqref{eqnsm}, we therefore write it as a fixed-point equation: \begin{align*} u = \mathcal L(u), \end{align*} where $\mathcal L$ is the nonlinear operator defined by \begin{align*} \mathcal L:\,u\mapsto T\left((A(v)-A(0))u+r_2 R_2(u,v,r_2)-r_2^2 P_2(u,v,r_2) \frac{du}{dv}\right) \end{align*} Consider the closed subset $\Omega_1(\nu,\sigma)\subset \Omega(\nu)$ defined by $\Vert m\Vert \le \sigma$, $m\in \Omega(\nu)$. \begin{lemma} There exist constants $\nu>0$, $\sigma>0$ and $r_{20}>0$ such that $\mathcal L$ is contraction on $\Omega_1(\nu,\sigma)$ for all $0<r_2\le r_{20}$. \end{lemma} \begin{proof} The proof follows the proof of \cite[Lemma 3.25]{de2020a}, and we therefore leave out the full details. In fact, without the last term $r_2^2 P_2(u,v,r_2) \frac{du}{dv}$ the result is straightforward and obvious. We therefore only describe how to estimate the last term \begin{align*} T\left(r_2^2 P_2(u,v,r_2)\frac{du}{dv}\right), \end{align*} cf. the linearity of $T$. For this we essentially follow the proof of \lemmaref{estTfp}: Write $P_2(u,v,r_2) = \sum_{n=0}^\infty P_{2,n}(r_2)v^n$ for $u=\sum_{k=0}^\infty h_k v^k$. Then \begin{align*} T\left(r_2^2 P_2\frac{du}{dv}\right) =r_2^2 \sum_{k=0}^\infty T_k\left(\sum_{l=0}^{k+1} l h_l P_{2,k+1-l}(r_2)\right)v^k, \end{align*} such that \begin{align*} \Vert T\left(r_2^2 P_2\frac{du}{dv}\right)\Vert &\le r_2 \sum_{k=0}^\infty \Vert T_k\Vert r_2(k+1)\sum_{l=0}^{k+1} \vert h_l\vert \vert P_{2,k+1-l}(r_2) \vert\nu^k\\ &\le r_2 C \sum_{k=0}^\infty \sum_{l=0}^{k+1} \vert h_l \vert \vert P_{2,k+1-l}(r_2) \nu^k\\ &\le r_2 C \nu^{-1} \sum_{n=0}^\infty P_{2,n}(r_2)\nu^n \sum_{k=0}^\infty h_{k} \nu^k\\ &= r_2 C \nu^{-1} \Vert P_2\Vert \Vert u\Vert, \end{align*} which provides the essential estimate for showing that there are $\sigma>0$ and $\nu>0$ small enough, such that $\mathcal L: \Omega_1(\nu,\sigma)\rightarrow \Omega_1(\nu,\sigma)$ is well-defined for all $0\le r_2\ll 1$. The Lipschitz constant of $\mathcal L$ can be bounded from above by a constant $0<c<1$ in a simlar way. \end{proof} By Banach's fixed point theorem, we obtain a unique fixed-point in $\Omega_1(\nu,\sigma)$. This fixed point $\tilde u=\tilde m(v,r_2,\mu_2)$ gives our desired locally invariant manifold of \eqref{tildeuv}. Transforming the result back to the $(x_2,y_2,z_2)$-variables gives the desired result and completes the proof of \propref{this}. \end{proof} The invariant manifold $Z_{2,r_2}$ is a slow manifold extending uniformly with respect to $r_2$ across the degenerate set $y_2=0$, $\mu_2=0$. It is a subset of the stable (unstable) set of \eqref{y2eq} for $\lambda<0$ ($\lambda>0$, respectively). \subsection{Melnikov theory} We now straighten out the slow manifold of \propref{slowmanifold} by writing: \begin{align} \begin{pmatrix} x_2\\ z_2 \end{pmatrix} = m(y_2,r_2,\mu_2) + \tilde u.\eqlab{tildeust} \end{align} Then the invariant manifold corresponds to $\tilde u=0$ and consequently: \begin{equation}\eqlab{tildeuy22} \begin{aligned} \dot{\tilde u} &= B(\tilde u,y_2) \tilde u + r_2 R_2(\tilde u,y_2,r_2,\mu_2) \tilde u,\\ \dot y_2 &=r_2\left(\frac12 \mu_2 + \lambda y_2+P_2(\tilde u,y_2,r_2,\mu_2)\right), \end{aligned} \end{equation} where $P_2(0,y_2,0,\mu_2)=0$ and \begin{align*} B(\tilde u,y_2) = \begin{pmatrix} 0 & -1\\ 1 & 2y_2+\tilde u_2 \end{pmatrix}, \end{align*} for $\tilde u=(\tilde u_1,\tilde u_2)$. For $r_2=y_2=0$, we have $m(0,0,\mu_2)=0$ and hence $\tilde u$ reduces to $(x_2,z_2)$ in this case. We now again drop the tildes. The $u$-plane is therefore for $r_2=0$ also filled with the periodic orbits $\phi_h(t)=(x_{2h}(t),z_{2h}(t))$, intersecting the negative $x_2$-axis in $(-h,0)$ with $h>0$. Moreover, due to the invariance of $u=0$, the return map from $u_1<0,u_2=0$ to itself, mapping $(-h,0,y_{20})$ to $(u_1(T),0,y_2(T)$, with $T=T(h,y_2,r_2,\mu_2)$ being the transition time, is well-defined for all $h>0$, $0\le r_2\ll 1$, $\mu_2 \in [-\delta,\delta]$. In fact, polar coordinates in the $u$-plane is well-defined for all $y_2\in [-\nu,\nu]$, $\mu_2 \in [-\delta,\delta]$, and as a result we obtain: \begin{lemma}\lemmalab{extension} The return map has a smooth extension to $h=0$ with $T(0,y_2,r_2,\mu_2)=2\pi+\mathcal O(y_2,r_2)$. \end{lemma} We now obtain fixed points of the return map using Melnikov theory to perturb away from the family of period orbits $u=\phi_h(t)$, $h>0$, within $y_2=0$ for $\mu_2=r_2=0$. For this purpose, let \begin{align*} A_h(t):=\begin{pmatrix} 0 & -1\\ 1 & 2z_{2h}(t) \end{pmatrix}, \end{align*} denote the linearization around $u=\phi_h(t)$, $y_2=r_2=0$. We then write \begin{align*} u &= \phi_h(t)+\tilde u,\quad y_2=\tilde y_2. \end{align*} and let $\tilde u(t,h,y_2,r_2,\mu_2), \tilde y_2(t,h,y_2,r_2,\mu_2)$ denote the solutions of the resulting differential equations: \begin{align*} \dot{\tilde u} &=A_h(t) \tilde u+\bigg\{B(\phi_h(t)+\tilde u,\tilde y_2)(\phi_h(t)+\tilde u)-A_h(t)\tilde \phi_h(t)- A_h(t)\tilde u\\ &+r_2 R_2(\phi_h(t)+\tilde u,\tilde y_2,r_2,\mu_2)(\phi_h(t)+\tilde u) \bigg\},\\ \dot{\tilde y}_2&=r_2\left(\frac12 \mu_2 + \lambda \tilde y_2+P_2(\phi_h(t)+\tilde u,\tilde y_2,r_2,\mu_2)\right), \end{align*} with initial conditions $\tilde u(0)=0,\tilde y_2(0)=y_2$. We have \begin{align*} \tilde u(t,0,y_2,r_2,\mu_2)\equiv 0, \end{align*} due to $\phi_0(t)\equiv 0$ and the invariance of $u=0$ for \eqref{tildeuy22}. Let $\Phi_h(t,s)$ denote the state-transition matrix associated with $A_h(t)$. Then by variation of constants we have that \begin{align*} \tilde u(T,h,y_2,r_2,\mu_2) &= \int_0^{T}\Phi_h(T,t) \left\{\cdots \right\}dt, \\ \tilde y_2(T,h,y_2,r_2,\mu_2)&=y_2+\int_0^T r_2\left(\cdots\right)dt, \end{align*} where \begin{align*} \left\{\cdots \right\}&=B(\phi_h(t)+\tilde u(T,h,y_2,r_2,\mu_2),\tilde y_2)(\phi_h+\tilde u(t,h,y_2,r_2,\mu_2))-A_h(t)\tilde \phi_h(t)\\ &- A_h(t)\tilde u(t,h,y_2,r_2,\mu_2)+r_2 R_2(\phi_h(t)+\tilde u(t,h,y_2,r_2,\mu_2),\tilde y_2(t,h,y_2,r_2,\mu_2),r_2,\mu_2) \\ &\times(\phi_h(t)+\tilde u(t,h,y_2,r_2,\mu_2)),\\ \left(\cdots\right)&=\frac12 \mu_2 + \lambda \tilde y_2(t,h,y_2,r_2,\mu_2)+P_2(\phi_h(t)+\tilde u(t,h,y_2,r_2,\mu_2),\tilde y_2(t,h,y_2,r_2,\mu_2),r_2,\mu_2). \end{align*} Recall that $T(h,y_2,r_2,\mu_2)$ denotes the transition time, with $T_0(h):=T(h,0,0,0)$ being the period of the periodic orbits $\phi_h(t)=(x_{2h}(t),z_{2h}(t))$. Since the eigenvalues of $B(0,0)$ are $\pm i$, we have \begin{align} \phi_h(t) = -h\begin{pmatrix} \cos t+\mathcal O(h)\\ \sin t+\mathcal O(h))\end{pmatrix},\eqlab{phih} \end{align} for $h\rightarrow 0^+$. Morever: \begin{align} T_0(0):=\lim_{h\rightarrow 0^+}T_0(h)= 2\pi,\eqlab{T00} \end{align} see also \lemmaref{extension}. We then consider the adjoint system: \begin{align*} \dot \psi &= -A_h(t)^T \psi, \end{align*} and let $\psi_{h}(t)$ denote the solution with $\psi_h(T)=(1,0)$. Then a simple calculation shows the following: \begin{lemma} \begin{align*} \psi_{h}(t) &= -h^{-1} e^{\int_t^T 2z_{2h}(s)ds} \begin{pmatrix} z_{2h}'(t)\\ -x_{2h}'(t) \end{pmatrix}. \end{align*} Moreover, \begin{align} \psi_{h}(T) \cdot \tilde u(T,h,y_2,r_2,\mu_2) = \int_0^{T}\psi_{h}(t) \cdot \left\{\cdots\right\} dt.\eqlab{phihdotu} \end{align} \end{lemma} \begin{proof} This is standard and follows from the theory of adjoint equations, see \cite{matsumoto1993a}. In particular, for \eqref{phihdotu} we use $\psi_{h}(T)^T \Phi_h(T,s) = \psi_{h}(s)^T$. \end{proof} We therefore define the Melnikov functions \begin{align*} \Delta_1(h,y_2,r_2,\mu_2) & = \left[x_{2h}(T)-h\right]+ \int_0^{T}\psi_{h}(t) \cdot \left\{\cdots\right\} dt,\\ \Delta_2(h,y_2,r_2,\mu_2) &=\int_0^T \left(\cdots\right)dt. \end{align*} so that roots of $\Delta:=(\Delta_1,\Delta_2)$ for $h>0$ correspond to fixed points of the return map and periodic orbits. We have: \begin{lemma}\lemmalab{Deltah} $\Delta$ also extends smoothly to $h=0$, with $\Delta_1(0,y_2,r_2,\mu_2)\equiv 0$ and $\Delta(h,0,0,0)\equiv 0$ for all $h\ge 0$. \end{lemma} \begin{proof} The extension of $\Delta$ follows from the invariance of $u=0$ of \eqref{tildeuy22} and the fact that $T$ extends smoothly to $h=0$, recall \lemmaref{extension}. The invariance of $u=0$ and the fact that $\phi_0(t)\equiv 0$ also gives $\{\cdots\}=0$ for $h=0$ and therefore $\Delta_1(0,y_2,r_2,\mu_2)\equiv 0$. Finally, $\Delta(h,0,0,0)=0$ since $u=(x_2,z_2)=\phi_h(t)$ is a solution of \eqref{layer2} within $y_2=0$ and $r_2=0$. \end{proof} Following this lemma, the function \begin{align*} \widehat \Delta:=(h^{-1} \Delta_1,\Delta_2), \end{align*} is well-defined and smooth on $h\ge 0$. Moreover: \begin{lemma} $\widehat \Delta(h,0,0,0)\equiv 0$ and \begin{align} \widehat \Delta_{y_2}'(h,0,0,0)&=\begin{pmatrix} -2 \int_0^{T_0(h)} e^{\int_t^T 2z_{2h}(s)ds} (h^{-1} z_{2h})(t)^2 dt\\ *\end{pmatrix},\eqlab{Deltay2}\\ \widehat \Delta_{\mu_2}'(h,0,0,0) &= \begin{pmatrix} 0 \\ \frac12 T_0(h) \end{pmatrix},\eqlab{Deltamu} \end{align} for all $h\ge 0$, with $*$ denoting a quantity that is not important. \end{lemma} \begin{proof} $\widehat \Delta(h,0,0,0)\equiv 0$ follows from \lemmaref{Deltah} and \eqref{Deltamu} is obvious. To show \eqref{Deltay2}, we compute for $h>0$: \begin{align*} \frac{\partial \Delta_1}{\partial y_2}(h,0,0,0)&=\int_0^{T_0} \psi_h(t) \cdot \begin{pmatrix} 0 & 0 \\ 0 & 2 \end{pmatrix}\phi_h(t) dt\\ &= 2h^{-1} \int_0^T e^{\int_t^T 2z_{2h}(s)ds} x_{2h}'(t) z_{2h}(t)dt \\ &=-2h^{-1} \int_0^T e^{\int_t^T 2z_{2h}(s)ds} z_{2h}(t)^2dt. \end{align*} $h^{-1}z_{2h}$ extends to all $h\ge 0$ by \eqref{phih}, which then completes the proof. \end{proof} The Jacobian matrix \begin{align*} \widehat \Delta_{(y_2,\mu_2)}(h,0,0,0): = \begin{pmatrix} \widehat\Delta_{y_2}'(h,0,0,0) & \widehat \Delta_{\mu_2}'(h,0,0,0) \end{pmatrix}, \end{align*} is therefore regular for all $h>0$, regardless of what $*$ is, also for $h=0$ due to \eqref{phih} and \eqref{T00}: \begin{align*} \widehat \Delta_{y_2}'(\textnormal{\textbf 0}) &= \begin{pmatrix} -2\pi \\ *\end{pmatrix},\quad \widehat \Delta_{\mu_2}'(\textnormal{\textbf 0}) = \begin{pmatrix} 0 \\ \pi \end{pmatrix}. \end{align*} \begin{proposition}\proplab{smallcycles} Fix any $h_1>0$. Then there exist smooth functions $\overline y_2(h,r_2),\overline \mu_2(h,r_2)$ with $\overline y_2(h,0)=0$, $\overline \mu_2(h,0)=0$ such that $\widehat \Delta(h,\overline y_2(h,r_2),r_2,\overline \mu_2(h,r_2))=0$ for all $h\in [0,h_1]$ and all $0\le r_2\ll 1$ \end{proposition} \begin{proof} Follows from the implicit function theorem. \end{proof} \begin{corollary} Let $\mu_{H}(r_2):=r_2 \overline \mu_2(0,r_2)$. Then \eqref{eqn2} undergoes a Hopf bifurcation at $\mu=\mu_{H}(r_2)$ for $0<r_2\ll 1$. \end{corollary} \begin{proof} The statement follows directly from \propref{smallcycles}. \end{proof} \begin{remark} We do not aim to describe the stability of the cycles. This is more difficult. However, sufficiently close to the Hopf bifurcation, one can describe stability in terms of the signs of $\lambda$ and $\frac{\partial^2 \overline \mu_2}{\partial h^2}(0,r_2)$; here the latter quantity relates directly to the first Lyapunov coefficient of the Hopf (within a center manifold). In particular, if $\lambda>0$ then $\mu_2\gtrless \mu_H(r_2)$ means that the equilibrium is on the attracting/repelling side of $Z_{2,r_2}$, respectively. Therefore sufficiently close to the equilibrium, the limit (Hopf) cycles are unstable/stable if $\frac{\partial^2 \overline \mu_2}{\partial h^2}(0,r_2)\gtrless$ for $\lambda>0$. If $\lambda<0$, then the inequalities for $\mu_2$ regarding the ``normal stability'' of \eqref{y2eq} become $\mu_2 \lessgtr \mu_H(r_2)$ and the Hopf cycles are unstable/stable if $\frac{\partial^2 \overline \mu_2}{\partial h^2}(0,r_2)\lessgtr 0$ in this case. It is possible to compute a leading order expression for $\frac{\partial^2 \overline \mu_2}{\partial h^2}(0,r_2)$ but we chose not to include the complicated expression in the present paper. Obviously, one could just compute the Lyapunov coefficient by applying the center manifold reduction, see \cite{guckenheimer2008a}. But -- due the zero-Hopf bifurcation for $r_2=0$ -- this approach does not guarantee that this quantity describe the stability of limit cycles in a uniform neighborhood. Our Melnikov approach using \propref{this} in the analytic setting does imply such uniformity. \end{remark} The family of periodic orbits of \eqref{eqn2} are small periodic orbits (with amplitude of order $o(1)$) of \eqref{fast} upon \textit{blowing down} using \eqref{bareps1}. By working in the chart $\bar x=-1$ in the following section, we are able to extend these cycles to ``intermediate'' cycles that include $\mathcal O(1)$ canards. \section{Existence of intermediate periodic orbits}\seclab{inter} Our strategy for extending the small periodic orbits to intermediate ones, that connect to cycles of size $\mathcal O(1)$, follows the approach for proving canard cycles in $\mathbb R^2$: Working in the entry chart $\bar x=-1$, we fix a section along $z_1=0$, in a neighborhood of $(r_1,y_1,\epsilon_1)=0$ with $\mu\sim 0$, and flow the points forward and backward and then measure their separation on a section $\{z_2=0\}$ in the scaling chart $\bar \epsilon=1$ transverse to $\gamma_{2,0}(0)$, recall \eqref{gamma20}. Based upon expansions from the solution of a Shilnikov problem \cite{deng1988a}, we define appropriate scalings that allow us to solve for roots of the separation function by applying the implicit function theorem. \subsection{Analysis in the $\bar x=-1$-chart} We first consider the $\bar x=-1$-chart. Inserting \eqref{barxneg1} into (\eqref{fast},$\dot \epsilon=0$) gives \begin{equation}\eqlab{barxneg1eqns} \begin{aligned} \dot \epsilon_1&=\epsilon_1^2 \left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right],\\ \dot r_1 &=-\frac12 r_1 \epsilon_1 \left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right],\\ \dot y_1&=\epsilon_1 \left(\frac12 \mu+r_1 G_1(\epsilon_1,r_1,y_1,z_1,\mu)\right)+\frac12 \epsilon_1 y_1\left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right] ,\\ \dot z_1&=-1+z_1^2+r_1 z_1 H_1(\epsilon_1,r_1,y_1,z_1,\mu)+\frac12 \epsilon_1 z_1 \left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right], \end{aligned} \end{equation} where $ F_1 = r_1^{-2} F$, $G_1=r_1^{-2} G$, and $H_1 = r_1^{-2} H$, each being smooth, also upon extending to $r_1=0$. The points $(0,0,y_1,- 1)$ are partially attracting, whereas $(0,0,y_1,1)$ are partially repelling. In particular, in each case the linearization has a single nonzero eigenvalue (positive/negative, respectively). Consequently, by center manifold theory we have the following. \begin{lemma} There exists a constant $\sigma>0$ sufficiently small such that the following holds: There exist two three-dimensional center manifolds $M_{1,a}$ and $M_{1,r}$ of the sets $(0,0,y_1,- 1)$ and $(0,0,y_1,1)$, $y_1\in [-\sigma,\sigma]$, having the following graph representations: \begin{align*} M_{1,a}:\quad z_1 &= m_{1,a}(\epsilon_1,r_1,y_1,\mu),\\ M_{1,r}:\quad z_1 &= m_{1,r}(\epsilon_1,r_1,y_1,\mu), \end{align*} for $0\le r_1,\epsilon_1,\vert y_1\vert ,\vert \mu\vert\le \sigma$. The functions $m_{1,a}$ and $m_{1,r}$ are smooth functions on the specified domain and satisfy $m_{1,a}(0,0,y_1,\mu)\equiv -1,m_{1,r}(0,0,y_1,\mu)\equiv 1$ for all $y_1,\mu\in [-\sigma,\sigma]$ as well as \begin{align} m_{1,a}(\epsilon_1,0,\pm \mu,\mu)\equiv -1,m_{1,r}(\epsilon_1,0,\pm \mu,\mu)\equiv 1.\eqlab{m1ar} \end{align} \end{lemma} The properties in \eqref{m1ar} are consequences of the invariant sets $\gamma_{2,0}(\mu)$ in the $\bar \epsilon=1$-chart, recall \eqref{gamma2}, which in the present chart become $r_1=0,y_1=\mp \mu,z_1=\mp 1$ and $\epsilon_1\ge 0$; we denote these sets by $\gamma_{1,0,a}(\mu)$ and $\gamma_{1,0,r}(\mu)$, respectively. We illustrate the dynamics in the $\bar x=-1$-chart in \figref{entry1}. \begin{figure}[h!] \begin{center} \subfigure[$\mu=0$]{\includegraphics[width=.25\textwidth]{./entry1.pdf}} \subfigure[$\mu=\mu_0(r_1,0)$]{\includegraphics[width=.25\textwidth]{./entry2.pdf}} \end{center} \caption{Dynamics in the entry chart $\bar x=-1$. In (a): for $\mu=0$. In (b): for $\mu=\mu_0(r_1,0)$. $\mu_0$ is defined later, but essentially $\mu_0(\tilde r_1,0)$ with $\tilde r_1>0$ is so that for $r_1=\tilde r_1$, $\mu=\mu(\tilde r_1,0)$ there is fast jump connecting $\gamma_0$ to itself.} \figlab{entry1} \end{figure} As is standard, the center manifolds $M_{1,a}$, $M_{1,r}$ provide extensions of the family of Fenichel slow manifolds $\{(S_{a,\epsilon},\epsilon):\epsilon\in [0,\epsilon_0)\}$ as (\eqref{fast},$\dot \epsilon=0$) as foliations (through $r_1^2\epsilon_1=\epsilon=\text{const}.$) of the center manifolds \cite{dumortier_1996}. On $M_{1,a}$ and $M_{1,r}$ we have the following desingularized reduced problems \begin{equation}\eqlab{redMa10} \begin{aligned} \epsilon_1'& = \epsilon_1,\\ r_1'&=-\frac12 r_1,\\ y_1'&=\frac{\frac12 \mu+\mathcal O(r_1)}{y_1+(\mu+1)+\mathcal O(\epsilon_1,r_1)}+\frac12 y_1, \end{aligned} \end{equation} and \begin{equation}\eqlab{redMr10} \begin{aligned} \epsilon_1'& =- \epsilon_1,\\ r_1'&=\frac12 r_1,\\ y_1'&=\frac{\frac12 \mu+\mathcal O(r_1)}{-y_1+(\mu+1)+\mathcal O(\epsilon_1,r_1)}-\frac12 y_1, \end{aligned} \end{equation} obtained by dividing the right hand sides by $\pm \epsilon_1 \left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right]$ and then substituting $z_1=m_{1,a}(\cdots)$ and $z_1=m_{1,r}(\cdots)$, respectively. Notice that the square brackets $\pm \left[y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)\right]$ are positive near $M_{1,a}$ and $M_{1,r}$, respectively, and for $\epsilon_1>0$ the divisions therefore correspond to time transformations. The system \eqref{redMa10} has a hyperbolic equilibrium at $(0,0,-\mu)$ with eigenvalues $-1/2,1,\lambda(\mu)$ where \begin{align} \lambda(\mu) = \frac12 (1-\mu).\eqlab{lambdamu} \end{align} Similarly, \eqref{redMr10} has a hyperbolic equilibrium at $(0,0,\mu)$ with eigenvalues $\frac12,-1,-\lambda(\mu)$. There are therefore strong resonances of both systems when $\mu=0$. In the following, we now describe two transition mappings $P_{1,a}$ and $P_{1,r}$ from $\{z_1=0\}$ to $\{\epsilon_1=\epsilon_{11}\}$ obtained by applying the forward and backward flow, respectively. See illustration in \figref{entry1}. We will describe each mapping in approriate ``normal form coordinates''. Since the two mappings are similar, we focus on $P_{1,a}$. \begin{lemma}\lemmalab{transformnearM1a} Consider \eqref{barxneg1eqns} near $M_{1,a}$. Then for $\sigma>0$ small enough, there exists a regular transformation of time and a smooth diffeomorphism $(y_1,z_1)\mapsto (y_{1,a},z_{1,a})$, $y_1,z_1+1\in [-\sigma,\sigma]$, depending smoothly on $(\epsilon_1,r_1,\mu)$ for $0\le r_1,\epsilon_1,\vert \mu\vert \le \sigma$, such that \begin{equation}\eqlab{y1az1a} \begin{aligned} y_{1,a} &=h_{1,a}^0(\epsilon_1,r_1,y_1,\mu)+\epsilon_1h_{1,a}^1(\epsilon_1,r_1,y_1,z_1,\mu),\\ z_{1,a}&= z_1-m_{1,a}(\epsilon_1,r_1,y_1,\mu), \end{aligned} \end{equation} where \begin{align}\eqlab{h1a0prop} h_{1,a}^0(\epsilon_1,0,-\mu,\mu)\equiv 0, \quad \frac{\partial h_{1,a}^0}{\partial y_1} (\textnormal{\textbf{0}})=1,\quad \frac{\partial h_{1,a}^0}{\partial \mu} (\textnormal{\textbf{0}})=1,\end{align} and \eqref{barxneg1eqns} becomes \begin{equation}\eqlab{eqnsredMa1new2} \begin{aligned} \dot \epsilon_1 &=\epsilon_1^2,\\ \dot r_1 &=-\frac12 r_1 \epsilon_1,\\ \dot{y}_{1,a} &=\left((\lambda(\mu) +r_1 L_0(r_1,y_{1,a},\mu) )y_{1,a}+r_1 \epsilon_1 L_1(\epsilon_1,r_1,y_{1,a},\mu)\right)\epsilon_1,\\ \dot{z}_{1,a} &= (-2+L_2(r_1,y_{1,a},z_{1,a},\epsilon_1,\mu))z_{1,a}, \end{aligned} \end{equation} with $L_0$, $L_1$, $L_2$ each smooth and $L_2(\textnormal{\textbf{0}})=0$. \end{lemma} \begin{proof} We first divide the right hand side by $$y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu),$$ which is positive near $M_{1,a}$. This defines our regular transformation of time. Consequently, \begin{equation}\nonumber \begin{aligned} \dot \epsilon_1&=\epsilon_1^2,\\ \dot r_1 &=-\frac12 r_1 \epsilon_1,\\ \dot y_1&=\frac{\epsilon_1 \left(\frac12 \mu+r_1 G_1(\epsilon_1,r_1,y_1,z_1,\mu)\right)}{ y_1-(\mu+1)z_1+r_1F_1(\epsilon_1,r_1,y_1,z_1,\mu)}+\frac12 \epsilon_1 y_1 ,\\ \dot z_1&=-1+z_1^2+r_1 z_1 H_1(\epsilon_1,r_1,y_1,z_1,\mu)+\frac12 \epsilon_1 z_1 ,\end{aligned} \end{equation} We then rectify $M_{1,a}$ to $z_{1,a}=0$ by $z_1=m_{1,a}(\epsilon_1,r_1,y_1,\mu)+z_{1,a}$ and subsequently straighten out the associated stable fibers of $M_{1,a}$ by a near-identity transformation $(\epsilon_1,r_1,y_1,z_{1,a},\mu)\mapsto \tilde y_1$ such that $(\epsilon_1,r_1,\tilde y_1)'$ becomes independent on $z_{1,a}$ \cite{kelley,jones_1995}. Then \begin{equation}\nonumber \begin{aligned} \dot \epsilon_1&=\epsilon_1^2,\\ \dot r_1 &=-\frac12 r_1 \epsilon_1,\\ \dot{\tilde y}_1&= \epsilon_1\left(\frac{\frac12 \mu+\mathcal O(r_1)}{\tilde y_1+(\mu+1)+\mathcal O(\epsilon_1,r_1)}+\frac12 \tilde y_1\right),\\ \dot z_{1,a}&=(-2+\mathcal O(\epsilon_1,r_1,\tilde y_{1},z_{1,a},\mu))z_{1,a}. \end{aligned} \end{equation} By construction $r_1,\epsilon_1,\tilde y_1$ decouples. For $\epsilon_1=0$ it easy to see that $\tilde y_1=y_1$. This shows the form of \eqref{y1az1a} with the $z_1$-dependency of $y_{1,a}$ only entering the $\mathcal O(\epsilon_1)$-term. Indeed, in the remainder of the proof we only transform $\tilde y_1$ in order to normalize the reduced problem on $z_{1,a}=0$: \begin{equation}\eqlab{redMa1} \begin{aligned} \epsilon_1'& = \epsilon_1,\\ r_1' &=-\frac12 r_1,\\ y_1'&=\frac{\frac12 \mu+\mathcal O(r_1)}{y_1+(\mu+1)+\mathcal O(\epsilon_1,r_1)}+\frac12 y_1, \end{aligned} \end{equation} after dividing the right hand side by $\epsilon_1$. We have here dropped the tilde on $y_1$ and the system then coincides (by construction) with \eqref{redMa10}. Put $y_1=-\mu + \tilde y_1$. Then \begin{align*} \epsilon_1'&=\epsilon_1,\\ r_1' &=-\frac12 r_1,\\ y_1' &=\lambda(\mu) y_1+L_1(\epsilon_1,r_1,y_1,\mu), \end{align*} after dropping the tildes, with $L_1(\cdot,\mu)$ being of second order. First consider the $r_1=0$ subsystem. This system can be smoothly linearized by a transformation of the form $(\epsilon_1,y_1,\mu) \mapsto \tilde y_1=h_{1,a}(\epsilon_1,y_1,\mu)$ with $h_{1,a}(0,0,\mu)= 0$, $\frac{\partial h_{1,a}}{\partial y_1}(0,0,\mu)=1$, see \cite{dumortier2006a,ilyashenko1991a}. Applying this transformation to the full system, we may take $L_1$ to be of the form $L_1=r_1\widetilde L_1$. Then within $\epsilon_1=0$, we have a saddle for $\mu\sim 0$ and we can straighten out the unstable manifold by an $r_1$-dependent transformation of $y_1$: $(r_1,y_1)\mapsto y_{1,a}$. This gives $\widetilde L=y_{1,a}\widetilde L_0 +\epsilon_1 \widetilde L_1$: \begin{equation}\eqlab{eqnsredMa1new} \begin{aligned} \epsilon_1' &=\epsilon_1,\\ r_1' &=-\frac12 r_1,\\ y_{1,a}' &=(\lambda(\mu) +r_1 L_0(r_1,y_{1,a},\mu) )y_{1,a}+r_1 \epsilon_1 L_1(\epsilon_1,r_1,y_{1,a},\mu), \end{aligned} \end{equation} upon dropping the tildes. This completes the proof. \end{proof} Since the flow is regular on sets that are uniformly bounded away from $M_{1,a}$ and $M_{1,r}$, we can actually extend the transformation leading to \eqref{eqnsredMa1new2} to $z_1 \in [-1-\sigma,0]$, say, by the flow-box theorem. We therefore describe $P_{1,a}$ using the coordinates of \eqref{eqnsredMa1new2} from $\{z_{1,a}=-\widetilde m_{1,a}(\epsilon_1,r_1,y_{1,a},\mu)\}$ to $\{\epsilon_1=\epsilon_{11}\}$. Here $\widetilde m_{1,a}(\epsilon_1,r_1,y_{1,a},\mu) :=m_{1,a}(r_1,y_1(\epsilon_1,r_1,y_{1,a},\mu),\epsilon_1,\mu)$ with the smooth function $y_1(\epsilon_1,r_1,y_{1,a},\mu)$ being obtained (by the implicit function theorem) from \eqref{y1az1a} with $z_1=0$: \begin{align*} y_{1,a}=h_{1,a}^0(\epsilon_1,r_1,y_1,\mu)+\epsilon_1 h_{1,a}^1(\epsilon_1,r_1,y_1,0,\mu). \end{align*} For simplicity, we denote the transformed mapping $$P_{1,a}(\epsilon_1,r_1,y_{1,a},\mu)=\begin{pmatrix} \epsilon_{11}\\ \sqrt{\epsilon_1 \epsilon_{11}^{-1}}r_1\\ *\\ * \end{pmatrix} ,$$ by the same symbol. Notice that the $r_1-$ and $\epsilon_1$-components of the mapping are just consequences of the conservation of $\epsilon=r_1^2 \epsilon_1$. We describe $P_{1,a}$ in \propref{P1a}, but in preparation we first state a result on the \textit{Shilnikov problem} associated with the $(\epsilon_1,r_1,y_{1,a})$-subsystem, written in the desingularized form \eqref{eqnsredMa1new}. For this, it will be convinient to work with functions $(u,v)\in U\times V \mapsto f(u,v)\in \mathbb R^l$, $U\subset \mathbb R^n,V\subset \mathbb R^m$, $l,n,m\in \mathbb N$, that are $C^{k}$-smooth, $k\in \mathbb N$, with respect to the $v\in V$, depending continously on $u\in U$. We write the function space of such functions by $C(u^0v^k)$ and let \begin{align*} \Vert f\Vert_{u^0v^k}:=\sup_{(u,v)\in U\times V,\,0\le \vert i\vert\le k}\vert D^i f(u,v)\vert, \end{align*} with $D^i$ being the partial derivative with respect to $v$ of order $i=(i_1,\ldots,i_m)\in \mathbb N_0^m$, denote the associated Banach norm. Notice that by $C(u^0v^k)$ and $\Vert \cdot\Vert_{u^0v^k}$ we suppress $l$, $n$, $m$, $U$ and $V$; it should be clear from the context what these are. On the other hand, we use $\Vert f \Vert_{C^k}:=\sup_{u\in U,\,0\le \vert i\vert\le k} \vert D^i f(u)\vert $ to denote the usual $C^k$-norm of a $C^k$-smooth function $f:U\rightarrow \mathbb R^l$. \begin{proposition}\proplab{this} Fix $k\in \mathbb N$ and consider \eqref{eqnsredMa1new}. Then there exist constants $\tau_0>0$, $K>0$ and $C>0$ such that the following holds for all $\delta>0$ small enough: There exists a $C^k$-smooth function $\underline y_{1,a}(t,\tau,\epsilon_{11},r_{10},y_{11},\mu)$, defined on the set given by: $(t,\tau): 0\le t\le \tau, \tau>\tau_0$ and $0\le \epsilon_{11},r_{10},\vert y_{11}\vert,\vert \mu\vert \le K \delta$, such that $y_{1,a}(t)=\underline y_{1,a}(t,\tau,\epsilon_{11},r_{10},y_{11},\mu)$ solves the \textnormal{Shilnikov problem} defined by \begin{align*} \epsilon_1(\tau)=\epsilon_{11},\quad r_1(0)=r_{10},,\quad y_{1,a}(\tau)=y_{11}. \end{align*} Moreover, $\underline y_{1,a}$ has the following expansion: \begin{align*} \underline y_{1,a}(t,\tau,\epsilon_{11},r_{10},y_{11},\mu) = e^{\lambda(\mu)(t-\tau)} (y_{11}+\phi(t,\tau,\epsilon_{11},r_{10},y_{11},\mu)), \end{align*} where $\Vert \phi\Vert_{C^k}\le C$; specifically \begin{align} \Vert \phi\Vert_{(t,\tau,\epsilon_{11},r_{10})^0 (y_{11},\mu)^k} \le \delta,\quad \phi(t,\tau,\epsilon_{11},0,0,\mu)\equiv 0.\eqlab{bphi} \end{align} \end{proposition} \begin{proof} The proof follows \cite{deng1988a} and we therefore delay the details to \appref{proofprop32}. \end{proof} \begin{proposition}\proplab{P1a} Fix $k\in \mathbb N$ and any $\delta>0$. Then there exist constants $c>0$, $\chi>0$ and $0<\epsilon_{10}\ll \epsilon_{11}$, $r_{10}>0$ such that the following holds: The transition map $P_{1,a}(\cdot,\mu)$ of \eqref{eqnsredMa1new2} is well-defined on a region $V_{1,a}(\chi)$ defined by \begin{align} \vert y_{1,a}\vert \le \chi (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu)},\epsilon_1 \in [0,\epsilon_{10}],r_1\in [0,r_{10}],\eqlab{V1a} \end{align} and on this region, the mapping takes the following form \begin{align*} P_{1,a}(\epsilon_1,r_1,y_{1,a},\mu) =\begin{pmatrix} \epsilon_{11}\\ \sqrt{\epsilon_1 \epsilon_{11}^{-1}}r_1\\ (\epsilon_{1}^{-1} \epsilon_{11})^{\lambda(\mu)} y_{1,a}+\psi_{1,a} (\epsilon_1,r_{1},(\epsilon_{1}^{-1} \epsilon_{11})^{\lambda(\mu)} y_{1,a},\mu)\\ \mathcal O(e^{-c/\epsilon_1}) \end{pmatrix}. \end{align*} Here $\psi_{1,a}$ is a $C({\epsilon_1}^0(r_1,u,\mu)^k)$-function on the set $r_1\in [0,r_{10}],\epsilon_1\in [0,\epsilon_{10}],0\le \vert u\vert,\vert \mu\vert \le \chi$ with \begin{align} \Vert \psi_{1,a}\Vert_{(\epsilon_1,r_1)^0(u,\mu)^k}\le \delta,\quad \psi_{1,a}(\epsilon_1,0,0,\mu)\equiv 0.\eqlab{psi1aprop} \end{align} The $\mathcal O(e^{-c/\epsilon_1})$ remainder term in the $z_{1,a}$-component is also a $C({\epsilon_1}^0(r_1,y_{1,a},\mu)^k)$-function with the order being unchanged upon differentiation up to order $k$ with respect to $(r_1,y_{1,a},\mu)$ for $(\epsilon_1,r_1,y_{1,a})\in V_{1,a}(\chi)$. \end{proposition} \begin{proof} Since $(\epsilon_1,r_1,y_{1,a})$-decouples, we use \propref{this} with $\tau$ defined by $\epsilon_{11} = e^{\tau}\epsilon_{1}$. Let \begin{align}\eqlab{y0eqn} y_{1,a} =\underline y_{1,a}(0,\tau,\epsilon_{11},r_{1},y_{11},\mu)=e^{-\lambda(\mu) \tau}[y_{11}+\phi(0,\tau,\epsilon_{11},r_{1},y_{11},\mu)]. \end{align} The $y_{1,a}$-component of the transition map is then defined implicitly by this equation as $y_{1,a}\mapsto y_{11}$. Consider therefore $F(u_0,\tau,y_{11},\mu)=0$ with \begin{align*} F(u_0,\tau,r_1,y_{11},\mu) := u_0-(y_{11}+\phi(0,\tau,\epsilon_{11},r_1,y_{11},\mu)), \end{align*} for $\epsilon_{11}>0$ fixed, defined by setting the square bracket in \eqref{y0eqn} equal to $u_0$. Then by \propref{this} and the implicit function theorem, we obtain a locally unique solution \begin{align}\eqlab{y11eqn} y_{11}=u_0+\widetilde \psi_{1,a} (\tau,r_{1},u_0,\mu), \end{align} of $F=0$ with $\widetilde \psi_{1,a}\in C^k$ for all $\tau>0$ large enough and $0\le \vert u\vert,r_1,\epsilon_{11},\vert \mu\vert \le \chi$. With $\epsilon_{11}>0$ fixed in the folllowing we suppress its dependency from $\widetilde \psi_{1,a}$. Inserting $u_0=e^{\lambda(\mu) \tau}y_{1,a}=(\epsilon_{11} \epsilon_1^{-1})^{\lambda(\mu)}y_{1,a}$ into the right hand side of \eqref{y11eqn} and defining \begin{align*} \psi_{1,a}(\epsilon_1,r_1,u_0,\mu):=\widetilde \psi_{1,a}(\log( \epsilon_1^{-1} \epsilon_{11}),r_{1},u_0,\mu) \end{align*} give the desired expression for the $y_{1,a}$-component of $P_{1,a}$ on the set $V_{1,a}(\chi)$. The function $\psi_{1,a}$ extends to a $C({\epsilon_1}^0(r_1,u_0,\mu)^k)$-function on the closed interval $\epsilon_1\in [0,\epsilon_{10}]$ due to the Arzela-Ascoli theorem, \propref{this} (with $k\rightarrow k+1$) and the uniform boundedness of $\phi$ in $C^{k+1}$. The bound in $C((\epsilon_1,r_1)^0(r_1,u_0,\mu)^k)$ in \eqref{psi1aprop} follows from \eqref{bphi}. Finally, the equality in \eqref{psi1aprop} follows from the invariance of $r_1=y_{1,a}=0$ for \eqref{eqnsredMa1new} (due to the existence of \eqref{gamma2}), see also the second equality in \eqref{bphi}. We subsequently use the transition time $\tau=\log (\epsilon_1^{-1} \epsilon_{11})$ to exponentially esimate the $z_{1,a}$-component of $P_{1,a}$. This is standard and can be done in $C^k$ with respect to $(r_1,y_{1,a},\mu)$ through variational equations. This completes the proof. \end{proof} We can do precisely the same thing for $P_{1,r}$ by working in backward time. We therefore state the following results without proof. \begin{lemma}\lemmalab{transformnearM1r} Consider \eqref{barxneg1eqns} near $M_{1,r}$. Then for $\sigma>0$ small enough, there exists a regular transformation of time and a smooth diffeomorphism $(y_1,z_1)\mapsto (y_{1,r},z_{1,r})$, $y_1,z_1-1\in [-\sigma,\sigma]$, depending smoothly on $(\epsilon_1,r_1,\mu)$ for $0\le r_1,\epsilon_1,\vert \mu\vert \le \sigma$, such that \begin{equation}\eqlab{y1rz1r} \begin{aligned} y_{1,r} &=h_{1,r}^0(\epsilon_1,r_1,y_1,\mu)+\epsilon_1 h_{1,r}^1(\epsilon_1,r_1,y_1,z_1,\mu),\\ z_{1,r}&= z_1-m_{1,r}(\epsilon_1,r_1,y_1,\mu), \end{aligned} \end{equation} where \begin{align}\eqlab{h1r0prop} h_{1,r}^0(\epsilon_1,0,\mu,\mu)\equiv 0, \quad \frac{\partial h_{1,r}^0}{\partial y_1} (\textnormal{\textbf{0}})=1,\quad \frac{\partial h_{1,r}^0}{\partial \mu} (\textnormal{\textbf{0}})=-1,\end{align} and \eqref{barxneg1eqns} becomes \begin{equation}\eqlab{eqnsredMr1new2} \begin{aligned} \dot r_1 &=\frac12 r_1 \epsilon_1,\\ \dot{y}_{1,r} &=\left((-\lambda(\mu) +r_1 L_0(r_1,y_{1,a},\mu) )y_{1,a}+r_1 \epsilon_1 L_1(\epsilon_1,r_1,y_{1,a},\mu)\right)\epsilon_1,\\ \dot{z}_{1,r} &= (2+L_2(r_1,y_{1,a},z_{1,a},\epsilon_1,\mu))z_{1,r},\\ \dot \epsilon_1 &=-\epsilon_1^2, \end{aligned} \end{equation} with $L_0$, $L_1$, $L_2$ each smooth and $L_2(\textnormal{\textbf{0}})=0$. \end{lemma} We can as above extend the normal form transformation to $z_1 \in [0,1+\sigma]$, say, by the flow-box theorem. Let $\widetilde m_{1,r}(\epsilon_1,r_1,y_{1,r},\mu) =m_{1,r}(r_1,y_1(\epsilon_1,r_1,y_{1,r},\mu),\epsilon_1,\mu)$ with the smooth function $y_1(\epsilon_1,r_1,y_{1,r},\mu)$ being obtained (by the implicit function theorem) from \eqref{y1rz1r} with $z_1=0$: \begin{align*} y_{1,r}=h_{1,r}^0(\epsilon_1,r_1,y_1,\mu)+\epsilon_1 h_{1,r}^1(\epsilon_1,r_1,y_1,0,\mu). \end{align*} \begin{proposition}\proplab{P1r} Fix $k\in \mathbb N$ and any $\delta>0$. Then there exist constants $c>0$, $\chi>0$ and $0<\epsilon_{10}\ll \epsilon_{11}$, $r_{10}>0$ such that the following holds: The transition map $P_{1,r}(\cdot,\mu)$ of \eqref{eqnsredMr1new2} from $\{z_{1,r}=-\widetilde m_{1,r}(\epsilon_1,r_1,y_{1,r},\mu)\}$ to $\{\epsilon_1=\epsilon_{11}\}$ is well-defined on a region $V_{1,r}(\chi)$ defined by \begin{align} \vert y_{1,r}\vert \le \chi (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu)},\epsilon_1 \in [0,\epsilon_{10}],r_1\in [0,r_{10}],\eqlab{V1r} \end{align} and on this region, the mapping takes the following form \begin{align*} P_{1,r}(\epsilon_1,r_1,y_{1,r},\mu) =\begin{pmatrix} \sqrt{\epsilon_1 \epsilon_{11}^{-1}}r_1\\ (\epsilon_{11} \epsilon_{1}^{-1})^{\lambda(\mu)} y_{1,r}+\psi_{1,r}(\epsilon_1,r_{1},(\epsilon_{1}^{-1} \epsilon_{11})^{\lambda(\mu)} y_{1,a},\mu)\\ \mathcal O(e^{-c/\epsilon_1})\\ \epsilon_{11} \end{pmatrix}. \end{align*} Here $\psi_{1,r}$ is a $C({\epsilon_1}^0(r_1,u,\mu)^k)$-function on the set $r_1\in [0,r_{10}],\epsilon_1\in [0,\epsilon_{10}],0\le \vert u\vert,\vert \mu\vert \le \chi$ with \begin{align} \Vert \psi_{1,r}\Vert_{(\epsilon_1,r_1)^0(u,\mu)^k}\le \delta,\quad \psi_{1,r}(\epsilon_1,0,0,\mu)\equiv 0.\eqlab{psi1rprop} \end{align} The $\mathcal O(e^{-c/\epsilon_1})$ remainder term in the $z_{1,r}$-component is also a $C((\epsilon_1,r_1)^0(y_{1,r},\mu)^k)$-function with the order being unchanged upon differentiation up to order $k$ with respect to $(y_{1,r},\mu)$ in $V_{1,r}(\chi)$. \end{proposition} \begin{remark} Notice that if we define $\hat y_{1,a}$ and $\hat y_{1,r}$ by \begin{align}\eqlab{haty1a} y_{1,i}= (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu)} \hat y_{1,i}, \quad i=a,r \end{align} for $\epsilon_1\in (0,\epsilon_{10}]$ then $V_{1,i}$ and $V_{1,r}$ are ``blown up'' to $\hat y_{1,i}\in [-\chi,\chi]$, $i=a,r$, and the $y_{1,i}$-components of $P_{1,i}$ \begin{align*} \hat y_{1,i} +\psi_{1,i} ( r_{1},\epsilon_1,\hat y_{1,i},\mu) \end{align*} become regular as functions of $(r_1,\hat y_{1,i},\mu)$ for $i=a,r$. We will use a similar -- but slightly different scaling -- later on. The reason why \eqref{haty1a} will not work directly, is that we also have to transform $\mu$ appropriately in order to relate $y_{1,a}$ and $y_{1,r}$ (through their relationship to $y_1$) in such a way that the application of $P_{1,a}$ and $P_{1,r}$ correspond to flowing the same point forward and backward in time. \end{remark} \subsection{Analysis in the $\bar \epsilon=1$-chart} Having now followed points on $\{z_1=0\}$ forwards and backwards until the section $\{\epsilon_1=\epsilon_{11}\}$, we proceed to extend these further into the scaling chart. Here the equations are given by \eqref{eqn2}. Notice specificaly that $\epsilon_1=\epsilon_{11}>0$ in the $\bar x=-1$-chart corresponds to $x_2=-\epsilon_{11}^{-1}$ in the scaling chart. Let $P_{2,a}$ and $P_{2,r}$ denote the mappings from $\{x_2=-\epsilon_{11}^{-1}\}$ to $\{z_2=0\}$ near $\gamma_{2,0}(0)$, obtained by the first intersection upon application of the forward respectively backward flow of \eqref{eqn2}. Each of these mappings are regular, i.e. smooth diffeomorphisms. Through the flow of \eqref{eqn2} we can also extend the center manifolds $M_{2,a}$ and $M_{2,r}$, being the coordinate transformations of $M_{1,a}$ and $M_{1,r}$, respectively, using \eqref{cc12}. The manifolds $M_{2,a}$ and $M_{2,r}$ within the $(x_2,y_2,z_2,r_2)$-space are foliated by constant values of $r_2\sim 0$. Let $M_{2,a}(r_2)$ and $M_{2,r}(r_2)$ denote the corresponding leafs of this foliation projecting onto the $(x_2,y_2,z_2)$-space. It is standard, see \cite{szmolyan_canards_2001}, that $M_{2,a}(0)$ and $M_{2,r}(0)$ intersect transversally along $\gamma_{2,0}(\mu)$ for all $\mu\sim 0$. This gives rise to a connecting orbit $\gamma_{2,r_2}(\mu)=M_{2,a}(r_2)\cap M_{2,r}(r_2)$ for all $0<r_2 = \sqrt \epsilon\ll 1$, see \cite{szmolyan_canards_2001}. In this way, one may obtain the perturbed (maximal) strong canard $\gamma_{\epsilon}(\mu)$ with $\lim_{\epsilon\rightarrow 0}\gamma_{\epsilon}(\mu)\rightarrow \gamma_0(\mu)$, see \cite{szmolyan_canards_2001}. \subsection{Putting it all together} We summarize the local findings in charts into a global diagram in \figref{blowup0}. Notice that the strong canard orbit $\gamma_0$ (in orange) is a heteroclinic orbit of partially hyperbolic points on the blowup sphere. For $\mu=0$, there is a fast jump (in black) that together with $\gamma_0$ gives rise to heteroclinic cycle. It is the perturbation of this cycle, that produce the intermediate periodic orbits that connect to the small ones (as perturbations of the green orbits, recall \figref{fig:x2z2}) and canard cycles. \begin{figure}[h!] \begin{center} {\includegraphics[width=.53\textwidth]{./blowup0.pdf}} \end{center} \caption{Dynamics for the blown up system for $\mu=0$. } \figlab{blowup0} \end{figure} We now define $\Delta_{a}(\epsilon_1,r_1,y_{1,a},\mu)$ in the following way: \begin{align*} \Delta_a(\epsilon_1,r_1,y_{1,a},\mu) =\left(\Pi \circ P_{2,a} \circ \widehat C_{21} \circ P_{1,a}\right)(\epsilon_1,r_1,y_{1,a},\mu), \end{align*} with $\Pi$ the projection onto the $(x_2,y_2)$-plane and where $\widehat C_{21}$ is the change of coordinates from $(\epsilon_1,r_1,y_{1,a},z_{1,a})$ to $(x_2,y_2,z_2,r_2)$ defined by \eqref{y1az1a} and \eqref{cc12}. $\Delta_a$ therefore describes the transition map from $\{z_1=0\}$ in chart $\bar x=-1$, using the coordinates $(\epsilon_1,r_1,y_{1,a})$, to $\{z_2=0\}$ in chart $\bar \epsilon=1$, using the coordinates $(x_2,y_2)$ for $r_2=\sqrt{\epsilon}=r_1\sqrt{\epsilon_1}$. We define $\Delta_r$ completely analagously: \begin{align*} \Delta_r(\epsilon_1,r_1,y_{1,r},\mu) =\left( \Pi \circ P_{2,r} \circ \widehat C_{21} \circ P_{1,r}\right)(\epsilon_1,r_1,y_{1,r},\mu). \end{align*} If $(\epsilon_1,r_1,y_{1,a})\in V_{1,a}(\chi)$ and $(\epsilon_1,r_1,y_{1,r})\in V_{1,r}(\chi)$, under the coordinate transformations \eqref{y1az1a} and \eqref{y1rz1r} for $z_1=0$, correspond to the same point $(\epsilon_1,r_1,y_1,0)$ on the section $\{z_1=0\}$ in the $\bar x=-1$-chart, then $\Delta_a(\epsilon_1,r_1,y_{1,a},\mu)=\Delta_r(\epsilon_1,r_1,y_{1,r},\mu)$ for $r_1\ge 0$, $\epsilon_1>0$ clearly implies existence of a closed orbit of (\eqref{fast},$\dot \epsilon=0$) system through the point given by $(\epsilon_1,r_1,y_1,0) \in \{z_1=0\}$ in the $\bar x=-1$-chart. Solving \eqref{y1az1a} and \eqref{y1rz1r} with $z_1=0$ for $y_1$, we find that \begin{align*} y_1 = \overline h_{1,a}(\epsilon_1,r_1,y_{1,a},\mu), \end{align*} and \begin{align*} y_1 =\overline h_{1,r}(\epsilon_1,r_1,y_{1,r},\mu), \end{align*} by applying the inverse/implicit function theorem. Here $\overline h_{1,i}$ have the same smoothness properties as $h_{1,i}$, $\overline h_{1,i}( \epsilon_1,\textnormal{\textbf 0}) \equiv 0,$ $i=a,r$, and \begin{align} \frac{\partial \overline h_{1,a}}{\partial y_{1,a}}(\textbf 0) = 1,&\quad \frac{\partial \overline h_{1,a}}{\partial \mu}(\textbf 0) = -1,\\ \frac{\partial \overline h_{1,r}}{\partial y_{1,r}}(\textbf 0) = 1,&\quad \frac{\partial \overline h_{1,r}}{\partial \mu}(\textbf 0) = 1. \end{align} Equating these expressions for $y_1$, we obtain the following equation \begin{align} L(\epsilon_1,r_1,y_{1,a},y_{1,r},\mu):=\overline h_{1,a}(\epsilon_1,r_1,y_{1,a},\mu)- \overline h_{1,r}(\epsilon_1,r_1,y_{1,r},\mu)=0,\eqlab{Ldef} \end{align} relating $y_{1,a}$ and $y_{1,r}$, as desired. We can use the implicit function theorem to solve this equation for $y_{1,r}$ as function of $(\epsilon_1,r_1,y_{1,a},\mu)$. Subsequently, since we are interested in applying both $P_{1,a}$ and $P_{1,r}$, we need to ensure that the image of the region $V_{1,a}(\chi)$ (or an appropriate subset hereof) under the associated mapping $(\epsilon_1,r_1,y_{1,a})\mapsto (\epsilon_1,r_1,y_{1,r})$ defined by \eqref{Ldef}, will be contained within $V_{1,r}(\chi)$. For this, we therefore first adjust $\mu$. \begin{lemma} There exists a continuous function $\mu_0(\epsilon_1,r_1)$ with $\mu_0(\epsilon_1,0)\equiv 0$ such that $L(\epsilon_1,r_1,0,0,\mu_0(\epsilon_1,r_1))=0$ for all $\epsilon_1\in [0,\epsilon_{10}]$, $r_1\in [0,r_{10}]$. \end{lemma} \begin{proof} Simple application of the implicit function theorem using $L(\epsilon_1,\textbf 0)=0$ and $L'_{\mu}(\textbf 0)=-2$. \end{proof} \begin{remark} Clearly, $\mu_0\in C({\epsilon_0}^0 {r_1}^k)$ but we will not use the smoothness in $r_1$ in the proof \thmref{main}. This relates to the fact that the statement is only that $\overline \mu(h,\epsilon)$ is continuous. In the author's opinion, the smoothness properties of this function with respect to $h$ is more involved and certainly more difficult to state. The interpretation of $\mu_0$ within $\epsilon_1=0$ is illustrated in \figref{entry1}(b). \end{remark} Following on from this, we then obtain: \begin{lemma}\lemmalab{hra} The equation \begin{align*} L(\epsilon_1,r_1,y_{1,a},y_{1,r},\mu_0(\epsilon_1,r_1)+\tilde \mu)=0, \end{align*} has a locally unique solution of the form \begin{align} y_{1,r} = h_{ra}(\epsilon_1,r_1,y_{1,a},\tilde \mu),\eqlab{hrasol} \end{align} with $h_{ra}\in C((\epsilon_1,r_1)^0(y_{1,a},\tilde \mu)^k)$ and \begin{align}\eqlab{hraprop} h_{ra}(\epsilon_1,r_1,0,0)\equiv 0, \quad \frac{\partial h_{ra}}{\partial y_{1,a}}(\textnormal{\textbf{0}})=1,\quad \frac{\partial h_{ra}}{\partial \tilde \mu}(\textnormal{\textbf{0}})=-2. \end{align} \end{lemma} \begin{proof} Follows from the implicit function theorem, using $$L(\epsilon_1,r_1,0,0,\mu_0(\epsilon_1,r_2))=0,\quad L'_{y_{1,r}}(\epsilon_1,r_1,0,0,\mu_0(\epsilon_1,r_1))\approx 1.$$ \end{proof} Due to \eqref{hraprop} and the mean value theorem, $h_{ra}(\epsilon_1,r_1,y_{1,a},\tilde \mu ) = y_{1,a}(\cdots)+\tilde \mu(\cdots)$ which leads to the following. \begin{lemma}\lemmalab{hathra} Let $\sigma>0$, $0<\chi_i<\chi$ for $i=a,r$. Define the scaled quatities $\hat y_{1,a},\hat y_{1,a}, \hat \mu$ by \begin{align} y_{1,a} = (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0)}\hat y_{1,a},\quad y_{1,r} = (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0)}\hat y_{1,r},\quad \mu = \mu_0(\epsilon_1,r_1)+ (\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0)}\hat \mu,\eqlab{hatexpressions} \end{align} for $\epsilon_1>0$. Then we have the following. \begin{enumerate} \item \label{item1} For $\epsilon_{10}>0$ small enough, with $y_{1,i}$ given by \eqref{hatexpressions} for \begin{align}\eqlab{hatyarsets} \hat y_{1,a}\in [-\chi_a,\chi_a],\quad \hat y_{1,r}\in [-\chi_r,\chi_r],\quad \hat \mu\in [-\sigma,\sigma], \end{align} we have $(\epsilon_1,r_1,y_{1,i})\in V_{1,i}(\chi)$, $i=a,r$ for all $\epsilon_1\in (0,\epsilon_{10}]$. \item \label{item2} \eqref{hrasol} becomes \begin{align*} \hat y_{1,r} = \hat h_{ra}(\epsilon_1,r_1,\hat y_{1,a},\hat \mu):=(\epsilon_1 \epsilon_{11}^{-1})^{-\lambda(\mu_0)}h_{ra}(\epsilon_1,r_1,(\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0)}\hat y_{1,a},(\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0)}\hat \mu), \end{align*} for $\epsilon_1\in (0,\epsilon_{10}]$. $\hat h_{ra}$ has a $C((\epsilon_1,r_1)^0(\hat y_{1,a},\hat \mu)^{k-1})$-extension to the closed interval $\epsilon_1\in [0,\epsilon_{10}]$ with $\hat h_{ra}(\epsilon_1,r_1,0,0)\equiv 0$. \item \label{item3} $\hat h_{ra}(0,r_1,\cdot,\cdot)$ is linear and for $r_1=0$: \begin{align*} \hat h_{ra}(0,0,\hat y_{1,a},\hat \mu)=\hat y_{1,a}-2\hat \mu. \end{align*} \item \label{item4} For $\sigma>0$, $r_{10}>0$ and $\epsilon_{10}>0$ all small enough, we have that: \begin{align*} \hat h_{ra}(\epsilon_1,r_1,\hat y_{1,a},\hat \mu)\in (-\chi,\chi), \end{align*} for all $\epsilon_1\in [0,\epsilon_{10}],r_1\in [0,r_{10}], \hat y_{1,a}\in [-\chi_a,\chi_a], \hat \mu\in [-\sigma,\sigma]$. \end{enumerate} \end{lemma} \begin{proof} Recall from \eqref{V1a} and \eqref{V1r} that $(\epsilon_1,r_1,y_{1,i})\in V_{1,i}(\chi)$ with $\epsilon_1>0$ if and only if \begin{align} (\epsilon_1^{-1} \epsilon_{11})^{\lambda(\mu)}\vert y_{1,i}\vert \le \chi.\eqlab{V1aragain} \end{align} Now, by \eqref{lambdamu} and \eqref{hatexpressions} for $\vert\hat y_{1,i}\vert \le \chi_i$ the left hand side of \eqref{V1aragain} becomes \begin{align} (\epsilon_1\epsilon_{11}^{-1})^{(\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0(\epsilon_1,r_1))} \hat \mu} \vert \hat y_{1,i} \vert \le e^{(\epsilon_1 \epsilon_{11}^{-1})^{\lambda(\mu_0(\epsilon_1,r_1))} \hat \mu \log (\epsilon_1 \epsilon_{11}^{-1})}\chi_{i} = (1+\mathcal O(\epsilon_1^{\lambda(\mu_0)}\log \epsilon_1))\chi_i,\eqlab{estimatepseps} \end{align} and the item \ref{item1} therefore follows. Items \ref{item2} and \ref{item3} follow from \lemmaref{hra}; notice we lose one degree of smoothness due to the application of the mean value theorem. Finally, item \ref{item4} follows from \ref{item2} and \ref{item3} using $0<\chi_a<\chi$ and $\epsilon_{10}>0$ small enough. \end{proof} The implications of item \ref{item1}. are the following: Let \begin{equation}\eqlab{widehatP1i} \begin{aligned} \widehat P_{1,i}(\epsilon_1,r_1,\hat y_{1,a},\hat \mu) &:= P_{1,i}(\epsilon_1,r_1,y_{1,a},\mu), \end{aligned} \end{equation} for $i=a,r$ and with $y_{1,a}$, $y_{1,r}$ and $\mu$ on the right hand side given by the expressions in \eqref{hatexpressions}. Then we have: \begin{lemma}\lemmalab{widehatPlemma} Fix $\delta>0$ small enough. Then for $r_{10}>0$, $\epsilon_{10}>0$ sufficiently small, $$\widehat P_{1,i}\in C((\epsilon_1,r_1)^0(\hat y_{1,a},\hat \mu)^{k}),$$ $i=a,r$, defined for $\hat y_{1,a}\in [-\chi_a,\chi_r],\hat y_{1,r}\in [-\chi_r,\chi_r]$, $\hat \mu\in [-\sigma,\sigma]$, $r_1\in [0,r_{10}]$, $\epsilon_1\in [0,\epsilon_{10}]$. In particular, the $y_{1,i}$ component of $\widehat P_{1,i}$ takes the following form: \begin{align} \hat y_{1,i}+\widehat \psi_{1,i}(\epsilon_1,r_1,\hat y_{1,i},\hat \mu),\eqlab{expansionP1y} \end{align} where \begin{align}\eqlab{widehatpsi} \Vert \widehat \psi_{1,i}\Vert_{(\epsilon_1,r_1)^0(\hat y_{1,i},\hat \mu)^k}\le \delta,\quad \widehat \psi_{1,i}(\epsilon_1,0,0,\hat \mu)\equiv 0.\end{align} \end{lemma} \begin{proof} We insert \eqref{hatexpressions} into the $y_{1,i}$-component of $P_{1,i}$. Proceeding as in \eqref{estimatepseps} for the expansion in $\epsilon_1$ we obtain \eqref{expansionP1y}. \eqref{widehatpsi} follows from \eqref{psi1aprop} and \eqref{psi1rprop}. \end{proof} Moreover, by item \ref{item4}. we have that the composed mapping \begin{align} (\epsilon_1,r_1,\hat y_{1,a},\hat \mu) \mapsto \widehat P_{1,r}(\epsilon_1,r_1,\hat h_{ra}(\epsilon_1,r_1,\hat y_{1,a},\hat \mu),\hat \mu) \end{align} is a $C((\epsilon_1,r_1)^0(\hat y_{1,a},\hat \mu)^{k-1})$-function defined for $y_{1,a}\in [-\chi_a,\chi_a]$ and $\hat \mu\in [-\sigma,\sigma]$ and all $r_1\in [0,r_{10}]$, $\epsilon_1\in [0,\epsilon_{10}]$ for $\sigma>0$ and $\epsilon_{10}>0$ small enough. We now return to the problem of solving $\Delta_a=\Delta_r$ for closed orbits. Define $$\widehat \Delta_i(\epsilon_1,r_1,\hat y_{1,i},\hat \mu)=\Delta_i(\epsilon_1,r_1,y_{1,i},\mu),\quad i=a,r$$ with $y_{1,i}$ and $\mu$ on the right hand side given by the expressions \eqref{hatexpressions}. By the properties of $\widehat P_{1,i}$ and the regularity of $C_{21}$ and $P_{2,i}$, we have $\widehat \Delta_{i}\in C((\epsilon_1,r_1)^0(\hat y_{1,i},\hat \mu)^{k})$ for $i=a,r$. Finally, let \begin{align}\eqlab{widehatDelta} \widehat \Delta(r_1,\hat y_{1,a},\epsilon_1,\hat \mu ) &:=\widehat \Delta_r\left(\epsilon_1,r_1,\hat h_{ra}(\epsilon_1,r_1,\hat y_{1,a},\hat \mu),\hat \mu\right)-\widehat \Delta_a\left(\epsilon_1,r_1,\hat y_{1,a},\hat \mu\right), \end{align} for $r_1\in [0,r_{10}], \hat y_{1,a}\in [-\chi_a,\chi_a], \epsilon_1\in [0,\epsilon_{10}]$, $\hat \mu\in [-\sigma,\sigma]$. By \lemmaref{hathra} item \ref{item2}, we finally have that $\widehat \Delta \in C((\epsilon_1,r_1)^0(\hat y_{1,a},\hat \mu)^{k-1})$. Clearly, \begin{align}\eqlab{DeltaEq} \widehat \Delta(\epsilon_1,r_1,\hat y_{1,a},\hat \mu)=\textnormal{\textbf{0}}, \end{align}defines closed orbits for $\epsilon_1>0$, $r_1\ge 0$ sufficiently small. We therefore proceed to solve this equation. We will do so by solving for $\hat y_{1,a}$, $\hat \mu$ as functions of $\epsilon_1$ and $r_1$ using the implicit function theorem. Henceforth we therefore take $k\ge 2$. \begin{lemma}\lemmalab{Delta0prop} \begin{align}\eqlab{WidehatDelta0} \widehat \Delta(\textnormal{\textbf 0})=\textnormal{\textbf 0}. \end{align} Moreover, there exists two nonzero tangent vectors $$v_i=(*,*,0)\in T_{\gamma_{2,0}(0)\cap \{z_2=0\}} M_{2,i}(0),$$ $i=a,r$, such that: \begin{align} \frac{\partial \widehat \Delta}{\partial \hat y_{1,a}} (\textnormal{\textbf 0}) &=\Pi v_r-\Pi v_a,\nonumber\\ \frac{\partial \widehat \Delta}{\partial \hat \mu} (\textnormal{\textbf 0}) &=-2\Pi v_r.\eqlab{hatDeltaHatMu} \end{align} \end{lemma} \begin{proof} First, \eqref{WidehatDelta0} follows from the connection $\gamma_{2,0}(0)$. Next, for the proof of the the partial derivatives, we first focus on the partial derivatives of $\widehat \Delta_a$. For this we differentiate $\widehat P_{1,a}$. Following \lemmaref{widehatPlemma}, see \eqref{widehatpsi}, we have $\frac{\partial \widehat P_{1,a}}{\partial \hat \mu}(\textbf 0)=\textbf 0$, and hence $\frac{\partial \widehat \Delta_a}{\partial \hat \mu}(\textbf 0)=\textbf 0$. Next, by \eqref{expansionP1y} \begin{align} \frac{\partial \widehat P_{1,a}}{\partial \hat y_{1,a}}(\textbf 0) = \begin{pmatrix} 0\\ 0\\ 1+\mathcal O(\delta)\\ 0 \end{pmatrix}.\eqlab{partialwidehatPy1} \end{align} Since $\widehat P_{1,a}(\textbf 0)=(\epsilon_{11},0,0,0)\in M_{1,a}$, the vector \eqref{partialwidehatPy1} gives a tangent vector to $M_{2,a}(0)$ at $x_2=-\epsilon_{11}^{-1}$ upon application of the tangent map $T_{(\epsilon_{11},0,0,0)}\widehat C_{21}$ of the change of coordinates $\widehat C_{21}$. Consequently, by applying $T P_{2,a}$ we have \begin{align*} \frac{\partial \widehat \Delta_a}{\partial \hat y_{1,a}}(\textbf 0)=\Pi v_a, \end{align*} for some nonzero tangent vector $v_a\in T_{\gamma_{2,0}(0)\cap \{z_2=0\}} M_{2,a}(0)$. $\widehat \Delta_r$ can be handled similarly: \begin{align*} \frac{\partial \widehat \Delta_r}{\partial \hat y_{1,r}}(\textbf 0)=\Pi v_r,\quad \frac{\partial \widehat \Delta_r}{\partial \hat \mu}(\textbf 0)=\textbf 0 \end{align*} for some nonzero tangent vector $v_r\in T_{\gamma_{2,0}(0)\cap \{z_2=0\}} M_{2,r}(0)$. The partial derivative \eqref{hatDeltaHatMu} with respect to $\hat \mu$ then follows from \lemmaref{hathra}, see item \ref{item3}, and the chain rule. \end{proof} Since the intersection of $M_{a,2}(0)$ and $M_{r,2}(0)$ is transverse along $\gamma_2$, the vectors $v_i$, $i=a,r$, as well as $\Pi v_i$, $i=a,r$, are linearly independent and consequently, \begin{align*} \text{det}\left(\frac{\partial \widehat \Delta}{\partial (\hat y_{1,a},\hat \mu)}(\textnormal{\textbf 0})\right) =\text{det}\begin{pmatrix} \Pi v_r-\Pi v_a & -2\Pi v_r \end{pmatrix} = 2\text{det}\begin{pmatrix} \Pi v_a&\Pi v_r \end{pmatrix}\ne 0. \end{align*} In this way, by the implicit function theorem we can solve $\widehat \Delta(\epsilon_1,r_1,\hat y_{1,a},\hat \mu)=0$ locally for $(\hat y_{1,a},\hat \mu)$ as continous functions of $(\epsilon_1,r_1)$. In this way, we conclude the following. \begin{proposition}\proplab{prop:large} There exist continuous functions $\bar y(\epsilon_1,r_1)$, $\bar \mu(\epsilon_1,r_1)$, $\epsilon_1\in [0,\epsilon_{10}],r_1\in [0,r_{10}]$ for $\epsilon_{10}>0$ and $r_{10}>0$ small enough, with $\bar y(0,0)=\bar \mu(0,0)=0$ such that there is a periodic orbit of (\eqref{fast},$\dot \epsilon=0$) through any point $(\epsilon_1,r_1, \bar y(\epsilon_1,r_1),0)$ in chart $\bar x=-1$ for $\mu=\bar \mu(\epsilon_1,r_1)$ for all $\epsilon_1\in (0,\epsilon_{10}],r_1\in [0,r_{10}]$. \end{proposition} \begin{proof} After having solved $\widehat \Delta=0$ for $\hat y_{1,a}$ and $\hat \mu$ as continous functions $(\epsilon_1,r_1)$ by the implicit function theorem using \lemmaref{Delta0prop}, we complete the result by mapping the solution back to the $(y_1,\mu)$-variables. \end{proof} \section{Completing the proof of \thmref{main}}\seclab{completing} To complete the proof of \thmref{main}, we first collect our results thus far: First, by \propref{smallcycles} we have for any $h_1>0$ a family of periodic orbits $\Gamma^{\textnormal{small}}_{2,h,\epsilon}$, $h\in (0,h_1]$, of \eqref{eqn2} with $\mu=r_2 \overline \mu_2(h,\epsilon)$ for all $r_2=\sqrt{\epsilon}>0$ small enough, parametrized in the $\tilde u$-variables, see \eqref{tildeust}, by their intersection $\tilde u=(-h,0)$, $y_2= \overline y_2(h,r_2)$ with $\tilde u_2=0$. \begin{lemma}\lemmalab{overlap} Consider \eqref{eqn2} for parameter values $\mu \in [-\delta,\delta]$, and let $D$ be any compact domain in the phase $(x_2,z_2)$-plane. Then there is an $h_1>0$ such that any periodic orbit of this system for $r_2>0$ small enough, that is contained within the set defined by $(x_2,z_2)\in D$, $y_2\in [-\delta,\delta]$ and intersects $\tilde u_1<0$, $\tilde u_2=0$ in a single point, belongs to the family $\Gamma^{\textnormal{small}}_{2,h,\epsilon}$, $h\in (0,h_1]$. \end{lemma} \begin{proof} The statement is clearly true for $\mu=r_2\mu_2$ with $\mu_2\in [-\delta,\delta]$ for $\delta>0$ small enough, since the family $\Gamma_{2,h,r_2}$, $h\in (0,h_1]$, is obtained by the implicit function theorem. Next, we realize that there are no periodic orbits for any $\vert \mu_2\vert>\delta$ and $r_2>0$ small enough since in this case, the equilibrium \eqref{y2eq} is hyperbolic; specifically, $C_2$ is normally hyperbolic in a neighborhood of \eqref{y2eq}. Finally, for $r_2c \le \vert \mu\vert \le \delta$ with $c>0$ large enough, $\dot y_2$ has a single sign in the set $(x_2,z_2)\in D$, $y_2\in [-\delta,\delta]$, completing the proof. \end{proof} Clearly, the family $\Gamma_{2,h,\epsilon}^{\textnormal{small}}$, $h\in (0,h_1]$, becomes a family of ``small'' periodic orbits $\Gamma_{h,\epsilon}^{\textnormal{small}}$ of \eqref{fast} upon blowing down, intersecting $z=0$ in $x\sim \sqrt{\epsilon} h$, $y=r_2 \overline y_2(h,\sqrt{\epsilon})$, $h\in (0,h_1]$. (Here we have used $\sim$ in the expression for $x$ since this only holds to leading order; the family is paramematrized by $\tilde u_1=h$ and $x_2$ and $\tilde u_1$ differ by $\mathcal O(\sqrt{\epsilon})$ in the relevant domain.) \begin{remark} As documented in \cite{guckenheimer2008a}, there may exist bifurcations along $\Gamma_{2,h,r_2}$ (periodic doubling- and torus bifurcations). This does not contradict \lemmaref{overlap}, since this lemma only addresses the uniqueness of the branch of periodic orbits intersecting $u_1<0$ once. \end{remark} Next, we turn to the intermediate periodic orbits obtained from \propref{prop:large}. Here we obtain a family of periodic orbits $\Gamma^{\textnormal{inter}}_{h,\epsilon}$, $h\in[\sqrt{\epsilon}/\sqrt{\epsilon_{11}},r_{10}]$, of \eqref{fast} with $\mu=\overline \mu(h^{-2}\epsilon,h)$ for all $\epsilon>0$ small enough, parametrized by their intersecting $(x,z)=(-h,0)$ and $y=h\overline y(h^{-2}\epsilon,h)$. This follows from \propref{prop:large} and \eqref{barxneg1}, specifically $\epsilon=r_1^2\epsilon_1$ and $r_1=h$. Now, $r_1=h=\sqrt{\epsilon} /\sqrt{\epsilon_{11}}$ corresponds to $\epsilon_1=\epsilon_{11}$ or $x_2=-1/\epsilon_{11}$ upon change of coordinates \eqref{cc12}. Consequently, upon taking $h_1> -1/\sqrt{\epsilon_{11}}$ the branch $\Gamma^{\textnormal{inter}}_{2,h,\epsilon}$, i.e. $\Gamma^{\textnormal{inter}}_{h,\epsilon}$ written in the coordinates of the $\bar \epsilon=1$-chart, overlap with $\Gamma_{2,h,\epsilon}^{\textnormal{small}}$ for all $0<\epsilon\ll 1$. By \lemmaref{overlap}, where they overlap, they coincide for all $0<\epsilon\ll 1$. In this way, we obtain the desired family by gluing the two branches together in the domain where they overlap. \section{Discussion}\seclab{discuss} Our main theorem generalizes the birth of canard cycles in $\mathbb R^2$, see e.g. \cite{krupa2001a}, to $\mathbb R^3$ through the folded saddle-node of type II and the strong canard. In contrast to the result in \cite{krupa2001a}, our results are -- however -- only local. Nevertheless, it is also possible to use our method to obtain more global results. We have illustrated a situation in \figref{fig:canardswheads}(a). Here we can also follow the forward and backward flow of the set of points on the section (in orange) that is indicated in the figure using the normal forms and the solutions of the Shilnikov problem, see \lemmaref{transformnearM1a} and \propref{this}. In contrast to $P_{2,i}$,$i=a,r$, above, we would just define transition maps from $r_1=r_{10}$ fixed to $\epsilon_1=\epsilon_{11}$. Otherwise, the proof for the existence of closed orbits, would proceed analagously, using that the center manifolds $M_{a,r}$ intersect transversally along the strong canard $\gamma_0(\mu)$. The same holds for the perturbation of the singular cycles in \figref{fig:canardswheads}(b) (however, the transition from (a) to (b) is more complicated). These extensions could even be done in the smooth setting. The cycles in \figref{fig:canardswheads}(c) mark the end of canard cycles and in $\mathbb R^2$ this is where classical relaxation oscillations appear. However, in contrast to $\mathbb R^2$ the folded singularity $p$ can be of different types, e.g. folded nodes or folded saddles. In future work, we will describe the transition to relaxation oscillations in $\mathbb R^3$. We prove our result in the analytic setting. In particular, we used analyticity to prove existence of a slow manifold, that is not normally hyperbolicy but acts as the center of (normal) oscillations. This in connection with Melnikov theory, enabled an extension of the Hopf cycles. Although results on normally elliptic slow manifolds \cite{balser1994a,braaksma1992a,de2020a,kristiansenwulff} are almost exclusively in the analytic setting, it seems plausable that the result could be extended to the smooth setting, perhaps using (approach (ii), mentioned in the introduction, and) the structure of \eqref{eqn2} more explicitly. Finally, we note that our main result does not rely upon analyticity with respect to $\epsilon$, only in the space variables. This suggests that our results do also apply to systems that have been reduced from higher dimensions. Indeed, slow manifolds in analytic systems have been shown to be analytic in space variables (and only Gevrey-1 smooth in $\epsilon$), see e.g. \cite{de2020a} and references herein. \begin{figure}[h!] \begin{center} \subfigure[]{\includegraphics[width=.49\textwidth]{./canardswoheads.pdf}} \subfigure[]{\includegraphics[width=.49\textwidth]{./canardswheads.pdf}} \subfigure[]{\includegraphics[width=.49\textwidth]{./canardsrelax.pdf}} \end{center} \caption{Canard cycles in $\mathbb R^3$ in the case of a $S$-shaped critical manifold. In (a): canard cycle without head, (b): canard cycles with head and (c): the end of the canard cycles, where transition to classical relaxation oscillations appear.} \figlab{fig:canardswheads} \end{figure}
\section{Introduction} \label{sec: Introduction} 3D cameras have been widely used for gait analysis applications. Most studies claim that these sensors are not accurate enough to be used for gait analysis \cite{Pfister:2014, Mentiplay:2015}. The low accuracy is more noticeable when analyzing gait kinematics than spatio-temporal parameters \cite{Springer:2016}. There are several factors that can influence the accuracy of these sensors: light conditions, sensor position, occlusion of different parts of the body, among others \cite{LemkensW:2013}. However, the main advantage of these sensors is undoubtedly their low cost and mode of operation. As they are markerless systems, they greatly facilitate the work of clinicians when analyzing gait and avoid disturbing patients when placing markers. Machine Learning (ML) is an alternative that has opened the gap to improve the accuracy of 3D cameras. However, current approaches have only applied ML together with 3D cameras for the classification of normal and pathological gait \cite{SHRIVASTAVA:2021}, gait recognition \cite{HZhen:2018}, or for the classification of different pathologies \cite{Li:2019}. On the contrary, they have not pursued an enhancement of the accuracy of these sensors. This is highly relevant because for clinicians dealing with this type of gait pathologies, the classification of the disease is far less relevant than the quantifiable study of its evolution. For this reason, as a contribution to current approaches, this study aims to improve the accuracy of the estimated kinematic gait signals (joint angles) and gait descriptors when using 3D cameras. For this purpose, artificial neural networks (ANNs) are used. ANNs allow the system to learn from a gold standard system through supervised learning. It should be emphasized that in this study, ANNs do not learn a specific gait pattern. On the contrary, this study attempts to correct the measurement errors of the 3D camera based on the examples provided by the gold standard system. To achieve this, a windowing process is applied to the data and each window represents a training sample. This means that the ANN never learns the gait pattern as such. Consequently, it is possible to apply the training result to the analysis of new gait patterns obtaining better estimations. In this study, neural networks were trained based on two criteria: training based on kinematic gait signals, and training based on gait descriptors. The former seeks to improve the waveforms of kinematic signals in sagittal, frontal and transverse planes, by reducing the error and increasing the correlation with respect to a Vicon system. The second is a more direct approach, focusing on training the networks using gait descriptors (like step width, stride length, cadence, max. flexions, max. extensions and rotations at the joints, among other descriptors) directly. In this second approach, 23 most important descriptors for distinguishing pathological from normal gait were selected for the analysis. This set of descriptors was obtained according to the recommendations of clinicians Molina and Carratal\'a \cite{FranciscoMolinaRueda:2020}. The outline of this paper is organized as follows. Section 2 presents a state-of-the-art review of current methods applied to improve the accuracy of 3D cameras. Section 3 is dedicated to the methodology. This section explains the data collection protocol, as well as the process followed for the definition of joint kinematics. Section 4 addresses the two training approaches and the system accuracy achieved in each. Section 5 is focused on the comparison of both training approaches. Section 6 discusses the results of this study by comparing them with the current approaches. Finally, Section 7 concludes this study. \section{State of the art} \label{sec:state_of_the_art} The accuracy of 3D cameras in gait analysis has been a highly discussed topic. For example, Springer and Seligmann \cite{Springer:2016} presented a review of 12 studies that assessed gait analysis with a Microsoft Kinect sensor and a gold standard system. Results of this review indicated good validity for only some spatio-temporal parameters, and not enough validity for gait kinematics variables. For this reason, different approaches have pursued to improve the accuracy of 3D cameras in gait analysis. These approaches have covered, among others, the fusion of 3D cameras with inertial sensors \cite{BersamiraJ:2019}, fusion of multiple 3D cameras\cite{YeungKwok:2013, Geerse:2015,Muller:2017}, or the use of regression equations \cite{AminiAmin:2019} to increase accuracy in the detection of foot-off and foot-contact events. In \cite{Muller:2017}, M\"uller et al. applied a multi-camera configuration that was intended to improve the estimation of spatio-temporal parameters. In the study developed by Matthew et al. \cite{MatthewRobertPeter:2019}, the low accuracy of 3D cameras is attributed to the lack of a gait model for the data retrieved by the sensor, therefore, the authors propose the improvement of the estimations applying a new model based on rigid bodies. Nichols et al. \cite{NicholsJulia:2016} presented another approach based on retro-reflective markers. The researchers found that retro-reflective markers could be used to enhance the motion capture process with 3D cameras. Another approach has been the fusion with inertial sensors. In the study presented by Alizadegan and Behzadipour \cite{ALIZADEGANALIREZA:2017}, the authors proposed a new method to improve accuracy and real-time performance of inertial joint angle estimation for upper limb rehabilitation applications. The position measurements retrieved from a Kinect sensor were used to correct for the sensor-to-segment misalignment of the inertial sensors. Also in the study presented by Destelle et al. \cite{DestelleFrancois:2014} the authors built an improved skeleton using information fused from a Kinect sensor and nine inertial sensors fixed to the subject's forearms, arms, thighs, shanks and chest. This method proved to obtain more accurate measurements of joint angles, despite a complex calibration process prior to each experiment. Previous studies have also attempted sensor latency compensation. Latency compensation allows for the suppression of time delays due to data processing time in 3D cameras. In our previous study \cite{GuffantisensorsDTW:2020}, the latency of the Kinect V2 sensor was measured and corrected using the Dynamic Time Warping (DTW) algorithm. The latency compensation process improves the estimation of joint angles during gait analysis. ANNs have been another method to improve the accuracy of 3D cameras for gait analysis. For example, in the study of Kidzi\'nski et al. \cite{Kidzin:2019}, ANNs were applied to train the detection of foot-contact and toe-off events. Detection of these events is the initial step in the post-processing of most quantitative gait analysis workflows. Also, in the study presented by Bersamira et al. \cite{BersamiraJ:2019}, the researchers applied ANNs for the integration of 3D cameras and inertial sensors. The authors claim that the fusion of these systems makes it possible to obtain gait data comparable to those produced by a Vicon system. However, as mentioned above, according to Destelle et al. \cite{DestelleFrancois:2014} the use of IMUs together with 3D cameras is an alternative that requires a complex calibration prior to each experiment. The low accuracy of the measurements obtained using 3D cameras in gait analysis has also been attributed to the joint estimation method. Commonly used Software Development Kits (SDKs) for skeleton tracking estimate the joint center using ML techniques. Firstly, ML is applied to label the pixels corresponding to each body segment. Afterwards, the intersection of these segments is identified to estimate the joint center location \cite{ZhangZheng:2012}. According to Matthew et al. \cite{MatthewRobertPeter:2019}, by applying this method, the estimated joint centers may be biologically inconsistent. This can lead to errors in the location of the ankle, knee, and hip, which complicates their use for later gait analysis \cite{Pfister:2014}. This problem encouraged the development of a series of deep learning methods for joint estimation \cite{ZhouXingyi:2016,HoangVan:2019}. These methods have been particularly applied in self-occlusive tasks such as sitting. Unfortunately, these methods do not study more sophisticated gait descriptors. In summary, previous studies have attempted to improve the accuracy of the estimation of joint location, but not the gait kinematics itself. The studies that have addressed gait analysis have only dealt with spatio-temporal parameters. However, achieving an accurate enough gait analysis system based on 3D cameras remains a challenge. In contrast to previous studies, this study aims to directly improve the estimation of kinematic gait signals and gait descriptors based on supervised learning through the application of two training approaches. \section{Methodology} \label{sec:methods} \subsection{Ethics statement} The current study was approved by the ethics committee of the Faculty of Physical Activity and Sport Sciences INEF (Polytechnic University of Madrid, Spain). The experiments were performed in the Sports Biomechanics Laboratory of INEF. All subjects provided written informed consent prior to participation. In addition, the participants completed an electronic questionnaire with basic questions related to health problems and musculoskeletal injuries suffered in recent years. \subsection{Data Collection} Thirty-seven healthy participants (16 female and 21 male, Table~\ref{tab:Demographics}) were involved in the study. A total of 207 gait recordings were processed. \begin{table}[htbp] \centering \caption{Demographics} \begin{adjustbox}{max width=0.47\textwidth} \begin{tabular}{ll} \hline \textbf{Participants:} & 37 \\ \textbf{Gait sequences:} & 207 \\ \textbf{Male:} & 21 \\ \textbf{Female:} & 16 \\ \textbf{Mean age:} & 21$\pm$2 years \\ \textbf{Mean weight:} & 67.41$\pm$10.28 kg. \\ \textbf{Mean height:} & 172.99$\pm$8.53 cm. \\ \hline \end{tabular}% \end{adjustbox} \label{tab:Demographics}% \end{table}% The skeleton tracking task was performed with an Orbbec Astra 3D camera using the 0.34.1 release of the Nuitrack SDK. For each gait sequence, 15 joint trajectories were recorded at 30 Hz. During the experiments, the 3D camera was mounted on a mobile robot Rosbot 2.0 from Husarion. The robot is configured to maintain 2.5 meters distance from the person while walking. To achieve this, a controller was implemented in the robot, which was presented in our former study \cite{GUFFANTIroboticsAS:2021}. We encourage the reader to review the aforementioned study for more information about the robot configuration and control. The use of the mobile robot allows to obtain a path long enough to analyze the human gait in spite of the limited range of the 3D sensor. Consequently, this avoids the use of multiple sensors and the calibration processes that these configurations require. Simultaneously, gait information was retrieved with a Vicon system, which was considered as the ground truth in this study. The Vicon system consisted of six M2 MCAM cameras recording at a sampling rate of 120 Hz. Fig.~\ref{fig:experimental_environment} shows the data collection environment. The experiments were conducted in two phases, which are summarised below: \begin{itemize} \item Participant preparation (15 minutes). Thirty-two reflective markers were placed on anatomical landmarks following the Plug-in-Gait model \cite{DavisRoy:1991}. The participant wore running shoes, short tights, a top (in the case of women) and socks below the ankles. In addition, the participant signed an informed consent and completed an electronic questionnaire. The height and weight of each participant was registered. Finally, a warm-up and gait speed normalization process was done whereby the subject was prepared to walk at an average speed of 1m/s. \item Gait data capture (15 minutes). Gait acquisition was performed in a 15x5-meter corridor, in a one-way straight line gait. Six iterations per participant were repeated, giving a total of 222 gait sequences. However, some of these were discarded due to the loss of markers in the Vicon system, whose trajectories could not be reconstructed by filling in gaps. As a result, a total of 207 gait recordings were used for the supervised learning stage. \end{itemize} \begin{figure}[t] \centering \includegraphics[width=0.45\textwidth]{experimental_environment.png} \caption{An overview of the experimental setup. Faculty of Physical Activity and Sport Sciences INEF (Polytechnic University of Madrid, Spain).} \label{fig:experimental_environment} \end{figure} \subsection{Data processing} The robot-mounted 3D camera and the Vicon system started recording data at different times and through different input streams. The timestamp of the first heel strike was used to synchronize the data. This point was retrieved from the anterior-posterior distance between ankle joints (the inter-ankle distance) according to Ceccato et al. \cite{Ceccato:2009}. This event represented the onset of the recording. The end of the recording is determined by the last timestamp registered by the Vicon system. The data from both systems was cut between these limits. Due to the difference in sampling frequencies, the Vicon time series were linearly interpolated using the timestamps of the 3D camera data. This process ensures the same query points. Finally, to minimize any fluctuation, a low pass filter with a cut-off frequency of 4 Hz was applied in the 3D camera data. \subsection{Data analysis} To relate as much as possible the gait model applied by the Nuitrack in the 3D camera with that applied by the Plug-in-Gait in the Vicon system, it was necessary to find an equivalence between the landmarks detected by both systems by defining certain mapping rules. This mapping process is not new. Mapping rules similar to those used in \cite{MaMengxuan:2018,TaoWei:2015} were adopted, and are shown in Table~\ref{tab:Mapping_rules}. \begin{table}[h] \caption{Mapping rules to compare the skeleton model of the 3D camera and the Vicon system.} \begin{adjustbox}{max width=0.47\textwidth} \centering \begin{threeparttable}[t] \centering \begin{tabular}{ll} \hline \textbf{3D Camera} & \textbf{Vicon Markers}\\ \hline Spine base & Midpoint [RPSI,LPSI,RASI,LASI]\\ Spine middle & Midpoint [T10,STRN] \\ Spine shoulder & Midpoint [C7,CLAV] \\ Left shoulder & Marker LSHO \\ Left elbow & Midpoint [LELB marker A,LELB marker B] \\ Left hand & Midpoint [LWR marker A,LWR marker B] \\ Right shoulder & Marker RSHO \\ Right elbow & Midpoint [RELB marker A,RELB marker B] \\ Right hand & Midpoint [RWR marker A,RWR marker B] \\ Left hip & Hip joint centering algorithm[LASI,LPSI] \tnote{1} \\ Left knee & Midpoint [LKNE marker A,LKNE marker B] \\ Left ankle & Midpoint [LANK marker A,LANK marker B] \\ Right hip & Hip joint centering algorithm[RASI,RPSI] \tnote{1} \\ Right knee & Midpoint [RKNE marker A,RKNE marker B] \\ Right ankle & Midpoint [RANK marker A,RANK marker B]\\ \hline \end{tabular}% \begin{tablenotes} \item[1] The hip joint center location was calculated with the hip joint centering algorithm recommended by \cite{DavisRoy:1991}. \end{tablenotes} \end{threeparttable}% \end{adjustbox} \label{tab:Mapping_rules}% \end{table}% Six Vicon markers (LTROC, LTHI, LTIB, RTROC, RTHI and RTIB) were not considered in this mapping process. For the analysis of gait in this study, the joint angles from elbow, knee, hip, shoulder, trunk, and pelvis were analyzed in sagittal plane. Furthermore, the joint angles from hip, and shoulder were analyzed in frontal plane. In addition, an examination of trunk and pelvis took place in the transverse plane. \subsection{Inverse kinematic model to compute joint angles with the 3D camera} \label{inverse_kinematics} \begin{figure*}[htbp] \centering \includegraphics[width=1.0\textwidth]{Coordinate_systems.jpg} \caption{Coordinate systems definition of the body segments used to apply the inverse kinematics. Four local coordinate systems are shown. For each joint the coordinate system is indicated where the X-axis is in blue, the Y-axis in green, and the Z-axis in red. } \label{fig:Coordinate_systems} \end{figure*} The inverse kinematics is the process to retrieve joint angles from 3D coordinates of landmarks. For the inverse kinematics process in the 3D camera, the coordinate systems definition of the body segments were compatible with the International Society of Biomechanics recommendations \cite{WuGe:1995}. First, four local coordinate systems are defined at the joints: $R1=$ $^{w}F_{Knee}$, $R2=$ $^{w}F_{Pelvis}$, $R3=$ $^{w}F_{Trunk}$, $R4=$ $^{w}F_{Elbow}$. These are the projected knee, pelvis, trunk, and elbow coordinate frames, on the world frame, and are defined as shown in Fig.~\ref{fig:Coordinate_systems}. To locate these coordinate systems, vectors V1 and V2 are defined in alignment with the body segments involved. In Fig.~\ref{fig:Coordinate_systems}, the coordinate axes are color-coded: X in blue, Y in green and Z in red. In this figure, at the bottom of each skeleton, the definition of each coordinate axis is indicated according to the order in which they are defined. For example, in the case of the knee coordinate system $R1$, first the Z axis is defined as parallel to the vector V2, which connects knee to hip. Then the Y-axis is defined as the cross product between V1 and V2, where V1 is the vector connecting knee to ankle. Finally, X is defined as the cross product between Y and Z. In some cases, in order to define the coordinate system, it was necessary to calculate a virtual landmark at the midpoint of two adjacent landmarks. These midpoints are colored in orange. For example, the midpoint of left and right hip was used to define the coordinate system $R2$ and the midpoint between left and right shoulder was used to define the coordinate system $R3$. Once the local coordinate systems are defined, the mathematical definitions of the joint angles correspond as closely as possible to the existing clinical terminology as follows: \begin{itemize} \item The hip joint angles were determined as the Cardan rotation angles in the sequence YXZ between the pelvic and knee coordinate systems, i.e. the relative rotation $Eyxz1$, where $Eyxz1=R1\cdot R2^{T}$. \item The pelvic angles were calculated using the Cardan rotation angles in the sequence YXZ between the world and pelvis coordinate frames. This corresponds to the absolute rotation of $R2$ with respect to the world frame. \item The Cardan rotation angles between the world coordinate and the trunk coordinate frames were used to determine the trunk angles. This corresponds to the absolute rotation of $R3$ with respect to the world frame. \item The angle between the two vectors crossing at the knee joint center was considered as the knee joint angle. These vectors were rotated according to $R1$. \item The angle between the two vectors crossing at the elbow joint center was considered as the elbow joint angle. These vectors were rotated according to $R4$. \item The shoulder joint angles were determined as the Cardan rotation angles in the sequence YXZ between the trunk and elbow coordinate systems, i.e. the relative rotation $Eyxz2$, where $Eyxz2=R4\cdot R3^{T}$. \end{itemize} \section{Learning process: Training, Validation and Testing} The large number of experiments performed together with the robot-mounted 3D camera and the Vicon system allowed the application of supervised learning algorithms to improve the resulting estimations. This application was only possible due to the nature of our study in which the signals can be post-processed before obtaining the final results. There are two basic possible approaches for training and both have been studied in order to see which one achieves a better result: training based on kinematic gait signals, and training based on gait descriptors. As mentioned above, 37 participants provided 207 gait recordings. Of these, the recordings corresponding to 20\% of the participants were excluded from the learning process (i.e. were not used for training, validation or testing). This represents the re-testing group of participants. The data from these participants was used to make a further comparison between both training approaches. \subsection{Training approach based on joint kinematics} \begin{figure*}[htbp] \centering \includegraphics[width=0.95\textwidth]{gait_cycles_V_R_RT_paper.png} \caption{Kinematic gait cycles retrieved from the re-testing group of participants. The gait cycles are normalized from 0-100\% and averaged for all the iterations. The figure shows the Vicon system (blue line), the raw estimation of 3D camera (continuous black line), and the learned estimation of 3D camera (dashed black line). } \label{fig:gait_cycles_V_R_RT} \end{figure*} Since the samples from the Vicon system were linearly interpolated using the timestamps of the 3D camera, the samples from both systems are temporally aligned. This allowed us to associate each sample of the camera with a corresponding label on the Vicon signal. Each kinematic signal was represented by a single vector of dimension Nx1, where N was the total number of samples collected from all the experiments. From this single vector, a training matrix was generated by applying a windowing process. The windowing process is used to split the signal into several frames, or windows. The skip factor between frames was set to one sample. Therefore, the final training matrix had a dimension of (N-W+1)xW, where W was the windowing width. On this basis, each training sample was a vector of 1xW that had a corresponding target of dimension 1x1 coming from the Vicon signal. This data set was randomly split into three classes: training, validation and testing. The partitioning followed the rule 60/20/20, as follows: \begin{itemize} \item 60\% of the data was used for training. \item 20\% of the data was used to validate the network and stop training before overfitting occurs. \item 20\% of data was used as a completely independent test of network generalization. \end{itemize} Here a two-layer feedforward backpropagation ANN was created and trained per each kinematic signal. The ANN architecture had 10 neurons per layer, an input layer with a number of entries equal to the windowing width factor and an output layer with a single outcome. Experimentally, a windowing width of 3 samples was determined to be the most effective. The training algorithm updated weight and bias values according to Levenberg-Marquardt optimization. Mean squared error (MSE) was used as the performance function. The maximum number of training iterations was set to 1000 epochs. Table~\ref{tab:Error_rates_kinematic_signals_test} summarizes the accuracy of the system before and after training. When analyzing the correlations shown in Table~\ref{tab:Error_rates_kinematic_signals_test}, it can be seen that before training, trunk tilt, pelvis tilt, hip add./abd., shoulder add./abd. and pelvis rotation had correlations lower than 0.70. We can attribute the low accuracy in shoulder and hip add./abd to the principle of operation of 3D cameras. Since the 3D camera does not provide useful information for the detection of motion in frontal plane, the SDK relies on the intrinsic parameters of the RGB camera for the estimation, so these motions are approximations of medium quality. However, after training, the improvements are remarkable. Just trunk tilt and pelvis tilt remain with correlation lower than 0.70. Despite this fact, considering the low range of motion of these joints (about 2 degrees), the results are satisfactory. \begin{table}[t] \centering \caption{Accuracy of the estimation provided by the 3D camera in sagittal, frontal and transverse planes before and after training. Results are evaluated using the root-mean-square error (RMSE) and the Pearson Correlation (r). Values are displayed in the format mean$\pm$SD.\\} \begin{adjustbox}{max width=0.47\textwidth} \begin{tabular}{lcccc} \cline{2-5} {} & \multicolumn{2}{c}{\textbf{BEFORE TRAINING}} & \multicolumn{2}{c}{\textbf{AFTER TRAINING}}\\ {} & {\textbf{RMSE}} & {\textbf{r}} & {\textbf{RMSE}} & {\textbf{r}}\\ \hline Elbow flex/ext.($\degree$) & 3.54 $\pm$ 0.18 & 0.79 $\pm$ 0.02 & 2.03 $\pm$ 0.08 & 0.80 $\pm$ 0.02\\ Knee flex/ext.($\degree$) & 6.89 $\pm$ 0.29 & 0.94 $\pm$ 0.01 & 5.94 $\pm$ 0.31 & 0.95 $\pm$ 0.01 \\ Hip flex/ext.($\degree$) & 11.38 $\pm$ 0.24 & 0.93 $\pm$ 0.01 & 4.52 $\pm$ 0.19 & 0.95 $\pm$ 0.00 \\ Shoulder flex/ext.($\degree$) & 10.03 $\pm$ 0.28 & 0.69 $\pm$ 0.03 & 6.48 $\pm$ 0.40 & 0.71 $\pm$ 0.03 \\ Trunk tilt($\degree$) & 5.39 $\pm$ 0.11 & 0.45 $\pm$ 0.04 & 1.94 $\pm$ 0.06 & 0.55 $\pm$ 0.03 \\ Pelvis tilt($\degree$) & 5.81 $\pm$ 0.11 & 0.43 $\pm$ 0.04 & 1.91 $\pm$ 0.06 & 0.55 $\pm$ 0.03 \\ Hip add/abd.($\degree$) & 4.79 $\pm$ 0.11 & 0.64 $\pm$ 0.03 & 2.59 $\pm$ 0.10 & 0.70 $\pm$ 0.03 \\ Shoulder add/abd.($\degree$) & 7.65 $\pm$ 0.22 & 0.68 $\pm$ 0.03 & 3.27 $\pm$ 0.25 & 0.73 $\pm$ 0.03 \\ Trunk rotation($\degree$) & 2.77 $\pm$ 0.15 & 0.82 $\pm$ 0.02 & 1.72 $\pm$ 0.11 & 0.87 $\pm$ 0.02 \\ Pelvis rotation($\degree$) & 6.95 $\pm$ 0.18 & 0.62 $\pm$ 0.03 & 2.71 $\pm$ 0.09 & 0.71 $\pm$ 0.02 \\ \hline \end{tabular}% \end{adjustbox} \label{tab:Error_rates_kinematic_signals_test}% \end{table}% On the other hand, joint kinematics could be affected by the difference in the gait models between Vicon and the 3D camera. Since the number of markers detected by the camera is smaller than those detected by the Vicon, the definition of the coordinate systems of the body segments is in fact different in each system. While the Vicon uses four markers to detect the kinematics of the pelvis, the 3D camera uses only two landmarks, plus one belonging to the middle spine. However, the improvement in the correlation of the signals shown in Table~\ref{tab:Error_rates_kinematic_signals_test}, reflects a fit of the models. Thus, the kinematic signals reported by the 3D camera more closely resembles that of the Vicon. Fig.~\ref{fig:gait_cycles_V_R_RT} shows the effects of the training process for kinematic signals of gait in the re-testing group of participants. From the observation of Fig.~\ref{fig:gait_cycles_V_R_RT}, another problem arising from the difference in gait models can be noted. In some of the signals, such as the hip flex./ext., shoulder flex./ext., trunk tilt, pelvis tilt and shoulder add./abd., there is an offset with respect to the signal of the Vicon system. The correction of this offset is another benefit of the learning process that provides a better interpretation of joint kinematics. In the case of the trunk and pelvis tilt, this offset could not be corrected completely. The randomness of this offset in the training data and the small range of motion of these joints (about 2 degrees) made it difficult for the ANN to correct it. However, a considerable improvement in the correlation of these signals was obtained. It is very important to emphasize that the training is not aimed at learning just the normal gait pattern performed by the participants. This would make the system unsuitable for the analysis of new gait patterns or pathological patterns. On the contrary, what the ANN actually learns is how to correct the measurement errors of the original gait model in the 3D camera, based on the examples provided by the Vicon system. To achieve this, the ANN does not receive as input the whole gait cycle in each sample. Instead, it receives each the information into data frames thanks to the windowing process. This is extremely important for generalizing learning. Consequently, it is possible to apply the training result to the analysis of new gait patterns. \subsection{Training approach based on gait descriptors} \begin{figure}[htb] \centering \includegraphics[width=0.47\textwidth]{Knee_and_Hip_Kinematics_in_the_Sagittal_Plane_paper.png} \caption{Identification of kinematic descriptors for left knee and hip in sagittal plane. The reference signal for the detection of heel strike and toe-off events was the inter-ankle distance \cite{Ceccato:2009}.} \label{fig:Knee_and_Hip_Kinematics_in_the_Sagittal_Plane_1} \end{figure} This second approach arises from the concern of whether it would be better to train the ANN directly with gait descriptors rather than with kinematic signals divided into data frames. This is a more direct way to teach the ANN to report gait descriptors more accurately. In order to apply this second approach, following an in-depth study of gait disturbances in neurological patients, and on the recommendation of clinicians Molina and Carratal\'a \cite{FranciscoMolinaRueda:2020}, 27 most important descriptors representing the main differentiators of normal and pathological gait were selected for the analysis. As listed in Table~\ref{tab:Error_rates_gait_events_test}, 11 were spatio-temporal descriptors and 16 were kinematic descriptors. Fig.~\ref{fig:Knee_and_Hip_Kinematics_in_the_Sagittal_Plane_1} illustrates the process followed for the identification of the set of kinematic gait descriptors. According to Ceccato et al. \cite{Ceccato:2009}, the reference signal for the detection of heel strike and toe-off events was the maximal anterior-posterior distance between ankle joints (or inter-ankle distance). In this signal there is a characteristic sinusoidal curve when the X coordinate of the foot markers is graphed versus time. In this curve, the valleys correlate to the time at which the foot comes into contact with the ground (the heel strike) and the peaks fit with the initiation of swing phase (the toe-off) \cite{ZeniJA:2007}. Fig.~\ref{fig:Knee_and_Hip_Kinematics_in_the_Sagittal_Plane_1} illustrates the process followed for the identification of kinematic descriptors of knee and hip in sagittal plane. The rest of the kinematic descriptors correspond to instants of maxima and minima in the corresponding kinematic signals. In addition, the set of spatio-temporal descriptors were calculated according to established definitions \cite{Winter:2009,BeauchetOlivier:2009,Muller:2017}. \begin{comment} \begin{table*}[h] \centering \caption{Accuracy of the system for the detection of gait descriptors. These events represent the main descriptors of gait disorders. The table shows the results before and after training. Results are evaluated using the root-mean-square error (RMSE). Values are displayed in the format mean$\pm$SD.\\} \begin{adjustbox}{width=1.0\textwidth,totalheight=5cm} \begin{tabular}{lcclcc} \hline {\textbf{EVENTS AND PARAMETERS}} & {\textbf{BEFORE TRAINING}} & {\textbf{AFTER TRAINING }} & {\textbf{EVENTS AND PARAMETERS}} & {\textbf{BEFORE TRAINING }} & {\textbf{AFTER TRAINING}}\\ \textbf{OF GAIT} & \textbf{RMSE} & \textbf{RMSE} & \textbf{OF GAIT} & {\textbf{RMSE}} & {\textbf{RMSE}}\\ \hline step width ($m$) & 0.022 $\pm$ 0.01 & 0.021 $\pm$ 0.01 & \multicolumn{1}{l}{pelvis min. tilt ($\degree$)} & 6.14 $\pm$ 0.64 & 1.67 $\pm$ 0.31 \\ left stride length ($m$) & 0.15 $\pm$ 0.02 & 0.07 $\pm$ 0.01 & \multicolumn{1}{l}{hip max. adduction ($\degree$)} & 3.86 $\pm$ 0.75 & 1.37 $\pm$ 0.35 \\ right stride length ($m$) & 0.13 $\pm$ 0.03 & 0.05 $\pm$ 0.01 & \multicolumn{1}{l}{hip min. abduction ($\degree$)} & 5.02 $\pm$ 0.54 & 2.22 $\pm$ 0.67 \\ left stride time ($s$) & 0.03 $\pm$ 0.01 & 0.04 $\pm$ 0.01 & \multicolumn{1}{l}{pelvis max. rotation ($\degree$)} & 6.72 $\pm$ 1.59 & 2.96 $\pm$ 0.97 \\ right stride time ($s$) & 0.06 $\pm$ 0.02 & 0.06 $\pm$ 0.02 & \multicolumn{1}{l}{pelvis min. rotation ($\degree$)} & 9.81 $\pm$ 0.98 & 2.02 $\pm$ 0.47 \\ right step time ($s$) & 0.04 $\pm$ 0.02 & 0.03 $\pm$ 0.01 & \multicolumn{1}{l}{hip max. extension during stance ($\degree$)} & 14.08 $\pm$ 1.53 & 3.60 $\pm$ 0.83 \\ left step time ($s$) & 0.0273 $\pm$ 0.01 & 0.0266 $\pm$ 0.01 & \multicolumn{1}{l}{hip max. flexion during swing ($\degree$)} & 12.22 $\pm$ 1.31 & 2.61 $\pm$ 0.54 \\ right cadence ($steps/min$) & 4.65 $\pm$ 1.08 & 4.60 $\pm$ 1.27 & \multicolumn{1}{l}{hip max. flexion during stance ($\degree$)} & 15.10 $\pm$ 1.47 & 3.70 $\pm$ 1.05 \\ left cadence ($steps/min$) & 3.60 $\pm$ 1.80 & 4.01 $\pm$ 1.51 & \multicolumn{1}{l}{knee initial contact position ($\degree$)} & 7.43 $\pm$ 1.97 & 4.91 $\pm$ 1.33 \\ percentage of foot stance ($\%$) & 3.74 $\pm$ 0.78 & 2.74 $\pm$ 0.73 & \multicolumn{1}{l}{knee position at toe-off ($\degree$)} & 9.85 $\pm$ 1.81 & 3.29 $\pm$ 1.03 \\ percentage of foot swing ($\%$) & 2.73 $\pm$ 1.41 & 1.32 $\pm$ 0.29 & \multicolumn{1}{l}{knee max. flexion in load response ($\degree$)} & 4.10 $\pm$ 1.02 & 4.52 $\pm$ 1.03 \\ trunk max. tilt ($\degree$) & 4.22 $\pm$ 0.87 & 1.68 $\pm$ 0.45 & \multicolumn{1}{l}{knee max. flexion during swing ($\degree$)} & 4.06 $\pm$ 1.21 & 3.39 $\pm$ 1.69 \\ trunk min. tilt ($\degree$) & 5.93 $\pm$ 1.21 & 2.18 $\pm$ 0.35 & \multicolumn{1}{l}{knee max. extension before heel strike ($\degree$)} & 5.76 $\pm$ 1.68 & 4.81 $\pm$ 1.52 \\ pelvis max. tilt ($\degree$) & 4.31 $\pm$ 0.57 & 1.33 $\pm$ 0.52 & & & \\ \hline \end{tabular}% \end{adjustbox} \label{tab:Error_rates_gait_events_test}% \end{table*}% \end{comment} \begin{table}[t] \centering \caption{Accuracy of the system for the detection of 11 spatio-temporal and 16 kinematic descriptors. The table shows the results before and after training. Results are evaluated using the root-mean-square error (RMSE). Values are displayed in the format mean$\pm$SD.\\} \begin{adjustbox}{max width=0.47\textwidth} \begin{tabular}{lcc} \hline \multirow{2}{*}{\textbf{GAIT DESCRIPTORS}} & {\textbf{BEFORE TRAINING}} & {\textbf{AFTER TRAINING }}\\ \textbf{} & \textbf{RMSE} & \textbf{RMSE}\\ \hline \textbf{SPATIO-TEMPORAL DESCRIPTORS}\\ step width ($m$) & 0.022 $\pm$ 0.01 & 0.021 $\pm$ 0.01 \\ left stride length ($m$) & 0.15 $\pm$ 0.02 & 0.07 $\pm$ 0.01 \\ right stride length ($m$) & 0.13 $\pm$ 0.03 & 0.05 $\pm$ 0.01 \\ left stride time ($s$) & 0.03 $\pm$ 0.01 & 0.04 $\pm$ 0.01 \\ right stride time ($s$) & 0.06 $\pm$ 0.02 & 0.06 $\pm$ 0.02 \\ right step time ($s$) & 0.04 $\pm$ 0.02 & 0.03 $\pm$ 0.01 \\ left step time ($s$) & 0.027 $\pm$ 0.01 & 0.026 $\pm$ 0.01 \\ right cadence ($steps/min$) & 4.65 $\pm$ 1.08 & 4.60 $\pm$ 1.27 \\ left cadence ($steps/min$) & 3.60 $\pm$ 1.80 & 4.01 $\pm$ 1.51 \\ percentage of foot stance ($\%$) & 3.74 $\pm$ 0.78 & 2.74 $\pm$ 0.73 \\ percentage of foot swing ($\%$) & 2.73 $\pm$ 1.41 & 1.32 $\pm$ 0.29 \\ \textbf{KINEMATIC DESCRIPTORS}\\ trunk max. tilt ($\degree$) & 4.22 $\pm$ 0.87 & 1.68 $\pm$ 0.45 \\ trunk min. tilt ($\degree$) & 5.93 $\pm$ 1.21 & 2.18 $\pm$ 0.35 \\ pelvis max. tilt ($\degree$) & 4.31 $\pm$ 0.57 & 1.33 $\pm$ 0.52 \\ pelvis min. tilt ($\degree$) & 6.14 $\pm$ 0.64 & 1.67 $\pm$ 0.31 \\ hip max. adduction ($\degree$) & 3.86 $\pm$ 0.75 & 1.37 $\pm$ 0.35 \\ hip min. abduction ($\degree$) & 5.02 $\pm$ 0.54 & 2.22 $\pm$ 0.67 \\ pelvis max. rotation ($\degree$) & 6.72 $\pm$ 1.59 & 2.96 $\pm$ 0.97 \\ pelvis min. rotation ($\degree$) & 9.81 $\pm$ 0.98 & 2.02 $\pm$ 0.47 \\ hip max. extension during stance ($\degree$) & 14.08 $\pm$ 1.53 & 3.60 $\pm$ 0.83 \\ hip max. flexion during swing ($\degree$) & 12.22 $\pm$ 1.31 & 2.61 $\pm$ 0.54 \\ hip max. flexion during stance ($\degree$) & 15.10 $\pm$ 1.47 & 3.70 $\pm$ 1.05 \\ knee initial contact position ($\degree$) & 7.43 $\pm$ 1.97 & 4.91 $\pm$ 1.33 \\ knee position at toe-off ($\degree$) & 9.85 $\pm$ 1.81 & 3.29 $\pm$ 1.03 \\ knee max. flexion in load response ($\degree$) & 4.10 $\pm$ 1.02 & 4.52 $\pm$ 1.03 \\ knee max. flexion during swing ($\degree$) & 4.06 $\pm$ 1.21 & 3.39 $\pm$ 1.69 \\ knee max. extension before heel strike ($\degree$) & 5.76 $\pm$ 1.68 & 4.81 $\pm$ 1.52\\ \hline \end{tabular}% \end{adjustbox} \label{tab:Error_rates_gait_events_test}% \end{table}% Each gait descriptor was represented by a single vector of dimension Px1, where P was the number of gait cycles collected from all the experiments. This means that one sample of the set of descriptors is extracted for each gait cycle. Given that 27 descriptors were analyzed, a training matrix of dimension Px27 was generated. A target matrix of the same dimensions was obtained with the estimations of the Vicon system. Corresponding columns were used to train independent neural networks. Each ANN used a feedforward backpropagation architecture, with two hidden layers, 10 neurons per layer, an input layer with a single entry and an output layer with a single outcome. The training algorithm updated weight and bias values according to Levenberg-Marquardt optimization, and MSE was used as the performance function. The maximum number of training iterations was set to 1000 epochs. Again the data set was split into three classes: training, validation and testing, following the rule 60/20/20 (60\% for training, 20\% for validation and 20\% for testing). Table~\ref{tab:Error_rates_gait_events_test} summarizes the results of this training approach. Results show that the spatio-temporal parameters have no further improvement after the training process. An exception occurs in the case of left/right stride length, where a noticeable correction is observed. Regarding the sixteen kinematic descriptors, the results are better after training. Especially the descriptors for hip, pelvis and trunk show remarkable improvements. \section{Comparison of training approaches} \label{sec:results} \begin{figure}[htb] \centering \includegraphics[width=0.45\textwidth]{Comparison_of_trainings_paper.png} \caption{Process followed to obtain the data for the comparison of the two training approaches. } \label{fig:Comparison_of_trainings} \end{figure} \begin{figure*}[htbp] \centering \includegraphics[width=0.95\textwidth]{kinematic_events_boxplot_RTevents_RTsignals_paper.png} \caption{Results of the comparison of both training approaches. The box plots show the 25th percentile, 75th percentile, median, minimum and maximum values. The mean value is represented with a black dot. The box plot closest to the Vicon is colored in blue, which is also colored in blue. } \label{fig:kinematic_events_boxplot_RTevents_RTsignals} \end{figure*} This section is focused on the comparison of both training approaches. Specifically, it compares the accuracy for the estimation of the sixteen kinematic descriptors of gait. For this purpose, the re-test group of participants was used. Fig.~\ref{fig:Comparison_of_trainings} shows the process followed to obtain the data for comparison. On the basis of this process, four estimations can be compared: the Vicon estimation (V); the 3D camera raw estimation (C); the 3D camera trained with kinematic signals (CTs); and the 3D camera trained with gait descriptors (CTd). Results of this comparison are shown in the Fig.~\ref{fig:kinematic_events_boxplot_RTevents_RTsignals}. A visual representation of the data is shown using box plots. These box plots summarize the 25th percentile, 75th percentile, median, minimum and maximum values of each group of data. The mean value is also shown with a black dot over the box plots. In this figure, the box plot belonging to the system with mean value closest to that of the Vicon is colored in blue, which is also colored in blue. When analyzing the sixteen kinematic descriptors shown in Fig.~\ref{fig:kinematic_events_boxplot_RTevents_RTsignals}, it is clear that the trained systems (CTd or CTs) are always better than the untrained system (C). When comparing the two training approaches, it is observed that in 62\% of the events analyzed (10/16) the approach based on gait descriptors (CTd) is better than the approach based on kinematic signals (CTs), where only in 38\% of the cases (6/16) it proved to be better. However, it can be observed that for events of large amplitude, such as maximum knee and hip flexions and extensions, training based on gait descriptors (CTd) is better. On the contrary, for kinematics of smaller range of motion, the training based on kinematic signals (CTs) is superior. This can be noted in the pelvis maximum and minimum tilt, and in the lower amplitude descriptors of knee kinematics such as the knee initial contact position, knee maximum flexion in load response, and in the knee maximum extension before heel strike. Therefore, it is not possible to define which training approach is completely better. Despite the fact that training based on gait descriptors (CTd) presented higher accuracy in most parameters, training based on kinematic signals (CTs) has an enormous advantage because it allows to know the complete kinematics of the joint, which is very useful when determining deviations between gait patterns for clinical purposes. Consequently, the selection of the training approach will depend on the purpose of the study conducted. \begin{comment} \begin{figure*}[h] \centering \includegraphics[width=0.7\textwidth,height=6.5cm]{spatio_temporal_boxplot_RTevents_RTsignals.png} \caption{An overview of the Mobile robotic platform designed in this study for gait analysis applications. } \label{fig:spatio_temporal_boxplot_RTevents_RTsignals} \end{figure*} \end{comment} \section{Discussion} \label{sec:discussion} It is interesting to compare the results obtained with other techniques for improving the estimations of 3D cameras. An in-depth comparison can be made by analyzing Table~\ref{tab:Comparison_of_accuracy_improvement}. Current approaches have attempted to improve the accuracy of 3D cameras in gait analysis through remodeling the skeleton, sensor fusion, or applying machine learning techniques. In the study presented by M\"uller et al. \cite{Muller:2017}, a validation of enhanced Kinect sensor for gait assessment was presented. The idea of that study was to compute a spatially averaged skeleton tracked by two Kinect sensors placed at both sides of the walking path. The study demonstrated that sensor setups tracking the person only from one-side is less accurate and should be replaced by two-sided setups. To fuse data from both sensors, the authors used the iterative closest point method, which adds imprecision to the model. Similarly, the use of more than one camera suggests a requirement of a dedicated room where the system can be setup and left undisturbed between sessions. Regarding the fusion of 3D cameras with inertial sensors (IMUs), it is interesting to mention the study presented by Destelle et al. \cite{DestelleFrancois:2014}. The authors built a fused Kinect/IMUs skeleton where the Kinect sensor provided the positions of the joints and the IMUs provided the orientation information. The use of IMUs together with 3D cameras required a complex calibration prior to each experiment. The authors reported a final accuracy of 9.00$\degree$ for elbow flex./ext., and 7.89$\degree$ for knee flex./ext. In contrast, in our study considerably higher accuracies were achieved after training (2.03$\degree$ for elbow flex./ext. and 5.94$\degree$ for knee flex./ext.). It should also be noted that the study of Destelle et al. \cite{DestelleFrancois:2014} was performed for a simulated kicking motion while in our study a real gait was analyzed. \begin{table*}[htbp] \centering \caption{Comparison of our study with current approaches.} \begin{adjustbox}{width=1.0\textwidth,totalheight=7cm} \begin{threeparttable}[t] \begin{tabular}{lcccccccccccccc} \cline{2-15} & \multicolumn{2}{c}{\textbf{OUR SYSTEM\tnote{1}}} & \multicolumn{2}{c}{\textbf{RIGID BODIES\tnote{2}}} & \multicolumn{2}{c}{\textbf{DTW\tnote{3}}} & \multicolumn{2}{c}{\textbf{MULTI-SENSOR\tnote{4}}} & \multicolumn{2}{c}{\textbf{MULTI-SENSOR\tnote{5}}} & \multicolumn{2}{c}{\textbf{3D CAMERA+IMU\tnote{6}}} & \multicolumn{2}{c}{\textbf{3D CAMERA+IMU\tnote{7}}} \\ \cline{2-15} & \multicolumn{14}{c}{\textbf{ROOT MEAN SQUARE ERROR}} \\ & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} & \multicolumn{1}{c}{\textbf{BEF.}} & \textbf{AFT.} \\ \cline{1-15} \textbf{Elbow flex./ext.} & 3.54 & 2.03 & - & - & - & - & - & - & - & - & - & 5.90 & 13.33 & 9.00 \\ \textbf{Knee flex./ext.} & 6.89 & 5.94 & 9.40 & 5.20 & 13.15 & 10.47 & - & - & - & - & - & - & 28.23 & 7.89 \\ \textbf{Hip flex./ext} & 11.38 & 4.52 & 11.10 & 5.60 & 6.20 & 2.65 & - & - & - & - & - & - & - & - \\ \textbf{Shoulder flex./ext.} & 10.03 & 6.48 & - & - & 9.53 & 5.80 & - & - & - & - & - & 4.70 & - & - \\ \textbf{Trunk forw./back. tilt} & 5.39 & 1.94 & 3.90 & 3.50 & 2.09 & 1.93 & - & - & - & - & - & - & - & - \\ \textbf{Pelvis ant./post. tilt} & 5.81 & 1.91 & - & 6.10 & 3.22 & 3.14 & - & - & - & - & - & - & - & - \\ \textbf{Hip add./abd.} & 4.79 & 2.59 & - & - & 5.83 & 4.70 & - & - & - & - & - & - & - & - \\ \textbf{Shoulder add./abd.} & 7.65 & 3.27 & - & - & 11.76 & 11.69 & - & - & - & - & - & - & - & - \\ \textbf{Trunk int./ext. rot.} & 2.77 & 1.72 & - & - & 5.26 & 5.09 & - & - & - & - & - & - & - & - \\ \textbf{Pelvis int./ext. rot.} & 6.95 & 2.71 & - & - & 5.88 & 4.81 & - & - & - & - & - & - & - & - \\ \hline \textbf{step width (m)} & 0.022 & 0.021 & - & - & - & - & - & 0.01 & 0.0066 & 0.0039 & - & - & - & - \\ \textbf{AV. stride length (m)} & 0.14 & 0.06 & - & - & - & - & - & 0.00 & 0.0004 & 0.0020 & - & - & - & - \\ \textbf{AV. stride time (s)} & 0.05 & 0.05 & - & - & - & - & - & 0.01 & - & - & - & - & - & - \\ \textbf{AV.step time (s)} & 0.034 & 0.028 & - & - & - & - & - & 0.01 & 0.000 & 0.000 & - & - & - & - \\ \textbf{AV. Cadence (steps/min)} & 4.12 & 4.31 & - & - & - & - & - & 0.90 & - & - & & & & \\ \hline \end{tabular}% \begin{tablenotes} \item[1] Our trained robot-mounted 3D camera. \item[2] Matthew et al.\cite{MatthewRobertPeter:2019}. Single Kinect, tested in a sit to stand task. The original joint data is transformed into a new model built with rigid bodies. \item[3] Guffanti et al.\cite{GuffantisensorsDTW:2020}. Correction of latency usind DTW(Dynamic Time Warping). \item[4] Geerse et al.\cite{Geerse:2015}. Multi-Kinect v2, 10-Meter Walkway for spatio-temporal gait analysis. \item[5] M\"uller et al. \cite{Muller:2017}. Multi-Kinect v2, 7-Meter Walkway for spatio-temporal gait analysis. \item[6] Alizadegan and Behzadipour\cite{ALIZADEGANALIREZA:2017}. Single Kinect to improve accuracy of inertial joint angle estimation in upper limb stroke rehabilitation. \item[7] Destelle et al.\cite{DestelleFrancois:2014}. Single Kinect in front of the person and fusion with IMUs. The RMSE values were averaged for the two experiments performed by the authors. \end{tablenotes} \end{threeparttable}% \end{adjustbox} \label{tab:Comparison_of_accuracy_improvement}% \end{table*}% In the study presented by Alizadegan and Behzadipour \cite{ALIZADEGANALIREZA:2017}, the authors proposed a new method to improve accuracy of IMUs for joint angle estimation in upper limb rehabilitation applications. To correct for the sensor-to-segment misalignment of the inertial sensors, position measurements retrieved from Kinect sensor were used. The experiment was performed on a mechanical upper limb stroke rehabilitation device also known as shoulder wheel. The shaft rotation angle was encoded and this measurement was used as ground truth. The authors reported accuracies of 5.90$\degree$ for elbow flex./ext. and 4.70$\degree$ for shoulder flex./ext. Conversely, in our study, the elbow is reported with higher accuracy (3.54$\degree$ for elbow flex./ext.) but the shoulder is slightly less accurate (6.48$\degree$ for shoulder flex./ext.). It should also be noted that the experimental basis in the study of Alizadegan and Behzadipour \cite{ALIZADEGANALIREZA:2017} is different from ours. While in that study the person remains static, in our study the person is in constant walking motion. Matthew et al. \cite{MatthewRobertPeter:2019} proposed the improvement of the estimations provided by 3D cameras by remodeling the skeleton to a new one based on rigid bodies. The idea of the authors was to rescale the skeleton detected by the 3D camera, using the height of the person. Then, the joint centers are recalculated based on anthropometric approximations. The method was tested only for a sit to stand task (STS), and the results obtained were satisfactory. However, it should be noted that, applying joint center relocation based on anthropometric approximations may not be a robust technique. This is because 3D cameras are affected by many other factors like light conditions, sensor position, occlusion, among others \cite{LemkensW:2013}. This makes it difficult to relocate joint centers using only anthropometric approximations. Thus, this method may not be effective in all experimental conditions. In addition, our method based on ML obtained better results mainly for trunk and pelvis tilt. While in the study of Matthew et al. \cite{MatthewRobertPeter:2019} errors of 3.50$\degree$ for trunk tilt and 6.10$\degree$ for pelvis tilt were obtained, in our study 1.94$\degree$ and 1.91$\degree$ were achieved respectively. \setlength{\parskip}{0cm} \setlength{\parindent}{1em} Finally, in our previous study \cite{GuffantisensorsDTW:2020}, latency correction was proposed using Dynamic Time Warping (DTW). Latency of 3D cameras produces non-symmetrical temporal variations in signals that directly affects the system performance. Results showed a noticeable improvement in system performance. However, these corrections only solve the problem of the temporal disturbances produced by the variable latency of the sensor. Instead, the application of machine learning not only corrects these distortions. This can be observed in the improvement of the correlation in Table~\ref{tab:Error_rates_kinematic_signals_test} and in the correction of waveforms in Fig.~\ref{fig:gait_cycles_V_R_RT}. In addition to the above, machine learning can also correct discrepancies in the gait model applied by the 3D camera with respect to the Plug-in-gait model applied by the Vicon system. This effect can be observed in Fig.~\ref{fig:gait_cycles_V_R_RT}, where the offsets with respect to the Vicon signals were corrected. \section{Conclusion} \label{sec:conclusion} In this study, it was possible to improve the accuracy of 3D cameras for human gait analysis applications. This was achieved by post-processing raw estimations of the 3D camera using ANNs trained with the data provided by a certified Vicon system. The neural networks were trained based on two different criteria: training based on kinematic gait signals, and training based on gait descriptors. In the latter approach, 11 spatio-temporal and 16 kinematic descriptors representing the main differentiators of normal and pathological gait were selected for the analysis. The accuracy of the estimations with the 3D camera was measured before and after training. Results showed lower errors and higher correlations with respect to the gold standard, thus, improving the obtained biomechanical model with the 3D camera. The accuracy obtained when detecting the main descriptors of a pathological gait also showed a substantial improvement, mainly for kinematic descriptors. When comparing both training approaches, it was not possible to define which was the absolute best. Regardless, training based on kinematic signals has an enormous advantage because it allows to know the complete kinematics of the joint. This is very useful when determining deviations between gait patterns along a long-term clinical analysis. Therefore, we believe that the selection of the training approach will depend on the purpose of the study to be conducted. This study reveals the great potential of 3D cameras and encourages the research community to continue exploring their use in gait analysis. \section*{Conflicts of interest} None of the authors have any conflicts of interest associated with this research. \section*{Acknowledgments} The research leading to these results has received funding from PID2020-118299RB-I00 - SISTEMA ROBOTICO NO INVASIVO PARA EL ANALISIS BIOMECANICO DE LA MARCHA HUMANA -, ``Convocatoria 2020 Proyectos de I+D+i - RTI Tipo B'', of the Government of Spain. The authors would like to thank Faculty of Physical Activity and Sports Sciences - INEF, UPM, for the use of Sports Biomechanics Laboratory.